{ "1112/1112.5249.txt": { "abstract": "The 3D structure of the solar wind and its evolution in time is needed for heliospheric modeling and interpretation of energetic neutral atoms observations. We present a model to retrieve the solar wind structure in heliolatitude and time using all available and complementary data sources. We determine the heliolatitude structure of solar wind speed on a yearly time grid over the past 1.5 solar cycles based on remote-sensing observations of interplanetary scintillations, \\textit{in~situ} out-of-ecliptic measurements from \\textit{Ulysses}, and \\textit{in~situ} in-ecliptic measurements from the OMNI-2 database. Since the \\textit{in~situ} information on the solar wind density structure out of ecliptic is not available apart from the \\textit{Ulysses} data, we derive correlation formulae between the solar wind speed and density and use the information on the solar wind speed from interplanetary scintillation observations to retrieve the 3D structure of solar wind density. With the variations of solar wind density and speed in time and heliolatitude available we calculate variations in solar wind flux, dynamic pressure and charge exchange rate in the approximation of stationary H atoms. ", "introduction": "The goal of this paper is to retrieve the solar wind structure at 1~AU as a function of time and heliolatitude based on available \\textit{in~situ} data sources and interplanetary scintillation observations for the time interval since 1990 until present that covers 1.5 solar activity cycles. The existence of the solar wind was predicted on a theoretical basis by \\citet{parker:57a} and discovered experimentally by \\textit{Lunnik}~II and \\textit{Mariner}~2 at the very beginning of the space age \\citep{gringauz_etal:60a, neugebauer_snyder:62a}. Regular measurements of its parameters began in the early 1960s and data from many spacecraft are now available, obtained using various techniques of observations and data processing \\citep[see for review][]{bzowski_etal:12b}. Shortly after the discovery of the solar wind (SW hereafter), a question of whether or not it is spherically symmetric was put forward. Most spacecraft with instruments to measure the SW parameters are at orbits close to the ecliptic plane and the information about the latitudinal structure of the SW is hard to obtain. There are a few sources of data on the out of ecliptic SW parameters, but only one of them from \\textit{in~situ} measurements, namely from \\textit{Ulysses}. While direct observations of the SW in the ecliptic plane have been collected for many years, information on its latitudinal structure had been available only from indirect observations of the cometary ion tails \\citep{brandt_etal:75a}, until radio-astronomy observations of interplanetary scintillation \\citep{hewish_etal:64a, coles_maagoe:72a} and spaceborne measurements of the Lyman-$\\alpha$ helioglow \\citep{lallement_etal:85a, bertaux_etal:95} became available. To our knowledge these two techniques remain the only source of global, time-resolved information on the the solar wind structure. The launch of the \\textit{Ulysses} spacecraft \\citep{wenzel_etal:89a} improved our understanding of the 3D behavior of the solar wind by offering direct \\textit{in~situ} observations and a very high resolution in latitude, but a poor resolution in time\\footnote{The same latitudes were visited only a few times during the $\\sim 20$-year mission.}. The solar wind structure varies in latitude with the solar activity cycle. Knowledge of its evolution is needed to construct credible models of the heliosphere and its boundary regions. With the history of the SW evolution based on a homogeneous series of data and retrieved using homogeneous analysis method, one obtains a tool to interpret both present heliospheric observations, such as the ongoing measurements of Energetic Neutral Atoms (ENAs) by the NASA \\textit{Interstellar Boundary Explorer}~\\citep[IBEX,][]{mccomas_etal:09a} and \\textit{in~situ} measurements of the heliospheric environment by the Voyagers, and to compare them with the results from past and current long-lived experiments such as \\textit{Solar Wind ANisotropy}~(SWAN) onboard \\textit{SOlar and Heliospheric Observatory}~(SOHO) etc. We assume that the solar wind expansion is purely radial, its speed does not change with solar distance, and density drops down quadratically with distance from the Sun. These assumptions are valid for close distances from Sun $\\left( r < 10~AU \\right)$. Their validity at farther distances will be considered in the Discussion section. In the following text ``CR-averaged'' data mean Carrington rotation (CR hereafter) averaged values, where Carrington rotation period is the synodic period of solar rotation equal to 27.2753 days \\citep{franz_harper:02a}. Throughout the paper, ``adjusted'' means scaled from the value $x_0$ measured at a heliocentric distance $r_\\mathrm{0}$ to the distance $r_\\mathrm{E} = 1$~AU, where the value specific for $r_\\mathrm{E}$ is calculated as $x_\\mathrm{E} = x_0 \\left( r_0 / r_\\mathrm{E} \\right)^2$. ", "conclusions": "To check the credibility of our results we compare them with the \\textit{Ulysses} data from all scans. The \\textit{Ulysses} data are prepared by splitting the hourly data into full Carrington rotations and calculating average values. Next we linearly interpolate our model values to the times and heliolatitudes corresponding to the Carrington rotation-averaged \\textit{Ulysses} data. A comparison is presented in Figure~\\ref{figCompareModelWithUlysses}. The agreement is quite good in the ecliptic parts of the \\textit{Ulysses} orbit and satisfactory for higher latitudes. The model retrieves the fast solar wind speed, but some discrepancies exist for the slow and variable solar plasma flows. We have a better agreement in solar wind speed than in density, which is understandable, since the density values were derived from the already approximate speed values using approximate statistically-derived solar wind density-speed relations. The worst agreement between \\textit{Ulysses} density measurements and our model is during the third slow scan (during descending phase of solar activity), when \\textit{Ulysses} \\textit{in~situ} measurements for this solar cycle phase give highly variable values, nearly $50\\%$ greater than our model predictions. The source of the disagreement might be connected with the density-speed correlation formula we adopted for this time interval: for solar maximum we use an average formula from the two fast scans during solar minimum assuming that the correlation does not change with the solar cycle. The overall agreement, however, is much better. The residuals of speed are typically $10\\%$, not exceeding $30\\%$, and typical residuals of density are $20-30\\%$, not exceeding $60\\%$. The sign of the residuals varies, which suggests there is no systematic global bias in our method. Given all the uncertainties in the absolute calibration of both \\textit{in~situ} measurements and IPS observations and the relative simplicity of our approach, we believe such a level of agreement between the model and the measurements is quite good and hard to improve on without an additional data source. Such a source of additional information might result from an inversion of photometric maps of the Lyman-$\\alpha$ helioglow obtained from SWAN/SOHO observations \\citep{lallement_etal:10a}, aimed at calculation of the total ionization rate of neutral interstellar hydrogen in the inner heliosphere as a function of heliolatitude and Carrington rotation. With this information, one might be able to follow the idea presented by \\citet{bzowski_etal:12b} and independently calculate the profiles of solar wind density. In this analysis we assumed that the solar wind parameters obtained from OMNI-2, \\textit{Ulysses} and IPS are directly comparable, i.e. that there is no systematic change in solar wind speed with the solar distance between the region from the solar wind acceleration region (a dozen solar radii) to 1~AU, from which the IPS solar wind speeds are retrieved, and \\textit{Ulysses}, which measured between $\\sim 1.4$ and $\\sim 5.5$~AU. Also, we assume there is no distance-related change in the density other than the simple $1/r^2$ scaling. The assumption of purely radial expansion of solar wind implies that the latitude structure we inferred in this paper does not change until the termination shock. This may not be exactly true, as pointed out by \\citep{fahr_scherer:04b}, who argue that the pickup ions (PUIs) pressure induces nonradial flows at large heliocentric distances. Such flows would cause changes both in the radial component of solar wind speed and in the local density. A more thorough study of this effect requires, in our opinion, using a multifluid, 3D and time dependent MHD modeling with turbulence and boundary/initial conditions taken from observations. Such a model, to our knowledge, is still in development \\citep{usmanov_etal:11a}. Our results seem to be well suited as the boundary conditions for such modeling. The assumption of a $1/r^2$ drop in density and constancy of speed is increasingly invalid with an increase of solar distance because of the interaction with neutral interstellar gas, which results in the creation of pickup ions and slowdown and heating of the distant solar wind. These phenomena were extensively discussed by \\citet{fahr_rucinski:99,fahr_rucinski:01a,fahr_rucinski:02a,fahr:07a,lee_etal:09a,richardson_etal:95a, richardson_etal:08a,richardson_etal:08b}. However, for the global modeling of the heliosphere and calculation of survival probabilities it is the total flux of solar wind, being a sum of the core solar wind and pickup ions, that counts most. In this respect, the total solar wind flux quite exactly follows the $1/r^2$ scaling due to the continuity conditions, as discussed by \\citet{bzowski_etal:12b}. This is because a vast majority of PUIs are created due to charge exchange and thus are not new members of the solar wind population. \\begin{figure*}[!h] \\centering \\begin{tabular}{cc} \\includegraphics[scale=0.35]{figCompareModelWithUlysses_speed1s} & \t\\includegraphics[scale=0.35]{figCompareModelWithUlysses_dens1s}\\\\ \\includegraphics[scale=0.35]{figCompareModelWithUlysses_speed1f} & \t\\includegraphics[scale=0.35]{figCompareModelWithUlysses_dens1f}\\\\ \\includegraphics[scale=0.35]{figCompareModelWithUlysses_speed2s} & \t\\includegraphics[scale=0.35]{figCompareModelWithUlysses_dens2s}\\\\ \\includegraphics[scale=0.35]{figCompareModelWithUlysses_speed2f} & \t\\includegraphics[scale=0.35]{figCompareModelWithUlysses_dens2f}\\\\ \\includegraphics[scale=0.35]{figCompareModelWithUlysses_speed3s} & \t\\includegraphics[scale=0.35]{figCompareModelWithUlysses_dens3s}\\\\ \\includegraphics[scale=0.35]{figCompareModelWithUlysses_speed3f} & \t\\includegraphics[scale=0.35]{figCompareModelWithUlysses_dens3f}\\\\ \\includegraphics[scale=0.35]{figCompareModelWithUlysses_speed4s} & \t\\includegraphics[scale=0.35]{figCompareModelWithUlysses_dens4s}\\\\ \\end{tabular} \\caption{Comparison of \\textit{Ulysses} \\textit{in~situ} measurements for slow and fast scans with model results for solar wind speed and adjusted density. Presented are Carrington rotation averaged data. Blue - model, red - \\textit{Ulysses}.} \\label{figCompareModelWithUlysses} \\end{figure*} \\clearpage" }, "1112/1112.4841_arXiv.txt": { "abstract": "Thermal instability (TI) can strongly affect the structure and dynamics of the interstellar medium (ISM) in the Milky Way and other disk galaxies. Thermal conduction plays an important role in the TI by stabilizing small scales and limiting the size of the smallest condensates. In the magnetized ISM, however, heat is conducted anisotropically (primarily along magnetic field lines). We investigate the effects of anisotropic thermal conduction on the nonlinear regime of the TI by performing two-dimensional magnetohydrodynamic simulations. We present models with magnetic fields of different initial geometries and strengths, and compare them to hydrodynamic models with isotropic conduction. We find anisotropic conduction does not significantly alter the overall density and temperature statistics in the saturated state of the TI. However, it can strongly affect the shapes and sizes of cold clouds formed by the TI. For example, for uniform initial fields long filaments of cold gas are produced that are reminiscent of some observed HI clouds. For initially tangled fields, such filaments are not produced. We also show that anisotropic conduction suppresses turbulence generated by evaporative flows from the surfaces of cold blobs, which may have implications for mechanisms for driving turbulence in the ISM. ", "introduction": "\\label{sec:Intro} The thermal instability (TI) plays an important role in controlling several different aspects of the interstellar medium (ISM) and star formation. For example, it has been invoked to explain the observed multiphase structure of the ISM \\citep{par53,spi58,fie65,fie69}. The linear stage of the TI in astrophysical gases was first studied in detail by \\citet{fie65}. He identified three unstable modes: an isobaric and two isentropic modes. In the nonlinear regime, the isobaric mode produces condensations that are fundamental to the classical two-phase model of the ISM \\citep{fie69}, as well as the extension to a three-phase model by \\citet{cox74,mck77}. The TI also regulates the mass flow between the different components of the ISM, and therefore affects the star formation rate \\citep{chi87}. Thus, it is important to investigate the role of TI in determining the distribution of density, temperature, and other physical variables in the multiphase ISM. For this reason, a variety of authors have studied the linear and nonlinear stages of the TI using numerical hydrodynamic simulations \\citep{hen99,kri02a,kri02b,kri04,vaz00,gaz01,san02,pio04,pio05,koy04,vaz07,kim08,ino08}. Thermal conduction is important to include in studies of the TI for two reasons. First, it suppresses the growth rate at small scales, in fact isobaric perturbations with wavelength smaller than the Field length $\\lambda_F$ \\citep{fie65} do not grow at all. Second, it produces evaporation from the surfaces of cold dense fragments, and the interaction of the evaporative flows can induce turbulence. Including thermal conduction is essential for numerical studies of the TI. Without explicit thermal conduction, perturbations at the grid scale grow fastest, and may eventually come to dominate the dynamics. For this reason, \\citet{koy04} concluded that numerical studies of the TI must satisfy a \"Field criterion\", that is the Field length must be resolved by at least a few cells to prevent artificial fragmentation at the grid scale, and to avoid the results being dominated by grid noise. Satisfying the Field criterion requires including explicit thermal conduction \\citep{pio04,pio05,koy04,bra07}. To highlight one example, \\citet{pio04,pio05} studied the interaction of the TI and the magnetorotational instability (MRI) in disks including isotropic thermal conduction; they found the MRI could drive turbulence and fragmentation in the diffuse ISM at amplitudes consistent with HI observations. To date most studies of the TI with conduction have assumed the conductivity is isotropic. However, in a magnetized plasma, electrons can flow more freely along magnetic field lines than across them, leading to anisotropic transport coefficients \\citep{spi62}. The degree of anisotropy is measured by the ratio of the electron gyro radius to the mean free path between collisions. For the warm medium, where typically $T=1500$ K, $n=2$ cm$^3$, and $B=1 \\mu$G, the Columb mean free path is $\\lambda_{\\rm mfp} \\sim 10^{10}$ cm, while the electron gyro radius is $r_g=10^{6}$ cm. Thus, in this medium the thermal conduction should be highly anisotropic, and primarily along magnetic field lines. The implications of anisotropic transport terms on the dynamics of astrophysical plasmas has begun to be explored recently \\citep{bal01,sha03,par05}. For example, in stratified atmospheres, anisotropic conduction can result in the magnetothermal and heat-flux buoyancy instabilities in the intracluster medium \\citep{par08,sha09}. In addition, it can have effects on the evolution of supernova remnants \\citep{bal08} and on magnetized spherical accretion flows \\citep{sha08}. Recently, \\citet{sha10} studied the TI in the intracluster medium, including heating by cosmic rays. Their primary result is that with anisotropic thermal conduction, the TI could produce filaments of cold gas orientated along magnetic field lines. In this study, we perform two-dimensional numerical hydrodynamic and magnetohydrodynamic simulations of TI in the ISM that include the anisotropic heat conduction. The purpose of this paper is to investigate the effects of anisotropic thermal conduction on the structure and dynamics of interstellar medium, and to compare the results to models which include only isotropic conduction. Since the geometry and strength of magnetic field in different regions of the ISM may vary, we perform simulations of the TI with various magnetic field strengths and two different initial geometries for the field. Ideally, the numerical studies of the thermal instability should include both isotropic and anisotropic conduction with the temperature dependency since the thermal instability develops density distribution with distinct peaks in the cold/dense phase dominated by isotropic conduction and in the diffuse/hot phase with higher ionization fraction and thus more affected by anisotropic conduction. However, including both isotropic and anisotropic conduction with a realistic temperature dependency is challenging, since it would decrease the Field length, and require higher resolution. In this study, we restrict our exploration to the simulations only with isotropic conduction or ones only with anisotropic conduction to better understand the effects of anisotropic conduction. This paper is organized as follows: our numerical methods and code tests are summarized in \\S~\\ref{sec:Model}. In \\S~\\ref{sec:iso}, we first discuss the effects of the conduction rate and resolution on the TI in hydrodynamical simulations, and then use these results to choose a specific set of model parameters for our simulations. Results from calculations of the nonlinear regime of the TI with anisotropic conduction are presented in \\S~\\ref{sec:aniso}. We summarize and discuss our results in \\S~\\ref{sec:Summary}. ", "conclusions": "\\label{sec:Summary} In this work, we have studied the nonlinear regime of the thermal instability (TI) and the effect of anisotropic thermal conduction by performing two-dimensional hydrodynamical and magnetohydrodynamic simulations incorporating radiative cooling and heating. Our main results can be summarized as the following: 1. As found in previous studies, it is crucial to include explicit thermal conduction in numerical studies of the TI so that the Field length is resolved, in order to prevent artificial fragmentation driven by numerical noise at the grid scale (e.g. Figure~\\ref{K_density}). 2. The amplitude of the thermal conductivity controls the rate of evaporation from the surfaces of dense clouds formed by the TI, and therefore strongly affects the amplitude of turbulent motions induced by the TI (e.g. Figure~\\ref{K_KEmean}). 3. Even when the Field length is resolved, explicit viscosity must be included to obtain numerical convergence of some quantities, for example the amplitude of turbulent motions driven by evaporative flows (e.g. Figure~\\ref{R_study_sigma}). 4. Although the statistics of the density and temperature are not strongly affected by anisotropic conduction, the geometry of structures formed by the TI are quite different. With anisotropic conduction and a uniform magnetic field, the TI saturates as long thin filaments of dense gas aligned with the field. In a tangled field, spherical clouds are formed in regions of closed field lines (e.g. Figure~\\ref{fig_logd}). 5. The combination of anisotropic conduction and MHD strongly suppresses the rate of evaporation of cold gas from the surfaces of dense structure in regions where the field is parallel to the interface. This reduces the amplitude of turbulence driven by the TI (e.g. Figure~\\ref{mhd_velo}). These results have a number of implications for observations of cold neutral gas in the ISM. In particular, the thin filaments along the magnetic field in the weak uniform magnetic field case agree well with recent observations of the Riegel-Crutcher cloud conducted by \\citet{mcc06}. This neutral hydrogen (HI) cloud lies on the edge of the Local Bubble \\citep{cru84} filled with a hot and diffuse gas where the anisotropic conduction can be expected. \\citet{mcc06} found a network of dozens of hairlike filaments of cold hydrogen with widths of less than $\\sim 0.1$ pc and up to 17 pc long. They also have found that the filaments are aligned with the magnetic field of the cloud which agrees well with our results. They also calculated the magnetic field strength by using the Chandrasekhar-Fermi method, finding $\\sim 60 \\mu$G. We find that compression and twisting of field during the formation of filaments by the TI with anisotropic conduction can amplify it by up to a factor of 100 in two-dimensions. In three-dimensions, the amplification is likely to be larger. This may be enough to explain the observed field. Recently, \\citet{sha10} have reported a study of the TI with anisotropic thermal conduction in the hot X-ray emitting plasma in clusters of galaxies. The heating and cooling processes in this regime are very different than those in the ISM studied here (equations \\ref{eq:E_heat} and \\ref{eq:E_cool} respectively), in particular there are no stable phases in the cooling curve they adopt, so that magnetic pressure sets the only limit on how cold and dense the gas becomes. Nonetheless, the structure of density condensations in this case are very similar to our results: filaments of cold gas along magnetic field lines. Our results show that in the case of anisotropic conduction, the geometry of the magnetic field with respect to the interface between cold and hot phases is very important. Only if the field is normal to the interface can thermal conduction drive evaporation and outflows. The inclusion of anisotropic conduction could have important implications for the structure of interfaces, and the interpretation of observations of these regions \\citep{ino06,sto09,sto10}. There are a number of limitations to our work that should be addressed in future investigations. Firstly, the ISM is highly turbulent \\citep{hei03}, with typical turbulent velocities approximately 7 $\\rm km s^{-1}$ \\citep{hei03,moh04}, i.e. more than 100 times larger than those produced by the TI with isotropic conduction and viscosity. In the traditional picture of the ISM, supernova driven turbulence leads to a hot, diffuse third phase \\citep{cox74,mck77}. More recently, \\citep{pio04} have considered the interaction of turbulence driven by the MRI and TI. We have not considered the effect of externally forced turbulence on the TI in this work, however this would be a productive direction for study in the future. Secondly, in this work we have adopted a constant conductivity $\\mathcal{K}$. However, in reality the conductivity is a function of the temperature \\citet{par53}, so that the rate of thermal conduction decreases as the gas cools. The amplitude of the conductivity can vary by 2 orders of magnitude between warm and cold phases at the end of our typical simulation at $t\\sim 100$~$\\rm Myr$. As discussed in Section \\ref{sec:coneffect}, the value of the conductivity can affect the rate of evaporation from dense clouds, and therefore the kinetic energy of turbulence driven by the TI, thus assuming a constant value may alter the result. Using a realistic temperature dependent conductivity is challenging, since it would decrease the minimum value of Field length and require much higher resolution. Since the geometry of the magnetic field limits the amount of evaporation to a very small surface area at the ends of the filaments in the case of anisotropic conduction, we do not expect the use of a more realistic conductivity to produce qualitative changes in our result. Nonetheless, more realistic studies which use temperature dependent conductivities would be fruitful. Finally, this study has considered flows in only two dimensions. In full three dimensions, the amplification of the magnetic field due to geometrical compression into filaments may be larger, and the turbulence driven by evaporative flows may be of a different character. Fully three-dimensional simulations of the TI with anisotropic conduction would also be interesting for future studies." }, "1112/1112.1463_arXiv.txt": { "abstract": "A wide field of view Cherenkov/fluorescence telescope array is one of the main components of the Large High Altitude Air Shower Observatory project. To serve as Cherenkov and fluorescence detectors, a flexible and mobile design is adopted for easy reconfiguring of the telescope array. Two prototype telescopes have been constructed and successfully run at the site of the ARGO-YBJ experiment in Tibet. The features and performance of the telescopes are presented. ", "introduction": "The energy spectrum of primary cosmic rays spans almost 12 orders of magnitude, from $10^{9}$~eV to $10^{21}$~eV, and can be well fitted by a simple power law except in several small energy regions. A region called the ``knee\" of the spectrum existing at around $10^{15}$~eV is one of these regions where the spectrum becomes steeper at higher energy side. Many experiments have observed this phenomenon; however, controversial arguments on its origin persist because of limited discrimination power on the primary cosmic ray composition and ambiguities in nucleus-nucleus interaction modeling. These two aspects are closely related to each other. Modern balloon borne experiments, such as ATIC~\\cite{ATIC} and CREAM~\\cite{CREAM}, have efficiently measured the energy spectra of individual elements at the top of the atmosphere. The energy spectra for all nuclei are measured up to $\\sim$100~TeV which is not far from the ``knee\". Because the detector area is constrained by the payload, the spectrum measurement has to be extended to a higher energy using a ground based air shower detector array. The spectrum should initially be measured well below 100~TeV to create an overlap with the balloon experiments which serve as absolute calibrations for the ground-based techniques. Identifying the individual components of cosmic rays continues to be a major challenge in ground-based experiments. Multiple parameter measurements on an air shower seem to be a plausible approach. The ultimate goal is to separate individual species out of the total observed-event samples and measure a clear individual ``knee\" for every single species, enabling the discovery of the origin of the ``knee\". As one of the major scientific goals of the Large High Altitude Air Shower Observatory (LHAASO) project~\\cite{LHAASO-caozh,LHAASO-hhh}, the energy spectrum for a separated composition will be measured at energies above dozens of ~TeV. To tag each primary particle that causes an air shower, the atmospheric depth of the shower maximum should be measured as one of the important parameters. The wide field of view Cherenkov/fluorescence telescope array (WFCTA), one of the components of the LHAASO project, is designed to accomplish this goal. A portable design of WFCTA telescopes is adopted to maximize the flexibility of changing the configuration of the array of telescopes. The elevations, pointing directions, and locations of the telescopes are then easily reconfigured. This is one way of using the same telescopes to serve as both fluorescence and Cherenkov detectors. In the fluorescence detector, the telescopes are tilted down to a horizontal position. In such an operational mode, which is analogous to the HiRes experiment~\\cite{Hires-detector}, most of the Cherenkov photons are avoided except those that are scattered onto the field of views (FOVs) of the telescope, such as in the fluorescence detector of the telescope array experiment~\\cite{TA-exp} and the fluorescence detector of the Pierre Auger experiment~\\cite{auger-fluorescence, auger-spectrum}. Only the fluorescence light from the shower is collected together with the scattered Cherenkov light to trigger the telescopes. This requires showers having much higher energy, usually above 100 PeV, such as in the HiRes prototype experiment~\\cite{Hires-hybrid}, because the fluorescence light by a single electron is considerably weaker and isotropic. In the Cherenkov detector, the telescopes run in high elevation mode to directly measure Cherenkov light from the showers, similar to what was done in the Dice experiment~\\cite{dice-exp}. A Cherenkov light radiation provides considerably more photons along the shower axis that are useful for lowering the shower energy. In 2007, two prototype Cherenkov telescopes~\\cite{CRTNT-caozh,CRTNT-hhh} were deployed at Yangbaijing (YBJ) Cosmic Ray Observatory near the ARGO-YBJ experiment~\\cite{ARGO-detector}. Moreover, two WFCTA telescopes have been successfully running in Cherenkov mode beginning August 2008. To date, millions of cosmic ray events that simultaneously trigger the telescopes and the ARGO-YBJ detector carpet array have been collected. An analysis of these events is carried out to study the performance of the telescopes. Detailed descriptions of the telescopes and the analysis of the findings are presented in this paper. Several details about the apparatus are presented in Section 2. The detector calibration is then discussed in Section 3. The test-run of the two telescopes and results are reported in Section 4 including summaries on the detector performance. The conclusions drawn are provided in the last section. ", "conclusions": "The telescopes were successfully run at YBJ from August 2008 up to July 2009. Millions of coincidence events with the ARGO-YBJ experiment have been collected. The performance of the telescopes was studied using these events. The trigger rate is about 0.5 Hz in stereo mode. Moreover, the mode energy of the telescope is 40 TeV when a pure proton composition is assumed. The features of the two WFCTA prototype telescopes are summarized as follows: \\begin{itemize} \\item a 4.7~$m^2$ spherical mirror, \\item a 16$\\times$16 PMT array covers an FOV of $14^{\\circ}\\times16^{\\circ}$ with $1^{\\circ}$ pixels, \\item dual gain system for a dynamic range to 3.5 orders of magnitude, \\item DC coupling and modulized design for electronics, \\item three-level online trigger logic: single channel trigger based on S/N ratio, telescope trigger based on pattern recognition, and event trigger for stereoscopic observation, \\item maximized mobility and the telescope can be uplifted from $0^{\\circ}$ to $60^{\\circ}$ in elevation. \\end{itemize} The absolute gains of the telescopes are calibrated using calibrated LEDs mounted at the centers of the mirrors. The systematic uncertainty of the calibration constant is about 7\\%. The pixel gains are monitored on a daily basis." }, "1112/1112.4070_arXiv.txt": { "abstract": "Radiative feedback is among the most important consequences of clustered star formation inside molecular clouds. At the onset of star formation, radiation from massive stars heats the surrounding gas, which suppresses the formation of many low-mass stars. When simulating pre-main-sequence stars, their stellar properties must be defined by a prestellar model. Different approaches to prestellar modeling may yield quantitatively different results. In this paper, we compare two existing prestellar models under identical initial conditions to gauge whether the choice of model has any significant effects on the final population of stars. The first model treats stellar radii and luminosities with a ZAMS model, while separately estimating the accretion luminosity by interpolating to published prestellar tracks. The second, more accurate prestellar model self-consistently evolves the radius and luminosity of each star under highly variable accretion conditions. Each is coupled to a raytracing-based radiative feedback code that also treats ionization. The impact of the self-consistent model is less ionizing radiation and less heating during the early stages of star formation. This may affect final mass distributions. We noted a peak stellar mass reduced by 8\\% from 47.3$\\Msun$ to 43.5$\\Msun$ in the evolutionary model, relative to the track-fit model. Also, the difference in mass between the two largest stars in each case is reduced from 14$\\Msun$ to 7.5$\\Msun$. The HII regions produced by these massive stars were also seen to flicker on timescales down to the limit imposed by our timestep ($<$ 560 years), rapidly changing in size and shape, confirming previous cluster simulations using ZAMS-based estimates for prestellar ionizing flux. ", "introduction": "The conversion of molecular gas into fully-formed stars is complex, involving several diverse processes. These different processes are linked to each other through feedback mechanisms that make isolating and understanding the contribution of each process a difficult task. A key point in this regards is that stars also rarely form in isolation, but instead are seen to be forming in clusters and subclusters within molecular clouds \\citep{Clarke+2000,Testi+2000}. In the cluster environment, the formation of a sufficiently massive star can affect all the others through the energy it radiates back into the cloud. Numerical simulations of star formation have made it very clear that the effects of stellar radiation cannot be neglected. Simulations including some form of radiative transfer show a dramatic reduction in the production of brown dwarfs and other low-mass stars \\citep{Offner2009}, due to an increase in gas temperatures reducing fragmentation \\citep{Krumholz+2007,Peters2010c}. More of the available gas mass ends up being accreted by the fewer, larger stars formed, and the fragmentation that does occur takes place in optically thick self-shielding discs \\citep{Krumholz+2007,Peters2010a,Peters2010c}. The fact that radiation affects the mass spectrum in simulations of molecular cloud clumps has obvious implications for the shape of the initial mass function, for example the suppression of excessive brown dwarf formation \\citep{Bate2009,Krumholz+2010,Peters2010c}. Massive stars also emit prodigious amounts of UV radiation \\citep{Hoare+2007,Beuther+2007} creating expanding HII regions. The hot ($10^4$ K) gas expands into the colder ($10^2$ K) surrounding low-pressure gas, creating another feedback mechanism and ionized region that may contribute to the destruction of molecular clouds \\citep{Keto2002,Keto2003,Keto2007,Matzner2002,Peters2010a,Peters2010c}. HII regions can be observed by their radio continuum emission \\citep{MezgerHenderson1967}, or by their recombination lines (e.g.\\ \\citet{WoodChurchwell1989} use the H76$\\alpha$ line). More recently, observations have shown time variability in HII regions \\citep{Franco-HernandezRodriguez2004,Rodriguez+2007,Galvan-Madrid+2008,gomezetal08}. \\citet{Franco-HernandezRodriguez2004} have suggested that such observed time-variability may be due to the changes occurring in the source of the ionizing radiation, though it may also be due to increased absorption in the rapidly-evolving core of the nebula. \\citet{Peters2010b} present a technique for using synthetic radio maps to study the time-evolution of stars forming in a cluster environment and variability in the morphology and size of HII regions. Analysis of these simulations by \\citet{Peters2010a} and \\citet{Galvan-Madrid+2011} confirmed variability in the flux and size measurements of HII regions, which in a few cases might be observable on timescales of $\\sim 10$ years. They also noted that positive changes were more likely to occur than negative changes, i.e.\\ that most of the flux variations were increases rather than decreases. To further explore the impact of radiative feedback and the possible variability in HII regions, simulations must be equipped with good protostellar models. These have been investigated by \\citet{PallaStahler1991}, \\citet{PallaStahler1992}, \\citet{Nakano2000}, \\citet{McKeeTan2003}, \\citet{Offner2009} and \\citet{HosokawaOmukai2009}, among others. It is clear from these models, that the evolution of a protostar depends heavily on the mass accretion rate. Among other things, they show that the radius of the protostar may grow or contract depending on the stellar evolutionary stage. With a radius that can change significantly during the pre-main-sequence lifetime of the star, the effective temperature can also be expected to vary significantly. To study this, we simulate the formation of a cluster of stars inside a molecular cloud. We equip the stars with one of two prestellar models based on the ones described in \\citet{Peters2010a} and \\citet{Offner2009}, each with its own characteristics. The \\citet{Offner2009} model has already been used to study star cluster formation in \\citet{Krumholz+2011}, though with different initial conditions. Ours is the first simulation with the protostellar model to also include the effects of ionizing radiation and HII region formation. We connect the model to a radiative transfer method that computes the heating and ionization due to radiation from the stars formed in the simulation. The differences between the two models is explained in \\ref{sec:protmod}, but the key difference is that the \\citet{Offner2009} model treats the evolution of the radius and luminosity self-consistently. The choice of stellar model affects the early evolution of stars in a cluster, and may have repercusions for the final mass spectrum. Though not entirely conclusive, we find that reduced heating and ionization in the early stages of star formation when using the \\citet{Offner2009} model resulted in a more equitable mass distribution. With the \\citet{Peters2010a} model, the cluster came to be dominated more by a single star about $14\\Msun$ more massive than the next largest, compared to a $\\sim7.5\\Msun$ gap in the \\citet{Offner2009} model simulations. Other effects of the self-consistent prestellar modeling are delayed ionization of the cluster gas by 3\\% of a freefall time (17.7 kyr), and delayed heating of the cluster gas by 1\\% of a freefall time (5.9 kyr). Our numerical approach is described in Section \\ref{sec:numerical_methods}. In Section \\ref{sec:results} we list our results for the early evolution of star clusters with massive stars. Our assessment of the impact of protostellar modeling we discuss in Section \\ref{sec:discussion} and summarize our findings in Section \\ref{sec:conclusion} with a view to future simulations. ", "conclusions": "\\label{sec:conclusion} Stars begin to affect their birth environments as soon as they are born through radiative feedback. We have considered the impact that pre-main-sequence modeling can have on a star cluster by comparing two different prestellar models already described in the literature. We did this by repeating the simulation of \\citet{Peters2010a}. We then upgraded the FLASH code to include a protostellar evolution module based on the one described in the appendices of \\citet{Offner2009}. Each model works by equipping the stars in the simulation (``sink particles'') with a stellar radius and luminosity. The greatest difference between the two models was self-consistency. The \\citet{Peters2010a} model calculated approximate stellar parameters on-the-fly, while the evolving protostar model evolved the stellar parameters self-consistently through the simulation as the stars grew and accreted mass. In terms of the overall gas structure, HII regions, temperature structure, mean ionization fraction, or stellar binarity, the two models produced qualitatively the same results. This is because a cluster comes to be dominated by its most massive stars, which are evolved, main-sequence, highly luminous stars, regardless of the choice of stellar model. These one or two massive stars control the overall dynamics. The differences exist in the early phase of star formation. Major ionization of the gas in the evolving protostar model lagged the \\citet{Peters2010a} model by about 3\\% of a freefall time, or about 17.7 kyr. Major heating of the gas lagged by about 1\\% of a freefall time, or about 5.9 kyr. The difference in heating and ionization was due to the fact in \\citet{Peters2010a}, the stellar radius was underestimated (a ZAMS-equivalent value was taken), when protostars have radii an order of magnitude larger than a zero-age main-sequence star of equal mass. The correspondingly higher surface temperatures resulted in excess heating and ionization in this model. When both models had stars converging onto the main sequence, the differences between the two models diminished. It is possible that these initial differences could have had a lasting effect on the stellar population. The most massive star at the end of each simulation was 43.5$\\Msun$ in the evolving protostar model, and 47.3$\\Msun$ in the \\citet{Peters2010a} case---a difference of 8\\%. The differences in mass between the most massive star and the next largest star was 7.5$\\Msun$ in the evolving protostar case and 14$\\Msun$ in the \\citet{Peters2010a} case. It would require further simulations, varying the initial conditions, to confirm that this is always the case. The cluster of stars is embedded in a rotating disc of gas approximately 0.2 pc in size. The expanding HII regions above and below the disc are rapidly changing in shape and size on timescales shorter than 570 years. The physical size of these HII regions in our simulation is at most about 0.2 pc. This flickering is observed regardless of the prestellar model used. Future simulations will have initial conditions including turbulence to model molecular clouds as realistically as possible. The stars will no longer be forming within a global disc, but rather along sheets and filaments in diverse parts of the cloud. With star formation thus spread out more in space and time, we expect the influence of individual young stellar objects on their environments to be more significant than when all stars form in a central cluster. It will be important to have the radiative feedback accurately modeled in these cases." }, "1112/1112.1655_arXiv.txt": { "abstract": "We present the results of the optical, X-ray and gamma-ray analysis of some recent GRBs. The data were obtained by the automated P60 telescope and the Swift telescope (UVOT, XRT and BAT). We present some example fits for the lightcurves. The data reduction and the investigations were made by the Konkoly Observatory HEART group (http://www.konkoly.hu/HEART/index.html). ", "introduction": "The UVOT photometry was done using NASA's HEASOFT software package. This package provides a complete assistance for doing photometry for measurements done by various telescopes (e.g. Swift, CGRO, INTEGRAL). For photometry, contrary to the suggestions of the software manual, we explored various aperture sizes between $1$\" and $10$\" (instead of keeping $5$\" at all times) to get the most usable data. Then we chose those aperture sizes, which provided the most accurate results. For a given filter band we used the best aperture size, meaning that in some cases when obtaining photometry for a given GRB we used various apertures depending on the filter band. All UVOT magnitudes are in the Standard UVOT Photometric System \\citep{uvot_phot}. Magnitudes obtained by ground based telescopes were taken from GCNs (in detail see at references \\cite{gcn}). When converting magnitudes to fluxes, we used the methods described in \\cite{bessel,p60,sdss}. Magnitudes are not corrected for galactic extinction. X-ray data is taken from the XRT observations available from \\cite{evans}. According to our results we suggest 19 magnitude as acceptable faintest limit in the UVOT photometry system. We used the \\textit{mpfit} package \\cite{mpfit} for fitting broken power-law functions to the lightcurves with the $F_{\\nu} \\propto t^{\\alpha}$ convention. ", "conclusions": "\\ In this work we aimed to produce photometric data with the highest possible accuracy. Our sample consisted of relatively bright and well observed bursts in order to achieve a reliable data set with our reduction method. We excluded those large error measurements which we had to disregard during our fitting procedure. According to our results we suggest 19 magnitude as the faintest limit when using the UVOT system. \\begin{theacknowledgments} Thanks to Bob Wiegand for the help us to make HEASOFT working correctly. This work was partially supported by OTKA grant K077795 (P.V.). \\end{theacknowledgments}" }, "1112/1112.0379_arXiv.txt": { "abstract": "\\himass\\ is one of the most \\HI-massive galaxies in the local ($z<0.1$) Universe. Not only are such galaxies extremely rare, but this ``coelacanth\" galaxy exhibits characteristics -- in particular its active, inside-out stellar disk-building -- that appear more typical of past ($z \\sim 1$) star formation, when large gas fractions were more common. Unlike most local giant \\HI\\ galaxies, it is actively star forming. Moreover, the strong infrared emission is not induced by a merger event or AGN, as is commonly found in other local LIRGs. The galaxy is suggestive of a scaled-up version of local spiral galaxies; its extended star formation activity likely being fueled by its large gas reservoir and, as such, can aid our understanding of star formation in systems expected to dominate at higher redshifts. The multi-wavelength imaging and spectroscopic observations that have led to these deductions will be presented. These include NIR ($JHK$) and MIR (Spitzer; $3-24\\micron$) imaging and photometry, MIR spectroscopy, ATCA \\HI-interferometry and Mopra CO line emission observations. But no optical data, as the galaxy is heavily obscured due to its location in Vela behind the Milky Way. ", "introduction": "The galaxy \\himass, originally discovered in the Parkes deep Multibeam \\HI-survey of the Zone of Avoidance (ZOA; e.g. Kraan-Korteweg et al. 2005), is a rapidly rotating disk galaxy ($v_{\\rm rot}=630$\\kms; $D_{\\rm HI} = 130$kpc) with an \\HI-mass of $M_{\\rm HI} = 7.5 \\cdot 10^{10}\\msun$ and a dynamical mass of $M_{\\rm tot} = 1.4 \\cdot 10^{12}\\msun$ (Donley et al. 2006). It has a prominent bulge and smallish stellar disk ($B/D \\sim 0.8$) of evolved stars of $M^* = 4.4 \\cdot 10^{10}\\msun$ embedded in the five times larger HI-disk, while its current hearty star forming activity makes it a luminous infrared galaxy (LIRG; next section). \\himass\\ is the most \\HI-rich galaxy known. Such galaxies are extremely rare ($\\sim 3 \\cdot 10^{-8}$/Mpc$^3$) according to the current best determined local galaxy HI-mass function (HIMF; Zwaan et al. 2005) and should not have been found at all in the explored volume. Interestingly, the recent, larger volume ALFALFA survey 40\\%-data release (Haynes et al. 2011) identify several, similarly extreme \\HI-massive galaxies in excess of the predicted HIMF number density. Independently, galaxies with such vast reservoirs of gas were more common in the past, as were LIRGs. Hence this enigmatic, relatively nearby ($v=10\\,689$\\kms) galaxy is an ideal {\\sl local} probe that enables detailed studies of evolutionary processes and the transformation of gas into stars. ", "conclusions": "" }, "1112/1112.5240_arXiv.txt": { "abstract": "We discuss a possible physical connection between helium-rich ($Y \\ge 0.35$) stellar populations of massive globular clusters (GCs) and the ultraviolet (UV) upturn of galactic spheroids by using analytical and numerical models. In our model, all stars are initially formed as bound or unbound star clusters (SCs) formed from giant molecular clouds (GMCs) and the SCs can finally become GCs, open clusters, and field stars depending on physical properties of their host GMCs. An essential ingredient of the model is that helium-rich stars are formed almost purely from gas ejected from massive asymptotic giant branch (AGB) stars. The helium-rich star formation is assumed to occur within massive SCs if the masses of the progenitor GMCs are larger than a threshold mass ($M_{\\rm thres}$). These massive SCs can finally become either massive GCs or helium-rich field stars depending on whether they are disintegrated or not. Using this model, we show that if the initial mass functions (IMFs) in galactic spheroids are mildly top-heavy, then the mass fractions of helium-rich main-sequence stars ($F_{\\rm He}$) can be as large as $\\sim 0.1$ for $M_{\\rm thres}=10^7 {\\rm M}_{\\odot}$. $F_{\\rm He}$ is found to depend on IMFs and $M_{\\rm thres}$ such that it can be larger for shallower IMFs and smaller $M_{\\rm thres}$. The inner regions of galactic spheroids show larger $F_{\\rm He}$ in almost all models. Based on these results, we suggest that if the UV upturn of elliptical galaxies is due to the larger fractions of helium-rich stars, then the origin can be closely associated with top-heavy IMFs in the galaxies. ", "introduction": "Recent observational and theoretical studies of the Galactic GCs have suggested that some of the massive GCs ($\\omega$ Cen and NGC 2808) have significant fractions of helium-rich stars (e.g., Bedin et al. 2004; Norris 2004; Lee et al. 2005a; Piotto et al. 2005, 2007; Renzini 2008). The observed unusual level of helium enhancement ($\\Delta Y/\\Delta Z \\approx 70$; Piotto et al. 2005) and sizable fractions of helium-rich stars in these GCs have driven many theoretical studies to make great efforts in understanding where and how helium-rich stars can be formed (e.g., Bekki \\& Norris 2006; D'Antona \\& Ventura 2007; D'Ercole et al. 2008). As discussed by Renzini (2008), helium-rich stars can be formed from gaseous ejecta from massive AGB stars with initial masses ($m_{\\rm I}$) ranging from $3{\\rm M}_{\\odot}$ to $8{\\rm M}_{\\odot}$, though the ejecta should not be diluted by interstellar gas (ISM) to keep the original high helium abundances of the ejecta. Code \\& Welch (1979) investigated the integrated light of seven elliptical and S0 galaxies for a wavelength range of 1550-4250 \\AA \\, and found that some of them show an increase in energy at the shortest wavelengths. Since this discovery of the UV upturn, the origin of the UV upturn has been extensively discussed both in observational and theoretical studies (e.g., Bertola et al. 1980; Burstein et al. 1988; Greggio \\& Renzini 1990; Horch et al. 1992; Dorman et al. 1995; Brown et al. 1997; O'Connell 1999; Yi et al. 1997, 1998, 2005). Although there can be a number of stellar candidates responsible for the UV upturn in galaxies, one of the most promising ones is old horizontal-branch stars (e.g., Yi 2008 for a recent review). It is suggested that enhanced helium abundances can play an important role in the formation of hot stars responsible for the UV upturn (Yi 2008). Brown et al. (2003) investigated the UV emission in eight early-type galaxies at $z=0.33$ (a look-back time of 3.9 Gyr) and found that the UV emission in these galaxies is significantly weaker than it is in the current epoch. Observational studies by {\\it Galaxy Evolution Explorer (GALEX)} investigated the UV properties of bright cluster galaxies (BCGs) in clusters at $z<0.2$ and compared them with those of nearby giant elliptical galaxies (e.g., Lee et al. 2005b; Ree et al. 2007). Ree et al. (2007) concluded that the observed evolution of FUV $-$ $V$ color with a model in which the dominant FUV source is hot horizontal-branch stars (e.g., Ree et al. 2007). Yi et al. (2011) investigated correlations between the strength of the UV upturn and global galactic properties (e.g., luminosity) in BCGs and did not find any remarkable correlations (see also Loubser \\& S\\'anchez-Bl\\'azquez 2011 for similar results). Yi et al. (2011) therefore concluded that the helium sedimentation scenario proposed by Peng \\& Nagai (2009) can not be supported by their observational results. These observational and theoretical studies appear to suggest that there can be helium-rich stellar populations in diverse objects with dramatically different masses and sizes ranging from GCs to BCGs. Although it would be likely that different astrophysical objects with helium-rich stars have different origins, it would be possible that they can have a common origin. Previous theoretical studies, however, have not yet provided an unified picture for the origins of helium-rich stars in GCs, galactic bulges, and elliptical galaxies with the UV-upturn. It is therefore worthwhile to construct a model to discuss the origins of helium-rich stars observed in GCs and galactic spheroids in a self-consistent manner. The purpose of this paper is to present a new model which can provide a possible explanation for the origins of helium-rich stars with $Y \\ge 0.35$ in GCs and galactic spheroids in a self-consistent manner. Based on the model, we mainly discuss what types of IMFs are required to explain the observed possible fraction of helium-rich stars in galactic spheroids with the UV upturn (e.g., Chung et al. 2011). The most important ingredient of the model is that helium-rich stars are formed almost purely from gas ejected from massive AGB stars with $3 {\\rm M}_{\\odot} \\le m_{\\rm I} \\le 8 {\\rm M}_{\\odot}$ with no/little dilution of the gas with ISM. This formation of new stars from AGB ejecta have been included also in recent theoretical models of GC formation (e.g., D'Ercole et al. 2008). In the present study, we assume that such formation of helium-rich stars almost directly from AGB ejecta can be possible only within SCs formed from massive GMCs. By considering that the vast majority of stars are observed to be initially formed as bound or unbound SCs (e.g., Lada \\& Lada 2003), we assume that all stars form as SCs from GMCs in the model. Some massive SCs can retain gaseous ejecta from AGB stars and consequently have new helium-rich stars formed form the ejecta, and the helium-rich stars can become field stars of their host galaxies when the SCs become disintegrated. Such massive SCs can finally become GCs or nuclear SC (or stellar galactic nuclei) depending on their birth places, if they are not destroyed by their host galaxies. The IMFs and GMCMFs are key parameter that can determine mass fractions of helium-rich stars in galaxies. By using this new model, we discuss the origin of physical properties of the Galactic GCs and the UV upturn in galactic bulges and elliptical galaxies. Although it remains less clear how much fractions of stars should be helium-rich stars in galactic spheroids with the UV upturn (Yi 2008), Chung et al (2011) have recently suggested that if about 11\\% of stellar populations in galactic spheroids are helium-rich stars, then they can show the UV upturn for a given IMF. In their models, they considered that the major source of far-UV flux originates from metal-poor and helium-enhances hot horizontal branch stars. We therefore consider that if the fractions of helium-rich stars in galactic spheroids are less than $\\sim 0.1$, the spheroids are unlikely to show the UV upturn in the present models. We investigate in what physical conditions (e.g., IMFs) the fractions of helium-rich stars can be above 0.1 for galactic spheroids to show clearly the UV upturn. A possible physical link between helium-rich stellar populations and the UV upturn of galactic spheroids has been already pointed out by a number of authors (e.g., Yi 2008). Thus the main point of the present study is {\\it not} to propose the importance of helium-rich stellar populations in the origin of the UV upturn {\\it but to discuss how galactic spheroids can have significant fractions of helium-rich stellar populations.} It should be also stressed that the present model is idealized and less realistic in some points (e.g., no chemical evolution of galaxies): the model should be regarded as a first step toward better understanding the origin of helium-rich stars in GCs and galaxies. More sophisticated numerical simulations including SC formation processes in galaxies will need to be done in our future studies to address the UV upturn problem in a much more quantitative way. The plan of the paper is as follows: in the next section, we describe a model which enables us to estimate (i) the mass fraction of helium-rich stars in a single GC and (ii) the mass fraction of helium-rich main-sequence (MS) stars in a galaxy for a given IMF and GMCMF. In \\S 3, we present the results on the number/mass fraction of massive GCs with helium-rich stars in the Galaxy and the mass fractions of helium-rich MS stars in galaxies with different IMFs and GMCMFs In \\S 4, we provide important implications of the present results in terms of the origin of helium-rich stars in GCs, the Galactic bulge, and elliptical galaxies. We also discuss the origin of the radial gradients of helium-rich stars in galactic spheroids in this section. We summarize our conclusions in \\S 5. ", "conclusions": "We have adopted a model in which (i) all stars are formed as bound or unbound SCs from GMCs and (ii) HRSs with $Y\\ge 0.35$ can be formed from gas ejected from AGB stars only in massive SCs with $m_{\\rm gmc} \\ge M_{\\rm thres}$. Some SCs can become field stars in their host galaxies after disintegration and others can become massive GCs or nuclear star clusters (or stellar nuclei) in galaxies. If massive SCs with HRSs are destroyed by their host galaxies, then HRSs become helium-rich field stars in the galaxies. We have investigated mass fractions of MS HRSs ($f_{\\rm He}$ and $F_{\\rm He}$) in GCs and galactic spheroids and their dependences on IMFs, $M_{\\rm thres}$, and GMCMFs based on the CDS (``cluster disintegration scenario''). We have also investigated the radial gradients of $F_{\\rm He}$ in galaxies with mass-size relations of GMCs by using N-body simulations on dynamical evolution of multiple cluster systems. We summarize our principle results as follows. \\\\ (1) The mass ratios ($f_{\\rm m, agb}$) of gaseous ejecta from massive AGB stars with $3{\\rm M}_{\\odot} \\le m_{\\rm I} \\le 8 {\\rm M}_{\\odot}$ to initial total masses of SCs are typically $\\sim 0.08$ for a canonical IMF (${\\alpha}_1 = 2.35$) with the reasonable lower and upper cut-off masses. If these gas can be efficiently converted into new stars with enhanced helium abundances (i.e., HRSs) in SCs, then the mass fractions of MS HRSs among all MS stars ($f_{\\rm He}$) are $\\sim 0.08$ for canonical IMFs and a SC age of 12 Gyr. Only if original SCs have top-heavy IMFs (e.g., ${\\alpha}_1 \\sim 1.5$), then $f_{\\rm He}$ can be larger than 0.3 for $\\alpha_2 \\ge 1.85$. \\\\ (2) The initial number and mass fractions of massive GCs (MGCs) with HRSs ($F_{\\rm n, mgc}$ and $F_{\\rm m, ngc}$, respectively) among the Galactic GCs formed from GMCs with $m_{\\rm gmc} \\ge 10^6 {\\rm M}_{\\odot}$ is $\\sim 0.16$ and $0.66$, respectively, for $M_{\\rm th}=10^7 {\\rm M}_{\\odot}$ and $\\beta =1.7$. Both $F_{\\rm n, mgc}$ and $F_{\\rm m, mgc}$ are lager for smaller $M_{\\rm th}$ and smaller $\\beta$. MGCs with $m_{\\rm gc}= 5 \\times 10^6 {\\rm M}_{\\odot}$ can sink into the Galactic center owing to dynamical friction to disappear from the halo region within 13 Gyr, if the initial radii from the Galactic center are less than 3.4 kpc. About 37\\% (49\\%) of the Galactic GCs with $m_{\\rm gc} \\ge 5 \\times 10^6 {\\rm M}_{\\odot}$ ($m_{\\rm gc} \\ge 10^7 {\\rm M}_{\\odot}$) have already been lost from the halo owing to efficient dynamical friction of these massive GCs. Thus, the Galactic halo initially could have a larger fraction of MGCs. \\\\ (3) The mass fractions of HRSs among all stars ($F_{\\rm He, t}$) in galactic spheroids are $\\sim 0.05$ for $M_{\\rm thres}=10^7 {\\rm M}_{\\odot}$, $\\beta =1.7$, $f_{\\rm m, agb}=0.1$, and ${\\epsilon}_{\\rm bsc}=1$. $F_{\\rm He, t}$ of galactic spheroids are larger for smaller $M_{\\rm th}$ and smaller $\\beta$ for a given IMF. The mass fractions of MS HRSs among all MS stars ($F_{\\rm He}$) in galactic spheroids are 0.06 for canonical IMFs ($\\alpha_1=\\alpha_2=2.35$), $M_{\\rm th}=10^7 {\\rm M}_{\\odot}$, $\\beta=1.7$, and a galaxy age of $\\sim 12$ Gyr. $F_{\\rm He}$ in galactic spheroids can be larger for smaller $M_{\\rm th}$ and smaller $\\beta$ for a given IMF, and it can be significant ($>0.1$), if original SCs (i.e., building blocks of the spheroids) have top-heavy IMF (e.g. $\\alpha_1 < 2$). These results suggest that if the observed UV upturn in bright elliptical galaxies is due to the larger fractions ($\\sim 10$\\%) of HRSs within them, the IMFs need to be top-heavy. If ${\\epsilon}_{\\rm bsc}$ is as small as $\\sim 0.1$, then bright elliptical galaxies are unlikely to show the UV upturn owing to the small fractions of HRSs. \\\\ (4) The Galactic bulge can have a larger $F_{\\rm He}$ if it had a top-heavy IMF at its formation in the CDS. Given a number of suggestions on the top-heavy IMF of the bulge by chemical evolution models of the bulge, the observed possible presence of helium-rich stars with $Y \\sim 0.35$ in the bulge (e.g., Nataf et al. 2011) can be understood in the context of the CDS. HRSs in MGCs that were initially located in the Galactic halo and had sunk into the bulge can be currently observed as field HRSs in the bulge, though the contribution of such stars to the entire helium-rich population is rather minor. It is doubtlessly worthwhile for observational studies to investigate whether or not the bulge has a larger fraction of HRSs in the inner regions. \\\\ (5) If $M_{\\rm th}$ and $\\beta$ are constant in elliptical galaxies with different physical properties, then elliptical galaxies with shallower (or more top-heavy) IMFs can show larger fraction of HRSs. Therefore, if more massive/luminous elliptical galaxies have shallower IMFs, then they can have larger fraction of HRSs. More massive elliptical galaxies can retain more efficiently the ejecta of supernovae so that they can finally have higher metallicities of their stars (e.g., Arimoto \\& Yoshii 1987). Therefore, if the Burstein relation (a correlation between UV flux and Mg index; Burstein et al. 1988) in bright elliptical galaxies is due to the correlation between $F_{\\rm He}$ and Mg indices, then the relation can be understood in terms of shallower IMF slopes ($\\alpha_1$) in more massive/luminous elliptical galaxies. \\\\ (6) Galactic spheroids can show negative radial gradients of $F_{\\rm He,t}$ and $F_{\\rm He}$ (i.e., higher in inner parts) for a reasonable set of model parameters. This is mainly because massive SCs with HRSs can sink rapidly into the inner regions of galaxies owing to dynamical friction and disintegrate there. The HRSs initially in the SCs can be dispersed to become field HRSs there after disintegration of the host SCs. Therefore they can show the stronger UV upturn in their inner regions, if $F_{\\rm He}$ can determine the strengths of the UV upturn. The observed spatial distribution of the FUV excess in elliptical galaxies can be understood in the context of formation and evolution of MGCs with HRSs. The central regions of galactic spheroids can be composed of two different stellar populations with normal helium abundances and enhanced ones (i.e., HNSs and HRSs).\\\\ Thus, an advantage of the CDS is that both (i) the absence or presence of the UV upturn in galactic spheroids and (ii) correlations between the strength of the UV upturn and galactic properties (e.g., Mg$_2$ index) can be discussed in the context of IMF slopes of galaxies in a self-consistent manner. A disadvantage of the CDS is that it is unclear whether and in what physical conditions ${\\epsilon}_{\\rm bsc}$ can be $\\sim 1$ for massive GMCs in galaxies. The formation of HRSs is possible, if helium-rich gas from AGB stars is converted into new stars without dilution of the gas by ISM with normal helium abundances. In the CDS, the inner regions of massive SCs are assumed to be the production sites of HRSs. It would be possible that AGB ejecta can be converted into new cold gas clouds outside SCs and consequently into new stars in deep potential wells of galactic spheroids, if there is no/little cold gas with normal helium abundances there. We need to investigate where and how AGB ejecta from field stars and SCs can be converted into new stars in galaxies using a more sophisticated chemodynamical simulation in our future studies." }, "1112/1112.2710_arXiv.txt": { "abstract": "Pristine stars with masses between $\\sim$140 and 260 $M_{\\odot}$ are theoretically predicted to die as pair-instability supernovae. These very massive progenitors could come from Pop III stars in the early universe. We model the light curves and spectra of pair-instability supernovae over a range of masses and envelope structures. At redshifts of reionization $z\\geq 6$, we calculate the rates and detectability of pair-instability and core collapse supernovae, and show that with the \\emph{James Webb Space Telescope}, it is possible to determine the contribution of Pop III and Pop II stars toward reionization by constraining the stellar initial mass function at that epoch using these supernovae. We also find the rates of Type Ia supernovae, and show that they are not rare during reionization, and can be used to probe the mass function at 4-8 $M_{\\odot}$. If the budget of ionizing photons was dominated by contributions from top-heavy Pop III stars, we predict that the bright end of the galaxy luminosity function will be contaminated by pair-instability supernovae. ", "introduction": "The life of a massive star ends in a supernova (SN). The detection of neutrinos from SN 1987A verified the idea that some SNe are set off by the gravitational collapse of the iron core of their progenitor star \\citep{Krauss1987}. However, theory predicts that very massive stars with helium cores between $\\sim$64 and 133 $M_{\\odot}$ could find another way to blow up, through the thermonuclear explosion of oxygen via the pair-production instability \\citep{Rakavy1967, Barkat1967, Heger2002}. The production of electron/positron pairs in the core softens the equation of state, leading to collapse and the ignition of explosive oxygen burning. The subsequent thermonuclear runaway reverses the collapse and ejects the entire star, leaving no remnant behind. Unlike iron core-collapse supernovae (CCSNe), which involves poorly constrained physical processes such as turbulence, pulsations, perhaps rotation and magnetic fields, the physics involved in pair-instability supernovae (PISNe) is fairly well understood and can be modeled with fewer uncertainties \\citep{Langer2009}. Due to the extremely large stellar mass required, the progenitors of PISNe are expected to be rare, and may only form under unusual conditions. One such condition existed in the early universe, when metal-free Population III stars were born \\citep{Loeb2010}. In star formation, it is the accretion process that ultimately sets the final mass of a star. From dimensional arguments, the mass growth rate is simply given by the Jeans mass $M_{J} \\sim c_{s}^3 G^{-\\frac{3}{2}} \\rho^{-\\frac{1}{2}}$ over the free-fall time $t_{ff} \\sim 1/\\sqrt{G\\rho}$, implying $dM/dt \\propto c_{s}^3/G \\propto T^{\\frac{3}{2}}$, where the sound speed $c_{s} \\sim \\sqrt{kT/m_{p}}$. In present day star-forming regions, heavy elements radiatively cool the gas to a temperature as low as $T\\sim 10$K. However, in primordial clouds, the primary coolant at low temperatures is molecular hydrogen, which can only cool the gas to $T\\sim 200-300$K, implying an accretion rate higher than present day by two orders of magnitude. Hence, theoretical studies suggest that the initial mass function (IMF) of Pop III stars might have been biased toward masses much higher than today, e.g. several hundred $M_{\\odot}$ \\citep{Bromm2004}. The nucleosynthesis imprints of this top heavy IMF have been seen in globular clusters and damped Lyman alpha systems \\citep{Cooke2011, Puzia2006}. Moreover, massive stars have strong winds driven by radiation pressure through spectral lines, with a mass loss rate scaling with stellar metallicity $\\dot{M}\\propto Z^{0.5\\sim 0.7}$ \\citep{Vink2001, Kudritzki2002}. Most PISNe should therefore be from Pop III stars, which have weak radiation-driven winds due to their extremely low metallicities, and retain enough of their initial masses at the end of their lives to undergo a pair-instability explosion. Naturally, studies of the rates and detectability of PISNe focused on high redshifts before reionization. \\citet{Mackey2003} found the PISNe rate to be $\\sim 50$ deg$^{-2}$ yr$^{-1}$ at $z > 15$, while \\citet{Weinmann2005}, using more conservative assumptions for the number of PISNe produced per unit Pop III stellar mass formed, found the PISNe rate to be $\\sim 4$ deg$^{-2}$ yr$^{-1}$ at similar redshifts. Assuming that only one supermassive Pop III star forms in unenriched minihalos, and that none form in protogalaxies, \\citet{Wise2005} found the PISNe rate be $\\sim 0.34$ deg$^{-2}$ yr$^{-1}$ at $z\\sim 20$. After our paper was submitted, \\citet{Hummel2011} presented a complimentary analysis of the source density of PISNe from pristine minihalos, and determined the observability of such events with the \\emph{James Webb Space Telescope} (JWST), finding approximately $\\sim 0.4$ PISNe visible per JWST field of view at any given time. PISNe after the epoch of reionization were also considered; \\citet{Scannapieco2005a} calculated a suite of PISNe model light curves with blackbody spectra, and analyzed the detectability and rates of PISNe from Pop III stars formed from leftover pristine gas at $z \\lesssim 6$. As for CCSNe during reionization, \\citet{Mesinger2006} presented detailed predictions for the number of core collapse SNe that JWST could observe as a function of different survey parameters. In this paper, we present light curves and spectral time series for PISNe from our multi-wavelength radiation-hydrodynamics simulations. As the stellar population responsible for reionization is currently unknown, instead of predicting a fixed SNe rate, we normalize the star formation rate by requiring that enough ionizing photons must be produced by either Pop III or Pop II stars in protogalaxies to complete reionization by $z \\sim 6$, and calculate the rates of pair-instability, core-collapse, and Type Ia SNe and their detectability with JWST for these two scenarios; the actual SNe rates will be in between these limiting cases. We show that using the observed rates of these SNe, it is possible to distinguish the contribution of Pop III and Pop II stars toward reionization by characterizing the IMF at that time. ", "conclusions": "We analyzed simulated light curves and spectra of pair-instability supernovae for a variety of progenitor masses and envelope types, and found that the supernovae from the more massive progenitors are super-luminous and have extended light curves, traits that would help photometrically distinguish pair-instability supernovae from other types of supernovae using repeated snapshots. We calculated the rates and detectability of pair-instability, core collapse, and Type Ia supernovae during the redshifts of reionization, and showed that it is possible to constrain the initial mass function of stars at that time, and identify the stellar population responsible for reionization. If Pop III stars made the dominant contribution of ionizing photons during reionization, the bright end of the galaxy luminosity function will be contaminated by pair-instability supernovae. \\bigskip" }, "1112/1112.0009_arXiv.txt": { "abstract": "We carry out adaptive mesh refinement (AMR) cosmological simulations of Milky-Way mass halos in order to investigate the formation of disk-like galaxies in a $\\Lambda$-dominated Cold Dark Matter model. We evolve a suite of five halos to $z=0$ and find gaseous-disk formation in all; however, in agreement with previous SPH simulations (that did not include a subgrid feedback model), the rotation curves of all halos are centrally peaked due to a massive spheroidal component. Our standard model includes radiative cooling and star formation, but no feedback. We further investigate this angular momentum problem by systematically modifying various simulation parameters including: (i) spatial resolution, ranging from 1700 to 212 pc; (ii) an additional pressure component to ensure that the Jeans length is always resolved; (iii) low star formation efficiency, going down to 0.1\\%; (iv) fixed physical resolution as opposed to comoving resolution; (v) a supernova feedback model which injects thermal energy to the local cell; and (vi) a subgrid feedback model which suppresses cooling in the immediate vicinity of a star formation event. Of all of these, we find that only the last (cooling suppression) has any impact on the massive spheroidal component. In particular, a simulation with cooling suppression and feedback results in a rotation curve that, while still peaked, is considerably reduced from our standard runs. ", "introduction": "\\label{intro} In cosmologies dominated by Cold Dark Matter (CDM), galaxy rotation is produced by gravitational tidal torques arising from the hierarchical collapse of structure. Analytic models and N-body simulations have shown that this can produce enough angular momentum to explain the observed sizes of disk galaxies \\citep{Fall:1980p1033, Mo:1998p1034}. However, computational models including gas dynamics have struggled to reproduce realistic disk galaxies. Such models initially produced undersized disks with low angular momentum \\citep{Navarro:1991p1002, Navarro:1994p1004}. Later work did generate disks with approximately the correct extent \\citep{Steinmetz:2002p1035, Abadi:2003p641}, but these had oversized stellar spheroidal components and therefore unnaturally large core circular velocities \\citep{Steinmetz:1999p635} far in excess of those found observationally. This \\emph{angular momentum problem} remains one of the major shortcomings of the CDM paradigm. The origin of the angular momentum problem is not entirely understood, but it probably stems from the fact that such simulations do not achieve sufficient spatial and mass resolution to correctly model the appropriate physical processes. For example, insufficient spatial resolution leads to the spurious mixture of hot and cold components of the ISM, producing an artificial warm component which is very efficient at radiating away energy \\citep{Katz:1992p1037, Katz:1996p1031, Steinmetz:1995p1005}. Thus, underresolved gas in such simulations can cool very quickly, and in cooling it loses its pressure support and collapses into dense knots of material. These knots interact with the galactic dark matter component through dynamical friction processes, and much of the angular momentum of the gas knots is transferred to the dark matter halo of a galaxy \\citep{Donghia:2006p1032}. Consequently, the knots tumble into the center of the galaxy to produce a dense cusp of material in the core of the simulated galaxy. Typically the buildup of these massive cores with cold gas and stars occurs rapidly, and even nascent simulated galaxies exhibit evidence of these cusps as early as $z \\sim 4$-5. Energetic feedback from star formation events has the ability, in theory, to alleviate this problem either by heating and ``puffing\" up the collapsing knots so they are more easily disrupted before losing their angular momentum \\citep{Weil:1998p1030}, and/or by preferentially ejecting low angular momentum gas \\citep[e.g.,][]{2010arXiv1010.1004B}. However, existing simulations are unable to resolve the detailed structures of star-forming regions. Individual star-forming and stellar feedback events occur on parsec scales. Stellar feedback processes can be resolved in simulations confined to local regions of the ISM \\citep[e.g.,][]{Joung:2006p1038} but not in cosmological simulations, which need to accurately co-evolve a galaxy's environment on scales of $ > 10$ Mpc. Therefore, cosmological simulations simplify and parameterize star formation and stellar feedback on scales more easily met by current computational resources (i.e. scales of $10^2 - 10^3$ pc). A variety of techniques have been suggested in order to achieve this puffiness within the confines of low-resolution models. In the most primitive prescription, stellar feedback is simply the return of energy from newly created star particles to their surrounding gas, usually in the form of thermal energy. Star ``particles'' in these simulations typically represent clusters of stars of mass $10^{2} - 10^{5}$ \\msun, so the thermal feedback is justified as the energy output from the most massive stars in the cluster becoming type II supernovae shortly after the creation of the star particle. In nature, this supernovae heating produces a small component of very hot gas surrounding the stellar population, so hot that its cooling time is very long. Unfortunately at the resolution currently achievable in cosmological simulations, this primitive thermal prescription for feedback deposits the same SN energy into a much larger reservoir of gas, which does not reach the same high temperature as it should. This now-warm component of gas can easily radiate the excess energy away, cool further and proceed with star-formation runaway thus defeating the purpose of the feedback \\citep[e.g.,][]{Steinmetz:1999p635}. Building from these failures, a number of research groups artificially turn off cooling in a gas parcel for a period of time ($t \\sim 10^7$ yr) after a cluster of stars has formed out of it \\citep[e.g.,][]{Gerritsen:1997p1039, Thacker:2000p1040, SommerLarsen:2003p1116, Stinson:2006p1023, Governato:2007p1022, Agertz:2010p461, Colin:2010p1053, Piontek:2011p1041, Guedes:2011p1080}. This method is justified as an application of the Sedov-Taylor blast wave solution for a Type II SN \\citep{Taylor:1950p1075, Sedov:1959p1074}, which blows out any cold media from the immediate environment of a star formation event. Using this prescription, any gas in a galaxy which starts to collapse into knots will reach the star formation criteria, form a star, and then heat up without any allowed cooling, thus preventing further collapse. Not surprisingly, these research groups have found some success with this method, yielding simulated galaxies with reduced inner rotation curves due to less massive bulge components; however, gas parcel masses and sizes in cosmological simulations of this sort are typically too large for the Sedov-Taylor solution to apply (see Section \\ref{code_coolingsuppression}). Thus despite the successes of the cooling suppression feedback model, the community continues to search for other more physically-motivated solutions. Another subgrid model for feedback (i.e. on scales smaller than the true resolution of the simulation) is to inject kinetic energy directly into the gas; this can alleviate the problem of thermal energy being radiated away. For example, some studies \\citep[e.g.][]{Springel:2003p1044, Scannapieco:2006p1118, Oppenheimer:2008bu} using Smoothed Particle Hydrodynamics (SPH) give some of the SN energy to individual gas particles in the form of momentum. This method can result in significant mass outflows (by design), but at the cost of decoupling wind particles from hydrodynamic interaction for a period of time. An alternate approach, to keep wind particles coupled to the disk gas was explored by \\citet{Schaye:2008p1045}. Both approaches help but, by themselves, do not appear to generate realistic rotation curves. In addition, \\citet{Truelove:1997p1046} showed that insufficient resolution in a simulation can lead to artificial fragmentation of the gas, perhaps resulting in a further overproduction of stars. One way to prevent artificial fragmentation is to add additional (numerical) pressure in high-density, low-temperature regions to ensure that the Jeans length is always resolved \\citep{Machacek:2001p1047, Robertson:2008p1017}. This can be achieved by modifying the equation of state (EOS) itself, making it stiffer in order to provide an additional source of pressure to gas in denser regions \\citep{Schaye:2008p1045, Ceverino:2009p1014, Agertz:2010p461}. A polytropic EOS ($P \\propto \\rho^{\\Gamma}$) with $\\Gamma = 4/3$ will keep the ratio of Jeans length to resolution length constant (assuming Lagrangian resolution such that the resolution length decreases as $\\rho^{-1/3}$ -- for fixed resolution $\\Gamma = 2$ is required), but even stiffer relations have been used. For example, \\citet{Agertz:2010p461} ran simulations with such an equation of state, where in low-density regions it behaved as an ideal gas, but in high-density (star forming) regions it followed a polytropic equation of state with $\\Gamma = 2$. In this paper, we undertake an investigation of galaxy formation using an Adaptive Mesh Refinement (AMR) hydrodynamics code. The majority of work in this field has used SPH codes, and so this allows us to investigate the problem from a new angle. Although there has been some work with AMR codes \\citep{Joung:2009p1010, Ceverino:2009p1014, Agertz:2010p461, Colin:2010p1053}, there has not been a clear demonstration that an equivalent AMR calculation (i.e. one without a subgrid feedback model) actually does reproduce the classic SPH result. We begin by simulating a set of five halos without any feedback or subgrid model (except a minimum pressure support to prevent artificial fragmentation). We find, in agreement with SPH codes that a large, concentrated bulge is produced, resulting in a rotation curve that rises to $\\sim 500$ km/s at $r \\sim 1$ kpc. We then vary a number of numerical and physical parameters in order to understand how sensitive the result is to our a choice of parameters. The paper is organized as follows. Section \\ref{method} describes the details of our hydrodynamics code, our initial conditions and the relevant parameters for this study. In Section \\ref{results}, we present the results of our simulations including the five canonical runs, our resolution study, and our modified runs. Section \\ref{discussion} is a discussion of our results and their implications. Finally, Section \\ref{summary} summarizes our conclusions and makes predictions for future solutions to the angular momentum problem. ", "conclusions": "\\label{discussion} These results demonstrate that it is challenging to generate disk systems with the correct mass distribution. Without effective feedback, the default outcome is for dense clumps to lose angular momentum and result in centrally-cusped rotation curves. In particular, the simulation with a very low star formation efficiency nicely demonstrates that this result is fundamentally a dynamical one, and does not depend on whether the clumps are primarily gas, or mostly stars. As long as they are concentrated, they will lose angular momentum and hence rotational support. Although this result is not new, and there is a long history of SPH simulations which found this result much earlier, we show it here clearly and systematically using a completely different numerical method (AMR). Therefore, the result is quite general. In addition, we went on to systematically vary our numerical parameters and investigate a range of resolution and feedback methods. We confirmed that only cooling-suppression feedback models are capable of significantly changing the mass distribution and hence the rotation curve. Cooling suppression models do effectively enhance feedback, although it is unclear how physically meaningful this technique is (see Section \\ref{code_coolingsuppression}), and a better approach might be to use high-resolution local models to generate subgrid models \\citep[see][for some attempts in this direction]{Yepes:1997p1093, Tasker:2006p1072, Ceverino:2009p1014, Joung:2009p1094}. Another constraint is the baryon content of galactic halos: a variety of techniques have been used to infer that the baryon-to-dark-matter ratio in galaxies is much smaller than the cosmic mean \\citep[e.g.][]{Moster:2010p1095, Behroozi:2010p1096}, implying that a significant amount of mass has been ejected from galactic systems (or never accreted in the first place). For example, Milky-Way massed halos only appear to host 20\\% of their baryons, with the fraction decreasing rapidly for smaller-mass systems \\citep{Behroozi:2010p1096}. We find that all of our simulations result in very high disk baryon fractions; even the run with feedback and cooling suppression has a baryon fraction of about 63\\% of the cosmic mean. This appears to be a general issue with cosmological galaxy simulations \\citep[see also the discussion in][]{AvilaReese:2011p1097}. We demonstrate that simulations must be run to $z\\sim0$ in order to gauge the efficacy of model parameters at minimizing the effects of the angular momentum problem. Many past studies \\citep[e.g.][]{Ceverino:2009p1014, Joung:2009p1010} have traded off simulation run time for increased resolution in their simulations. While there are some early indicators of the angular momentum problem like bursts of star formation and peaked rotation curves at redshifts as early as $z=5$, successful results during this epoch do not guarantee successful results at $z\\sim0$. We specifically demonstrate this in the case of the low star formation efficiency run H26SPML, which staved off early bursts of star formation but eventually succumbed to the same fate as runs with a normal star formation efficiency. \\subsection{Comparison to Previous Work} In agreement with previous work using SPH, we find that unless we include efficient feedback, the resulting systems are dominated by a too-large spheroidal component, and the resulting rotation curve is peaked in the center \\citep[e.g.][]{Navarro:1991p1002, Navarro:1997p1098, Weil:1998p1030, Abadi:2003p641, Donghia:2006p1032, Zavala:2008p1099, Piontek:2011p1041}. \\begin{figure*} \\begin{center} \\plotone{figure14a.png} \\plotone{figure14b.png} \\end{center} \\caption{We render faceon and edgeon views for the last four of our modified halos at $z = 0$. The top two rows show each galaxy in gas, whereas the bottom two rows display each galaxy's stellar component. Each postage stamp has a width of 25 kpc. For the gas, edgeon views show a volume rendering with isocontours of the gas density, whereas faceon views are column density bricks. For the stars, we use the column density of stars in the faceon and edgeon views respectively. The color of a star particle is representative of its age where blue represents young stars and red represents old stars according to the color bars on the right. Notably, these modified runs produce young stellar disks and have a much higher density of gas in their cores than the canonical runs.} \\label{fig:modified_runs2_renderings} \\end{figure*} A number of reasons have been suggested for this in the past, including purely numerical causes, such as angular momentum transfer between the disk and the hot halo \\citep[e.g.][]{Okamoto:2005p1101}, or between cold gas clumps and the hot halo \\citep[e.g.][]{Thacker:2000p1040}. The concern was particularly that SPH simulations might be susceptible to this issue because of smoothing between hot and cold phases. However, the fact that AMR simulations -- which use a completely different numerical method to solve the fluid equations -- find the same result, is an indication that these effects do not dominate. Another suggested source of numerical angular momentum loss is in the form of gravitational instabilities which arise from inadequate resolution of the Jeans length in the disk \\citep{Truelove:1997p1046, Robertson:2004p1102}. We tested this idea by running with and without an additional numerical (``Jeans\") pressure designed specifically to ensure that the Jeans length was adequately resolved, finding no difference in our results. A third numerical reason is the lack of resolution \\citep{Governato:2004p1024, Kaufmann:2007p1103}; however, we specifically test this over the computational range available for us, and find no significant difference from 200-1700 pc. This is in agreement with the SPH simulations of \\citet{Piontek:2011p1041}, who also varied their numerical resolution over a wide range, and found no difference. There have been a number of recent cosmological AMR simulations which we can compare to. \\citet{Colin:2010p1053} used the \\emph{ART} code \\citep{Kravtsov:1997p1104} to simulate a halo which is smaller by an order of magnitude (about $10^{11}$ M$_\\odot$), finding peaked rotation curves (decreasing as the star formation density criterion was reduced). Although the halo masses are quite different, the essential result seems to be in agreement. \\citet{Agertz:2010p461} used the \\emph{RAMSES} AMR code \\citep{Teyssier:2002p1105} to simulate a Milky-Way mass galaxy with similar resolution to that found here. They argued that a low star formation efficiency by itself was enough to produce nearly flat rotation curves (and that feedback was only efficient if extreme amounts of energy were injected). We have not been able to confirm the first suggestion -- using a range of low efficiencies for star formation, we find that clumps lose angular momentum at high-redshift, and generate steep rotation curves, whether they are in gas or stellar form. The efficiency controls when the gas is converted to stars, but has little impact on the distribution of the material. In this, we are in agreement with previous SPH work \\citep[e.g.][]{Weil:1998p1030, Donghia:2006p1032, Piontek:2011p1041}. Finally, we have found that the only way to significantly decrease the peak of the rotation curve was to introduce a sub-grid model which enhanced the efficiency of stellar feedback. We briefly explored the cooling suppression model and found this to be effective. This agrees with a substantial number of SPH simulations which adopt this mechanism \\citep{Gerritsen:1997p1039, Thacker:2000p1040, SommerLarsen:2003p1116, Stinson:2006p1023, Governato:2007p1022, Agertz:2010p461, Colin:2010p1053, Piontek:2011p1041, Guedes:2011p1080}. In addition, \\citet{AvilaReese:2011p1097} used AMR simulations (with the \\emph{ART} code) and also found cooling suppression to be effective in obtaining approximately flat rotation curves. \\citet{Ceverino:2009p1014} also used the \\emph{ART} code, but with a different sub-grid model, arguing that a model in which stars are born with significant velocities relative to the nascent gas will feed energy into low-density regions, producing efficient feedback and flat rotation curves, although the simulation is only run to $z \\sim 3$." }, "1112/1112.1419_arXiv.txt": { "abstract": "We obtained $\\rm{U_{330}}$ and B band images of the M31 nucleus using the High Resolution Camera of the Advanced Camera for Surveys on board the Hubble Space Telescope (HST). The spatial resolution in the $\\rm{U_{330}}$-band, $0\\asec03$ FWHM, or 0.1 pc at M31, is sufficient to resolve the outskirts of the compact cluster (P3) of UV-bright stars surrounding the M31 black hole. The center of the cluster is marked by an extended source that is both brighter and redder than the other point sources within P3; it is likely to be a blend of several bright stars. We hypothesize that it marks the location of the M31 black hole. Both stellar photometry and a surface brightness fluctuation analysis, show that the P3 stellar population is consistent with early-type main sequence stars formed in a $\\sim100~-~200$ Myr old starburst population. Evolutionary tracks of post early asymptotic giant-branch stars, associated with late-stage evolution of an old population, also traverse the U and ${\\rm U-B}$ domain occupied by the P3 stars; but we argue that only a few stars could be accounted for that way. PEAGB evolution is very rapid, and there is no progenitor population of red giants associated with P3. The result that P3 comprises young stars is consistent with inferences from earlier HST observations of the integrated light of the cluster. Like the Milky Way, M31 harbors a black hole closely surrounded by apparently young stars. ", "introduction": "There is a highly-compact cluster of blue stars \\citep{l98, brown} at the heart of the M31 double nuclear-cluster of old stars. \\citet{kb} showed that this blue cluster in turn hosts the $\\sim10^8M_\\odot$ super-massive black hole also known to reside in the nucleus \\citep{d84, dr, k88, r90}. \\citet{bend} argue that the integrated spectrum and spectral-energy distribution of the stars are consistent with their formation in a burst 200 Myr ago. Young stars are also closely bound to the nuclear black hole in our galaxy \\citep{forrest, allen, krabbe}, suggesting that this may not be a rare phenomenon. Understanding how apparently young stars were formed deeply interior to the tidal field of a supermassive black hole probes several issues of how stars and gas interact within such an extreme environment (see \\citealt{alex} for a review). Study of the blue cluster in M31 provides an additional context in which to test theories for the formation of the unusual population surrounding the Milky Way black hole. There is a diversity of ideas for explaining the origin of the M31 blue cluster. \\citet{chang}, for example, argue that the cluster is the expected consequence of the non-axisymmetric structure of the surrounding M31 nucleus funneling gas from stellar mass-loss into orbit around the black hole, where it periodically reaches surface densities sufficient to induce collapse and star formation. Likewise, \\citep{levin}, \\citet{br}, and \\citet{ward} have advanced the formation of compact disks of young stars around the Milky Way black hole by massive central accretion of gas. In contrast, \\citet{demar} argue that the blue stars in M31 result from stellar collisions and tidal stripping acting on an old population of stars interacting within the high-velocity orbits and strong tidal-field associated with the black hole. We present new observations of the blue cluster obtained at the maximum angular resolution offered by the {\\it Hubble Space Telescope} to provide better constraints on the origin of the stars residing in this unusual environment. \\subsection{The Discovery of the Blue Cluster} The discovery of the blue cluster proceeded in stages. \\citet{king} imaged the nucleus of M31 with the HST Faint Object Camera at 1750\\AA, seeing a double structure similar to that discovered in the optical by \\citet{l93}. By the simple expedient of checking the position-angle of their images, however, they discovered that the optically-dimmer peak, designated P2 by Lauer et al., was actually brighter in the UV than P1. \\citet{king} lacked $S/N$ sufficient to resolve the spatial structure of the UV source at P2, noting only that it was highly-compact and possibly consistent with a point source, such as low-level emission associated with the weak AGN detected in the radio \\citep{crane}, or emission from a single PAGB star, such as those seen further out from the nucleus \\citep{k92,bert}. In retrospect, the discovery of a UV-bright nuclear source explains the observations of \\citet{nieto}, who found the nucleus to have color gradient becoming {\\it bluer} towards its center, using CFHT images obtained at 3750\\AA. Discovery of the cluster may also have been presaged by an enigmatic reference in \\citet{red} to third-party unpublished ``Mt. Wilson material'' purporting to show that ``the spectrum of the nucleus is of a peculiar dwarf A type,'' a result supported by no other work prior to \\citet{bend} to the best of our literature research.\\footnote{We have also been unable to find photographic spectra of the M31 nucleus prior to 1937 in catalogues of Mt.\\ Wilson observations. We thank Dr. Christopher Burns of the Carnegie Observatories for searching for this material.} Later HST observations by \\citet{l98} and \\citet{brown} were able to resolve the UV-source; both papers argued that it is a cluster of stars. WFPC2/PC images at 3000\\AA\\ obtained by \\citet{l98} showed that the half-power radius of the blue source is $\\sim0.2$ pc. \\citet{l98} further combined their U and V band fluxes of the source with the \\citet{king} fluxes at 1750\\AA\\ to conclude that the source is consistent with an A-star spectral energy distribution. \\citet{bend} refined this picture further by obtaining HST STIS spectroscopy of the nucleus over $\\lambda\\approx3600-5100$\\AA, and reanalyzing the F300W images presented in \\citet{l98}. The spectroscopy shows strong Balmer lines, including a strong Balmer break, consistent with either A0 main-sequence or giant stars, but not a population of white dwarfs of the same photospheric temperature. The best match to the spectrum is provided by a $200\\pm50$ Myr population formed in a single burst. Populations of less than half this age would exhibit too much UV-flux to the blue side of the Balmer break. \\citet{bend} further show that the stars are distributed in a flat disk with an exponential profile of scale-length of $0.37\\pm0.04$ pc in rapid rotation about the black hole.\\footnote{We use 770 Kpc as the distance to M31 \\citep{dm31}.} The rotational broadening of the STIS spectrum implies a black hole mass of $M_\\bullet=1.4\\times10^8M_\\odot,$ a significant upwards revision from earlier determinations. \\subsection{The Blue Cluster in the Context of the M31 Nucleus} The M31 black hole and its cluster of blue stars mark the center of a much larger nuclear star-cluster of complex structure. The surface brightness of the nuclear stellar system begins to rise above the underlying M31 bulge interior to $r\\sim3''$ or $\\sim10$ pc from the center. The cluster becomes increasingly elliptical at smaller radii, with $1-b/a >0.3$ for $r<1\\asec7,$ but its isophotes remain concentric with the photo-center of the bulge for $r>1\\asec4$ \\citep{l93}. At yet smaller radii the nuclear cluster exhibits the double-peaked structure discovered by \\citet{l93}. The optically-brighter peak, P1, is offset by $0\\asec49$ from the secondary brightness peak, P2, and a slightly lesser amount from the M31 bulge photo-center. The best explanation for the double morphology is that both peaks result from the line-of-sight projection of an eccentric disk of stars roughly $\\sim2$ pc in diameter bound to the black hole \\citep{trem}. The black hole strongly dominates the potential, thus the stars in the disk follow Keplerian orbits, spending most of their time at the orbital apo-center, creating an enhancement of light at P1. Significantly, the P1---P2 line twists away from the major axis of the outer nucleus by $\\sim20^\\circ.$ P1 and P2 are both redder in $V-I$ than the surrounding nucleus \\citep{l98}. High spatial-resolution spectroscopy shows that both P1 and P2 comprise old stellar populations with characteristics more like each other, than the underlying bulge \\citep{kb}. The ``Tremaine Disk'' thus appears to be a distinct component embedded in a much larger nuclear-cluster. The initial model of the disk provided by \\citet{trem} has been refined by the more detailed analysis presented by \\citet{pt03}. An interesting feature of the refined model is that the disk requires a central hole in order to generate the apparent minimum of stellar emission that occurs between P1 and P2. The cluster of blue stars occurs at still smaller radii, $r<0\\asec1,$ and must be closely bound to the black hole, which was first inferred to exist from ground-based spectroscopy \\citep{d84, dr, k88, r90}. \\citet{kb} carefully registered their high spatial-resolution spectroscopy to the WFPC2 imagery of \\citet{l98}, concluding that the M31 black hole was coincident with the cluster and that both were within $0\\asec07$ of the M31 bulge photo-center. The conclusion that the cluster hosts the black hole was further confirmed by \\citet{bend}, who used STIS spectroscopy to show that the central velocity dispersion in the cluster rises to $1183\\pm200~\\rm{km~s^{-1}},$ with organized rotation peaking at $618\\pm81~\\rm{km~s^{-1}}.$ Building on the analysis of \\citet{kb}, they reduced the offset of the black hole from the bulge center to $0\\asec033$ in the ``anti-P1'' direction along the P1-P2 line. They further demonstrated that this slight offset is balanced by the asymmetry of the Tremaine disk --- the center of mass of the complete nuclear system is coincident with the bulge photo-center to $<0\\asec01.$ \\citet{bend} also designated the cluster as a third component, P3, of the inner nucleus. This designation includes a subtle redefinition of P2 as it previously had been introduced by \\citet{l93} and used by \\citet{l98}. The center of the cluster is coincident with the peak of a shallow surface-brightness cusp seen in V-band that falls within $0\\asec1$ of the bulge center; it was this location that the two Lauer et al.\\ papers adopted as the center of P2. \\citet{bend}, however, use P2 to denote the elongated and more diffuse concentration of older stars that extends from the cluster on the side opposite from P1. In this schema, P1 and P2 explicitly correspond to the apo- and peri-center apses of the Tremaine disk, which has a mean peri-center that is {\\it not} coincident with the cluster. The present environment of P3 appears to be largely quiescent, despite the immediate proximity of the black hole \\citep[see the extensive discussion of this topic by][]{li}. \\citet{crane} find a weak radio source coincident with P3, and \\citet{garcia, garcia2} find low-level X-ray emission associated with P3, as well \\citep[however, see][]{li}. On the other hand, a high spatial-resolution map of emission from ionized gas within in the nucleus finds a few sources within an arcsecond of P1, but nothing coincident with P2 or P3 \\citep{delb}. While central accretion of gas may have created the P3 stars in the past, there is no evidence that any significant reservoir of cold gas exists there at present. \\subsection{A Closer Look\\dots} Under the \\citet{bend} description of P3, the cluster is a disk comprising only a few hundred A-stars. Given the scale length of the disk, we concluded that it should begin to show resolution into discrete sources in images with slightly higher angular resolution than were obtained by \\citet{l98}, such as could be provided by the HST ACS/HRC in the blue. The photometry of the brighter sources might be obtained directly, but in any case, the full image of the cluster could also be analyzed by surface-brightness fluctuation (SBF) analysis \\citep{sbf} to constrain the P3 stellar population. If the cluster indeed comprises main sequence A stars, it would look ``lumpy'' in HRC images. SBF could quantify the amplitude of the lumpiness, and in turn the typical luminosity of the stars that the cluster comprises. ", "conclusions": "We began the paper with a history of earlier work on P3, which already built a strong case that it is a cluster of young stars. We have now partially resolved the cluster into individual stars, and characterized its form with significantly higher spatial resolution than was provided by the WFPC2/PC observations of \\citet{l98}; however, in broad detail the conclusions of that paper, and the later work of \\citet{bend} remain unaltered. Both a color-magnitude diagram (CMD) obtained of the resolved stars, and an SBF analysis performed on the total image, show that P3 plausibly comprises a population formed in a burst 100 to 200 Myr ago, consistent with the P3 SED and the spectrum obtained by \\citet{bend}. The result may also be supported by the \\citet{sag} spectroscopy observation of the inner bulge, which suggest that a burst of star formation occurred within the nucleus about $\\sim100$ Myr ago. One small modification of the earlier results provided by the present work is that that P3 may be slightly younger than the 200 Myr age derived by \\citet{bend}, but that work really only ruled out an age much younger than 100 Myr. Another new result is that the color of S11 suggests that a single-age starburst model for the cluster does not capture the full story of P3. At some level, the large luminosity of P3 blends with the much bluer stars around it to generate an overall A-like SED. It is intriguing that S11 provides the only visible trace of P3 in the V-band. If it does mark the location of the M31 black hole, then it suggests that there may yet be another change in the stellar population of the nucleus within $\\sim0\\asec03$ or $\\sim0.1$ pc of the black hole. Although we have concluded that the P3 stars are young, \\citet{king} and \\citet{brown} argue that the UV-bright stars in the bulge of M31, which of course surrounds the nucleus, represent the final stages of stellar evolution in an old metal-rich stellar population. If there are truly two different populations contributing UV-bright stars to the bulge and nucleus of M31, it still begs the question of whether or not old remnants may be present in P3, as well as how far out the putative young stars in P3 may be traced out into the surrounding nucleus or even bulge. We show that PEAGB tracks traverse the U vs.\\ ${\\rm U-B}$ domain defined by the P3 stars, thus it is possible that a few of the P3 stars are indeed old-evolved stars, rather than being young massive main sequence stars. However, since this is a very rapid evolutionary phase, it is extremely unlikely that the entire P3 cluster could be accounted for by PEAGB stars. A strong factor in constraining the population of P3 is the compactness of P3, itself. The UV-stars outside the nucleus do trace the bulge-light at long wavelengths, but yet are not strongly concentrated within the outer nucleus; their abundance by total light or number is but a small faction of the bulge. Requiring any sort of HP-HB stars as the dominant P3 population requires an associated massive population of red giants also closely bound to the M31 black hole to continually generate the short-lived PEAGB stars. There is no evidence for such a population in the V and I nuclear images of \\citet{l98}. This same consideration would appear to also apply to the hypothesis of \\citet{demar}, who suggest that the P3 UV-bright stars may be the remnant cores of giants stripped by close encounters with the black hole. The high velocities associated with the black hole are such that stars passing close to the black hole from the bulge, or even the outer nucleus will not dwell in the vicinity of P3. If the P3 stars are really processed giants, they still must have been initially closely bound to the black hole, which again begs the question of where the progenitor giants are. The only way that this mechanism might work is if the stripping of the P3 stars takes place just as they begin their first ascent up the red giant branch. This might produce long-lived UV-bright stripped-cores, and would account for lack of progenitor giants as well. Further investigation of the origin of P3 might be linked to understanding the early-type stars closely bound to the black hole in our own galaxy. The existence of P3 in a second Local Group galaxy suggests that this is not a rare phenomenon. The present observations offer an additional site to test mechanisms that can form stars within the strong tidal field of a supermassive black hole. We finish by noting that as with the galactic center, it may be possible to measure the proper motions of stars closely bound to the M31 black hole. Figure \\ref{fig:stars} shows that circular velocities in the plane of the sky around the M31 black hole will exceed $1000~{\\rm km~s^{-1}}$ over the entire extent of P3. It is possible that proper motions may be detected in about a decade after the present observations were made." }, "1112/1112.4716_arXiv.txt": { "abstract": "We show that, even for extremely stiff systems, explicit integration may compete in both accuracy and speed with implicit methods if algebraic methods are used to stabilize the numerical integration. The required stabilizing algebra depends on whether the system is well-removed from equilibrium or near equilibrium. This paper introduces a quantitative distinction between these two regimes and addresses the former case in depth, presenting explicit asymptotic methods appropriate when the system is extremely stiff but only weakly equilibrated. A second paper \\cite{guidQSS} examines quasi-steady-state methods as an alternative to asymptotic methods in systems well away from equilibrium and a third paper \\cite{guidPE} extends these methods to equilibrium conditions in extremely stiff systems using partial equilibrium methods. All three papers present systematic evidence for timesteps competitive with implicit methods. Because an explicit method can execute a timestep faster than an implicit method, algebraically-stabilized explicit algorithms might permit integration of larger networks than have been feasible before in various disciplines. ", "introduction": "In many scientific and technical contexts one encounters phenomena that may be modeled by fluxes transferring population between sources and sinks for various species. Examples include kinetic processes that modify abundances and transfer energy in atomic, molecular, and nuclear systems; geochemical, climate, and other environmental systems; electrical circuits; economic models; and population dynamics. Terminology varies but let us refer generically to these sources and sinks as {\\em boxes,} and term the resulting systems of boxes connected by fluxes {\\em reaction networks.} Such systems are commonly modeled by a coupled set of differential equations that describe a continuous flow of population through the boxes. The reaction network is often classified as a {\\em stiff system}, which we shall define to be a system of equations containing multiple timescales ranging over many orders of magnitude \\cite{gear71,lamb91,press92,oran05}. Most physical systems involve important processes operating on very different timescales, so realistic problems tend to be at least moderately stiff. Some, such as those encountered in many astrophysics applications, are extremely stiff, with fastest and slowest timescales in the problem differing by as much as 10--20 orders of magnitude. In stiff systems the timestep constraints are set by numerical stability requirements rather than accuracy considerations. Hence, explicit numerical integration of stiff systems is usually impractical because the maximum stable timestep is far too small for efficient solutions (see, for example, Refs.\\ \\cite{press92,oran05}). This is commonly addressed by employing implicit or semi-implicit stiff solvers that are stable, but that require time-consuming iterative matrix solutions. A given box in a reaction network often is connected strongly only to a few other boxes. For example, the explosive burning conditions encountered in astrophysical novae, X-ray bursts, or supernovae may require reaction networks with hundreds to thousands of nuclear isotopes. Yet individual isotopes are typically connected directly to other isotopes through (at most) $\\sim$10 reactions of consequence, and under many conditions no more than 2--3 reactions are important for a given isotope. Such restrictions on the direct box reaction coupling imply that the matrices appearing in the iterative implicit solution are {\\em sparse.} Although various methods are available to deal with sparse matrices, in practice many codes for solving large reaction networks have not exploited sparseness in particularly effective ways. For example, in astrophysical calculations with implicit solvers in large networks (say $\\sim 150$ species or more), one finds often that greater than 90\\% of the processor time is consumed in matrix operations \\cite{timmes,hix05}. Efficient algorithms exist for the required matrix algebra (with incremental improvements in them over time), but the matrix nature of the core problem implies that the time required for implicit solution grows non-linearly with the size of the network. In typical working codes for large-scale applications, increasing the size of the network increases the time for solution, often quadratically, sometimes as much as cubically, until there are enough boxes in the network to justify the overhead of sparse-matrix methods with more favorable scaling. In applications in thermonuclear networks, for example, it is often found that the overhead required to implement sparse-matrix iterative solutions is not justified until there are several hundred boxes in the network. Thus, many present implicit stiff-network algorithms do not scale very gracefully to larger networks. We are primarily interested in the most ambitious applications of large networks, where the reaction network is only a portion of a larger problem. Let us take as representative astrophysical thermonuclear reaction networks, where a proper description of the overall problem typically requires multi-dimensional hydrodynamics or radiation hydrodynamics coupled tightly to a large thermonuclear reaction network. The hydrodynamical evolution controls the conditions in the network such as temperature and density, and the network influences the hydrodynamic evolution strongly through energy production and modification of composition variables. As a consequence of the limitations discussed in the preceding paragraphs, the solution of large networks by the usual approaches is time-consuming and few calculations have attempted to couple the element and energy production strongly to the hydrodynamics with a network of realistic complexity. The most ambitious approaches use very small networks, perhaps tuned empirically to get critical quantities like energy production correct on average, coupled to the hydrodynamical simulation. In many calculations even this is not done and the network is replaced entirely by parameterization. Then a more complete network is run in a separate ``post-processing'' step, where fixed hydrodynamical profiles computed in the hydrodynamical simulation with the small network are used to specify the variation of thermodynamic variables such as temperature and density with time. Astrophysical thermonuclear networks have been used for illustration, but many problems of scientific and technical interest exhibit similar complexity. Examples include astrochemical kinetics, where one must model large chemical evolution networks in contracting molecular clouds, or combustion chemistry, where chemical burning networks are strongly coupled to simulations of the dynamics of the air and fuel mixture. Physically-realistic networks in such contexts would often be quite large. In combustion of larger hydrocarbon molecules or studies of soot formation, hundreds to thousands of reacting species undergoing as many as 10,000 reactions may be encountered \\cite{oran05}, and in supernova explosions hundreds to thousands of nuclear isotopes with tens of thousands of reaction couplings make non-zero contributions \\cite{hix05}. For such cases one finds that current techniques do not allow for a coupling of realistic reaction networks to the full dynamics of the problem and often severely truncated or highly schematic reaction networks have been used in even the most realistic simulations. ", "conclusions": "Explicit numerical integration can compute a timestep faster then implicit methods, and the time to compute a network explicitly scales linearly and therefore more favorably with network size than for implicit codes. Nevertheless, previous discussions of numerical integration for very stiff systems have concluded rather uniformly that explicit methods are not competitive with implicit methods for stiff networks because they are unable to take large enough stable timesteps. To quote {\\em Numerical Recipes} \\cite{press92}, ``For stiff problems we {\\em must} use an implicit method if we want to avoid having tiny stepsizes.'' Improvements in explicit methods based on using asymptotic and steady-state limiting solutions to remove stiffness from the network have had some success for systems of moderate stiffness such as various chemical kinetics problems. However, it has been concluded in the previous literature that such methods are not competitive, failing even to give correct results, with timesteps that are far too short to be useful even if they gave correct results, for extremely stiff networks such as those encountered commonly in astrophysical thermonuclear networks \\cite{oran05,mott99}. This paper has presented evidence strongly challenging all of these conclusions. We have cleanly identified three fundamentally different sources of stiffness in large networks, only the first of which is commonly emphasized in the literature: \\begin{enumerate} \\item Situations where small populations can become negative if the explicit timestep is too large, with the propagation of this anomalous negative population leading to exponentially growing terms that destabilize the network. \\item Situations where the right sides of the differential equations expressed as $dY = F = \\fplus{} - \\fminus{}$ approach a constant derived from the difference of two large numbers (the total flux in $\\fplus{}$ and total flux out $\\fminus{}$), and numerical errors in taking this difference destabilize the network if the timestep is too large. \\item Situations where on the right sides of the differential equations expressed in the form of Eq.~(\\ref{equilDecomposition}) the net flux in specific forward-reverse reaction pairs $(f^+_i - f^-_i)$ tends to zero as the system approaches equilibrium, leading to large errors if the timestep is too large because the net flux is derived from the difference of two large numbers and the timescale equilibrating the populations is short compared with the desired numerical timestep. \\end{enumerate} We have shown that these distinctions are important because different sources of stiffness require different approximations for their removal in an algebraically-stabilized explicit integration. Using the extremely stiff systems characteristic of astrophysical thermonuclear networks as a stringent test, we have shown that asymptotic methods are very successful at removing the first two types of stiffness, and give correct results, even for the stiffest of thermonuclear networks, provided that adequate attention is paid to conservation of probability in the network. Furthermore, we have shown various examples of stable and accurate timestepping with these methods in extremely stiff systems that are competitive with that of standard implicit codes, demonstrating in some simple but physically-important networks timesteps that are as much as 20 orders of magnitude larger than the maximum timestep that would be stable in a standard explicit method. Asymptotic methods are adept at removing the first two types of stiffness listed above, permitting explicit numerical timesteps that are competitive with implicit methods even in the stiffest networks. However, we have also shown that such methods give correct results but fail to exhibit competitive timestepping when the system approaches microscopic equilibrium and the third type of stiffness instability begins to dominate. In a following paper \\cite{guidPE}, we shall provide evidence for competitive timestepping, even in the approach to equilibrium, if the explicit asymptotic method is supplemented by partial equilibrium approximations designed specifically to deal with the third type of stiffness instability. Taken together, this paper and the following ones on quasi-steady-state methods \\cite{guidQSS} and partial equilibrium methods \\cite{guidPE} present compelling evidence that algebraically-stabilized explicit integration methods are capable of timesteps competitive with implicit integration methods for a variety of highly-stiff reaction networks. Since explicit methods can execute a timestep faster than an implicit method in a large network, our results suggest that algebraically-stabilized explicit algorithms may be capable of performing as well as, or even substantially outperforming, implicit integration in a variety of moderate to extremely stiff applications. Because of the highly-favorable linear scaling for explicit methods, this fundamentally new view of the efficacy of explicit integration for stiff equations may be particularly important for applications in any field where it is imperative that more realistic---and therefore much larger---networks be used in complex physical simulations. \\begin{ack} We thank Tony Mezzacappa for useful discussions, Austin Harris for help with some of the calculations, Eric Lingerfelt for programming assistance, and Christian Cardall for a careful reading of the manuscript. Research was sponsored by the Office of Nuclear Physics, U.S. Department of Energy. \\end{ack} \\clearpage" }, "1112/1112.3300_arXiv.txt": { "abstract": "We measure the gas-phase oxygen abundances of $\\sim3000$ star-forming galaxies at $z=0.05-0.75$ using optical spectrophotometry from the AGN and Galaxy Evolution Survey (AGES), a spectroscopic survey of $I_{\\rm AB}<20.45$ galaxies over $7.9$~deg$^{2}$ in the NOAO Deep Wide Field Survey (NDWFS) \\bootes{} field. We use state-of-the-art techniques to measure the nebular emission lines and stellar masses, and explore and quantify several potential sources of systematic error, including the choice of metallicity diagnostic, aperture bias, and contamination from unidentified active galactic nuclei (AGN). Combining volume-limited AGES samples in six independent redshift bins and $\\sim75,000$ star-forming galaxies with $r_{\\rm AB}<17.6$ at $z=0.05-0.2$ selected from the Sloan Digital Sky Survey (SDSS) that we analyze in the identical manner, we measure the evolution of the stellar mass-metallicity (\\mz) between $z=0.05$ and $z=0.75$. We find that at fixed stellar mass galaxies at $z\\sim0.7$ have just $30\\%-60\\%$ the metal content of galaxies at the present epoch, where the uncertainty is dominated by the strong-line method used to measure the metallicity. Moreover, we find no statistically significant evidence that the \\mz{} relation evolves in a mass-dependent way for $\\mass\\simeq10^{9.8}-10^{11}$~\\msun{} star-forming galaxies. Thus, for this range of redshifts and stellar masses the \\mz{} relation simply shifts toward lower metallicity with increasing redshift without changing its shape. ", "introduction": "\\label{sec:intro} Like stellar mass, the gas-phase metallicity of a galaxy is a sensitive observational diagnostic of its past star formation history and present-day evolutionary state for the simple reason that in a closed system metallicity increases monotonically with each successive generation of massive stars.\\footnote{Throughout this paper we use the terms {\\em metallicity} and {\\em abundance} interchangeably to mean the heavy-element content of the warm ($T\\approx10^{4}$~K) interstellar medium of galaxies. In particular, we make the reasonable assumption that the nebular oxygen abundance, historically written as \\logoh, traces the total gas-phase metallicity, $Z_{\\rm gas}$.} In detail, however, galaxies are not closed systems: infall of cold, metal-poor gas from the intergalactic medium, rapid gas accretion via minor and major mergers, and supernova-driven winds of gas and metals can modulate the metallicity of individual galaxies according to their large-scale environment, gas supply, and assembly history. Therefore, accurate abundance measurements provide valuable insight into the interplay between many fundamental processes in galaxy evolution, including star formation, gas accretion, and supernova-driven feedback across cosmic time \\citep{tinsley80a, pagel97a, pettini04b}. Nebular emission lines---ubiquitous in the rest-frame optical spectra of star-forming galaxies---provide a particularly powerful way to study the chemical abundances of both nearby and distant galaxies. Among the most commonly observed lines are the Balmer \\halam, \\hblam, and \\hglam{} hydrogen recombination lines, and the collisionally excited \\oiidoublet, \\neonlam, \\oiiidoublet, \\niidoublet, and \\siidoublet{} forbidden lines. Because these lines originate principally in star-forming (\\hii) regions, they trace the physical conditions in the gas from which the current generation of massive stars is forming. The metal lines in particular are the principal coolants in \\hii{} regions, making them sensitive to the total abundance of oxygen and other heavy elements in the interstellar medium. Moreover, the relative line-strengths can be used to infer the interstellar pressure, density, temperature, ionizing radiation field strength, dust reddening, and the presence of an active galactic nucleus (AGN). Finally, from an observational standpoint, many of the nebular lines are intrinsically strong, making them measurable in even relatively low signal-to-noise spectra over a broad range of redshifts. Following the discovery of the correlation between dynamical mass and oxygen abundance in dwarf irregular galaxies (\\citealt{lequeux79a}; see also \\citealt{garnett87a, oey93a}), numerous subsequent studies demonstrated that star-forming galaxies ranging from the lowest-luminosity dwarfs to massive disk galaxies obey a well-defined luminosity-metallicity (\\lz) correlation spanning several orders of magnitude in $B$-band luminosity \\citep{garnett87a, skillman89a, zaritsky94a, pilyugin04a, lama04a, salzer05a, lee06b, moustakas10a}. Metallicity was also found to correlate with morphological type \\citep{edmunds84a}, surface mass density \\citep{mccall82a, ryder95a, garnett97a}, and maximum rotational velocity \\citep{zaritsky94a, garnett02a, dalcanton07a}. However, because these global properties all correlate with one another, the underlying physical driver of the \\lz{} relation remained elusive. \\citet{tremonti04a} was the first to leverage the tremendous statistical power of the Sloan Digital Sky Survey \\citep[SDSS;][]{york00a} spectroscopic database to show that the gas-phase metallicity of a galaxy correlates best with stellar mass, \\mass, now known as the stellar mass-metallicity (\\mz) relation. The \\mz{} relation reveals that the gas-phase metallicity of star-forming galaxies increases monotonically with stellar mass and then remains relatively constant above $\\mass\\approx3\\times10^{10}~\\msun$. Although subsequent studies have found weak residual correlations from the \\mz{} relation with stellar mass density \\citep{tremonti04a, liang10a}, size \\citep{ellison08a}, large- and small-scale environment \\citep{mouhcine07a, cooper08a, ellison08b, ellison09a, peeples09a}, star formation rate \\citep[SFR;][]{mannucci10a, lara-lopez10a, yates11a, cresci11a}, and the existence of bars \\citep{ellison11a}, the intrinsic dispersion in the \\mz{} relation is $\\lesssim0.1$~dex, making it among the tightest empirical correlations known. The physical origin of the \\mz{} relation remains under debate. One possibility is that low-mass galaxies started forming stars later than massive galaxies \\citep[i.e., they are ``younger'';][]{noeske07b, leitner11b} and have been less efficient at synthesizing metals via star formation \\citep{brooks07a, mouhcine08a, calura09a}. This interpretation is qualitatively consistent with the measured low gas fractions in massive galaxies relative to lower-mass galaxies \\citep{mcgaugh97a, geha06a, garcia-appadoo09a}, and with the observed correlation between stellar mass and SFR \\citep[the {\\em star formation sequence};][]{brinchmann04a, salim07a, elbaz07a, noeske07a}. Another popular interpretation is that supernova-driven galactic winds preferentially expel metals from low-mass galaxies \\citep{larson74a, garnett02a, tremonti04a, dalcanton07a}. Indeed, state-of-the-art hydrodynamic and semianalytic theoretical models \\emph{require} metal-enriched outflows to match many observed galaxy properties, including the \\mz{} relation \\citep{kobayashi07a, finlator08a, dutton09a, peeples11a, dave11b}. A related outstanding question is whether gas and metals ejected in previous star formation episodes are re-accreted onto galaxies \\citep{delucia04a, oppenheimer10a}, or whether galaxies are predominantly fed by infall of cold, metal-poor gas from the intergalactic medium \\citep{koppen99a, faucher11a}. Alternatively, \\citet{koppen07a} show that a SFR-dependent, and therefore mass-dependent, stellar initial mass function (IMF) naturally explains the observed \\mz{} relation without needing to invoke metal-enriched outflows from galaxies. One way of gaining insight into the complex and multifaceted interrelationship between chemical abundance measurements, star formation, gas accretion, and the role of supernova feedback during galaxy growth is to measure the evolution of the \\mz{} relation. The time since $z=1$, spanning $\\sim60\\%$ of the age of the Universe, is important for many reasons. Measurements of the star formation sequence \\citep{noeske07a, wuyts11a} suggest that star-forming galaxies over this redshift range evolve smoothly as a population, driven by continuous, secular processes like gas consumption and {\\em in situ} star formation \\citep{bell05a, noeske09a, leitner11b}. Bolstered by measurements of their morphological distribution \\citep{bell05a, konishi11a} and clustering properties \\citep{coil08a, zehavi11a}, these results suggest that we can draw direct evolutionary connections linking star-forming galaxies across this redshift range. In addition, the metallicity-sensitive \\pagel{} parameter (defined in \\S\\ref{sec:oh}) can be measured from ground-based optical spectroscopy of galaxies at least to $z\\sim0.8$, ensuring that metallicities can be derived using a single consistent abundance calibration. Unfortunately, previous studies have reported widely varying results on the evolution of both the shape and normalization of the \\mz{} relation since $z=1$. \\citet{savaglio05a} and \\citet{zahid11a} find that the mean metallicity of galaxies at $z\\sim0.8$ with $\\mass\\sim10^{10.3}$~\\msun{} differs by just $\\sim0.05$~dex (factor of $\\sim1.12$) relative to similarly massive star-forming galaxies at $z\\sim0.1$, whereas they find that galaxies with $\\mass\\sim10^{9.5}$~\\msun{} undergo a factor of $2-3$ more chemical evolution over the same redshift interval. On the other hand, \\citet{liang06a} and \\citet{cowie08a} report $0.2-0.3$~dex (factor of $1.6-2$) of chemical evolution for galaxies with $\\mass\\sim10^{10.3}$~\\msun{} since $z\\sim0.7$ and no statistically significant evidence for evolution in the shape of the \\mz{} relation \\citep[see also][]{rodrigues08a}. \\citet{lama09a} and \\citet{perez-montero09a} find an even greater amount of metallicity evolution, $\\sim0.35$~dex (factor of $2.2$) since $z=0.6-1$, and a \\emph{flattening} of the \\mz{} relation at intermediate redshift. Other studies based on the evolution of the $B$-band \\lz{} relation report that at fixed luminosity star-forming galaxies at $z=0.5-1$ are $0.1-0.7$~dex (factor of $1.25-5$) more metal-poor than local star-forming galaxies \\citep{kobulnicky99b, carollo01a, lilly03a, kobulnicky03a, kobulnicky04a, maier04a, maier05a, maier06a, lama09a}. However, the significant amount of luminosity evolution experienced by blue, star-forming galaxies since $z=1$ \\citep{blanton06a, faber07a, cool12a} renders the interpretation of the \\lz{} relation less straightforward. The origin of these widely varying results on the evolution of the \\mz{} relation can be attributed to various issues, including cosmic variance, small sample size, heterogenous selection criteria, systematic differences in the methods used to infer nebular abundances, an inconsistent analysis of the local \\mz{} relation, spectroscopy with insufficient spectral coverage, signal-to-noise (S/N) ratio, or instrumental resolution, and uncertain stellar mass estimates due to limited broadband photometric coverage. Consequently, the efficacy of existing observational metallicity constraints on state-of-the-art theoretical models of galaxy formation \\citep{delucia04a, brooks07a, dutton09a, dave11b} has been fairly limited. From an observational standpoint, there exists a clear need for a large, homogeneously selected sample of galaxies with high-quality spectroscopy and reliable stellar mass estimates to better constrain the evolution of the \\mz{} relation at intermediate redshift. To address this need, we measure the evolution of the \\mz{} and $B$-band \\lz{} relations at intermediate redshift using oxygen abundances of $\\sim3000$ star-forming galaxies at $z=0.05-0.75$ observed as part of the AGN and Galaxy Evolution Survey \\citep[AGES;][]{kochanek11a}, and $\\sim75,000$ galaxies at $z=0.05-0.2$ selected from the SDSS. The AGES main galaxy survey consists of optical spectrophotometry for $\\sim12,000$ galaxies in the $\\sim9$~deg$^{2}$ NOAO Deep Wide Field Survey \\citep[NDWFS;][]{jannuzi99a, brown03a, brown07a, brown08a} \\bootes{} field at a median redshift of $z\\sim0.3$. This sample is statistically complete over $7.9$~deg$^{2}$ for galaxies with $I_{\\rm AB}<20.45$ \\citep{kochanek11a}, reaching $\\sim2$~mag deeper than the SDSS main galaxy sample \\citep{strauss02a} over a considerably larger area than other recent or ongoing surveys of intermediate-redshift galaxies such as DEEP2 \\citep{davis03a}, AEGIS \\citep{davis07a}, VVDS \\citep{lefevre04a, lefevre05a, garilli08a}, and zCOSMOS \\citep{lilly09a}.\\footnote{For galaxies at $z\\lesssim0.5$, the Galaxy and Mass Assembly (GAMA) survey will ultimately supplant all these surveys by obtaining optical spectroscopy for several hundred thousand galaxies brighter than $r_{\\rm AB}\\approx19.8$ over $360$~deg$^{2}$ \\citep{driver11a}.} The statistical completeness, large sample size, and availability of high-quality optical spectroscopy and deep optical and near-infrared photometry enables us to investigate the evolution of both the \\mz{} and optical \\lz{} relations of star-forming galaxies at $z=0.05-0.75$ using a single, self-consistent abundance diagnostic over the entire redshift range. The plan of the paper is as follows. In \\S\\ref{sec:data} we present the ground-based optical and near-infrared imaging we use, summarize the AGES observations, and describe how we measure the nebular emission lines for the galaxies in our sample. In \\S\\ref{sec:sed} we describe the methodology used to derive rest-frame luminosities, colors, and stellar masses, and in \\S\\ref{sec:sample} we select a subset of the AGES galaxies for chemical abundance analysis. We present the methods we use to derive oxygen abundances in \\S\\ref{sec:oh}, and our principal results in \\S\\ref{sec:evol}, where we quantify the mass-dependent evolution of the \\mz{} relation for star-forming galaxies since $z=0.75$. In \\S\\ref{sec:syseffects} we address the effect of various potential sources of systematic uncertainty on our results, and in \\S\\ref{sec:discussion} we discuss recent theoretical work on the origin and evolution of the \\mz{} relation. Finally, we summarize our principal conclusions in \\S\\ref{sec:summary}. We adopt a concordance cosmology with $\\Omega_{\\rm m}=0.3$, $\\Omega_{\\Lambda}=0.7$, and $h_{70}\\equiv H_{0}/100=0.7$, the AB magnitude system \\citep{oke83a}, and the \\citet{chabrier03a} initial mass function (IMF) from $0.1-100~\\mathcal{M}_{\\sun}$ unless otherwise indicated. For reference, the conversion from Vega to AB for the $I$-band filter used to select AGES targets is $+0.45$~mag, that is $I_{\\rm AB}=I_{\\rm Vega}+0.45$. ", "conclusions": "" }, "1112/1112.1844_arXiv.txt": { "abstract": "We present the result of a systematic study of pileup phenomena seen in the X-ray Imaging Spectrometer, an X-ray CCD instrument, onboard the Suzaku observatory. Using a data set of observed sources in a wide range of brightness and spectral hardness, we characterized the pileup fraction, spectral hardening, and grade migration as a function of observed count rate in a frame per pixel. Using the pileup fraction as a measure of the degree of pileup, we found that the relative spectral hardening (the hardness ratio normalized to the intrinsic spectral hardness), branching ratio of split events, and that of detached events increase monotonically as the pileup fraction increases, despite the variety of brightness and hardness of the sample sources. We derived the pileup fraction as a function of radius used for event extraction. Upon practical considerations, we found that events outside of the radius with a pileup fraction of 1\\% or 3\\% are useful for spectral analysis. We present relevant figures, tables, and software for the convenience of users who wish to apply our method for their data reduction of piled-up sources. ", "introduction": "\\label{section:1} Charge coupled devices (CCDs) have been playing an important role in the modern X-ray astronomy (Lumb et al. 1991) since the SXT (Soft X-ray Telescope; \\cite{Tsuneta1991}) mounted on YOHKOH. As a photon-counting detector, the first successful use of it in space was realized with the SIS (Solid-state Imaging Spectrometer; Burke et al. 1991) on board the ASCA satellite. The fifth Japanese X-ray satellite Suzaku (\\cite{Mitsuda2007}) also carries the CCD devices, the X-ray Imaging Spectrometer (XIS; \\cite{Koyama2007}), located at the foci of the X-ray Telescope (XRT; \\cite{Serlemitsos2007}) modules, as well as a non-imaging hard X-ray instrument, the Hard X-ray Detector (HXD; \\cite{Takahashi2007}). The XIS covers soft X-rays from 0.2 to 12 keV, while the HXD covers from 10 to 600 keV. The wide-band coverage by a combination of these instruments is a key competence of Suzaku. Sources that have rich statistics in the HXD are usually bright in the XIS as well. In observations of such a bright source, more than one photon can strike the same CCD pixel, or one of its immediate neighbors during the exposure time per frame. It causes a complicated and irreversible phenomenon called ``pileup''. For example, when two photons with energies of $E_1$ and $E_2$ fall in the same pixel within an integration period, it is impossible to distinguish them from one with an energy of $E_1 + E_2$. It affects obtained spectra, images, and light curves. Incorrect evaluation of pileup can lead to wrong interpretations of the data. This is especially the case for bright sources because their small statistical uncertainty can be overwhelmed by systematic uncertainties brought by pileup. Therefore, it is important to understand pileup effects in order to maximize the wide-band capability of Suzaku. In general, when pileup occurs, flux and spectra become lower and harder, respectively. The pileup effects of the SISs of ASCA, have been studied by \\citet{Ebisawa1996}, \\citet{Ueda1997}, and Kotani et al. (2000). Recently, a detailed study for the CCD detector of Chandra, the ACIS, as well as a generalized formula of the pileup, has been compiled by Davis (2001). However, the telescopes used for the XIS and the ACIS are entirely different in the shape and size of the point spread function (PSF), which adds non-linear complications to the effect. Therefore, it is difficult to simply apply methods developed for the ACIS to the XIS. Likewise, the way of estimating pileup\\footnote{http://xmm.esac.esa.int/sas/current/documentation/ \\\\ threads/epatplot.shtml} for the CCD detector of XMM-Newton, the EPIC, can not be directly applied to the XIS. Thus, studies on the pileup effects should be performed specifically for the combined system of XISs and XRTs. One of the practical solutions to mitigate the pileup in the XIS is the \"core-exclusion\" method, in which the central region of the PSF is excluded to avoid the CCD region severely affected by pileup. This method was adopted in the analysis of 4U~1630$-$472 (Kubota et al. 2007), Cygnus X-1 (Makishima et al. 2008), GX 339$-$4 (Yamada et al. 2009), and others. Based on these experiences, we released a recipe to apply the method for XIS data as an official document\\footnote{http://www.astro.isas.ac.jp/suzaku/analysis/xis/pileup/ \\\\ HowToCheckPileup\\_v1.pdf}. In the document, it is poorly stated how to estimate the size of the core to be excluded. This is often difficult to do by using data that are affected by pileup. In this paper, we present the results of a systematic analysis for several bright sources affected by pileup in the XIS. A brief summary on pileup characteristic for phenomenological specific to the XIS+XRT system is presented in section 2. After describing target selection and data processing in section 3, systematic studies on pileup effects are presented in section 4. In section 5, we discuss a criterion that utilizes pileup fraction and show concrete examples for the application. The main results of the paper are summarized in section 6. ", "conclusions": "We have comprehensively studied pileup phenomena of the XIS, by analyzing the actually observed data for several bright point sources. Using the surface brightness, the pileup fraction as the primary measure, we characterized grade-branching ratios of split events and detached events. We examined the effects on spectral analysis when excluding the central regions within $R_{3\\%}$ and $R_{1\\%}$. Our results are summarized as follows: \\begin{itemize} \\item Pileup causes spectral hardening in FI and BI CCDs, as well as changes in the sharp spectral features such as atomic edges of instrumental response. The changes in hardness is not dependent on intrinsic spectral hardness. \\item Pileup increases the ratio of split events by $\\sim$ 20\\% from a pileup fraction of $\\sim$ 1\\% to $\\sim$ 10\\% in both FI and BI CCDs. On the other hand, the ratio of the detached events increases proportionally with the pileup fraction by $\\sim$ several orders of magnitude from $\\sim$ 0.1 and $\\sim$ 10 \\% of pileup fraction. \\item When surface brightness, spectral hardness, and split event ratio begin to increase, the pileup fraction starts to exceed $\\sim$ 1\\%, which illustrate that the pileup fraction can be a good measure for estimating pileup. \\item The relation between the measured surface brightness and the radius corresponding to pileup fraction of 3\\% and 1\\%, is useful to estimate the extent of pileup in a practical manner. \\item Based on the observation of Cygnus X-1, the differences between the spectra accumulated outside $R_{3\\%}$ and $R_{1\\%}$ are less than 5\\% in the 2--10 keV and $\\sim$ 20 \\% in 0.5--2 keV. The latter might be caused by uncertainties of contamination modeling in the XIS. \\end{itemize} Thus, we have phenomenologically compiled the pileup effects specific to the XIS + XRT systems and found a practical way of estimating pileup effects, which would be useful for those who analyze the bright sources taken with Suzaku, and even with similar CCD sensors to the XIS+XRT system, such as Soft X-ray Imager on board ASTRO-H (Tsunemi et al. 2010). \\vspace{0.5cm} The authors would like to express their thanks to Suzaku and the XIS team members, and anonymous referee for helpful comments and suggestions. The present work was supported by Grant-in-Aid for JSPS Fellows and RIKEN." }, "1112/1112.4466_arXiv.txt": { "abstract": "In this paper we investigate scaling relations between star formation rates and molecular gas masses for both local Galactic clouds and a sample of external galaxies. We specifically consider relations between the star formation rates and measurements of dense, as well as total, molecular gas masses. We argue that there is a fundamental empirical scaling relation that directly connects the local star formation process with that operating globally within galaxies. Specifically, the total star formation rate in a molecular cloud or galaxy is linearly proportional to the mass of dense gas within the cloud or galaxy. This simple relation, first documented in previous studies, holds over a span of mass covering nearly nine orders of magnitude and indicates that the rate of star formation is directly controlled by the amount of dense molecular gas that can be assembled within a star formation complex. We further show that the star formation rates and total molecular masses, characterizing both local clouds and galaxies, are correlated over similarly large scales of mass and can be described by a family of linear star formation scaling laws, parameterized by $f_{DG}$, the fraction of dense gas contained within the clouds or galaxies. That is, the underlying star formation scaling law is always linear for clouds and galaxies with the same dense gas fraction. These considerations provide a single unified framework for understanding the relation between the standard (non-linear) extragalactic Schmidt-Kennicutt scaling law, that is typically derived from CO observations of the gas, and the linear star formation scaling law derived from HCN observations of the dense gas. ", "introduction": "Knowledge of the physical factors that control the conversion of interstellar gas into stars is of fundamental importance for both developing a predictive physical theory of star formation and understanding the evolution of galaxies from the earliest epochs of cosmic history to the present time. An essential first step to obtaining such knowledge is to establish empirically the underlying relation or relationships that most directly connect the rate of star formation in a galaxy to some general physical property of the interstellar gas from which stars form. A little more than a half-century ago, Schmidt (1959) conjectured that this might take the form of a scaling relation between the rate of star formation and some power, n, of the surface density of atomic (HI) gas. From evaluation of the distributions of local HI gas and stars orthogonal to the Galactic plane, he suggested that n$\\approx$ 2. Subsequent studies comparing the surface densities of OB stars and HII regions with those of atomic gas within nearby external galaxies produced scaling laws with similar, super-linear, power-law indices (e.g., Sanduleak 1969; Hamajima \\& Tosa 1975). By the 1980s it became clear that molecular, not atomic, clouds were the sites of star formation in galaxies. The ability to make sensitive CO molecular-line observations enabled, for the first time, the measurement of total gas surface densities ($\\Sigma_{HI + H_2}$) in external galaxies while advancements in infrared and ultraviolet observations led to significant improvements in the measurements of star formation rates. Significant effort was then expended by a number of researchers to systematically measure star formation rates and total gas surface densities in increasingly large samples of galaxies (e.g., Kennicutt 1989 and references therein). These efforts culminated in the study of Kennicutt (1998a) who compiled galaxy averaged measurements of star formation rates and gas surface densities for a large sample of galaxies including normal spirals and starbursts. He derived a scaling relation between the star formation rate surface density ($\\Sigma_{SFR}$) and total gas surface density ($\\Sigma_{HI + H_2}$) that was characterized by a power-law index of n $\\approx$ 1.4. This value was shallower than that Schmidt and others found for individual galaxies using only atomic gas but still super-linear. Wong and Blitz ( 2002), employing spatially resolved observations of seven nearby, molecular rich, spiral galaxies, showed that the star formation rate was better correlated with the molecular hydrogen surface density, $\\Sigma_{H_2}$, than with the atomic surface density, but still obtained n $\\approx$ 1.4. More recently, Bigiel et al. (2008) analyzed spatially resolved observations of 18 nearby galaxies containing both atomic rich and molecular rich objects and confirmed that $\\Sigma_{SFR}$ was better correlated with $\\Sigma_{H_2}$ than $\\Sigma_{HI}$, but they determined that n $=$ 1.0 ($\\pm$ 0.2) for the $\\Sigma_{SFR}$ -- $\\Sigma_{H_2}$ relation. However, recent observations of M 101 and M 81 have suggested that the index of the scaling law can vary within a galaxy with values of n ranging between 1 and 2 (Suzuki et al. 2010). Among the more interesting investigations of the extragalactic scaling laws for star formation was that of Gao and Solomon (2004) who used molecular-line emission from HCN, rather than CO, to trace the molecular gas. They found a linear (n $=$ 1) correlation between the total far-infrared luminosities and the HCN molecular-line luminosities of a large sample of star forming galaxies including normal spirals and luminous and ultra-luminous infrared galaxies. Since the total infrared luminosity is a good proxy for the total star formation rate (SFR) and the HCN luminosity a good proxy for the total amount of dense (i.e., n(H$_2$) $\\geq$ 3 $\\times$ 10$^4$ cm$^{-3}$) gas in a galaxy, this also implied a linear correlation between the SFR and the mass of dense molecular gas. % The various determinations of differing power-law indices for the extragalactic star formation scaling relations present a somewhat confused and problematic picture. Particularly since the difference between a linear and non-linear scaling relation can have significant consequences for the theoretical understanding of the star formation process in galaxies. Therefore it is important to understand the nature of such differences. Are the different scaling relations consistent with each other? Are the differences due to such effects as the choice of the samples studied (e.g., normal spirals vs starbursts, CO rich vs. HI rich galaxies, distant vs. nearby systems, etc.) or the different quantities actually measured (e.g., SFR vs. $\\Sigma_{SFR}$, $\\Sigma_{HI + H_2}$ vs. $\\Sigma_{H_2}$, or CO vs. HCN, etc.), or the systematic uncertainties in the quantities measured (e.g., observational tracers or IMFs adopted for SFR determinations, conversion factor for transforming CO measurements into H$_2$ masses, etc.), or some linear combination of all these effects? Do any of these scaling relations represent the fundamental underlying physical relationship that most directly connects star formation activity with interstellar gas? Schmidt's original scaling law was determined from observations of the local region of the Galaxy. Since our knowledge of the local Milky Way has improved profoundly over the last half century, it would seem that important insights into the relation between star formation and interstellar gas could and should be derived from observations of local star formation activity. In a previous paper (Lada et al. 2010; hereafter Paper I) we presented a study of the star formation activity in a sample of local (d $<$ 0.5 kpc) molecular clouds with total masses between 10$^3$ and 10$^5$ \\msun. We employed infrared extinction measurements derived from wide-field surveys to determine accurate cloud masses and mass surface densities, and compiled from the literature both ground and space-based infrared surveys of young stellar objects to construct complete inventories of star formation within the clouds of our local sample. We found the specific star formation rates (i.e., the star formation rates per unit cloud mass) in these clouds to vary by an order of magnitude, independent of total cloud mass. However, we also found the dispersion in the specific star formation rate, to be minimized (and reduced by a factor of 2-3) if one considers only the mass of molecular gas characterized by high extinction in calculating the specific star formation rates. As a result we showed that the (total) star formation rate in local clouds is linearly proportional to the cloud mass contained above an extinction threshold of A$_K$ $\\geq$ 0.8 magnitudes, corresponding to a gas surface density threshold of $\\Sigma_{H_2}$ $\\approx$ 116 \\msun pc$^{-2}$. Similar surface density thresholds for star formation in local clouds have been suggested in other recent studies (e.g., Goldsmith et al 2008; Heiderman et al. 2010). Given the density stratification of molecular clouds, we argued that such surface density thresholds also correspond to volume density thresholds of n(H$_2$) $\\approx$ 10$^4$ cm$^{-3}$. These findings are consistent with and reinforce those of Wu et al. (2005) who had already demonstrated a linear correlation between far-infrared luminosity and HCN luminosity (i.e., between SFR and dense gas mass) for more massive and distant star formation regions in the Milky Way. The correspondence between these results and those obtained by Gao and Solomon (2004) for external galaxies is intriguing and especially striking because the scalings of the Galactic and extragalactic power-law relations, that together span more than nine orders of magnitude in cloud mass, agree to within a factor of 2-3. This suggested to us that the close relationship between the star formation rates and the {\\it dense} gas masses of molecular clouds could be the underlying physical relation that connects star formation activity with interstellar gas over vast spatial scales from the immediate vicinity of the sun to the most distant galaxies. However, if this is so, how does one understand these observations in the context of the classical Schmidt-Kennicutt scaling relations based on CO observations? These classical relations are often super linear and moreover, as Heiderman et al. (2010) point out, they under predict the $\\Sigma_{SFR}$ in local regions by factors of 17 - 50 (see also Evans et al. 2009). In this paper we attempt to address this issue by re-examining the extinction observations of local clouds to include low extinction material and re-examining the CO observations of the clouds studied by Gao and Solomon. We show that all the observations can be understood within a self-consistent framework in which the differences are primarily due to the dense gas fractions that characterize the molecular gas being observed, supporting a hypothesis originally put forward by Gao and Solomon (2004). ", "conclusions": "The SFR-Molecular Mass diagram of Figure 2 provides a physical context for understanding the star formation scaling laws over spatial scales ranging from those of local molecular clouds to those of entire galaxies. The close correlation of the star formation rate with the mass of dense gas over these immense scales has been established in previous studies (Wu et al. 2005, Paper I). Here we find that a close relation also appears to hold between the SFR and the total molecular mass over a similarly large range, 8-9 orders of magnitude in both quantities. Both the local clouds and galaxies appear to scatter around the linear relation given by Equation 1 for $f_{DG}$ $=$ 0.1 and $M_G = M_{TG}$. From extrapolation of the results for local clouds we suggest that this particular line corresponds to the case where 10\\% of the measured gas mass is in the form of dense, star forming material for the galaxies as well as for the local clouds. The smaller scatter of the galaxies around this relation compared to that of the local clouds is likely the result of the fact that the galaxy measurements are averages over entire systems. These results indicate that, similar to the situation for dense gas, the star formation scaling law for total (H$_2$ + He) gas mass is likely linear across all scales for molecular clouds with similar dense gas fractions. This notion is reinforced by the recent observations of Daddi et al. (2010) who studied infrared-selected BzK galaxies at $z\\sim1.5$ and found evidence for unusually high gas fractions and extended molecular reservoirs in these distant systems. Using the star formation rates and CO gas masses provided by Daddi et al. (2010), we plot these six galaxies (open triangles) on Figure 2 and find that the BzK galaxies occupy an area in the SFR-Molecular Mass plot that is close to the linear relation described by Equation 1, consistent with the locations of Gao and Solomon galaxies and the extrapolation of the local Galactic cloud sample. These results lead us to the conclusion that there is a basic and universal physical process of star formation that presently operates in our Milky Way galaxy and is also responsible for the bulk of star forming activity occurring in external galaxies both in the present epoch (z $\\approx$ 0; GS04) and perhaps at much earlier (z $\\approx 1-2$; Daddi et al. 2010) cosmic times. It is a process in which the rate of star formation is simply and directly controlled by the amount of dense molecular gas that can be assembled within a star forming complex. In most situations massive molecular clouds appear to be able to convert only about 10\\% or less of their total mass into a sufficiently dense (n(H$_2$) $\\ge$ 10$^4$ \\cc) form to actively produce stars. This may be considered as the normal process of star formation in GMCs. Closer inspection of Figure 2 suggests that for starburst galaxies, particularly the ULIRGS, this standard scenario may be modified. As the SFRs for starbursts (i.e., LIRGs and ULIRGs in Figure 2) increase with gas mass, the open symbols (CO derived gas masses) appear to approach and then merge with the filled symbols (HCN derived gas masses), almost overlapping at the highest SFRs. As originally hypothesized by Gao and Solomon (2004), we interpret this to indicate that these galaxies are characterized by an increasingly high dense gas fraction and consequently, the CO observations begin to trace nearly the same material as the HCN observations. Nonetheless, the star formation rate is still dictated by the amount of dense gas within the galaxies. This interpretation is also favored by Heiderman et al. (2010) who suggested that the maximal starburst activity occurs when $f_{DG} = \\rm 1$ which they posit to happen when the mass surface density exceeds values $\\sim$ 10$^4$ \\msun\\ pc$^{-2}$. ULIRGS (e.g., Arp 200) are believed to be experiencing major mergers and we suggest that this extreme process likely produces conditions (e.g., high pressures) that could increase the dense gas fractions of the molecular clouds within these systems (e.g., Blitz \\& Rosolowsky 2006). In contrast the BzK galaxies studied by Daddi et al. (2010) have similarly high SFRs but lower dense gas fractions. Their high star formation rates appear to result from high global molecular gas mass fractions (i.e., M$_{H_2}$/M$_*$), as might be expected for very young galaxies. We note that a linear relation in the SFR-Mass plane should transform to a linear relation in the $\\Sigma_{SFR}$-$\\Sigma_{g}$ plane (provided the surface densities for the galaxies are global averages) and we can express our star formation scaling law in this latter plane as: \\begin{equation} \\Sigma_{SFR} \\propto f_{DG}\\Sigma_{g} \\end{equation} \\noindent where $\\Sigma_{g}$ refers to the H$_2$ gas mass. Moreover, the Spitzer study of Galactic clouds by Heiderman et al. (2010) suggested a linear star formation law in the $\\Sigma_{SFR}$-$\\Sigma_{g}$ plane that holds for gas above a threshold surface density of $\\sim$ 130 \\msun\\ pc$^{-2}$ (i.e., A$_K$ $>$ 0.9 mag) and extrapolates smoothly to the GS04 galaxies. Our result is apparently not consistent with the standard Schmidt-Kennicutt, super-linear, scaling law (Kennicutt 1998a). Both are based on valid empirical relations. However, here we argue that the underlying scaling law for star formation is linear over all scales for all clouds and galaxies, provided they are characterized by the same dense gas fraction. Kennicutt (1998a) uses total (HI $+$ H$_2$) gas mass surface densities with CO derived molecular masses and combines results for both normal star-forming disk galaxies and starburst galaxies to derive his star formation scaling law. Note that for these latter galaxies the total gas surface densities are dominated by the molecular component. The starbursts dominate his relation for $\\Sigma_{gas}$ $>$ 100 \\msun pc$^{-2}$. It is possible that the fit of a single relation to the combined sample with CO determined masses is inappropriate and skewed by the starbursts because $f_{DG}$ for starbursts is higher than that for normal star forming spirals. Indeed, Gao and Solomon (2004) showed that using the masses calculated from the CO observations produced a super-linear (n $\\approx$ 1.7) scaling law (in the SFR vs M$_G$ plane) for a sample that included their galaxies and an additional number of luminous starbursts drawn form the literature. Using gas masses derived solely from HCN observations, however, produces a linear star formation law connecting both normal star forming galaxies and starbursts. The standard Schmidt-Kennicutt relation may also be skewed at low mass surface densities. For galaxies in this portion of the diagram, the HI surface density is a large fraction of the total gas surface density and thus the measured total gas surface density, $\\Sigma_{HI + H_2}$, contains a large component of inert, non-star forming, (HI) gas; this dilutes and lowers the SFR corresponding to a fixed mass surface density, resulting in a steepening of the slope of the $\\Sigma_{SFR}$ vs $\\Sigma_{gas}$ relation. These two effects, the increasing dense gas fraction for the starbursts and the dilution of the dense gas fraction by HI gas at low gas surface densities, which act together to steepen the slope of the Schmidt-Kennicutt relation, can also explain the finding of Heiderman et al. (2010) and Evans et al. (2009) that the extrapolation of the extragalactic scaling relations to local scales (i.e., low mass surface densities) lies below the data for Galactic clouds. It can also be shown that our scaling law (Equation 1) is consistent with a volumetric scaling law, $\\dot{\\rho_*} \\propto \\rho_G^n$ if and only if $n = 1$ and $\\rho_G \\geq \\rho_{thres}$, where $\\rho_{thres}/\\mu$ corresponds to the threshold volumetric number density for star formation for a mean particle mass given by $\\mu$ (i.e., n(H$_2$) $\\geq$ 10$^4$ cm$^{-3}$). As discussed earlier, taking the empirical, linear star-forming scaling relations at face value leads to a simple interpretation of the observations in Figures 1 \\& 2. Namely, that the total rate of star formation, $\\dot{M}_*$, is directly proportional to the mass of dense molecular gas above a threshold density, $M_{DG} = \\int_{\\rho_{thres}}^\\infty M(\\rho) d\\rho$. Moreover, once the gas has reached this threshold density, the SFR does not depend on the exact value of the density but only on the total mass of gas whose density has exceeded the threshold. This interpretation of the observations differs from those that explain the observed non-linear index of the Schmidt-Kennicutt law as resulting from star formation timescales dictated by the free-fall time, e.g., SFR $\\sim$ M/$\\tau_{ff}$ $\\sim$ $\\rho/ \\rho^{-0.5}$ $\\sim$ $\\rho^{1.5}$ since $\\tau_{ff} \\sim \\rho^{-0.5}$ (e.g., Elmegreen 1994; Krumholz \\& Thompson 2007; Narayanan et al. 2008). A recent variant of such a model has been studied by Krumholz et al. (2011). They propose that the underlying physical law governing the relation between star formation rates and cloud properties is given by $\\dot{\\rho_*} \\propto \\rho_G/ \\tau_{ff}$. They find that the standard Schmidt-Kennicutt law can be linearized if the data are plotted in the $\\Sigma_{SFR}-\\Sigma_G/\\tau_{ff}$ plane as long as the free-fall times are measured using the typically higher densities of the star forming gas rather than those derived from the mean densities averaged over kpc scales. Their interpretation differs from the one in this paper in that Krumholz et al. (2011) posit that the positions of galaxies in the standard $\\Sigma_{SFR}$-$\\Sigma_{G}$ plane are a consequence of both their gas surface densities and their local free-fall times, while here we posit that the locations of these galaxies instead depend on their gas surface densities and their dense gas fractions. Although both interpretations are consistent with the observations, they appear not to be consistent with each other. However, we point out that Figure 1 empirically demonstrates that the locations of Galactic clouds in the SFR-Mass diagram are in fact a result of their dense gas fractions. Therefore it seems reasonable to infer that the locations of galaxies in the diagram are due to the same cause. Finally, we reiterate our point that the linear scaling law of Equation 1 implies that the process of star formation across entire galaxies as well as individual local clouds is governed by a very similar and simple physical principle: the rate at which molecular gas is turned into stars depends on the mass of dense gas within a molecular cloud or cloud population. The underlying star formation scaling law is linear over all scales for all clouds and galaxies characterized by the same dense gas fraction. The star formation rate appears therefore to be controlled by local processes and not by global, galactic scale mechanisms, except to the extent that such mechanisms can alter the dense gas fractions in the molecular gas. If this interpretation is correct, then the key problem that needs to be addressed in future studies is that of the origin of the dense gas component of molecular clouds." }, "1112/1112.6149_arXiv.txt": { "abstract": "We consider inflationary models in which vector fields are responsible for part or eventually all of the primordial curvature perturbation $\\zeta$. Such models are phenomenologically interesting since they naturally introduce anisotropies in the probability distribution function of the primordial fluctuations that can leave a measurable imprint in the cosmic microwave background. Assuming that non-Gaussianity is generated due to the superhorizon evolution, we use the $\\delta N$ formalism to do a complete tree level calculation of the % non-Gaussianity parameters $f_{\\rm NL}$ and $\\tau_{\\rm NL}$ in the presence of vector fields. % We isolate the isotropic pieces of the non-Gaussianity parameters, which anyway have contributions from the vector fields, and show that they obey the Suyama-Yamaguchi consistency relation $\\tau^{\\rm iso}_{\\rm NL}\\geqslant(\\frac{6}{5}f^{\\rm iso}_{\\rm NL})^2$. Other ways of defining the non-Gaussianity parameters, which could be observationally relevant, are stated and the respective Suyama-Yamaguchi-like consistency relations are obtained. ", "introduction": "The study of non-Gaussianities in the primordial curvature perturbation $\\zeta$ is a subject of major interest in modern cosmology because the evaluation of the non-Gaussianity (NG) parameters provide criteria to discriminate among the many models proposed to explain the origin of the large-scale structure that we observe today\\cite{obs,obs1,obs2}. Particular attention has been paid to the study of a consistency relation between the $f_{\\rm NL}$ and $\\tau_{\\rm NL}$ parameters\\cite{cr10,cr11,cr12,cr13,cr14,cr20,cr21}, the Suyama-Yamaguchi (SY) consistency relation ($\\tau_{\\rm NL} \\geqslant (\\frac{6}{5} f_{\\rm NL})^2$ at least at tree level), because its violation would rule out many of the popular models of inflation. So far, it has been shown that this consistency relation works, in principle, on models that only include scalar fields\\cite{cr10,cr11,cr12,cr13,cr14} under just a few assumptions, but there is no any conclusive result for models that include other type of fields, for instance, vector fields (VF)\\footnote{VF naturally generate scale-dependence and, therefore, the proofs in Ref. \\cite{cr20,cr21} do not apply to those scenarios where VF contribute to the generation of the primordial curvature perturbation.}. Vector field models are particularly interesting because they are suitable candidates to explain the apparent violation of statistical isotropy observed in recent analysis of data from the WMAP satellite\\cite{gr1,gr2}, given that VF define inherently a preferred direction for the expansion, for the distribution of primordial fluctuations, or for both\\cite{vi1,vi2,vi3,vi4,vi5,vi6,ValenzuelaToledo:2011fj,dklr}. Because of this reason, it is pertinent to study consistency relations in models that include scalar fields as well as vector fields as these not only include NG parameters but also the amount of primordial statistical anisotropy $g_\\zeta$ \\cite{varios1,varios2,varios3,varios4}. The main purpose of this paper is to find out how the well known SY consistency relation, valid for models involving only scalar fields under just a few assumptions, is modified when including VF. To this end, we do a complete tree level calculation of the $f_{\\rm NL}$ and $\\tau_{\\rm NL}$ parameters based on the $\\delta N$ formalism\\cite{dklr}, and concentrate on the isotropic pieces of these parameters since these are the ones that are currently constrained by observations. Under almost the same few assumptions made in the scalar fields case, and although the VF do contribute to the isotropic pieces of the NG parameters, we prove that the SY consistency relation is still obeyed, at least at tree level. Finally, the modified SY consistency relations for other ways of defining the NG parameters, which could be observationally relevant when taking into account the level of statistical anisotropy, are obtained. ", "conclusions": "In this paper, we have studied under which conditions new kinds of the well known Sayama-Yamaguchi consistency relation are allowed when vector fields are included in the inflationary dynamics. In particular we have shown that the isotropic pieces of the non-gaussianity parameters obey the Suyama-Yamaguchi consistency relation $\\tau^{\\rm iso}_{\\rm NL}\\geqslant(\\frac{6}{5}f^{\\rm iso}_{\\rm NL})^2$ when its calculation is performed at tree level. We have also derived a modified consistency relation for the complete non-gaussianity parameters $f_{\\rm NL}$ and $\\tau_{\\rm NL}$ and for a set of new parameters $f'_{\\rm NL}$ and $\\tau'_{\\rm NL}$ that could be observationally relevant; these relations are given by the Eqs. \\eqref{cr1} and \\eqref{cr2}. Since the latter depend on the level of statistical anisotropy and the configuration of the wavevectors, the naive relations $\\tau_{\\rm NL}\\geqslant(\\frac{6}{5}f_{\\rm NL})^2$ and $\\tau'_{\\rm NL}\\geqslant(\\frac{6}{5}f'_{\\rm NL})^2$ could eventually be violated in some particular inflationary model that includes vector fields. On this basis, we expect that in a near future the derived relations could provide a criteria to discriminate among the different models for the generation of the primordial curvature perturbation when vector fields are involved." }, "1112/1112.2076_arXiv.txt": { "abstract": "\\begin{center} {\\bf Abstract} \\end{center} We calculate the $^4$He abundance in a universe of Bianchi type I whose cosmic anisotropy is dynamically generated by a fluid with anisotropic equation of state. Requiring that the relative variation of mass fraction of $^4$He is less than $4\\%$ with respect to the standard isotropic case to be consistent with astrophysical data, we constrain the parameter of cosmic anisotropy, the shear $\\Sigma$, as $|\\Sigma(T_f)| \\lesssim 0.4$, where $T_f$ is the freeze-out temperature of the weak interactions that interconvert neutrons and protons. Anisotropic fluids, whose energy density is subdominant with respect to the energy content of the Universe during inflation and radiation era, generate much smaller shears at the time of freeze-out and then do not appreciably affect the standard $^4$He production. This is the case of anisotropic dark energy, and of a uniform magnetic field with energy density much smaller than about $1.25$ times the energy density of neutrinos. ", "introduction": "} \\renewcommand{\\thesection}{\\arabic{section}} The high level of isotropy of the cosmic microwave background (CMB) radiation is the most convincing justification of the Cosmological Principle: the Universe is homogeneous and isotropic at large cosmological scales~\\cite{Weinberg}. However, tiny deviations from perfect isotropy are not excluded by present CMB data. Indeed, a particular anisotropic cosmological model of Bianchi type I, known as {\\it ellipsoidal universe}~\\cite{ellipsoidal1,ellipsoidal2}, can better match CMB data and solve the so-called ``quadrupole problem,'' namely the lack of CMB power detected on large angular scales. Various mechanisms could give rise to an ellipsoidal universe, such as a uniform cosmological magnetic field~\\cite{Barrow1,ellipsoidal1,ellipsoidal2}, topological defects (e.g. cosmic stings, domain walls)~\\cite{Barrow1}, or a dark energy fluid with anisotropic equation of state~\\cite{Barrow1,Koivisto-Mota}. Independently on the nature of the mechanism, however, a modification of the standard picture of primordial nucleosynthesis can occur if universe anisotropization takes place during the early Universe~\\cite{R1,R2,R3,R4}. The aim of this paper is, indeed, to constrain the level of cosmic anisotropy, so as to be consistent with observational bounds on primeval $^4$He abundance. ", "conclusions": "} \\renewcommand{\\thesection}{\\arabic{section}} In this paper, we have analyzed the effects caused by cosmic anisotropy on the primordial production of $^4$He. We worked in the context of a cosmological model of Bianchi type I where the anisotropy of spatial geometry, the shear $\\Sigma$, is generated by a fluid with anisotropic equation of state. We found that in such an anisotropic universe there is an overproduction of $^4$He with respect to the standard isotropic case. Imposing that the relative increase of $^4$He abundance is below the $4\\%$ to be consistent with observational data, we constrained the absolute value of the shear to be less than $0.4$ at the time of freeze-out. This limit does not depend on the equation(s) of state of the anisotropic fluid and has been obtained assuming that the energy density of the anisotropic fluid is small compared to that of radiation. Moreover, we showed that anisotropic fluids generated at inflation, such as dark energy with anisotropic equation of state and a uniform magnetic field, create anisotropies much smaller than the above limit if their energy densities are subdominant with respect to that of the Universe during inflation and radiation era. In particular, the existence of a uniform magnetic field at the time of nucleosynthesis is compatible with astrophysical data if its energy density is much smaller than about $1.25$ times the energy density of neutrinos." }, "1112/1112.0303_arXiv.txt": { "abstract": " ", "introduction": "\\label{sec:intro} \\subsection{Background and Motivation} The recent discovery of the accelerating expansion of the Universe \\cite{Perlmutter1999, Riess1998} has prompted many theoretical speculations about the underlying mechanism. The most likely mechanism is a cosmological constant, which is the simplest model and is in good agreement with observational data \\cite{Blake2011}. More complicated models involve new dynamical sources of gravity that act as dark energy, and/or modifications to general relativity on large scales. A plethora of models have been postulated and explored in recent years, including Quintessence, K-essence \\cite{Armendariz-Picon2000, Chiba2000}, Ghost Condensates \\cite{Arkani-Hamed2004}, DGP gravity \\cite{Dvali2000}, and $f(R)$ gravity, to name but a few. See Refs.\\ \\cite{Peebles2003, Nobbenhuis2006, Copeland2006, Caldwell:2009ix,Silvestri:2009hh,Skordis2011,Amendola2010} for detailed reviews of these and other models. A common feature of the majority of dark energy and modified gravity models is that in the low energy limit, they are equivalent to general relativity coupled to one or more scalar fields, often called quintessence fields. Therefore it is useful to try to construct very general low energy effective quantum field theories of general relativity coupled to light scalar fields, in order to encompass broad classes of dark energy models. Considering dark energy models as quantum field theories is useful, even though the dynamics of dark energy is likely in a classical regime, because it facilitates discriminating against theories which are theoretically inconsistent or require fine tuning. A similar situation occurs in the study of models of inflation, where it is useful to construct generic theories using effective field theory. Cheung \\textit{et al.} \\cite{Cheung2008} constructed a general effective field theory for gravity and a single inflaton field, for perturbations about a background Friedman-Robertson-Walker cosmology in unitary gauge. This work was later generalized in multiple directions \\cite{Senatore:2010wk,2011arXiv1106.2189F} and has been very useful. An alternative approach to single field inflationary models was taken by Weinberg \\cite{Weinberg2008}, who constructed an effective field theory to describe both the background cosmology and the perturbations. This theory consisted at leading order of a standard single field inflationary model with a potential, together with higher order terms in a covariant derivative expansion up to four derivatives. More detailed discussions of this type of effective field theory were given by Burgess, Lee and Trott \\cite{Burgess:2009ea}. When one turns from inflationary effective field theories to quintessence effective field theories, the essential physics is very similar, but there are three important differences that arise: \\begin{itemize} \\item First, the hierarchy of scales is vastly more extreme in quintessence models. The Hubble parameter $H$ is typically several orders of magnitude below the Planck scale $\\mpl \\sim 10^{28}$ eV in inflationary models, whereas for quintessence models $H_0 \\sim 10^{-33} \\, $eV is $\\sim 60$ orders of magnitude below the Planck scale. Quintessence fields must have a mass that is smaller than or on the order of $H_0$. It is a well-known, generic challenge for quintessence models to ensure that loop effects do not give rise to a mass much larger than $H_0$. Because of the disparity of scales, this issue is more extreme for quintessence models than inflationary models. \\item In most inflationary models, it is assumed that the dynamics of the Universe are dominated by gravity and the scalar field (at least until reheating). By contrast, for quintessence models in the regime of low redshifts relevant to observations, we know that cold dark matter gives an $O(1)$ contribution to the energy density. Therefore there are additional possible couplings and terms that must be included in an effective field theory. \\item For any effective field theory, it is possible to pass outside the domain of validity of the theory even at energies $E$ low compared to the theory's cutoff $\\Lambda$, if the mode occupation numbers $N$ are sufficiently large (see Sec.\\ \\ref{sec:validity} below for more details). This corresponds to a breakdown of the classical derivative expansion. For quintessence theories, mode occupation numbers today can be as large as $N \\sim (\\mpl/H_0)^2$ and it is possible to pass outside the domain of validity of the theory. By contrast in inflationary models, this is less likely to occur since mode occupation numbers for the perturbations are not large before modes exit the horizon. Thus, the effective field theory framework is less all-encompassing for quintessence models than for inflation models. This issue seems not to have been appreciated in the literature and we discuss it in Sec. \\ref{sec:validity} below. \\end{itemize} Several studies have been made of generic effective field theories of dark energy. Creminelli, D'Amico, Nore\\~{n}a and Vernizzi \\cite{Creminelli2009} constructed a the general effective theory of single-field quintessence for perturbations about an arbitrary FRW background, paralleling the similar construction for inflation \\cite{Cheung2008}. Park, Watson and Zurek constructed an effective theory for describing both the background cosmology and the perturbations, following the approach of Weinberg \\cite{Weinberg2008} but generalizing it to include couplings to matter \\cite{Watson2010}. The two approaches to effective field theories of quintessence -- specialization to perturbations about a specific background, and maintaining covariance and the ability to describe the dynamics of a variety of backgrounds -- are complementary to one another. The dynamics of the cosmological background FRW solution can be addressed in the covariant approach of Weinberg, but not in the background specific approach of Creminelli et al., which restricts attention to the dynamics of perturbations about a given, fixed background. On the other hand, a background specific approach can describe a larger set of dynamical theories for the perturbations than can a covariant derivative expansion\\footnote{To see this, consider for example a term in the Lagrangian of the form $f(\\phi) (\\nabla \\phi)^{2 n}$, where $\\phi$ is the quintessence field. Such a term would be omitted in the covariant derivative expansion for sufficiently large $n$. However, upon expanding this term using $\\phi = \\phi_0 + \\delta \\phi$, where $\\phi_0$ is the background solution, one finds terms $\\sim (\\nabla \\phi_0)^{2n-2} (\\nabla \\delta \\phi)^2$ which are included in the Creminelli approach of applying standard effective field theory methods to the perturbations.}. \\subsection{Approach and Assumptions} \\label{sec:approach} The purpose of this paper is to revisit, generalize and correct slightly the covariant effective field theory analysis of Park, Watson and Zurek \\cite{Watson2010}. Following Weinberg and Park et al., we restrict attention to theories where the only dynamical degrees of freedom are a graviton and a single scalar. We allow couplings to an arbitrary matter sector, but we assume the validity of the weak equivalence principle, motivated by the strong experimental evidence for this principle. We assume that the theory consists of a standard quintessence theory coupled to matter at leading order in a derivative expansion, with an action of the form \\begin{align} S[g_{\\alpha\\beta},\\phi,\\psi_{\\rm m}] = \\int d^4 x \\sqrt{-g} \\left\\{ \\frac{\\mpl^2}{2} R - \\frac{1}{2} (\\nabla \\phi)^2 - U(\\phi) \\right\\} + S_{\\rm m} \\left[ e^{\\alpha(\\phi)} g_{\\mu \\nu}, \\psi_{\\rm m}\\right]. \\label{eq:intro:leadingorder} \\end{align} Here $\\psi_{\\rm m}$ denotes a set of matter fields, and $\\mpl$ is the Planck mass. The factor $e^{\\alpha(\\phi)}$ in the matter action provides a leading-order non-minimal coupling of the quintessence field to matter, in a manner similar to Brans-Dicke models in the Einstein frame. Our analysis then consists of a series of steps: \\begin{enumerate} \\item We add to the action all possible terms involving the scalar field and metric, in a covariant derivative expansion up to four derivatives. We truncate the expansion at four derivatives, as this is sufficient to yield the leading corrections to the action (\\ref{eq:intro:leadingorder}). As described by Weinberg \\cite{Weinberg2008} there are ten possible terms, with coefficients that can be arbitrary functions of $\\phi$ [see Eq.\\ (\\ref{eq:S1}) below]. Section \\ref{sec:scaling} below describes one possible justification of this covariant derivative expansion from an effective field theory viewpoint, starting from a set of ultralight pseudo Nambu-Goldstone bosons (PNGBs). It is likely that the same expansion can be obtained from other, more general starting points. \\item We allow for corrections to the coupling to matter by adding to the metric that appears in the matter action all possible terms involving the metric and $\\phi$ allowed by the derivative expansion, that is, up to two derivatives. There are six such terms [see Eq.\\ (\\ref{eq:JordanMetric}) below.] We also add to the action terms involving the stress energy tensor $T_{\\mu\\nu}$ of the matter fields, up to the order allowed by the derivative expansion using $T_{\\mu\\nu} \\sim \\mpl^2 G_{\\mu\\nu}$ [see Eq.\\ (\\ref{eq:S1}) below]. Including such terms in the action seems poorly motivated, since {\\it a priori} there is no reason to expect that the resulting theory would respect the weak equivalence principle (see Appendix \\ref{app:WEP}). However we show in Appendix \\ref{app:WEP} that the weak equivalence principle is actually satisfied, to the order we are working to in the derivative expansion. In addition, all the terms in the action involving $T_{\\mu\\nu}$ can be shown to have equivalent representations not involving the stress energy tensor, using field redefinitions (see Appendix \\ref{app:WEP}). \\item The various correction terms are not all independent because of the freedom to perform field redefinitions involving $\\phi$, $g_{\\mu\\nu}$ and the matter fields, again in a derivative expansion. In Sec.\\ \\ref{sec:techs} we explore the space of such field redefinitions, finding eleven independent transformations and tabulating their effects on the coefficients in the action (see Table \\ref{tab:transformations} below). \\item Several of the correction terms that are obtained from the derivative expansion are ``higher derivative'' terms, by which we mean that they give contributions to the equations of motion which involve third-order or higher-order time derivatives of the fields\\footnote{The precise definition of higher derivative that we use, which is covariant, is that an equation will be said not to contain any higher derivative terms if there exists a choice of foliation of spacetime for which any third-order or higher-order derivatives contain at most two time derivatives. Theories which are higher derivative in this sense are generically associated with instabilities (Ostragradski's theorem) \\cite{Woodard2007}, although the instabilities can be evaded in special cases, for example $f(R)$ gravity. For most of this paper (except for the Chern-Simons term), a simpler definition of higher derivative would be sufficient: a term in the action is ``higher derivative'' if it gives rise to terms in the equation of motion that involve any third-order or higher order derivatives.}. Normally, such higher derivative terms give rise to additional degrees of freedom. However, if they are treated perturbatively (consistent with our derivative expansion) additional degrees of freedom do not arise. Specifically, one can perform a {\\it reduction of order} procedure on the equations of motion \\cite{1971ctf..book.....L,Parker:1993dk,Flanagan:1996gw}, substituting the zeroth-order equations of motion into the higher derivative terms in the equations of motion to eliminate the higher derivatives\\footnote{This is more general than requiring the solutions of the equation of motion to be analytic in the expansion parameter, as advocated by Simon \\protect{\\cite{Simon1990}}; see Ref.\\ \\protect{\\cite{Flanagan:1996gw}}.}. We actually use a slightly different but equivalent procedure of eliminating the higher derivative terms directly in the action using field redefinitions\\footnote{This procedure is counterintuitive since normally field redefinitions do not change the physical content of a theory; here however they do because the field redefinitions themselves involve higher derivatives.} (see Appendix \\ref{app:backsubs}). Weinberg \\cite{Weinberg2008} and Park {\\it et al.} \\cite{Watson2010} use a slightly different method, consisting of substituting the leading order equations of motion directly into the higher derivative terms in the action. This method is not generally valid, but it is valid up to field redefinitions that do not involve higher derivatives, and so it suffices for the purpose of attempting to classify general theories of dark energy (see Appendix \\ref{app:backsubs}). \\item Another issue that arises with respect to the higher derivative terms is the following. Is it really necessary to include such terms in an action when trying to write down the most general theory of gravity and a scalar field, in a derivative expansion? Weinberg \\cite{Weinberg2008} suggested that perhaps a more general class of theories is generated by including these terms and performing a reduction of order procedure on them, rather than by omitting them. However, since it is ultimately possible to obtain a theory that is perturbatively equivalent to the higher derivative theory, and which has second order equations of motion, it should be possible just to write down the action for this reduced theory. In other words, an equivalent class of theories should be obtained simply by omitting all the higher derivative terms from the start. We show explicitly in Sec.\\ \\ref{sec:details} that this is the case for the class of theories considered here. \\item We fix the remaining field redefinition freedom by choosing a ``gauge'' in field space, thus fixing the action uniquely (see Sec. \\ref{sec:details:final}). \\end{enumerate} \\subsection{Results and Implications} Our final action is [Eq.\\ (\\ref{eq:details:finalaction}) below] \\begin{align} S =& \\int d^4 x \\sqrt{-g} \\left\\{ \\frac{\\mpl^2}{2} R - \\frac{1}{2} (\\nabla \\phi)^2 - U(\\phi)\\right\\} + S_{\\rm m}[e^{\\alpha(\\phi)} g_{\\alpha\\beta},\\psi_{\\rm m}] \\nonumber \\\\ & + \\epsilon \\int d^4 x \\sqrt{-g} \\bigg\\{a_1 (\\nabla \\phi)^4 + b_2 T (\\nabla \\phi)^2 + c_1 G^{\\mu \\nu} \\nabla_\\mu \\phi \\nabla_\\nu \\phi \\nonumber \\\\ & + d_3 \\left( R^2 - 4 R^{\\mu \\nu} R_{\\mu \\nu} + R_{\\mu \\nu \\sigma \\rho} R^{\\mu \\nu \\sigma \\rho} \\right) + d_4 \\epsilon^{\\mu \\nu \\lambda \\rho} \\tensor{C}{_{\\mu \\nu}^{\\alpha \\beta}} C_{\\lambda \\rho \\alpha \\beta} + e_1 T^{\\mu \\nu} T_{\\mu \\nu} + e_2 T^2 + \\ldots \\bigg\\}. \\label{eq:details:finalaction0} \\end{align} Here the coefficients $a_1$, $b_2$ etc.\\ of the next-to-leading order terms in the derivative expansion are arbitrary functions of $\\phi$, and the ellipsis $\\ldots$ refers to higher order terms with more than four derivatives. The corresponding equations of motion do not contain any higher derivative terms. This result generalizes that of Weinberg \\cite{Weinberg2008} to include couplings to matter. \\begin{figure}[h!] \\centering \\ifx\\dofigures\\undefined \\else \\includegraphics[width=0.55\\textwidth]{graph3.pdf} \\fi \\caption{\\sl The parameter space of fractional density perturbation $\\delta \\rho/\\rho$ for perturbations to the quintessence field, and cutoff scale $M$ for the effective field theory, illustrating the constraint (\\ref{eq:ddd}) on the domain of validity. Near the boundary of the domain of validity the higher derivative terms in the action are potentially observable, this is labeled the ``interesting regime''. Further away from the boundary the higher derivative terms are negligible and the theory reduces to a standard quintessence model with a matter coupling.} \\label{fig:scales0} \\end{figure} We can summarize our key results as follows: \\begin{itemize} \\item The most general action contains nine free functions of $\\phi$: $U, \\alpha, a_1, b_2, c_1, d_3, d_4, e_1, e_2$, as compared to the four functions that are needed when matter is not present \\cite{Weinberg2008}. \\item There are a variety of different forms of the final theory that can be obtained using field redefinitions. In particular some of the matter-coupling terms in the action can be re-expressed as terms that involve only the quintessence field and metric. Specifically, the term $T (\\nabla \\phi)^2$ term could be eliminated in favor of $\\square \\phi (\\nabla \\phi)^2$, the $(\\nabla \\phi)^4$ could be eliminated in favor of a term $T^{\\mu\\nu} \\nabla_\\mu \\phi \\nabla_\\nu \\phi$, or the $G^{\\mu\\nu} \\nabla_\\mu \\phi \\nabla_\\nu \\phi$ term could be eliminated in favor of a term $T^{\\mu\\nu} \\nabla_\\mu \\phi \\nabla_\\nu \\phi$ (see Sec.\\ \\ref{sec:details:final}). \\item As mentioned above, one obtains the correct final action if one excludes throughout the calculation all higher derivative terms. \\item The final theory does contain terms involving the matter stress-energy tensor. Nevertheless, the weak equivalence principle is still satisfied (see Appendix \\ref{app:WEP}). It is possible to eliminate the stress-energy terms, but only if we allow higher derivative terms in the action (where it is assumed that the reduction of order procedure will be applied to these higher derivative terms). Thus, for a fully general theory, one must have either stress-energy terms or higher derivative terms; one cannot eliminate both (see Sec.\\ \\ref{sec:details:final}). \\item We can estimate how all the coefficients $a_1$ etc. scale with respect to a cutoff scale $M$ for an effective field theory as follows (see Sec.\\ \\ref{sec:scaling}). We assume that several ultralight scalar fields of mass $\\sim H_0$ arise as pseudo Nambu-Goldstone bosons (PNGBs) from some high energy theory \\cite{1995PhRvL..75.2077F,2003JCAP...07..003A}, and are described by a nonlinear sigma model at low energies. We then suppose that all but one of the these PNGB fields have masses $M$ that are somewhat larger than $\\sim H_0$, and integrate them out. This will give rise to a theory of the form discussed above for the single light scalar, where the higher derivative terms are suppressed by powers of $M$. The scalings for each of the coefficients in the action are summarized in Table \\ref{tab:scalings}. We find that the fractional corrections to the cosmological dynamics due to the higher derivative terms scale as $H_0^2/M^2$, as one would expect. \\item Finally, we can use these scalings to estimate the domain of validity of the effective field theory (see Sec.\\ \\ref{sec:validity}). We find that cosmological perturbations with a density perturbation $\\delta \\rho$ in the quintessence field must have a fractional density perturbation that satisfies \\begin{align} \\frac{\\delta \\rho}{\\rho} \\ll \\frac{M^2}{H_0^2}. \\label{eq:ddd} \\end{align} Thus perturbations can become nonlinear, but only modestly so, if $M$ is close to $H_0$. The parameter space of fractional density perturbation $\\delta \\rho/\\rho$ and cutoff scale $M$ is illustrated in Fig.\\ \\ref{fig:scales0}. In addition there is the standard constraint for derivative expansions \\begin{align} E \\ll M \\label{eq:ddd1} \\end{align} where $E^{-1}$ is the length-scale or time-scale for some process. We show in Fig.\\ \\ref{fig:scales1} the two constraints (\\ref{eq:ddd}) and (\\ref{eq:ddd1}) on the two dimensional parameter space of energy $E$ and mode occupation number $N$. \\end{itemize} \\begin{figure}[h!] \\centering \\ifx\\dofigures\\undefined \\else \\includegraphics[width=0.80\\textwidth]{graph2.pdf} \\fi \\caption{\\sl The domain of validity of the effective field theory in the two dimensional parameter space of energy $E$ per quantum of a mode of the quintessence field, and mode occupation number $N$. The cutoff scale $M$ must be larger than the Hubble parameter $H_0$ in order that the background cosmology lie within the domain of validity. Perturbation modes on length-scales that are small compared to $H_0^{-1}$ but large compared to $M^{-1}$ can be described, but only if the mode occupation number and fractional density perturbation are sufficiently small. See Sec.\\ \\ref{sec:validity} for details.} \\label{fig:scales1} \\end{figure} Finally, in Appendix \\ref{app:compare} we compare our analysis to that of Park, Watson and Zurek \\cite{Watson2010}, who perform a similar computation but in the Jordan frame rather than the Einstein frame (see also Ref.\\ \\cite{Jimenez:2011nn}). The main difference between our analysis and theirs is that they use a different method to estimate the scalings of the coefficients, and as a result their final action differs from ours, being parameterized by three free functions rather than nine. \\ifx\\doejects\\undefined \\else \\eject \\fi ", "conclusions": "\\label{sec:conclusion} In this paper, we have investigated effective field theory models of cosmic acceleration involving a metric and a single scalar field. The set of theories we considered consists of a standard quintessence model with matter coupling, together with a general covariant derivative expansion, truncated at four derivatives. We showed that this class of theories can be obtained from a PNGB scenario, where one of the PNGB fields is lighter than all the others, and the heavier fields are integrated out. We showed that in constructing this class of theories, including higher derivative terms in the action, as suggested by Weinberg \\cite{Weinberg2008}, does not give any increased generality. We also showed that complete generality requires one to include terms in the action that depend on the stress-energy tensor of the matter fields. We now turn to a discussion of some of the advantages and shortcomings of the approach adopted here to describe models of dark energy. Some of the shortcomings are: \\begin{itemize} \\item By construction, our approach excludes theories where nonlinear kinetic terms in the action give an order unity contribution to the dynamics, such as K-essence, ghost condensates etc., since such theories do not arise from the PNGB construction used in this paper. On the other hand, such theories are less natural than the class of theories considered here, from the point of view of loop corrections: they require very nontrivial physics at the scale $\\sim H_0$, instead of at the scale $\\sim \\sqrt{H_0 \\mpl}$ required in the PNGB approach. The most general class of theories of this kind is that of Horndeski \\cite{Horndeski1974}, which contains four free functions of $\\phi$ and $(\\nabla \\phi)^2$ \\cite{Skordis2011}, and which is the most general class of theories of a metric and a scalar field for which the equations of motion are second order. As discussed in the Introduction, these theories are included in the alternative, background-dependent approach to effective field theories of quintessence of Creminelli et al. \\cite{Creminelli2009}. \\item Our class of theories will be observationally distinguishable from vanilla quintessence theories only if the cutoff $M$ is near the Hubble scale $H_0$. In this regime, our framework cannot be used to analyze Solar System tests of general relativity, since they are outside the domain of validity of the effective field theory. Also, when the background cosmology is evolved backwards in time it passes outside the domain of validity at fairly low redshifts. (This is not a serious disadvantage since dark energy dominates only at low redshifts.) \\item We have restricted attention to theories with a metric and a single scalar field, with the only symmetry being general covariance. Thus, our analysis does not include models with several scalar fields, vector fields etc. In addition, our analysis excludes an interesting class of models that one obtains by imposing that the action be invariant under $\\phi \\to f(\\phi)$, where $f$ is any monotonic function, as such a symmetry cannot be realized with our derivative expansion. This class of models includes Horava-Lifshitz gravity and has the same number of physical degrees of freedom as general relativity \\cite{Skordis2011,Blas:2010hb}. It would be interesting to explore the most general dark energy models of this kind. \\end{itemize} Some of the advantages of the approach used here are: \\begin{itemize} \\item Our class of theories is generic within the PNGB construction, which itself is a well motivated way to obtain the ultralight fields needed for cosmic acceleration. The theories are fairly simple and it should be straightforward to confront them with observational data. \\item Our class of theories allow for a unified treatment of the cosmological background and perturbations, unlike the background-dependent approach of Ref.\\ \\cite{Creminelli2009}. \\end{itemize} Finally, we list some possible directions in which the approach used here could be extended: \\begin{itemize} \\item It would be interesting to compute the relation between the nine free functions used in our theories to the free functions of the post-Friedmannian approach to parameterizing dark energy models \\cite{Skordis2011}. \\item It would be interesting to explore the phenomenology of the various higher order terms in our action, for the cosmological background evolution and perturbations. Many of the terms have already been explored in detail, see for example Refs.\\ \\cite{Deffayet:2010qz,Charmousis2011}. \\item Either by using the post-Friedmannian approach, or more directly, it would be useful to compute the current observational constraints on the free functions in the action. \\item An interesting open question is the extent to which our final action is generic. That is, is there a class of theories more general than nonlinear sigma model PNGB theories for which our action is obtained by integrating out some of the fields? \\end{itemize} \\subsubsection*" }, "1112/1112.2901_arXiv.txt": { "abstract": "{Supernova remnants (SNRs) have emerged as one of the largest source classes in very-high-energy (VHE; E$>$0.1\\,TeV) astronomy. Many of the now known VHE $\\gamma$-ray emitting SNRs have been discovered by the H.E.S.S. imaging Cherenkov telescope array, thanks to its unique access to the inner galaxy. Statistically-significant emission of VHE $\\gamma$-rays has now been detected from the direction of the supernova remnant \\g. While the centroids of the H.E.S.S. source and the shell-type SNR are compatible, the VHE morphology suggests a center-dominated source at TeV energies, something which is at odds with the shell-like morphology observed at radio frequencies. This suggests that H.E.S.S. may be observing TeV emission from a previously unknown pulsar wind nebula (PWN) located within the boundaries of the radio shell. If this interpretation is correct, \\g\\ would in fact be a composite SNR, the first case in which an SNR is identified as a composite on the basis of VHE $\\gamma$-ray observations. Archival data from MAGPIS gives exciting hints that there is radio emission from the central parts of the remnant, giving support to this hypothesis. Unfortunately, image artefacts from a nearby strong radio source produce considerable uncertainties in the radio analysis. Additional observations in both the radio and X-ray are needed to confirm the composite nature of \\g\\ suggested by H.E.S.S.} ", "introduction": "Very-high-energy (VHE; E$>$0.1\\,TeV) $\\gamma$-ray emission has now been detected from more than 60 Galactic sources, and the vast majority of these are in some way connected to supernova remnants (SNRs) \\cite{Gast2011}. TeV $\\gamma$-rays have been observed from SNR shells themselves and also from the pulsar wind nebulae (PWNe) at the center of many SNRs. While the former gives rise to VHE $\\gamma$-ray emission tracing the shock-front of the SNR, the latter shows a centre-filled morphology, i.e. a central nebula, generally visible from low-energy radio wavelengths up to $\\gamma$-ray photon energies of tens of TeV in the most extreme cases. Young PWNe typically have a rather symmetric morphology and are centrally located inside the host SNR. Such PWNe show a distinct spectral fingerprint with a roughly flat radio spectral index and a much steeper spectrum at higher energies \\cite{pwnevolution}. The pulsar powering the nebula is usually found close to its birthplace at the centre of the SNR in such systems. However, only in a few cases has the pulsar actually been detected. One prominent example of a young PWN is found in the SNR G\\,0.9$+$0.1 which was first classified as a composite SNR on the basis of observations in radio \\cite{helfland1987} then subsequently detected in TeV $\\gamma$-rays \\cite{hessg0.9}, before ultimately the energetic pulsar was found in a deep radio observation of the source \\cite{radiog0.9}. The SNR \\g\\ is a poorly studied object which was initially discovered in a 90-cm survey of the inner Galaxy conducted by the Very Large Array (VLA) \\cite{brogan2006}. The SNR was reported to have a shell-like morphology with a size about $14^{\\prime} \\times 15^{\\prime}$ and an average spectral index $\\alpha = -0.6 \\pm 0.2$, indicating that the radio emission is dominated by non-thermal synchrotron emission from the shell. For the reasons outlined above, this source is a prime candidate for H.E.S.S. observations. Furthermore, since \\g\\ is larger than the H.E.S.S. point-spread-function (PSF; $\\sim6^{\\prime}$), morphological studies with H.E.S.S. have the capability to distinguish between a shell-type SNR or PWN-like origin of the VHE $\\gamma$-ray emission, thus shedding more light on this little known SNR. ", "conclusions": "Since a chance positional coincidence of the centroid of the VHE $\\gamma$-ray emission with the center of the radio shell of \\g\\ seems highly unlikely, an association between \\hessj\\ and the SNR is assumed here. The question is then whether the VHE emission originates from the shell of the SNR or from a putative PWN embedded inside the SNR. However, since the extension of the H.E.S.S. source is significantly smaller than the extension of the SNR as seen in the archival VLA data, it is clear that the bulk of the VHE emission originates from a yet unknown source located within the shell. The discovery of a radio or X-ray PWN counterpart to \\hessj\\ would confirm this scenario. High-quality X-ray observations of the region are unfortunately lacking, but preliminary analyses of 20-cm and 90-cm radio observations suggest that the SNR is composed of two distinct regions with different spectra. The central region is found to be characterised by a relatively flat radio spectrum, characteristic of the central PWN of a composite SNR \\cite{pwnevolution}, while the shell is characterised by a steeper spectrum, indicative of the non-thermal synchrotron emission expected from an SNR shell. \\g\\ thus appears to be an example of a composite SNR, similar to G\\,0.9$+$0.1, the first composite SNR to be detected at TeV energies. If the nature of \\g\\ can be confirmed, \\hessj\\ would be the first case in which a SNR is identified as a composite SNR on the basis of VHE $\\gamma$-ray observations. Additional supporting evidence for a PWN origin of the TeV emission would be the detection of a pulsar within the central regions of the SNR. Unfortunately, no pulsar is currently known, in radio or X-rays, inside the shell of \\g, nor are there any known hard X-ray point sources. However, a typical young radio pulsar located at $\\sim$9\\,kpc with an assumed luminosity L$_{21\\mathrm{\\,cm}} \\sim 56$\\,mJy\\,kpc$^2$ (i.e. similar to the median of high-luminosity, young, rotation-powered pulsars estimated by \\cite{camilio}) would have a flux density of S$_{21\\mathrm{\\,cm}} \\sim 0.08$\\,mJy, considerably lower than the sensitivity of the current VLA data. Beaming effects could also make the pulsar virtually undetectable in radio. If future observations of \\g\\ are able to detect the putative central pulsar associated with \\hessj, it is expected to have a very high spin-down luminosity ($\\dot{\\mathrm{E}} \\sim 4 \\times 10^{37}$\\,erg\\,s, as seen in other TeV $\\gamma$-ray emitting PWNe \\cite{radiog0.9}. Little is known about the distance and age of \\g. Using the uncertain $\\Sigma-D$ relationship \\cite{sigmaD}, a distance of $10 \\pm 3$\\,kpc is derived. Adopting 10\\,kpc as the nominal distance of the SNR would imply a Sedov age of $\\sim$9\\,kyr, assuming an ambient ISM density of 1\\,cm$^{-3}$. However, varying the density by $\\pm$50\\% and taking into account uncertainties in the $\\Sigma-D$ relationship, an age as low as 4\\,kyr can be derived. These age estimates would make SNR \\g\\ older than similar systems previously detected in VHE $\\gamma$-rays." }, "1112/1112.5197_arXiv.txt": { "abstract": "A plausible scenario for the gamma ray and the hard x-ray burst in a strongly magnetized plasma, based on the collective plasma maser instability, is proposed. The physical parameters with which this scenario becomes relevant are estimated. The attractive feature of this scenario over the conventional cyclotron radiation theory is discussed. ", "introduction": "According to a recent study, % the thermal electron gyro-motion in a strongly magnetized plasma could lead to instabilities of electromagnetic (E\\&M) waves~\\citep{sonmaxwell}. This instability, so-called the maser instability, has significant implications on strongly magnetized plasmas~\\citep{Stone,Biskamp, Innes,Fishman, Nakar, Horse}. The goal of this paper is to estimate its relevance to the gamma ray or hard x-ray burst in the strongly magnetized astrophysical plasmas~\\citep{Fishman, Nakar}. The gamma ray burst in the absence of the magnetic field~\\citep{Mendon\u00e7a} is not relevant to our analysis. Our analysis shows that the photons of 10 keV to 1 MeV can be emitted from a relativistic plasma with the electron density $ 10^{19} - 10^{26}~\\mathrm{cm^{-3}} $ and the magnetic field of order of $10^{8}$ to $10^{13}$ gauss. In contrast to the conventional cyclotron radiation where the photons are emitted rather uniformly, the angular distribution of the radiated photons could be sharply concentrated. Advantages of this possible phenomenon over the ones based on the incoherent cyclotron radiation are discussed. The rest of the paper is organized as follows. The instability growth rate is presented in Sec. 2, based on the Landau damping theory, which is analyzed when the photon wave vector is parallel to the magnetic field (Sec. 3) and when the photon wave vector is not parallel to the magnetic field (Sec. 4). The gamma ray burst is discussed in Sec. 5, and the difference between this theory and the conventional theories is presented Sec. 6, which is followed by summary (Sec. 7). ", "conclusions": "A scenario of the gamma ray burst, based on the recent radiation theory~\\citep{sonmaxwell}, is proposed and examined. The previous analysis~\\citep{sonmaxwell} is generalized to an arbitrary angle. The estimation shows that the coherent burst of 10 keV to 1 MeV photons is plausible in a relativistic plasma when $\\gamma_0 = 10\\sim 1000$ and the electron density is higher than some critical value, $n_e > 10^{18} \\sim 10^{26} \\ \\mathrm{cm}^{-3}$. It is shown that a rather compact dense object with less available energy could cause the short gamma ray burst and that the observed gamma rays would be coherent rather than incoherent. In addition, the coherent photons of lower frequency comparable to the plasma frequency might be observed due to the Weibel instability. This is particularly relevant to the case when $\\theta = \\pi/2$ because the low frequency coherent photons (high frequency coherent photons) from the Weibel instability (from the instability studied here) can be observed simultaneously. The above conjecture might be useful in verifying whether the scenario proposed here would account for some of the short gamma ray burst events observed in the satellites~\\citep{Nakar}. While our estimation is rather focused on the gamma and hard x-rays, a similar mechanism would manifest in generating soft x-rays in the inertial confinement fusion plasma~\\citep{tabak}. The electron beam of $\\gamma > 10 \\sim 100$ and the magnetic field of $10^8$ gauss can be readily generated in laboratories. Even a magnetic field of $10^9$ gauss might be possible~\\citep{sonprl}. Then, the photon generated from the instability may have energy between 10 eV and 1 keV. Complications would be the electron quantum diffraction effect and the degeneracy~\\citep{sonpla, sonprl, sonlandau}. The plausibility study is in progress." }, "1112/1112.3122_arXiv.txt": { "abstract": "The initial shear field, characterized by a primordial perturbation potential, plays a crucial role in the formation of large scale structures. Hence, considerable analytic work has been based on the joint distribution of its eigenvalues, associated with Gaussian statistics. In addition, directly related morphological quantities such as ellipticity or prolateness are essential tools in understanding the formation and structural properties of halos, voids, sheets and filaments, their relation with the local environment, and the geometrical and dynamical classification of the cosmic web. To date, most analytic work has been focused on Doroshkevich's unconditional formulae for the eigenvalues of the linear tidal field, which neglect the fact that halos (voids) may correspond to maxima (minima) of the density field. I present here new formulae for the constrained eigenvalues of the initial shear field associated with Gaussian statistics, which include the fact that those eigenvalues are related to regions where the source of the displacement is positive (negative): this is achieved by requiring the Hessian matrix of the displacement field to be positive (negative) definite. The new conditional formulae naturally reduce to Doroshkevich's unconditional relations, in the limit of no correlation between the potential and the density fields. As a direct application, I derive the individual conditional distributions of eigenvalues and point out the connection with previous literature. Finally, I outline other possible theoretically- or observationally-oriented uses, ranging from studies of halo and void triaxial formation, development of structure-finding algorithms for the morphology and topology of the cosmic web, till an accurate mapping of the gravitational potential environment of galaxies from current and future generation galaxy redshift surveys. ", "introduction": "The large-scale spatial distribution of dark matter, as revealed from numerical simulations, shows a characteristic anisotropic web-like structure. This \\textit{cosmic web}, which arises through the gravitational clustering of matter, is mainly due to the effects of the tidal field: in fact, the competition between cosmic expansion, the trace, and the traceless part of the tidal field imprints anisotropies in the large-scale matter distribution in much the same way that gravity and radiation pressure imprints baryonic acoustic oscillations (BAO) on the Cosmic Microwave Background (CMB) sky (Hu \\& Sugiyama 1995; Lee \\& Springel 2010). Hence, the initial shear field plays a crucial role in the formation of large scale structures, and a number of studies in the literature have been devoted to this subject -- among the plethora of papers, see for example the classic works by Zeldovich (1970), Icke (1973), Peebles (1980), White (1984), Bardeen et al. (1986), Kaiser (1986), Bertschinger (1987), Bond \\& Myers (1996), Bond, Kofman \\& Pogosyan (1996) and van de Weygaert \\& Bertschinger (1996). In addition, if the cosmic web originates from primordial tidal effects and its degree of anisotropy increases with the evolution of the Universe, then studying this initial field is crucial in understanding the subsequent nonlinear evolution of cosmic structures (Springel et al. 2005; Shandarin et al. 2006; Desjacques 2008; Desjacques \\& Smith 2008; Pogosyan et al. 2009), the alignment of shape and angular momentum of halos (West 1989; Catelan et al. 2001; Lee \\& Springel 2010; Rossi, Sheth \\& Tormen 2011), the statistical properties of voids (Lee \\& Park 2006; Platen, van de Weygaert \\& Jones 2008), and more generally for characterizing the geometry and morphology of the cosmic web (Shen et al. 2006; van de Weygaert \\& Bond 2008; Forero-Romero et al. 2009; Aragon-Calvo et al. 2010a,b; Shandarin et al. 2010). The basic theory for the formation and evolution of structures is now well-understood, thanks to the pioneering work of Doroshkevich \\& Zeldovich (1964), Doroshkevich (1970), Zeldovich (1970), and Sunyaev \\& Zeldovich (1972) -- the latter in the context of galaxy formation. In particular, Doroshkevich (1970) derived the joint probability distribution of an ordered set of eigenvalues in the tidal field matrix -- at random positions -- given the variance of the density field, corresponding to a Gaussian potential; we will refer to it as the \\textit{unconditional} probability distribution of eigenvalues. In addition, Zeldovich (1970) provided the fundamental understanding of anisotropic collapse on cosmological scales, and recognized the key role of the large scale tidal force in shaping the cosmic web. Subsequently, Doroshkevich and Shandarin (1978) calculated some statistical properties of the maxima of the largest eigenvalue of the shear tensor. Their study proved that the most probable formation process starts first with a one-dimensional collapse (cosmic pancake formation). The directions (orientations) for the one-dimensional collapsed sheets are determined by the largest eigenvalue of the deformation tensor, which can be attributed to the initial linear density perturbations; moreover, the probability that two or even three of the initial eigenvalues are identical or nearly equal is extremely small, indicating that the collapse is triaxial. Later on, following a constrained field approach pioneered by Bertschinger (1987), van de Weygaert \\& Bertschinger (1996) developed an algorithm for setting up tailor-made initial conditions for cosmological simulations, which addressed the role of the tidal fields in shaping large-scale structures. In the same period, Bond, Kofman \\& Pogosyan (1996) developed a cosmic web theory which naturally explains the filamentary structure present in the Cold Dark Matter (CDM) cosmology, due to the coherent nature of the primordial tidal field. They realized that an ``embryonic'' cosmic web is already present in the primordial density field, and explained why in overdense regions sheet-like membranes are only marginal features. Since then, because of the correspondence between structures in the evolved density field and local properties of the linear tidal field pointed out by the same authors, the statistics of the shear has received more attention in the literature. For example, Lee \\& Shandarin (1998) computed some probability distributions for individual shear eigenvalues and obtained an analytic approximation to the halo mass function, and Catelan \\& Porciani (2001) explored the two-point correlation of the tidal shear components. However, a variety of studies in the physics of halo formation and cosmic web classification are based on Doroshkevich's unconditional formulae for the ordered eigenvalues of the initial shear field associated with Gaussian statistics (Doroshkevich 1970), but those formulas cannot differentiate between random positions and peak/dips as they neglect the fact that halos (voids) may correspond to maxima (minima) of the density field. According to Bardeen et al. (1986), if one assumes the cosmological density fluctuations to be Gaussian random fields, the local maxima of such fields are plausible sites for the formation of nonlinear structures. Hence, the statistical properties of the peaks can be used to predict the abundances and clustering properties of objects of various types, and in studies of the non-spherical formation of large-scale structures. The study of Bertschinger (1987) goes in this direction, by generalizing the treatment of Bardeen et al. (1986) and proposing a path integral method for sampling constrained Gaussian random fields. The method allows one to study the density field around peaks or other constrained regions in the biased galaxy formation scenario, but unfortunately it is too elaborate and inefficient in its implementation. Instead, by applying the prescription of Hoffman \\& Ribak (1991) to construct constrained random fields, van de Weygaert \\& Bertschinger (1996) were able to show that it is possible to generate efficiently initial Gaussian random density and velocity fields, and specify the presence and characteristics of one or more peaks and dips at arbitrary locations -- with the gravity and tidal fields at the site of the peaks having the required strength and orientation. Some other studies on constrained initial conditions were also pursued by van Haarlem \\& van de Weygaert (1993), and by van de Weygaert \\& Babul (1994). The idea has been expanded in Bond \\& Myers (1996), who presented a \\textit{peak-patch} picture of structure formation as an accurate model of the dynamics of peaks in the density field. Their approach goes further, as it involves the explicit formalism for identifying objects in a multiscale field (i.e. it does not restrict to a single scale). The model is even more precise for void patches, the equivalent framework for studying voids (Sahni et al. 1994; Sheth \\& van de Weygaert 2004; Novikov, Colombi \\& Dore 2006; Colberg et al. 2008). In general, density peaks define a well-behaved point-process which can account for the discrete nature of dark matter halos and galaxies, and on asymptotically large scales are linearly biased tracers of the dark matter field (Desjacques \\& Sheth 2010). Therefore, it would be desirable to incorporate the peak (or dip) constraint in the statistical description of the initial shear field, in order to characterize more realistically the geometry and dynamics of the cosmic web. The main goal of this paper is to do so, by providing a set of analytic expressions which extend the work of Doroshkevich (1970) and Bardeen et al. (1986), and are akin in philosophy to that of van de Weygaert \\& Bertschinger (1996). This is achieved by constraining the Hessian of the displacement field (the matrix of the second derivatives) to be positive (negative) definite, which is the case in the vicinity of minima (maxima) of the source of the displacement field. The new \\textit{conditional} probability distributions derived in this study include the correlation between the potential and density fields through a reduced parameter $r$, and naturally recover Doroshkevich's (1970) \\textit{unconditional} formulae in the absence of correlation. Hence, the main focus of this work is to derive explicitly the joint probability distribution of the eigenvalues of the shear field, given the fact that positions are peaks or dips in the corresponding density field -- and not random locations. In this sense the field is termed \\textit{constrained} (i.e. it has \\textit{constrained} eigenvalues, which are the result of looking only at peak/dip regions), and it is statistically described by a \\textit{conditional} probability distribution. Note that even though the formalism is essentially restricted to one scale, the formulae derived here are useful in a variety of applications -- some of which will be discussed at the end of this paper. The layout is organized as follows. Section \\ref{new_formula} provides the derivation of the new analytic expressions for the conditional distributions of eigenvalues of the initial shear field. In particular, Section \\ref{notation} illustrates the basic notation adopted; Section \\ref{joint_eigen_new} contains the derivation of the joint distribution of eigenvalues in the peak/dip picture, while Section \\ref{bbks_connection} shows the reverse conditional distribution function. As a direct application of the new formulae, Section \\ref{individual_eigenvalues} presents the individual distributions of eigenvalues subjected to the extremum constraint, along with some other related conditional probabilities. This part extends previous work by Lee \\& Shandarin (1998), which is briefly summarized in Appendix \\ref{Lee_Shandarin}, and complements the study of van de Weygaert \\& Bertschinger (1996). Finally, Section \\ref{conclusion} highlights the main results and discusses ongoing and future applications, which will be presented in forthcoming publications. ", "conclusions": "\\label{conclusion} Since the initial shear field associated with Gaussian statistics plays a major role in the formation of large scale structures, considerable analytic work has been based on the joint distribution of its eigenvalues -- i.e. Doroshkevich's formulae (Equation \\ref{doro_standard}). However, Doroshkevich's equations neglect the fact that halos (voids) may correspond to maxima (minima) of the density field. The main goal of this work was to provide new analytic expressions, in the context of the peak/dip picture (Bardeen et al. 1986; Bond et al. 1991), which include the fact that the eigenvalues of the linear shear field are related to regions where the source of the displacement is positive (negative). These new conditional probabilities, derived in Section \\ref{new_formula}, are Equations (\\ref{doro_inter_extended}), (\\ref{doro_eigen}), (\\ref{bbks_inter_extended}) and (\\ref{bbks_eigen}): they represent the main results of this paper. Written in Doroshkevich-like format, they naturally reduce to Doroshkevich's (1970) unconditional relations in the limit of no correlation between the potential and the density fields (i.e. when $r=0$). As a first direct application of (\\ref{doro_eigen}), the individual conditional distributions of eigenvalues were obtained in Section \\ref{individual_eigenvalues}; these relations extend some previous work by Lee \\& Shandarin (1998) -- see Appendix \\ref{Lee_Shandarin} for a compact summary of their results. Much more analytic work can be carried out using these new formulae, especially in connection with the statistics of peaks developed by Bardeen et al. (1986): their calculations can be extended within this framework (see Section \\ref{bbks_connection}), and results of this extension will be presented in a forthcoming publication -- along with more insights on the equations derived in Section \\ref{joint_eigen_new}. To obtain the conditional distribution $p(\\tilde{T}|r,\\tilde{H}>0)$ (i.e. Equation \\ref{what_we_want}), in principle one needs to sample numerically the probability (\\ref{cond}) as done in Lavaux \\& Wandelt (2010). However, the new analytic results of this paper suggest a simpler generalized excursion set algorithm, which will be also presented in a forthcoming study. The algorithm allows for a fast sampling of (\\ref{cond}), and permits to test the new relations derived here (i.e. Sections \\ref{new_formula} and \\ref{individual_eigenvalues}, and Appendix \\ref{Lee_Shandarin}) against mock data. In addition, along these lines it is possible to extend the main shape distributions involved in the triaxial formation of nonlinear structures (i.e. ellipticity $e$, prolateness $p$, axis ratios $\\nu$ and $\\mu$, etc.). Halos and voids are in fact triaxial rather than spherical (i.e. Rossi, Sheth \\& Tormen 2011), and the initial shear field plays a crucial role in their formation. For example, in the ellipsoidal collapse framework the virialization condition depends on ellipticity and prolateness, which are directly related to the eigenvalues of the external tidal field. Starting from $p(\\zeta_1,\\zeta_2,\\zeta_3|r)$, it is then straightforward to turn this joint conditional distribution into $p(e,p,\\delta|r)$ -- to include the peak constrain, -- and then to characterize $p(e|r)$, $p(p|r)$, $p(\\mu|r)$, $p(\\nu|r)$ and so forth. Applications to halos and voids will be also presented next, including implications for the skeleton of the cosmic web. The extension to non-Gaussian fields, along the lines of Lam et al. (2009), is a natural follow-up of this work and will be presented in a forthcoming publication as well. In the context of cosmic voids, other interesting uses of the new formulae (\\ref{doro_inter_extended}), (\\ref{doro_eigen}), (\\ref{bbks_inter_extended}) and (\\ref{bbks_eigen}) involve the Monge-Amp\\`{e}re-Kantorovitch reconstruction procedure -- see for example Lavaux \\& Wandelt (2010). The analytic framework described here can be also useful in several observationally-oriented applications, and in particular for developing algorithms to find and classify structures in the cosmic web. For example, Bond, Strauss \\& Cen (2010) presented an algorithm that uses the eigenvectors of the Hessian matrix of the smoothed galaxy distribution to identify individual filamentary structures. They used the distribution of the Hessian eigenvalues of the smoothed density field on a grid to study clumps, filaments and walls. Other possibilities include a web classification based on the multiscale analysis of the Hessian matrix of the density field (Aragon-Calvo et al. 2007), the skeleton analysis (Novikov et al. 2006), as well as a morphological (Zeldovich-based) classification (for instance Klypin \\& Shandarin 1983; Forero-Romero et al. 2009). More generally, the fact that the eigenvalues of the Hessian matrix can be used to discriminate different types of structure in a particle distribution is fundamental to a number of structure-finding algorithms (Forero-Romero et al. 2009), shape-finders algorithms (Sahni et al. 1998), and structure reconstruction on the basis of tessellations (Schaap \\& van de Weygaert 2000; Romano-Diaz \\& van de Weygaert 2007), etc. In addition, the classification of different environments should provide a framework for studying the environmental dependence of galaxy formation (see for example Blanton et al. 2005). To this end, recently Park, Kim and Park (2010) extended the concept of galaxy environment to the gravitational potential and its functions -- as the shear tensor. They studied how to accurately estimate the gravitational potential from an observational sample finite in volume, biased due to galaxy biasing, and subject to redshift space distortions, by inspecting the dependence of dark matter halo properties on environmental parameters (i.e. local density, gravitational potential, ellipticity and prolateness of the shear tensor). It would be interesting to interpret their results with the theoretical formalism developed here, and ultimately to study the gravitational potential directly from a real dataset such as the SDSS Main Galaxy sample within this framework. The formalism presented in this paper is restricted to one scale (i.e. peaks and dips in the density field, as in Bardeen et al. 1986), but the extension to a multiscale \\textit{peak-patch} approach along the lines of Bond \\& Myers (1996) is duable and subject of ongoing work. This will allow to account for the role of the peculiar gravity field itself, an important aspect not considered here but discussed for example in van de Weygaert \\& Bertschinger (1996). In fact, these authors introduced the peak constraints to describe the density field in the immediate surroundings of a peak, and then addressed the constraints on the gravitational potential perturbations; in particular, they constrained the peculiar gravitational acceleration at the position of the peak itself, in addition to characterizing the tidal field around the peak. Including all these effects in our formalism is ongoing effort. Finally, the interesting and more complex question of the local expected density field alignment/orientation distribution as a function of the local field value (or the other way around -- see Bond 1987; Lee \\& Pen 2002; Porciani et al. 2002; Lee 2011) can be addressed within this framework, and is left to future studies." }, "1112/1112.1916_arXiv.txt": { "abstract": "The new multi-color $BVRI$ photometric light curves of the short-period eclipsing binary \\astrobj{GSC 3576-0170} were obtained on two consecutive nights (October 5 and 6, 2009). With the 2003 version of Wilson-Devinney program, the precise photometric solutions are derived for the first time. The result shows that \\astrobj{GSC 3576-0170} is a semi-detached binary system with a large temperature difference of approximately 1490~K. The light-curve distortions are further explained by a hot spot on the secondary component through mass transfer via a stream hitting the facing surface of the secondary component. By analyzing all available light minimum times, we also derived an update ephemeris and found for the first time a possible periodic oscillation with an amplitude of 0.0038 days and a period of 4.3 years. The periodic oscillation could be explained either by the light-time effect due to a presumed third component or by magnetic activity cycle of the system. ", "introduction": "\\label{intro} \\astrobj{GSC 3576-0170}(P$_{orb}$=$0^{d}$.405, G1 V) is a new near-contact solar-type eclipsing binary, which displays asymmetric light curves (two different light maxima) (Nelson et al. 2006). Therefore, it is a very intriguing target for understanding the property of the system.\\\\ \\indent \\astrobj{GSC 3576-0170} was discovered to be variable by Nelson et al. (2006). They attempted a preliminary light-curve synthesis with Wilson-Devinney modeling and preferred a detached model with a mass ratio in the range $0.15-0.35$ and an orbital inclination of 65-$70^{\\circ}$. For the period variation, they invoked a quadratic function to fit the values of the observational times of light minimum - calculational times of light minimum $(O-C)$. Subsequently, new minimum times of GSC 3576-0170 were published by several astronomers (Nelson et al. 2006; H$\\ddot{u}$bscher \\& Walter 2007; H$\\ddot{u}$bscher et al. 2006; Br\\'{a}t 2007; etc.).\\\\ \\indent In this paper, we present our new $B, V, R,$ and $I$ LCs of \\astrobj{GSC 3576-0170} which were analyzed using the 2003 version of Wilson-Devinney code (Wilson \\& Devinney 1971; Wilson 1979, 1990, 1994; Wilson \\& Van Hamme 2004). We accumulated all available times of light minimum and discussed the period change.\\\\ \\begin{table} \\caption{Minimum times of \\astrobj{GSC 3576-0170}.} \\tabcolsep 0.30truecm \\renewcommand\\arraystretch{0.6} \\begin{tabular}{lrrc} \\hline \\hline \\multicolumn{1}{l}{JD(Hel.)} & \\multicolumn{1}{c}{Cycle} & \\multicolumn{1}{c}{(O-C)} & \\multicolumn{1}{c}{References} \\\\ \\hline 2452794.863 & $-2.5$ & $-0.0001$ & 1\\\\ 2452795.8716 & $0.0$ &$-0.0038$ & 1\\\\ 2452799.9230 & $10.0$ & $-0.0024$ & 1 \\\\ 2452802.5542 &$16.5$ &$-0.0039$ &1,4\\\\ 2452806.8076 &$27.0$ &$-0.0029$ &1\\\\ 2452807.8210&$29.5$&$-0.0020$& 1\\\\ 2452812.4781&$41.0$&$-0.0026$& 1,4\\\\ 2452826.8600&$76.5$&$0.0017$&1\\\\ 2452829.4887&$83.0$&$-0.0021$&4\\\\ 2452831.5151&$88.0$&$-0.0007$&4\\\\ 2452863.5105&$167.0$&$-0.0007$&4\\\\ 2452864.5304&169.5&0.0067&4\\\\ 2452867.5607&177.0&$-0.0006$&4\\\\ 2452868.3701&179.0&$-0.0012$&4\\\\ 2452946.3385&371.5&$0.0037$&4\\\\ 2453215.4640&1036.0&0.0034&4\\\\ 2453216.4767&1038.5&0.0036&4\\\\ 2453217.4888&1041.0&$0.0032$&4\\\\ 2453221.5370&1051.0&0.0013&4\\\\ 2453263.8659&$1155.5$&0.0072&1 \\\\ 2453264.6735&$1157.5$&$0.0048$&1\\\\ 2453305.7787 &$1259.0$&0.0020&1\\\\ 2453612.3645&2016.0&$-0.0011$&4\\\\ 2453612.5733&2016.5&0.0052&4\\\\ 2453621.4824&2038.0&0.0042&4\\\\ 2453837.9506&$2573.0$&$-0.0028$&1\\\\ 2453852.9370&$2610.0$&$-0.0016$&1\\\\ 2453900.7278&$2728.0$&$-0.0014$&1\\\\ 2453941.8337&$2829.0$&$-0.0035$&1\\\\ 2453943.8605&$2834.5$&$-0.0017$&1\\\\ 2454073.2580&$3154.0$&$-0.0033$&2\\\\ 2454270.4947&$3641.0$&$-0.0041$&3\\\\ 2455109.0636&5711.5&$0.0019$&5\\\\ 2455110.0739&5714.0&$-0.0004$&5\\\\ \\hline \\end{tabular} \\small \\renewcommand\\arraystretch{0.1} \\\\References - (1) Nelson et al. 2006; (2) H$\\ddot{u}$bscher \\& Walter 2007; (3) Br\\'{a}t 2007; (4) H$\\ddot{u}$bscher et al. 2006; (5) Present paper. \\end{table} ", "conclusions": "Our new LCs and the period variation of \\astrobj{GSC 3576-0170} are included in present investigation. \\subsection{Cyclic magnetic activity or the light-time effect due to the third Body?} For \\astrobj{GSC 3576-0170}, there seems to be the sinusoidal oscillation of the O-C residuals. The oscillating characteristic may be caused by the light-time effect due to the existence of the third body orbiting the eclipsing binary or a possible result of magnetic activity cycles of the system.\\\\ \\indent On one hand, the period oscillation may be explained by the light-time effect. Assuming that the orbital of the presumed third body is circular, we can obtain the mass function for the third body f($M_{3}$)=0.016($\\pm$0.001) M$\\odot$ using the following equation: \\begin{equation} f(m) =\\frac{4\\pi^{2}}{G{T}^{2}}\\times(a\\prime_{12}\\sin{i^{\\prime}})^{3}=\\frac{(M_{3}\\sin{i_3^{\\prime}})^{3}} {(M_{1}+M_{2}+M_{3})^{2}}. \\end{equation} where $M_{1}$, $M_{2}$ and $M_{3}$ are the masses of the binary, and the third body, respectively. The mass of the third component can also be calculated with the above equation, which depends on the orbital inclination. The minimal mass $M_{3,min}$ is 0.38 M$\\odot$ when $i^{\\prime}$=$90^{\\circ}$. \\\\%Because the third mass is much \\indent On the other hand, the period oscillation may be accounted by magnetic cycle of the system (Applegate 1992; Lanza et al. 1998; Lanza \\& Rodon$\\grave{o}$ 1999). The orbital period change corresponding to a variation of the quadrupole moment is given by $\\Delta Q$~=~-($\\Delta p$/p)$\\times$($Ma^{2}$/9) (Applegate 1992) where a is the semi-major axis of the binary orbit and M is the mass of the active component. Since no spectroscopic solutions are available in literature \\astrobj{GSC 3576-0170}, its absolute parameters can not be directly determined. We estimated the primary mass 1.024 as M$\\odot$, and the radius as 1.06 R$\\odot$ by assuming the primary component to be a normal and main-sequence GIV star. Based on our photometric solutions, the mass of the secondary component is $M_{2}$ = 0.478 M$\\odot$ and the separation between the two components is a~=~2.37 R$\\odot$. Therefore, the quadrupole moment $\\Delta Q$ is calculated to be 0.5$\\times$$10^{50}$ g $cm ^{2}$ for the primary.% This value is similar to the typical values for active binaries (Lanza \\& Rodono 1999). Therefore, the magnetic activity cycle is a possible mechanism to explain the variation of period.\\\\ \\subsection{Long-term change and their evolutionary status} For our observations, the two light maxima are basically equal. However, the light curves show that the first quadrature is brighter than the second one in 2003 (Nelson et al. 2006). These indicate a variability on a time scale of about six years. This variation might be attributed to surface magnetic activity of one or both stars. More importantly, GSC 3576-0170 is perhaps associated with possibly thick deep convection according to their spectral types G1 V. Thus, it confirms that magnetic activities may produce the sinusoidal variation of the period.\\\\ \\indent Our result showed that \\astrobj{GSC 3576-0170} is a marginal-contact binary system with a large temperature difference of about 1490~K between the two components, which belongs to the subclass of V1010 Oph binaries (Shaw 1994). It is similar to several semi-detached systems with a possible third body, such as \\astrobj{GW Tau} (Zhu \\& Qian 2006). They may be potential candidates of binaries that will become overcontacted or remain in the broken-contact in accordance with the thermal relaxation oscillation theory of overcontact binaries (Lucy 1976; Flannery 1976; Robertson \\& Eggleton 1977; etc.).\\\\ Of course, our solutions are based on photometric observations only. For better understanding of the properties and the evolutionary state of \\astrobj{GSC 3576-0170}, spectroscopic observations are needed. Moreover, the O-C residual of \\astrobj{GSC 3576-0170} is about seven years and there is an observational gap. Therefore, it is too early to decide the character of the periodic variation, and it might take another 10-20 years are needed for confirmation.\\\\ {\\bf Acknowledgements} The authors gratefully acknowledge the assistance provided for the 85 cm telescope at Xinglong station. We also thank Profs. X., Zhou, A. Y. Zhou, X.J. Jiang, and Y. H. Zhao, for allocation of time for observation and their kind helps during the visit to NAOC. This work was partly supported by GuiZhou University under Grant No. 2008036, GuiZhou Natural Science Foundation 20092263, Shandong Natural Science Foundation (ZR2009AM021) and the Joint Fund of Astronomy of the National Natural Science Foundation of China and the Chinese Academy of Sciences Grant No. 10978010, Natural and Scientific funding supported by Department of Education of Guizhou Province No. 20090130, and Dezhou University Foundation (402811)." }, "1112/1112.5915_arXiv.txt": { "abstract": "We report on the first simultaneous observation of an H$\\alpha$ Moreton wave, the corresponding EUV fast coronal waves, and a slow and bright EUV wave (typical EIT wave). Associated with an X6.9 flare that occurred on 2011 August 9 at the active region NOAA 11263, we observed a Moreton wave in the H$\\alpha$ images taken by the Solar Magnetic Activity Research Telescope (SMART) at Hida Observatory of Kyoto University. In the EUV images obtained by the Atmospheric Imaging Assembly (AIA) on board the {\\it Solar Dynamic Observatory} ({\\it SDO}) we found not only the corresponding EUV fast ``bright'' coronal wave, but also the EUV fast ``faint'' wave that is not associated with the H$\\alpha$ Moreton wave. We also found a slow EUV wave, which corresponds to a typical EIT wave. Furthermore, we observed, for the first time, the oscillations of a prominence and a filament, simultaneously, both in the H$\\alpha$ and EUV images. To trigger the oscillations by the flare-associated coronal disturbance, we expect a coronal wave as fast as the fast-mode MHD wave with the velocity of about 570 -- 800~km~s~$^{-1}$. These velocities are consistent with those of the observed Moreton wave and the EUV fast coronal wave. ", "introduction": "Moreton waves, flare-associated waves seen in H$\\alpha$, have been observed \\citep{mor1960,smi1971,shi2011} to propagate in restricted angles with the velocity of about 500 -- 1500~km~s$^{-1}$. They sometimes show arc-shaped fronts, and are often associated with type-II radio bursts \\citep{kai1970}. They are transient, and appear only for about 10 minutes. Associated with flares, remote filaments and prominence are sometimes activated or excited to oscillate. These ``winking filaments'' are also thought to be caused by flare-associated waves, and are called as invisible Moreton waves \\citep{smi1971,tri2009,her2011}. After the findings, \\citet{uchi1968} suggested that Moreton waves are the intersection of the fast-mode magnetohydrodynamic (MHD) shock propagating in the corona with chromosphere. This model has been widely accepted, and the coronal counterparts have been surveyed for a few decades. Moreton waves are rare to be observed even for large flares \\citep{shi2011}. After the launch of the {\\it Solar and Heliospheric Observatory} ({\\it SOHO}), the EUV Imaging Telescope (EIT) found wavelike phenomena associated with flares, which are called ``EIT waves'' \\citep{thom1999,thom2000}. Although EIT waves were expected to be the coronal counterpart of Moreton waves, they show different physical characteristics from those of Moreton waves: the propagating velocity is much slower than that of Moreton wave and is about 200 -- 400~km~s$^{-1}$, the lifetime is much longer and is about 45 -- 60 minutes, they can show isotropic propagation, while Moreton waves propagate with restricted angles \\citep{kla2000,war2007,thom2009}. There have been, therefore, remained a question whether EIT waves are really coronal counterparts of Moreton waves or no. As for searching for a coronal counterpart of Moreton waves, the Soft X-ray Telescope (SXT) on board {\\it Yohkoh} found wavelike phenomena in soft X-rays, called X-ray waves \\citep{khan2000,khan2002}. X-ray waves are confirmed to be a real counterpart of Moreton waves by simultaneous observations of X-ray waves and Moreton waves \\citep{naru2002,naru2004}. Then, we come to an issue what EIT waves are. \\citet{eto2002} clearly showed that an EIT wave is different from a Moreton wave, based on simultaneous observations of them. On the other hand, \\citet{war2004a,war2004b} argue that the velocity discrepancy of EIT and Moreton waves can be explained by the deceleration of coronal waves. The mechanism of EIT waves remains, therefore, very controversial \\citep{war2007,will2009,gall2010}. \\citet{del1999} and \\citet{chen2002,chen2005} proposed the field-line stretching model for EIT waves. They suggested that EIT bright fronts were not ``waves'' at all, but instead plasma compression at stable flux boundaries due to rapid magnetic field expansion. This model can also resolve the puzzle why EIT waves often stop at magnetic separatrices. Recently, by the Atmospheric Imaging Assembly (AIA; Title \\& AIA team 2006, Lemen et al. 2011) on board the {\\it Solar Dynamic Observatory} ({\\it SDO}), fast coronal waves have been observed associated with flares \\citep[e.g.][]{liu2010,ma2011}. These waves (hereafter called ``EUV fast coronal waves'') are thought to be the fast-mode MHD waves. Coronal X-ray waves and EUV fast coronal waves have been also observed spectroscopically with the EUV Imaging Spectrometer (EIS) on board {\\it Hinode} \\citep{asa2008,harra2011}. \\citet{chen2011} found two different coexisting coronal waves, slow coronal wave (i.e. EIT wave) and fast coronal wave, from EUV observations taken by {\\it SDO}/AIA. Although the fast coronal wave seems to be the coronal counterpart of a Moreton wave, they used only EUV images, and it remained to be confirmed whether it is a classical H$\\alpha$ Moreton wave. This letter presents the first simultaneous observation of EUV waves and a Moreton wave by using EUV and H$\\alpha$ images with high spatial and temporal resolutions. Moreover, we found not only a winking filament on the disk, but also an oscillating prominence on the limb, triggered by the coronal wave (Moreton wave). ", "conclusions": "We simultaneously observed the H$\\alpha$ Moreton wave and the corresponding EUV fast coronal wave. The Moreton wave front was well consistent with the fast-bright-sharp EUV wave front (F2b). Even after the Moreton wave disappeared, we identified the propagation of the fast-faint EUV wave (F2f). Even along the Line 3 (Figure~\\ref{shock}a), which is the direction without the Moreton wave, we found that the fast EUV waves (F3b, F3f). The fast EUV waves (F1, F2f, F3b, and F3f) are thought to be the fast-mode MHD waves (coronal waves). The EUV fast coronal waves appear more frequently than Moreton wave, although they are very faint and have not been observed until the launch of {\\it SDO}. Especially, only when the shock strongly contacts with the chromosphere, the intersection is observed as the Moreton wave (F2b). The temporal evolutions of the H$\\alpha$ Moreton wave and the EUV waves are summarized in Figure~\\ref{hei_tim}(b). In Figure~\\ref{hei_tim} we also show the temporal evolutions of a type-II radio burst associated with the flare. In the metric radio spectrogram (25 -- 2500~MHz) observed with the Hiraiso Radio Spectrograph (HiRAS; Kondo et al. 1995), we identify the type-II radio burst from 08:02:40 to 08:06:30~UT. Assuming the coronal density model proposed by \\citet{new1961} and \\citet{mann1999}, we derived the propagation velocity of about 850~km~s$^{-1}$. The observed type-II radio burst seems to be consistent with the Moreton wave and fast bright EUV wave, which also supports the interpretation of the wave as a fast-mode MHD shock. We also found oscillations of a prominence and a filament.% To trigger the oscillations by a flare-associated coronal disturbance, we expect a coronal wave as fast as the fast-mode MHD wave with the velocity of about 570 -- 800 km~s~$^{-1}$. These velocities are consistent with the propagation velocities of the observed Moreton wave and the EUV fast coronal wave. An invisible Moreton wave could be such an EUV fast-faint coronal wave. It is known that a typical slow EIT wave sometimes causes filament oscillations \\citep{oka2004}, or such filament oscillations triggered by an EIT wave are expected to be stronger than those by an invisible Moreton wave (P. F. Chen, private communication). In the current case, however, the role of the EIT wave on the filament/prominence oscillations is unclear. Along the Line 2 and 3, we identified slow-bright EUV waves (S2b, S3b) behind the fast-faint EUV waves. From the propagating features (i.e. the velocity and the isotropic propagation; Figure~\\ref{euv}), we think it is a typical EIT wave. We simultaneously observed the EIT wave and the EUV fast wave (fast-mode MHD wave) as reported by \\citet{chen2011}. This means that the EIT wave is different from a fast-mode MHD wave, which supports the field-line stretching model \\citep{chen2002,chen2005}. It is, however, difficult to clearly identify both features separately in the very initial phase of the flare, and it is unclear the relation between the fast-bright (F3f) and the slow-bright (S3b) EUV waves along the Line 3. There is an alternative possibility that we observed a single coronal disturbance, and the two different waves (F2f/F3f and S2b/S3b) correspond to the fronts of two different heights (faster ones are higher) of the disturbance \\citep{vero2008,war2011}. We, however, do not think the possibility due to the following reasons: First, the slow bright waves (S2b, S3b) stopped propagating at small active regions. This conflicts with the features of a fast-mode wave, % while this is possibly reconsiled by considering the stopping front as CME flanks as reported by \\cite{pat2009}. Second, we observed the prominence/filament oscillations. Since they are located low in the corona, we can derive the velocity of the shocks/waves there, from the distances and the times of the oscillations. In the current case we need waves with velocities of 570~km~s$^{-1}$ or more. Especially, the direction of the filament is close to the Line 3, and the required velocity is much faster than the bright slow wave (S3b), while it is consistent with the fast faint wave (F3f)." }, "1112/1112.2018_arXiv.txt": { "abstract": " ", "introduction": "The outer crust of a neutron star, below densities of $\\rho \\sim 4 \\times 10^{11}$ g cm$^{-3}$ consists of matter in a state not too far removed from that found in white dwarfs: a lattice of nuclei permeated by a relativistic, degenerate electron gas which gives the dominant contribution to the pressure of the matter \\cite{BPS1971}. As pressure increases with depth, equilibrium with respect to weak interactions drives the nuclei to become more neutron rich. There comes a point when the intra-nuclear forces can no longer bind all the neutrons, and \\emph{neutron drip} occurs. Above $\\rho \\sim 4 \\times 10^{11}$ g cm$^{-3}$, a new regime is entered in which the nuclear lattice is bathed in a fluid of (`dripped') neutrons. These neutrons are delocalized much like conduction band electrons in metals. From this density inwards, the \\textbf{equation of state} (\\textbf{EOS}) is dominated by pressure arising from nucleon-nucleon interactions. \\begin{figure}[!tb]\\label{fig:1} \\begin{center} \\includegraphics[width=14cm,height=8cm]{Fig1.pdf} \\caption{Cartoon representation of the inner neutron star crust.} \\end{center} \\end{figure} A cartoon representation of the crustal layers below the outer crust is shown in Fig.~1. Models predict two distinct layers. (1) The \\textbf{inner crust} between densities $\\rho \\sim 4 \\times 10^{11}$ g cm$^{-3}$ and $\\rho \\sim 10^{14}$ g cm$^{-3}$ is an elastic solid consisting of a lattice of heavy, neutron rich nuclei surrounded by fluid neutrons, with the nuclei increasing in size and mass with density while the inter-nuclear spacing and nuclear proton fraction decrease \\cite{BBP1971}. (The presence of a background electron gas will be taken as given from now on). The dripped neutrons are expected to become superfluid shortly after the neutron star is formed as it rapidly cools below temperatures of $10^8 - 10^9$ K \\cite{Dean2003}. (2) The \\textbf{mantle} \\cite{Gusakov2004} between $\\rho \\sim 10^{14}$ g cm$^{-3}$ and the crust-core transition density consists of frustrated matter: the competition between the nuclear surface energy and the nuclear and lattice Coulomb energies over similar length scales drives the formation of exotic nuclear geometries termed nuclear `pasta' \\cite{Ravenhall1983, Oyamatsu1984} which proceed through a canonical sequence of phases: cylindrical (spaghetti) $\\to$ slab (lasagna) $\\to$ cylindrical bubble $\\to$ spherical bubble. The latter three (`bubble') phases are distinguished by the delocalization of the \\textbf{charged nuclear component} of the matter (containing the protons) in one or more dimensions, and corresponding localization of the \\textbf{charge-neutral nuclear component} (fluid neutrons). Similar microscopic structures are observed in terrestrial soft condensed matter systems such as surfactants \\cite{Jones2002,Watanabe2005}; by analogy, we can expect rich mechanical properties, intermediate between liquid and elastic solid, to emerge in the mantle \\cite{Pethick1998}. Some crust models predict the absence of the mantle \\cite{Douchin2001}; its presence depends sensitively on the nuclear microphysics of the crust. Fig.~1 gives a range of widths for the inner crust and mantle taken from the model presented in this chapter and encompassing a range of neutron star masses from 1-2 $M_{\\odot}$ and a range of equations of state as discussed later. To describe the states of matter in a neutron star, one needs a model for the nucleon-nucleon interactions as they are manifested in a many-nucleon context. A useful concept that bridges the gap between \\emph{ab initio} nucleon-nucleon calculations, nuclear experimental observables, and neutron star matter is that of uniform \\textbf{nuclear matter (NM)}. This is an idealized system, homogeneous and infinite in extent, of neutrons and protons interacting solely via the strong force. The energy per particle of such a system at a density $\\rho$ and proton fraction $x$, $E(\\rho, x)$, is referred to as the nuclear matter equation of state (\\textbf{NM EOS}). In the regions of the neutron star core where protons and neutrons exist, the NM EOS can be combined with the electron energy, and under conditions of charge neutrality and beta-equilibrium gives an EOS for the core. In the inner crust, the NM EOS can be used to describe the dripped neutrons ($x=0$) and the bulk matter in the nuclear clusters. A consistent model for the EOS of crust and core necessarily uses a unique NM EOS, and one should expect parameters characteristic of a given NM EOS to correlate with both crust and core properties. Nuclear matter with equal numbers of neutrons and protons ($x=0.5$) is referred to as \\textbf{symmetric nuclear matter (SNM)}; nuclear matter with $x=0.0$ is naturally referred to as \\textbf{pure neutron matter (PNM)}. Nuclei on Earth contain closely symmetric nuclear matter at densities close to nuclear saturation density $\\rho_0 \\approx 2.7 \\times 10^{14}$ g cm$^{-3} \\equiv 0.16$ fm$^{-3} = n_0$, where we use $n$ to refer to baryon number density. Thus experiment has constrained the properties of $E(\\sim n_0,\\sim0.5)$ to within relatively tight ranges, but the properties of PNM remain uncertain from an experimental standpoint. In the past decade, much experimental activity has been devoted to extending our knowledge of nuclear interactions to more neutron-rich systems and to higher and lower densities. Although we cannot produce pure neutron matter in the laboratory, we can produce matter with proton fractions as low as $x \\approx 0.3$ in certain neutron rich isotopes and in the products of heavy ion collisions. This allows us to obtain information on how $E(\\sim n_0, x)$ changes as $x$ decreases. By expanding $E(n, x)$ about $x = 0.5$ using the isospin asymmetry variable $\\delta = 1-2x$, we can define a useful quantity called the \\emph{symmetry energy} S(n), \\be\\label{eq:eos1} E(n,\\delta) = E_{\\rm 0}(n) + S(n)\\delta^2 + ...; \\;\\;\\;\\;\\;\\;\\;\\; S(n) = {1 \\over 2}{\\partial^2 E(n,\\delta) \\over \\partial \\delta^2}\\bigg|_{\\delta=0}, \\ee \\noindent which encodes the change in the energy per particle of NM as one moves away from isospin symmetry. This allows extrapolation to the highly isospin asymmetric conditions in neutron stars. The simplest such extrapolation, referred to as the \\textbf{parabolic approximation (PA)}, truncates the expansion to second order, giving \\be \\label{PAapp} E_{\\rm PNM}(n) \\equiv E(n, \\delta=1) \\approx E_{\\rm 0}(n) + S(n) \\ee for the PNM EOS. Expanding the symmetry energy about $\\chi=0$ where $\\chi = \\frac{n-n_{\\rm 0}}{3n_{\\rm 0}}$ we obtain \\be\\label{eq:eos3} S(n) = J + L \\chi + \\half K_{\\rm sym} \\chi^{2} + ..., % \\ee \\noindent where $J$, $L$ and $K_{\\rm sym}$ are the symmetry energy, its slope and its curvature at saturation density. Since neutron star matter contains a low fraction of protons, many inner crust and global stellar properties are sensitive to the symmetry energy parameters $J$,$L$, etc. To give a simple example, the pressure of PNM at saturation density is given by $P_{\\rm PNM}(n_0)$=$n_0L/3$. The pressure in the inner crust and outer core is dominated by neutron pressure so a strong correlation exists between the pressure in neutron stars near saturation density and $L$. Neutron star EOSs which have higher pressures are often referred to as `stiff'; lower pressure EOSs are referred to as `soft'. Thus, around 1 to 2$n_0$, `stiff' EOSs are associated with high values of $L$ and `soft' EOSs with low values of $L$. That large uncertainties exist in characteristic NM parameters such as $J$ and $L$ is one reason why many model predictions of potential neutron star observables span such wide ranges. On the other hand, observations of neutron stars offer the opportunity to obtain constraints on NM parameters and hence on the underlying models of the nucleon-nucleon interactions \\cite{Lattimer2001}. The following is a non-exhaustive list of (potentially) observable neutron star phenomena whose precise properties depend on the properties of the inner crust and mantle and of the star as a whole; some of these will be described in detail elsewhere in this book. \\vspace{\\baselineskip} $\\bullet$ \\hspace{5pt} \\textbf{Pulsar glitches.} Young pulsars spin down under the action of magnetic torque, their rotational energy powering the radiation beam. Many are observed to undergo occasional, sudden, spin-ups called glitches \\cite{Espinoza2011}. Proposed mechanisms include crust-cracking as the star attempts to adjust its shape to become more spherical \\cite{Baym1969} and angular momentum transfer from one internal component to another such as some part of the crust superfluid neutrons to the rigid part of the crust \\cite{Anderson1975, Link1999}, or a combination of both \\cite{Ruderman1998}. In such models, the size, frequency and post-glitch relaxation of the spin period depend on, among other microscopic properties, the crust and core sizes, moments of inertia and composition. $\\bullet$ \\hspace{5pt} \\textbf{Free precession.} Certain pulsars exhibit long timescale periodic variation in their timing residuals, with periods of order years, suggestive of free precession of the star \\cite{Stairs2000,Shabanova2001}. Free precession is can arise from mechanically or magnetically supported crustal deformation \\cite{Cutler2002,Wasserman2003,Cutler2003}, and the period depends also on details of crust-core coupling, notably through the properties of the crust and core superfluid \\cite{Jones2001,Link2003,Link2006,Glampedakis2009}. $\\bullet$ \\hspace{5pt} \\textbf{QPOs from SGR giant flares.} Quasi-periodic oscillations (QPOs) in the tails of light curves of giant flares from soft gamma-ray repeaters (SGRs) have been observed \\cite{Israel2005, Watts2006, Strohmayer2005, Strohmayer2006}, and their frequencies lie in the range of possible torsional vibrations of the crust. The crust thickness, composition (through, e.g., the shear modulus) and the stellar size all affect the frequencies of such modes \\cite{SteinerWatts2009,Samuelsson2007,Andersson2009,Gearheart2011,Sotani2011}. $\\bullet$ \\hspace{5pt} \\textbf{Neutron star cooling.} The crust thermalization timescale depends on the crust thickness as well as the thermal conductivity and specific heats arising from the heat transport mechanisms operating in the crust \\cite{Lattimer1994, Gnedin2001}. One intriguing possibility is the operation of the direct Urca process in the bubble phases of the mantle, where the delocalization of the protons may allow it \\cite{Gusakov2004}. The thickness of this layer plays an important role in determining how effective a cooling mechanism this might be. $\\bullet$ \\hspace{5pt} \\textbf{Gravitational waves (GWs) from neutron stars.} A rich array of stellar oscillation modes are possible, some of which might generate GWs detectable on Earth \\cite{Andersson2011,Abbott2010}, or lead to other observational signatures such as limiting neutron star spin-up \\cite{Bildsten1998,Andersson1999}. Stability of modes can depend sensitively on the physics at the crust core interface and crust thickness, \\cite{Bildsten00,Andersson00,Lindblom00,Rieutord01,Peralta06,Glampedakis06,Wen11}. GWs can also be generated by a quadrupole deformation in the stellar shape supported, among other possibilities, by the elastic crust \\cite{Ushomirsky2000, Haskell2006}. Whether the crust is strong enough to support a large enough deformation to produce detectable gravitational waves depends on the shear modulus throughout the crust, and thus its composition (especially in the mantle where the mechanical properties are particularly uncertain) and the crust thickness and stellar size \\cite{Gearheart2011}. \\vspace{\\baselineskip} In this chapter we will review the dependences of the composition, thickness of the crust and the mantle, and certain global neutron star properties on the symmetry energy parameters $J$ and $L$. The interplay of such relationships in modeling neutron star observables and the potential for obtaining astrophysical constraints on $J$ and $L$ will be illustrated by constructing consistent crust and core models based on a model of uniform nuclear matter whose symmetry energy parameters can be smoothly varied. We shall use the \\textbf{compressible liquid drop model (CLDM)} of crustal matter for its expediency; an outline of more sophisticated models will be given in the discussion at the end of the chapter. ", "conclusions": "We have constructed sets of neutron star EOSs that consistently encompass the inner crust and core, and include the crust composition and transition densities, using the compressible liquid drop model for the crust. These have been used to demonstrate the effect of the magnitude $J$ and slope $L$ of the symmetry energy at nuclear saturation density on microscopic and global crustal properties, and potential neutron star observables. The crust-core transition density and pressure, crustal composition, stellar and crustal mass, thickness and moment of inertia, torsional crust oscillation frequencies and maximum crust deformation all depend sensitively on both $J$ and $L$ within their experimentally constrained ranges. One of the dominant neutron star model dependences is therefore the correlation between $J$ and $L$ which constrains their possible values in $J-L$ space. Experimental and theoretical information about $L$ and $J$ and their correlation will continue to improve; in order to add neutron star observations to this investigation, consistent explorations of neutron star properties over the constrained ranges is of great importance. The sets of EOSs used in this paper are available to interested parties \\cite{mywebsite}. The simplicity of the CLDM allows useful exploration of the dependence of composition and transition density on $J$ and $L$, but it possesses several drawbacks. Firstly shell effects within the nuclei, or arising from scattering of dripped neutrons off of nuclear clusters, are ignored. Such shell effects can dominate the determination of nuclear geometry, the equilibirium size of the nuclear clusters (which will proceed in discrete jumps corresponding to changes in the nuclear `magic' numbers with density) and the ordering of the pasta phases, as well as transport properties such as contributions to heat transport from nuclear components \\cite{Sandulescu2007} and entrainment of dripped neutrons by clusters \\cite{Chamel2005b}. Secondly, the WS approximation is expected to break down at when the nuclear separation becomes comparable to the cell size \\cite{Chamel2007}, which occurs in the mantle. Thirdly, effects that act over ranges greater than the unit cell are not consistently accommodated in the CLDM; longer range electron screening, larger scale self-organization of pasta phases and long wavelength transport effects are all unaccounted for. \\begin{figure}[!t]\\label{fig:12} \\begin{minipage}{6.6cm} \\begin{center} \\includegraphics[width=3.3cm,height=3cm]{Fig10a.pdf}\\includegraphics[width=3.3cm,height=3cm]{Fig10b.pdf}\\\\ \\includegraphics[width=3.3cm,height=3cm]{Fig10c.pdf}\\includegraphics[width=3.3cm,height=3cm]{Fig10d.pdf} \\end{center} \\caption{Integrated neutron density in a cubic cell at densities of $n$=0.04, 0.06, 0.08 and 0.1 fm$^{-3}$ (top left to bottom right) calculated using the 3DHF method \\cite{Newton2011} with the SLy4 Skyrme parameterization.} \\end{minipage} \\hfill \\begin{minipage}{6.6cm} \\begin{center} \\includegraphics[width=6.6cm,height=6cm]{Fig10e.pdf} \\end{center} \\caption{Transition densities obtained using the 3DHF method compared to those obtained using the dynamical method of \\cite{Xu2009} for the SLy4, SII, SkM* and SkMp Skyrme parameterizations.} \\end{minipage} \\vspace{10pt} \\end{figure} Some of these effects can be taken into account by more sophisticated crust models. The Thomas-Fermi and Extended Thomas-Fermi methods (e.g \\cite{Buchler1971,Oyamatsu1993,Cheng1997}) are semi-classical models employing the local density approximation which allows the nuclear surface energy to be calculated self-consistently with the bulk nuclear energy. Shell corrections can be self-consistently added using the Strutinsky integral method \\cite{Onsi2008}. The 1D-Hartree-Fock method is fully microscopic and accounts for surface and shell energies self-consistently (e.g. \\cite{Sandulescu2007,Negele1973,Montani2004,Baldo2005}), but it describes only spherically symmetric configurations and is thus constrained by the spherical Wigner-Seitz approximation. This restriction is lifted in the more computationally demanding 3D-Hartree-Fock (3DHF) method \\cite{Magierski2002,Gogelein2008}, allowing a self-consistent probe of the shape-phase-space of pasta layers with shell effects included. As example, some results of a 3DHF model \\cite{Newton2009} are shown in Figs.~10 and ~11. The former shows the local neutron density is plotted over the unit cell at four different densities encompassing the crust-core transition, showing the evolution of nuclear shape from spherical through bubble to uniform matter. The latter shows the resulting transition densities for four different parameterizations of the Skyrme interaction, compared to the dynamical method of \\cite{Xu2009}. The Hartree-Fock method is naturally extended to include pairing effects self-consistently (Hartree-Fock-Bugoliubov). Longer range effects can be simulated via the semi-classical quantum molecular dynamics method \\cite{Maruyama1998,Horowitz2004,Watanabe2001,Sonoda2007}. Between these models, a complete physical description of the crust can be built up. Many of the methods mentioned are much more time-consuming than the CLDM, making a wide-ranging survey of nuclear matter parameters unwieldy. However, if we know the quantities for which the CLDM provides a reasonable estimate, and to what densities it remains reasonable, one can use the CLDM as a useful guide for more realistic calculations. Ultimately, the goal should be to have all relevant microscopic inputs to neutron star models calculated consistently with the nuclear matter EOS; much work still needs to be done in this direction." }, "1112/1112.5312_arXiv.txt": { "abstract": "Recent $\\gamma$-ray observations suggest that the $\\gamma$-ray millisecond pulsar (MSP) population is separated into two sub-classes with respect to the pair multiplicity. Here, we calculate the cosmic ray electron/positron spectra from MSPs. Based on the assumption of the equipartition in the pulsar wind region the typical energy of electrons/positrons ejected by a MSP with the pair multiplicity of order unity is $\\sim50$ TeV. In this case, we find that a large peak at 10 - 50 TeV energy range would be observed in the cosmic ray electron/positron spectrum. Even if the fraction of pair starved MSPs is 10\\%, the large peak would be detectable in the future observations. We also calculate the contribution from MSPs with high pair multiplicity to the electron/positron spectrum. We suggest that if the multiplicity of dominant MSP population is $\\sim 10^3$, electrons/positrons from them may contribute to the observed excess from the background electron/positron flux and positron fraction. ", "introduction": "The {\\it Fermi} Gamma-Ray Space Telescope has detected $\\gamma$-ray pulsed emissions from more than twenty millisecond pulsars (MSPs) \\citep{Ab11}, which have a rotation angular frequency $\\Omega\\sim 10^3$s$^{-1}$ and a stellar surface magnetic field $B_s\\sim 10^{8.5}$G. The detection of the GeV emissions from a pulsar magnetosphere means that electrons and positrons are accelerated to more than $\\sim$ TeV by the electric field parallel to the magnetic field, which arises in a depleted region of the Goldreich-Julian (GJ) charge density \\citep{GJ69}. The $\\gamma$-ray light curve is an important tool for probing the particle acceleration process in the pulsar magnetosphere. Therefore, the $\\gamma$-ray emission region has been explored by comparing theoretical models such as polar cap \\citep{DH96}, outer gap \\citep{CHR86} and slot gap models \\citep{MH04} with the observed light curve (e.g., Venter, Harding \\& Guillemot 2009; Romani \\& Watters 2010; Kisaka \\& Kojima 2011). \\citet{VHG09} fitted the pulse profiles of the {\\it Fermi} detected MSPs with the geometries of $\\gamma$-ray emission region predicted by different theoretical models. They found that the pulse profiles of six of eight MSPs could be fitted by the geometries of either the outer gap or the slot gap model, as was the case of canonical pulsars. They interpreted that copious pairs are produced in the magnetosphere of these MSPs. However, \\citet{VHG09} also found that the pulse profiles of remaining two MSPs show the unusual behavior in the $\\gamma$-ray light curves and could not be fitted by the geometry of either the outer gap or the slot gap models. They proposed that these unusual light curves could be fitted by the pair starved polar cap model \\citep{MH04b}, in which the multiplicity of the pairs is not high enough to completely screen the electric field above the polar cap, and the particles are continuously accelerated up to high altitude over the entire open field line region. Thus, from the model fitting of the $\\gamma$-ray light curves, \\citet{VHG09} suggested that the $\\gamma$-ray MSP population is separated into two sub-classes. The important fact is that radio pulsed emission is also detected from all currently detected $\\gamma$-ray MSPs and remarkably similar to that from canonical pulsars. The pulsar radio emission is a highly coherent process because the brightness temperature is extremely high. In the theoretical models of the radio emission mechanisms, some authors have believed the conditions that there are a highly relativistic primary beam with the large Lorentz factor ($\\sim 10^7$) and the number density nearly equal to the GJ density, and the secondary electron/positron plasma with relatively small bulk streaming Lorentz factor ($\\sim 10$ - $10^3$) and the large pair multiplicity ($\\sim 10^3$ - $10^5$) in the radio emitting region (e.g., Melrose 1995; Lyutikov, Blandford \\& Machabeli 1999; Gedalin, Gruman \\& Melrose 2002). However, the existence of pair starved MSPs suggests that the radio emission mechanisms should be insensitive to the particle number density down to sub-GJ number density. The pulsar radio emission mechanism is still poorly understood, so that the observationally-based constraints are valuable \\citep{M95}. Therefore, another verification for the extent of the MSP multiplicity, especially the existence of the pair starved MSPs is important for the pulsar radio emission mechanisms. Recently, HESS has discovered a new TeV source, which is located in the close vicinity of the globular cluster Terzan 5 \\citep{HE11}. Several globular clusters, including Terzan 5 also emit GeV $\\gamma$-ray \\citep{Ab09b, KHC10, Ab10b, Ta11}, which may plausibly be due to a number of MSPs residing in these clusters \\citep{HUM05, VDC09}. Thus, inverse Compton scattering by the high-energy particles ejected from MSPs are proposed for the origin of the observed TeV emission \\citep{BS07, VDC09}. The high-energy electron/positron spectrum ejected from MSPs would be a useful probe for the multiplicity of the MSPs. However, only from the TeV spectra, we cannot distinguish two models \\citep{BS07, VDC09}, which assume different pair multiplicities \\citep{HE11}. Another way to investigate the electron/positron spectrum ejected from MSPs is its direct measurement. Since high-energy electrons/positrons can propagate only about a few kpc due to the energy losses by the synchrotron and the inverse Compton emission, the direct detection of the electrons/positrons ejected from MSPs in the globular clusters is unlikely. However, for the following reasons, we may detect those from nearby MSPs. MSPs have much lower spin-down luminosity than canonical pulsars. \\citet{BVD08} investigated the possible contribution of the nearby MSP, PSR J0437-4715 to the cosmic ray electron/positron spectrum. They concluded that unlike canonical pulsars such as Geminga pulsar, the contribution from a MSP to the observed electron/positron flux is negligible. However, since the lifetime of MSPs is much longer than canonical pulsars ($>10^{10}$ yr), there should be much more nearby active MSPs. Furthermore, Kashiyama, Ioka \\& Kawanaka (2011; hereafter KIK11) pointed out that since white dwarf pulsars have long lifetime and continue to inject the electrons/positrons after the nebulae stop expanding, the adiabatic energy losses of electrons/positrons in the pulsar wind nebula region are negligible. Also the synchrotron cooling of electrons/positrons is so small and the high-energy electrons/positrons can escape the nebulae without losing much energy. Although they consider the case of the white dwarfs, their results are also applicable to MSPs. Therefore MSPs could potentially contribute to the observed high energy cosmic ray electrons/positrons and will be detectable by the next generation experiments, such as CALET \\citep{To08} and CTA \\citep{CTA}. In this paper, we investigate the contribution of electrons/positrons ejected from the MSPs to the observed cosmic ray spectrum. In section 2, we apply KIK11 model to the case of MSPs. We estimate the typical energy of electrons/positrons from the MSPs and show that during the propagation in pulsar wind nebulae, the adiabatic losses and radiative cooling of electrons/positrons are not so large. We also describe the propagation in the interstellar medium (ISM). In section 3, we calculate the energy spectrum of cosmic ray electrons/positrons from the MSPs and show the possibility that the electrons/positrons from these MSPs are detectable for the future observations. ", "conclusions": "First, we calculate the cosmic ray electron/positron spectra from the pair starved MSPs. We set the pair multiplicity $\\kappa= 1$, the lifetime $\\tau= 5\\times 10^{10}$yr, the total energy $E_{\\rm rot}= 10^{52}$ erg, the local birth rate $R = 3\\times 10^{-9}$ yr$^{-1}$ kpc$^{-2}$ and the fraction of the lost energy due to synchrotron emission 30\\% for each MSP. We assume that each MSP has the same value of the parameters ($B_0=10^{8.5}{\\rm G}, \\Omega=10^3{\\rm s}^{-1}, R=10^6{\\rm cm}$), because most MSPs have the almost same spin-down luminosity. For the injection distribution function, we assume mono-energetic distribution eq.(\\ref{mono}) with the energy $\\varepsilon_e=50$ TeV. Even if we consider the power-law distribution eq.(\\ref{pl}), the energy range of the distribution is small because the allowed range of the cutoff energy should be only 50 - 80 TeV due to the observed constraint by KASKADE/GRAPES/CASA-MIA \\citep{KY09}. Thus the distribution should be nearly mono-energetic distribution. Note that our local MSP birth rate is based on the MSP local surface density of 38$\\pm$ 16 pulsars kpc$^{-2}$ for 430 MHz luminosity above 1 mJy kpc$^2$ \\citep{L08}. Actually, now 23 MSPs are detected within 1kpc \\citep{ATNF}. We can only detect MSPs that have the radio beam directed toward us and the radio flux larger than the threshold of detectors. However, the cosmic ray electrons/positrons ejected from MSPs are distributed isotropically because of the effect of Galactic magnetic field, so that a large number of MSPs will contribute to the observed electron/positron spectrum. Therefore, our local MSP birth rate corresponds to the lower limit for the current radio observations. In figure \\ref{fig:3.1}, the electron/positron flux from multiple pair starved MSPs is shown. Thick solid, dashed and dash-dotted lines show the total electron/positron spectra if the fraction of the pair starved MSPs is 100\\%, 25\\% and 10\\%, respectively. Thin solid line shows the contribution of electrons/positrons from the pair starved MSPs when the fraction is 100\\%. The background flux is shown as a thin dashed line. We adopt the background model of an exponentially cutoff power law with an index of -3.0 and a cutoff at 1.5 TeV, which is similar to that shown in \\citet{Ah08} and reproduces the data in $\\sim$10 GeV-1 TeV well. It is very interesting that there is a large peak at 10-50 TeV energy range. The existence of this peak cannot be ruled out from the current observations. The high-energy component is more enhanced for the long-duration sources \\citep{AAV95, KIN10}. This is because the longer the duration of injection is, the larger fraction of fresh electrons/positrons are expected to reach the Earth without losing their energy so much during the propagation. MSPs are the continuously injecting sources with the duration as long as the cosmic age, so that the spectrum from them has nearly the same shape with the injection spectrum with the soft energy tail component. This is the difference from other sources such as young pulsars, whose typical duration is only $\\sim 10^4$-$10^5$ yrs. Fitting result of \\citet{VHG09} showed that the fraction of the pair starved MSPs is 25\\%, although they have only eight samples. Even if the fraction is 10\\%, the flux is $\\sim 20$m$^{-2}$ s$^{-1}$ sr$^{-1}$ GeV$^2$ at 10 TeV. In this case, we can detect the electron/positron flux with near future missions such as CALET (we assume the geometrical factor times the observation time $\\sim 5$ yrs as $\\sim 220$ m$^2$ sr days) because the predicted electron/positron flux is sufficiently large. It was considered that the number of astrophysical sources contributing to the above several TeV energy range is quite small according to the birth rate of the supernovae and the canonical pulsars in the vicinity of the Earth \\citep{Ko04, KIN10, KIOK11}. However, we find that it is possible for multiple pair starved MSPs to contribute to the 10 TeV energy range in the electron/positron spectrum. Therefore, if the anisotropy of the observed electrons/positrons are weak in the 10 TeV energy range, we suggest that the pair starved MSPs may contribute to the spectrum significantly. Next, we also investigate the contribution of MSPs with high pair multiplicity to the observed cosmic ray electrons/ positrons. In this model, we assume that the injection function of these MSPs is power-law distribution with index $\\alpha=2$, the cutoff energy $\\varepsilon_{e, {\\rm cut}}=1$TeV and the minimum energy $\\varepsilon_{e, \\min}=1$GeV. In this case, the pair multiplicity is $\\sim 2000$. Other parameters are the same values as in the case of the pair starved MSPs. We assume that the fraction of MSPs with high multiplicity is 100\\%. The results are shown in figure \\ref{fig:3.2} for the electron/positron spectrum and in figure \\ref{fig:3.3} for the positron fraction. Both figures show that electrons/positrons ejected from MSPs with high multiplicity partially contribute to the excess from background flux observed by PAMELA, HESS and {\\it Fermi}. In this energy range, the other sources such as canonical pulsars \\citep{S70, AAV95, CCY96, ZC01, Ko04, G07, Bu08, P08, HBS09, YKS09, MCG09, Gr09, KIN10, HGH10} would also contribute to the observed electron/positron spectrum. Note that even if other sources are dominant for the observed excess, the total spectrum added to the contribution of MSPs with high multiplicity does not significantly exceed the observed electron/positron spectrum." }, "1112/1112.5548_arXiv.txt": { "abstract": "{Rotation is known to have a strong impact on the nucleosynthesis of light elements in massive stars, mainly by inducing mixing in radiative zones. In particular, rotation boosts the primary nitrogen production, and models of rotating stars are able to reproduce the nitrogen observed in low-metallicity halo stars. } {Here we present the first grid of stellar models for rotating massive stars at low metallicity, where a full s-process network is used to study the impact of rotation-induced mixing on the neutron capture nucleosynthesis of heavy elements. } {We used the Geneva stellar evolution code that includes an enlarged reaction network with nuclear species up to bismuth to calculate 25~$M_\\odot$ models at three different metallicities ($Z = 10^{-3}, 10^{-5}$, and $10^{-7}$) and with different initial rotation rates. } {First, we confirm that rotation-induced mixing (shear) between the convective H-shell and He-core leads to a large production of primary $^{22}$Ne (0.1 to 1\\% in mass fraction), which is the main neutron source for the s process in massive stars. Therefore rotation boosts the s process in massive stars at all metallicities. Second, the neutron-to-seed ratio increases with decreasing $Z$ in models including rotation, which leads to the complete consumption of all iron seeds at metallicities below $Z = 10^{-3}$ by the end of core He-burning. Thus at low $Z$, the iron seeds are the main limitation for this boosted s process. Third, as the metallicity decreases, the production of elements up to the Ba peak increases at the expense of the elements of the Sr peak. We studied the impact of the initial rotation rate and of the highly uncertain $^{17}$O$(\\alpha,\\gamma)$ rate (which strongly affects the strength of $^{16}$O as a neutron poison) on our results. This study shows that rotating models can produce significant amounts of elements up to Ba over a wide range of $Z$, which has important consequences for our understanding of the formation of these elements in low-metallicity environments like the halo of our galaxy and globular clusters. Fourth, compared to the He-core, the primary $^{22}$Ne production induced by rotation in the He-shell is even higher (greater than 1\\% in mass fraction at all metallicities), which could open the door for an explosive neutron capture nucleosynthesis in the He-shell, with a primary neutron source. } {} ", "introduction": "Massive stars in the early universe are different from the ones observed today. Their low metal content makes them more compact and rotate faster than their equivalents found in the Milky Way. This view is supported by observations of an increasing Be/B-type star ratio with decreasing metallicity \\citep{2007A&A...472..577M} and by more rapidly rotating, massive stars in the SMC compared to the Milky Way \\citep{2008A&A...479..541H}. Rapidly rotating stellar models at low $Z$ have been calculated by \\citet{2006A&A...447..623M} and \\citet{2007A&A...461..571H}. In these models, primary nitrogen yields are much larger than in non-rotating models. When yields from these rotating models are used as input in chemical evolution models, a nice fit of the N/O in very metal poor halo stars \\citep[see e.g.][]{2005A&A...430..655S} can be obtained \\citep{2006A&A...449L..27C}. This provides a strong support for the occurrence of rotation-induced mixing at low $Z$ and for our models. The primary nitrogen production in rotating low-$Z$ stellar models is accompanied by the primary production of other isotopes like $^{13}$C, and especially $^{22}$Ne, which is the neutron source for s process in massive stars \\citep[e.g.][and references therein]{2011RvMP...83..157K}. A first attempt to assess the impact of rotation on the s-process nucleosynthesis in low-$Z$ massive rotating stars was made by \\citet{2008ApJ...687L..95P}, who investigated the impact of primary $^{22}$Ne in a parametrised way. The recent observation of large s-process enhancements in one of the oldest globular clusters in the bulge of our galaxy supports the view that massive stars could indeed also be important sources for these elements \\citep{2011Natur.472..454C}, highlighting the need for comprehensive calculations of the s process in low-$Z$ massive rotating stars. In this letter, we present the first such low-$Z$, massive rotating star models including a full s-process network. We describe our models in Sect.~\\ref{sec:mixing}. The impact of rotation on the s process in massive 25~$M_\\odot$ stars at different metallicities is discussed in Sect.~\\ref{sec:sproc}. We give our conclusions in Sect.~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} We have calculated the first complete stellar models including a full s-process network to study the effects of rotation on the s process in massive stars in low-metallicity environments. Our models confirm that rotation-induced mixing in radiative zones leads to a primary production of $^{22}$Ne in the He core over the wide $Z$ range studied ($\\mathrm{[Fe/H]}=-5.8$, $-3.8$, and $-1.8$). From $0.1$ to $1$\\% of $^{22}$Ne in mass fraction is produced in central He-burning. The large primary production of $^{22}$Ne also occurs in the He-shell in all the rotating models calculated, i.e. 1-2\\% in mass fraction at all $Z$. This can have a strong impact on the supernova nucleosynthesis in He-burning shell, since it could provide a primary neutron source for neutron capture nucleosynthesis at all values of Z. We investigated the impact of the $^{17}$O$(\\alpha,\\gamma)$-rate on the strength of $^{16}$O as a neutron poison and find the impact of current uncertainties still significant, thus deserving further experimental studies. The s process in massive stars with rotation contributes significantly to the production of elements up to Sr above $\\mathrm{[Fe/H]}\\approx-2$, the main limiting factor being the iron seeds at low $Z$. The production of elements up to Ba is sensitive to the initial rotation rate, and we generally obtain $\\mathrm{[Sr/Ba]}>0$ from low-$Z$ massive rotating star s process. Pb is only produced in very extreme cases, but with $\\mathrm{[Pb/Sr]}\\apprle-1$ massive stars cannot be responsible for CEMP-s stars. A similar spread for [Sr/Ba] to the one used in \\citet{2011Natur.472..454C} was obtained to explain the unique abundance of several stars in one of the oldest globular clusters in the galactic bulge, thus supporting the conclusions of that study. Further stellar and galactic chemical evolution models will assess the full impact of this boosted s~process." }, "1112/1112.4778_arXiv.txt": { "abstract": "In contrast to the prevailing view in the literature, it is shown that even extremely stiff sets of ordinary differential equations may be solved efficiently by explicit methods if limiting algebraic solutions are used to stabilize the numerical integration. The stabilizing algebra differs essentially for systems well-removed from equilibrium and those near equilibrium. Explicit asymptotic and quasi-steady-state methods that are appropriate when the system is only weakly equilibrated are examined first. These methods are then extended to the case of close approach to equilibrium through a new implementation of partial equilibrium approximations. Using stringent tests with astrophysical thermonuclear networks, evidence is provided that these methods can deal with the stiffest networks, even in the approach to equilibrium, with accuracy and integration timestepping comparable to that of implicit methods. Because explicit methods can execute a timestep faster and scale more favorably with network size than implicit algorithms, our results suggest that algebraically-stabilized explicit methods might enable integration of larger reaction networks coupled to fluid dynamics than has been feasible previously for a variety of disciplines. ", "introduction": "Introduction} Problems from many disciplines require solving large coupled reaction networks. Representative examples include reaction networks in combustion chemistry \\cite{oran05}, geochemical cycling of elements \\cite{magick}, and thermonuclear reaction networks in astrophysics \\cite{hix05,timmes}. The differential equations used to model these networks usually exhibit stiffness, which arises from multiple timescales in the problem that differ by many orders of magnitude \\cite{oran05,gear71,lamb91,press92}. Sufficiently complex physical systems often involve important processes operating on widely-separated timescales, so realistic problems tend to be at least moderately stiff. Some, such as astrophysical thermonuclear networks, are extremely stiff, with 10--20 orders of magnitude between the fastest and slowest timescales in the problem. Our concern here is with stiffness as a numerical issue, but we remark that stiffness can have important physical implications because complex processes often function as they do precisely because of the coupling of very slow and very fast scales within the same system. Books on numerical and computational methods routinely state \\cite{oran05,lamb91,press92} that stiff systems cannot be integrated efficiently using explicit finite-difference methods because of stability issues: for an explicit algorithm, the maximum stable timestep is set by the fastest timescales, even if those timescales are peripheral to the main phenomena of interest. The standard resolution of the stiffness problem uses implicit integration, which is stable for stiff systems but entails substantial computational overhead because it requires the inversion of matrices at each integration step. Because of the matrix inversions, implicit algorithms tend to scale from quadratically to cubically with network size unless favorable matrix structure can be exploited. Thus, implicit methods can be expensive for large networks. A simple but instructive example of stiffness is provided by the CNO cycle for conversion of hydrogen to helium, which powers main-sequence stars more massive than the Sun (Fig.~\\ref{fig:cnoCycle}). In the CNO cycle the fastest rates under characteristic stellar conditions are $\\beta$-decays with half-lives $\\sim 100$ seconds, but to track the complete evolution of main-sequence hydrogen burning may require integration of the network for hydrogen burning over timescales as large as billions of years ($\\sim 10^{16}$ seconds). If one tries to implement this integration using explicit forward differencing, the largest stable integration timestep will be set by the fastest rates and will be of order $10^2$ seconds. Thus $10^{14}$ or more explicit integration steps could be required to integrate the CNO cycle to hydrogen depletion. Conversely, typical implicit integration schemes can take stable and accurate timesteps equal to 1-10\\% of the local time over most of the integration range, and would compute the above numerical integration in a few hundred implicit steps. By virtue of examples such as this, it is broadly accepted that explicit methods are not viable for stiff networks. To quote the authoritative reference {\\em Numerical Recipes} \\cite{press92}, ``For stiff problems we {\\em must} use an implicit method if we want to avoid having tiny stepsizes.'' \\putfig {cnoCycle} {0pt} {\\figdn} {0.90} {fig1.eps} {The CNO (carbon--nitrogen--oxygen) cycle. On the left side the main branch of the cycle is illustrated with solid arrows and a side branch is illustrated with dashed arrows. On the right side, the main branch of the CNO cycle is illustrated in more detail.} Our main interest lies in simulations where the reaction network is one part of a broader problem. Let us take as representative astrophysical thermonuclear networks coupled to multidimensional hydrodynamics. The hydrodynamical evolution controls network conditions (temperature, density, \\ldots), and the network influences the hydrodynamics through energy production and modification of composition. Solution of large networks by the usual means is costly in this context and even ambitious simulations use only small networks, or replace the network entirely by parameterization. Then a more realistic network is used in a separate ``post-processing'' step, where fixed hydrodynamical profiles computed in the original simulation specify the variation of temperature and density with time. Such approximations are especially at issue for problems like Type Ia supernovae, which are 3D asymmetric explosions powered by a complex reaction network releasing energy greater than that of a large galaxy on a timescale of order 1 s. Astrophysical reaction networks have been used to illustrate, but problems in various fields exhibit a similar complexity. For example, in astrochemical kinetics large chemical evolution networks must be modeled in dynamical environments such as contracting molecular clouds, or in combustion chemistry the burning networks are strongly coupled to dynamical simulations of the air and fuel mixture. Realistic networks in all such applications may be quite large. Modeling combustion of larger hydrocarbon molecules or soot formation can require hundreds to thousands of reacting species with up to 10,000 reactions \\cite{oran05}, and realistic networks for supernova explosions imply hundreds to thousands of nuclear isotopes with tens of thousands of reaction couplings \\cite{hix05}. In all such cases current techniques do not permit the coupling of realistic reaction networks to the full dynamics and highly-schematic reaction networks are used in even the most realistic contemporary simulations. ", "conclusions": "Conclusions} This paper demonstrates that algebraically-stabilized explicit integration is capable of timesteps competitive with those of implicit methods for various extremely-stiff reaction networks. Since explicit methods can execute a timestep faster than an implicit method in a large network, our results suggest that algebraically-stabilized explicit algorithms may be capable of performing as well as, or even substantially outperforming, implicit integration in a variety of moderate to extremely stiff applications. Because of the linear scaling with reaction network size for explicit methods, this fundamentally new view of explicit integration for stiff equations is particularly important for applications in fields where more realistic---and therefore larger---reaction networks are required for physical simulations. Arguably, this means almost all scientific and technical disciplines, since the sizes of reaction networks being used in simulations to this point have been dictated more often by what was feasible than by what was physical. Of particular significance is that these new explicit methods might permit coupling of more physically-realistic reaction kinetics to fluid dynamics simulations in a variety of disciplines. \\begin{ack} I thank Jay Billings, Reuben Budiardja, Austin Harris, Elisha Feger, and Raph Hix for help with some of the calculations. Discussions with Raph Hix, Bronson Messer, Kenny Roche, Jay Billings, Brad Meyer, Friedel Thielemann, Michael Smith, and Tony Mezzacappa have been useful in formulating the ideas presented here, and I thank Bronson Messer for a careful reading of the manuscript. Research was sponsored by the Office of Nuclear Physics, U.S. Department of Energy. \\end{ack} \\vfill \\clearpage \\setcounter{section}{0} \\appendix" }, "1112/1112.1688_arXiv.txt": { "abstract": "Astronomy has long had a working network of archives supporting the curation of publications and data. The discipline has already created many of the features which perplex other areas of science: (1) data repositories: (supra)national institutes, dedicated to large projects; a culture of user-contributed data; practical experience of long-term data preservation; (2) dataset identifiers: the community has already piloted experiments, knows what can undermine these efforts, and is participating in the development of next-generation standards; (3) citation of datasets in papers: the community has an innovative and expanding infrastructure for the curation of data and bibliographic resources, and through them a community of authors and editors familiar with such electronic publication efforts; as well, it has experimented with next-generation web standards (e.g. the Semantic Web); (4) publisher buy-in: publishers in this area have been willing to innovate within the constraints of their commercial imperatives. What can possibly be missing? Why don't we have an integrated framework for the publication and preservation of all data products already? Are there technical barriers? We don't believe so. Are there cultural or commercial forces inhibiting this? We aren't aware of any. This Birds of a Feather session (BoF) attempted to identify existing barriers to the creation of such a framework, and attempted to identify the parties or groups which can contribute to the creation of a VO-powered data-publishing framework. ", "introduction": "This BoF session provided a forum for data providers, publishers, librarians and scientists to explore the issues surrounding the preservation, identification and citation of data products in astronomy. These are small but critical steps towards the ultimate goal of identifying the right practices, resources, and incentives to ensure that the entire research lifecycle in astronomy is properly captured and described in the coming era of data-intensive astronomical research. The authors of this paper (and BoF panelists) were chosen to represent varied perspectives from different corners of our community, including: the ADS, the primary literature portal in astronomy; the CDS, one of the largest data curation hubs in the Virtual Obsevatory (VO); the AAS, the largest publisher of scholarly literature in astronomy; and the University of Glasgow, an institution involved in data management and resource discovery for the VO. The following sections highlight some of the topics covered in the BoF. ", "conclusions": "That there is a need for curated datasets, preserved indefinitely for scholarly purposes, is axiomatic by now. The details of a fully-fledged data preservation environment are difficult to discern at this point, but there are sensible places to begin, and from such places the mechanisms can evolve safely. The scientific and scholarly communities are vigorously pursuing possibilities, and there are many fruitful opportunities, such as this session at ADASS, for interested parties to engage in discussions about the promises and the challenges of broadly-scaled data preservation." }, "1112/1112.3191_arXiv.txt": { "abstract": "We announce the identification of a proper motion companion to the star HII~1348, a K5\\,V member of the Pleiades open cluster. The existence of a faint point source 1$\\farcs$1 away from HII~1348 was previously known from adaptive optics imaging by Bouvier et al. However, because of a high likelihood of background star contamination and in the absence of follow-up astrometry, Bouvier et al.\\ tentatively concluded that the candidate companion was not physically associated with HII~1348. We establish the proper motion association of the pair from adaptive optics imaging with the Palomar 5~m telescope. Adaptive optics spectroscopy with the integral field spectrograph OSIRIS on the Keck 10~m telescope reveals that the companion has a spectral type of M8$\\pm$1. According to substellar evolution models, the M8 spectral type resides within the substellar mass regime at the age of the Pleiades. The primary itself is a known double-lined spectroscopic binary, which makes the resolved companion, HII~1348B, the least massive and widest component of this hierarchical triple system and the first substellar companion to a stellar primary in the Pleiades. ", "introduction": "As one of the nearest young \\citep[125~Myr;][]{stauffer1998} open clusters, the Pleiades have long been recognized as an important astrophysical laboratory for studying stellar evolution and the dynamics of stellar associations. Multiplicity studies of the Pleiades have focused both on stellar \\citep[e.g.,][]{stauffer1984, mermilliod1992, bouvier1997} and substellar \\citep[e.g.,][]{martin2000, bouy2006} members of the cluster. However, mixed star--brown dwarf systems have not been identified. One of the most extensive studies is the adaptive optics imaging survey of \\citet{bouvier1997}. Conducted in the near-IR, it covered 144 G and K stars to a relatively shallow depth. As a result, its sensitivity encompassed only stellar and massive substellar companions. A systematic high-contrast imaging survey of the Pleiades on a similar scale but at a higher sensitivity has not been performed since. Consequently, no substellar companions are known to $>$0.2~\\Msun\\ stars in the Pleiades. With the frequency of wide substellar companions to field-aged Sun-like stars now estimated at $\\approx$\\,3\\,\\% \\citep[0.012\\,--\\,0.072~\\Msun\\ brown dwarfs in 28\\,--\\,1590\\,AU orbits;][]{metchev2009}, the frequency of brown dwarf secondaries around Sun-like stars in the Pleiades is expected to be comparable. In the present paper, we announce the identification of a low mass companion to the Pleiad HII~1348, a K5\\,V double-lined spectroscopic binary \\citep[hereafter refered to as the primary or HII~1348A]{queloz1998}. The faint companion, HII~1348B, was already detected by \\citet{bouvier1997}. However, without follow-up astrometric observations and due to a non-negligible probability of background star contamination, \\citet{bouvier1997} conservatively assumed that the candidate companion was an unrelated background star. The astrometric measurements confirm the proper motion association of the pair, and AO spectra obtained at Palomar and Keck reveal that the companion has a spectral type of M8. % ", "conclusions": "HII~1348B is a new M8 brown dwarf member of the Pleiades, and the first substellar companion discovered around a Pleiades star. Given that no other substellar companions were discovered in the \\citet{bouvier1997} survey at similar or wider separations, it is worth considering whether HII~1348B may be unusually weakly bound, compared to other binary systems in the Pleiades or in the field. \\citet{bouvier1997} found a total of 28 stellar binaries in the Pleiades in their CHFT AO survey. HST surveys of very low mass stars and brown dwarfs conducted by \\citet{martin2003} and \\citet{bouy2006} revealed three additional binaries. In Figure \\ref{fig:Ebind} we compare the binding energies of these systems, and those of field A--M binaries \\citep{close1990, close2003, close2007}, to the binding energy of HII~1348A/B. As can be seen, HII~1348A/B sits in the middle of the locus for stellar binaries, and is comparably or more tightly bound even than the three very low mass Pleiades binaries. What is more, substellar companions up to 10 times further away from their primaries would still be well above the minimum stellar binding energy in the Pleiades. The dearth of known brown dwarf companions to stars in the Pleiades may thus be attributable to the lack of a follow-up sensitive and comprehensive high-contrast imaging survey of the cluster. Small samples of Pleiades stars have since been observed in deep surveys by \\citet[23 stars;][]{metchev2009} and \\citet[14 stars;][]{tanner2007}, with no new brown dwarf companion detections. However, a much more comprehensive survey is needed to reveal brown dwarf companions with any statistical confidence \\citep{metchev2009}. Current AO systems at Keck or Gemini North should allow the detection of companions with masses down to $\\sim$0.03\\,$\\Msun$ at separations larger than $\\sim$60AU \\citep[0$\\farcs$5;][]{lafreniere2007, metchev2009b}. Using the Gemini North AO system Altair in combination with NIRI, \\citet{lafreniere2007} showed that companions up to 9.5\\,mag fainter can be detected at separations of $>$0$\\farcs$5 from the primary. Accounting for the narrow band filter used during the observation, brown dwarfs down to 0.03\\,$\\Msun$ should be detectable around bright (V$<$12\\,mag) Pleiades stars. Likewise, using angular differential imaging \\citep[ADI,][]{marois2006} in combination with the Keck AO system, \\citet{metchev2009b} demonstrated that companions up to $\\Delta$H=10.5\\,mag fainter than the primary can be detected at separations larger than 0$\\farcs$5. Thus, at the distance of the Pleiades brown dwarfs down to 0.02\\,$\\Msun$ should be detectable at separations larger than $\\sim$60AU. Upcoming instruments like the Gemini planet imager \\citep{macintosh2008, mcbride2011} and the extreme AO system at Palomar \\citep[PALM3000,][]{bouchez2008} will enable detections of brown dwarf companions at even smaller separation down to 0$\\farcs$2. Given the importance of the Pleiades as an age and a dynamical benchmark for stellar and substellar evolution, a deep AO survey could have a high impact in delivering substellar objects much fainter and cooler than the ones presently known from wide-area surveys. While wide-area surveys are more efficient in discovering large numbers of substellar candidate members, the limited precision of seeing-limited astrometry, and the increasingly challenging radial velocity measurements at fainter magnitudes, prevent the unequivocal confirmation of the lower mass candidates as bona-fide members. The factor of $\\sim$100 higher precision attainable over narrow angles with AO leaves little doubt about the astrometric association of candidate binaries, and is thus also an excellent tool for confirming the cluster membership of even the lowest-mass, faintest companions." }, "1112/1112.3969_arXiv.txt": { "abstract": "{ During the last decade, the FORS1 instrument of the ESO Very Large Telescope has been extensively used to study stellar magnetism. A number of interesting discoveries of magnetic fields in several classes of stars have been announced, many of which obtained at a $\\sim 3\\,\\sigma$ level; some of the discoveries are confirmed by measurements obtained with other instruments, some are not. } { We investigate the reasons for the discrepancies between the results obtained with FORS1 and those obtained with other instruments. } { Using the ESO FORS pipeline, we have developed a semi-automatic procedure for magnetic field determination. We have applied this procedure to the full content of circular spectropolarimetric measurements of the FORS1 archive (except for most of the observations obtained in multi-object spectropolarimetric mode). We have devised and applied a number of consistency checks to our field determinations, and we have compared our results to those previously published in the literature. } { We find that for high signal-to-noise ratio measurements, photon noise does not account for the full error bars. We discuss how field measurements depend on the specific algorithm adopted for data reduction, and we show that very small instrument flexures, negligible in most of the instrument applications, may be responsible for some spurious field detections in the null profiles. Finally, we find that we are unable to reproduce some results previously published in the literature. Consequently, we do not confirm some important discoveries of magnetic fields obtained with FORS1 and reported in previous publications. } { Our revised field measurements show that there is no contradiction between the results obtained with the low-resolution spectropolarimeter FORS1 and those obtained with high-resolution spectropolarimeters. FORS1 is an instrument capable of performing reliable magnetic field measurements, provided that the various source of uncertainties are properly taken into account. } ", "introduction": "During a full decade of operations, the FORS1 instrument of the ESO Very Large Telescope has collected a large number of magnetic field measurements in various kinds of stars. Together with the ESPaDOnS instrument of the Canada-France-Hawaii Telescope, and with the MuSiCoS and NARVAL instruments of the 2\\,m Telescope Bernard Lyot of the Pic-du-Midi Observatory, FORS1 has been one of the workhorse instruments for the observational studies of stellar magnetism. Several important detections obtained with FORS1 have led to the conclusion that magnetic fields are quite common in a variety of stars across the Hertzsprung-Russell diagram, including for instance central stars of planetary nebulae \\citep{Joretal05}, hot subdwarfs \\citep{Otoetal05}, $\\beta$\\,Cephei and slowly pulsating B stars \\citep{Hubetal09a}, B stars with emission lines \\citep{Hubetal09b}, and normal O-type stars \\citep{Hubetal08b}. In practice, a close inspection to the published results shows a number of problems: \\noindent \\textit{i)} Inconsistencies between field measurements obtained with FORS1 and field measurements obtained with other instruments. For instance, \\citet{Hubetal04a} reported the discovery of a magnetic field in the Herbig Ae/Be star HD\\,139614, while repeated ESPaDOnS measurements failed to confirm the magnetic nature of that star \\citep{Wadetal05}. Similarly, \\citet{Siletal09} failed to confirm the presence of a magnetic field in several of the $\\beta$\\,Cep and SPB stars observed by \\citet{Hubetal06a} and \\citet{Hubetal09a}. \\noindent \\textit{ii)} Inconsistencies between the analysis of the same FORS dataset performed by different authors. For instance, \\citet{McSwain08} observed the normal B stars NGC 3766 MG 111 and NGC 3766 MG 176, and the Be star NGC 3766 MG 200, and reported no field detection. Using the same data, \\citet{Hubetal09b} reported new field detections for all three stars. \\noindent \\textit{iii)} Inconsistencies between the analysis of the same FORS dataset performed by the same authors but at different epochs. For instance, \\citet{Wadetal05} reported a possible detection in the young Ap star HD\\,72106A, which was not confirmed by the later analysis of the same data performed by the same group \\citep{Wadetal07a}. Note that the magnetic nature of that star was established with data independently obtained with ESPaDOnS \\citep[see][]{Foletal08}. \\noindent \\textit{iv)} Inconsistencies between field measurements obtained from different subsets of an observing series of frames. Magnetic field measurements are often obtained by combining a number of pairs of frames obtained at two different position angles of the retarder waveplate. This redundancy is mainly motivated by the need to reach a very high signal-to-noise ratio. In some rare cases, a magnetic field may be firmly detected in a pair of frames, but not in the remaining pairs. This is for instance the case of the measurements of HD\\,139614 reported by \\citet{Hubetal04a}, where the magnetic field is detected only in a subset of frames, and in a couple of H Balmer lines. \\noindent \\textit{v)} Finally, there are some global inconsistencies of the full FORS dataset, revealed for instance by the high incidence of field detection in the null profiles; this kind of problem was not previously reported in the literature, and will be discussed in Sect.~\\ref{Sect_Internal_Tests}. Issues \\textit{ii)} and \\textit{iii)} point to possible glitches in the data reduction method, while issues \\textit{i)}, \\textit{iv)}, and \\textit{v)} might be the symptom of a wider, possibly instrumental, problem. The release of the FORS pipeline for spectropolarimetric data \\citep{Izzetal10,Izzetal11} gave us the opportunity to develop an accurate and efficient reduction method. The FORS pipeline is a software tool specifically designed for that instrument, which makes it easier to handle some characteristics that are specific to the data obtained with FORS1. Furthermore, the FORS pipeline allows a high degree of automatization in the data reduction process. Complemented with a suite of software tools for data pre-processing (for frame classification and quality check) and data post-processing (for the computation of the magnetic field), we were able to build up a nearly automatic tool for data analysis which allowed us to treat the entire FORS archive data in a homogeneous way. Using our tool suite, the difference in terms of effort required to perform the analysis of a single series of observations compared to performing the analysis of the entire archive of FORS1 data consists mainly in the amount of the necessary disk space and CPU time. Compared to the reduction of raw data coming from individual observing runs, the mass-production of reduced spectra offers the possibility to perform a quality check of the final products on a very large scale. The aim of this paper is to present: (1) a preliminary discussion of the methods, (2) the results of our quality checks, and (3) to caution the reader about the lack of robustness of some of the results previously published in the literature. A deeper analysis and a comprehensive and homogeneous catalogue of FORS1 magnetic field measurements will be published in a forthcoming paper. ", "conclusions": "" }, "1112/1112.4408_arXiv.txt": { "abstract": "We investigate the observability of cold accretion streams at redshift 3 via Lyman-alpha (\\lya{}) emission and the feasibility of cold accretion as the main driver of \\lya{} blobs (LABs). We run cosmological zoom simulations focusing on 3 halos spanning almost two orders of magnitude in mass, roughly from $10^{11}$ to $10^{13}$ solar masses. We use a version of the \\ramses{} code that includes radiative transfer of ultraviolet (UV) photons, and we employ a refinement strategy that allows us to resolve accretion streams in their natural environment to an unprecedented level. For the first time in a simulation, we self-consistently model self-shielding in the cold streams from the cosmological UV background, which enables us to predict their temperatures, ionization states and \\lya{} luminosities with improved accuracy. We find the efficiency of gravitational heating in cold streams in a $\\sim10^{11}$ solar mass halo to be around 10-20\\% throughout most of the halo but reaching much higher values close to the center. As a result most of the \\lya{} luminosity comes from gas which is concentrated at the central 20\\% of the halo radius, leading to \\lya{} emission which is not extended. In more massive halos, of $\\ga 10^{12}$ solar masses, cold accretion is complex and disrupted, and gravitational heating does not happen as a steady process. Ignoring the factors of \\lya{} scattering, local UV enhancement, and SNe feedback, the cold `messy' accretion alone in these massive halos can produce LABs that largely agree with observations in terms of morphology, extent, and luminosity. Our simulations slightly and systematically over-predict LAB abundances, perhaps hinting that the interplay of these ignored factors may have a negative net effect on extent and luminosity. We predict that a factor of a few increase in sensitivity from current observational limits should unambiguously reveal continuum-free accretion streams around massive galaxies at $z=3$. ", "introduction": "\\label{Intro.sec} The last decade has seen a shift in the way galaxies are thought to have assembled. In the classic theory \\citep{Rees:1977p4388, Silk:1977p4383,White:1978p4389}, galaxies collect their baryons via so-called hot mode accretion where diffuse gas symmetrically falls into dark matter (DM) halos and is shock-heated as it hits the gas residing in them. Depending on the mass of the halo, the gas may or may not eventually settle into the galaxy. However, it has become increasingly apparent through theoretical work and simulations that at high redshift ($z\\ga 2$), galaxies get their baryons primarily via accretion of relatively dense, cold ($10^4$ K) and pristine gas which penetrates in the form of \\textit{streams} through the diffuse shock-heated medium \\citep{Fardal:2001p3736, Birnboim:2003p3602, Keres:2005p3601, Dekel:2006p4450, Birnboim:2007p4448, Ocvirk:2008p2688, Dekel:2009p1318, Brooks:2009p3604, vandeVoort:2011p5673, FaucherGiguere:2011p5611, vandeVoort:2011p5669}. Simulations consistently show these streams to exist and peak in activity around redshift 3, though it appears that their widths are still dictated mostly by resolution. The problem is that cold accretion streams have never been directly observed, though we are starting to see some hints, both in emission \\citep{Rauch:2011p5439} and absorption \\citep{Ribaudo:2011p5454}. Is this lack of observational evidence consistent with the existence of cold accretion streams? Do we not observe them because they're not easily observable or simply because they don't exist? \\cite{FaucherGiguere:2011p3606} showed that the streams are hard to detect directly via absorption due to their small covering factor and surrounding galactic winds that overwhelm their signature. \\cite{Kimm:2011p4491} came to the same conclusion, adding that the low metallicity in streams ($\\leq 10^{-3}$ solar) further inhibits their detection via metal line absorption. Even so, \\cite{Fumagalli:2011p2943} and \\cite{vandeVoort:2011p5667} have argued that a large fraction of observed metal-poor Lyman-limit systems (LLSs) make up for indirect detections of cold streams. Furthermore, we may possibly have been directly observing the tips of these streams during the last decade in the form of Lyman-alpha blobs (LABs). \\vsk LABs are extremely bright ($\\ga 10^{43} \\; \\mathrm{erg \\; s^{-1}}$) and extended ($\\ga 30 \\;\\mathrm{kpc}$ in diameter) \\lya{} nebulae \\citep[e.g.][]{Francis:1996p4544, Keel:1999p4529, Steidel:2000p2213, Matsuda:2004p3081, Palunas:2004p4550, Nilsson:2006p3525, Smith:2007p4610, Prescott:2009p3951, Yang:2010p3447, Erb:2011p5386}. They have a slight tendency to be filamentary in structure \\citep[][hereafter M11]{Matsuda:2011p5426}, and often have short limbs protruding from the main body. They often coincide with galactic sources that give hints about their physical origin but the mechanism by which the emission becomes so strong and extended is a matter of debate. A subset of LABs however have no apparent coinciding galactic sources \\citep[e.g.][]{Steidel:2000p2213, Weijmans:2010p5527, Erb:2011p5386}. Up until now about two hundred LABs have been discovered, including about fifteen giant ones ($>100$ kpc). Smaller extended \\lya{} emitters exist in large quantities over a continuous range of sizes down to point sources. LABs appear to be specific to the high-redshift Universe \\citep{Keel:2009p4857} and most of them have been detected at $20.3\\ \\cci$) consistently accounts for $40\\%$ of the total \\lya{} luminosities of our halos, so we can estimate the total \\lya{} luminosities to be uncertain by (very) roughly $50\\%$, and even more if we exclude still more diffuse gas than the CGM. We have however shown that our results and conclusions regarding LAB areas are not sensitive to the density threshold applied (i.e. above which densities we ignore \\lya{} emissivity). \\vsk Our main results are the following: \\begin{itemize} \\item{ Cold accretion streams in halos more massive than $\\sim 10^{12}$~M$_\\odot$ produces extended and luminous \\lya{} nebulae which are by large compatible with LABs observed at $z\\sim 3$, in terms of morphology, luminosity and extent. Gravity alone provides most of the energy, and we find that extra sources such as UV fluorescence, \\lya{} scattering or superwinds are not necessary. This clearly doesn't rule out these other processes though, as they are likely all significant in the case of LABs, and further work is needed to study their complex interplay.} \\item{In our simulations, LAB area and luminosity are reasonably well-behaved functions of halo mass. We use these relations to compute the cumulative luminosity and area distributions, and find that they are in reasonable agreement with observations given the relatively large uncertainties. This comparison however suggests that the combined effects of SN feedback, \\lya{} scattering and an enhanced local UV field may possibly have a negative impact on the luminosity and extent of simulated LABs, when conjoined with cold accretion. } \\item{ The model of gravitational heating as a driver of extended \\lya{} emission works according to our results, but we need to alter our notion of \\textit{how} it works: It is inefficient in the classic sense where gas accretion is smooth. Rather the accretion is messy and disrupted in massive halos and probably involves some mass loss to the surrounding hot diffuse medium. } \\item{Our examination of maximum photofluorescence hints that in extreme cases local UV enhancement, e.g. near quasars, can boost the \\lya{} luminosity of LABs and to a lesser degree their extent. As demonstrated by \\citet[][]{Cantalupo:2005p4317} and \\citet{Kollmeier:2010p3256}, this means that large accretion flows may be more easily observed in the proximity of quasars than elsewhere.} \\item{We find that cold accretion streams should be unambiguously observable via direct \\lya{} emission for the first time in the near future, on upcoming instruments such as MUSE and K-CWI which will allow to probe emission at surface brightnesses as low as $\\sim 10^{-19}\\IS$.} \\end{itemize} Although we have significantly improved on previous work, a large number of theoretical issues remain to be addressed. In forthcoming papers, we plan to investigate the effects of \\lya{} scattering SNe-driven winds and local UV enhancement from star formation." }, "1112/1112.5455_arXiv.txt": { "abstract": "Existing photometry for NGC 2264 tied to the Johnson \\& Morgan (1953) {\\it UBV} system is reexamined and, in the case of the original observations by Walker (1956), reanalyzed in order to generate a homogeneous data set for cluster stars. Color terms and a Balmer discontinuity effect in Walker's observations were detected and corrected, and the homogenized data were used in a new assessment of the cluster reddening, distance, and age. Average values of $E_{B-V}= 0.075 \\pm0.003$ s.e. and $V_0${\\it --M}$_V=9.45 \\pm 0.03$ s.e. ($d = 777 \\pm12$ pc) are obtained, in conjunction with an inferred cluster age of $\\sim 5.5 \\times 10^6$ yr from pre-main-sequence members and the location of the evolved, luminous, O7 V((f)) dwarf S Mon relative to the ZAMS. The cluster main sequence also contains gaps that may have a dynamical origin. The dust responsible for the initial reddening towards NGC 2264 is no more than 465 pc distant, and there are numerous, reddened and unreddened, late-type stars along the line of sight that are difficult to separate from cluster members by standard techniques, except for a small subset of stars on the far side of the cluster embedded in its gas and dust and background B-type ZAMS members of Mon OB2. A compilation of likely NGC 2264 members is presented. Only 3 of the 4 stars recently examined by asteroseismology appear to be likely cluster members. NGC 2264 is also noted to be a double cluster, which has not been mentioned previously in the literature. ", "introduction": "} ``How often have I said to you that when you have eliminated the impossible, whatever remains, however improbable, must be the truth?'' Sherlock Holmes to Dr. Watson (Conan~Doyle 1890). In the study of open clusters a similar totalogy could be paraphrased as ``when you have eliminated the effects of extinction and differential reddening for cluster stars, the resulting cluster color-magnitude diagram, however unusual, must represent an accurate picture of the temperatures and luminosities of cluster stars.'' That philosophy was demonstrated to be the case for relatively young (10$^7$--10$^8$ yr old) open clusters by Turner (1996), as well as for the extremely young ($3.5 \\times 10^6$ yr old) cluster IC 1590 (Guetter \\& Turner 1997). The last study demonstrated how closely pre-main-sequence members of IC 1590 matched model isochrone predictions by Palla \\& Stahler (1993) converted to the Johnson \\& Morgan (1953) {\\it UBV} system. The situation for other young clusters is more complicated, primarily because most of the original {\\it UBV} studies of the brighter members of the class (NGC 2264, NGC 6530, IC 1546, and NGC 6611) were made by Merle Walker (Walker 1956, 1957, 1959, 1961, respectively) using a detector system that was not an ideal match to the Johnson system, although that was not realized at the time. For example, later studies of NGC 6611 (Hiltner \\& Morgan 1969) and NGC 2264 (e.g., Mendoza \\& G\\'{o}mez 1980) noted differences between Walker's photometry and observations tied more closely to the {\\it UBV} system. The origin of such differences can be explained by the work of Moffat \\& Vogt (1977) and Guti\\'{e}rrez-Moreno, Moreno \\& Cort\\'{e}z (1981), who noted that systematic errors, specifically in {\\it U--B} measures, can arise from the manner in which the Balmer discontinuity in the continua of early-type stars is sampled by non-standard telescope/filter/detector systems, as well as by the treatment of atmospheric extinction (see Cousins \\& Caldwell 2001). The differences in the case of NGC 6611 are fairly extreme, amounting to offsets of $0.10$ in {\\it U--B} and $0.03$ in {\\it B--V} (Hiltner \\& Morgan 1969). The purpose of the present study is to redo Walker's original study of NGC 2264 (Walker 1956) in order to generate a new reddening-free and extinction-corrected color-magnitude diagram for the cluster. The cluster has been studied many times previously, but never for the purpose of improving upon Walker's results. Recent detections of non-radial pulsation via asteroseismology in many of the pre-main-sequence members of NGC 2264 (e.g., Zwintz 2008; Kallinger, Zwintz \\& Weiss 2008; Guenther et al. 2009) make it imperative to have a clear picture of the evolutionary status and exact H-R diagram location of cluster stars. ", "conclusions": "} One advantage of removing the effects of interstellar reddening and extinction from cluster color-magnitude diagrams is that the resulting distribution of data points must be closely linked to the true variations in effective temperature and luminosity for cluster members. In the case of NGC 2264, the color-magnitude diagram presented here in Fig.~\\ref{fig5} provides a number of important insights into the evolutionary status of cluster stars. The main sequence gaps, for example, imply an advanced dynamical state, provided they arise from close binary mergers, as argued by Turner (1996). And the close coincidence in the cluster age inferred from both pre-main-sequence stars and the slightly-evolved S Mon imply that the formation of cluster stars was not spread out greatly in time, the creation of cluster stars probably consuming no more than a few hundred thousand years. \\subsection*{{\\rm \\scriptsize ACKNOWLEDGEMENTS}} \\scriptsize{The present study has used the open cluster data compilation maintained by WEBDA, operated at the Institute for Astronomy of the University of Vienna.}" }, "1112/1112.0046_arXiv.txt": { "abstract": " ", "introduction": "Small bodies in the Solar System are usually divided into stable populations (like the main asteroid belt, Jupiter Trojans or the Kuiper belt) and the unstable, planet-crossing populations (like the near-Earth asteroids, comets or the Centaurs). Stable populations are seen as fossilized remnants from the formation of the Solar System, while the planet-crossing small bodies are short-lived and are constantly replenished from these stable reservoirs. There are some exceptions to this classification, like the Trans-Neptunian Scattered Disk, which is very long-lived but evolves appreciably over the age of the Solar System \\citep{vol08}. The small Hungaria asteroid group, on low-eccentricity, high-inclination orbits just beyond Mars, has also been suggested as a constantly eroding population \\citep{mce10}, but the full implications of the Hungarias' long-term decline have not been explored yet. The issue of small body stability and instability is important for determining the impact history of planets and satellites. The Moon is the best studied body in the Solar System after Earth, and reveals a record of the system's early history that has since been erased on our planet. The lunar samples returned by the Apollo missions firmly established that the Moon was subject to intense bombardment at 3.85 Gyr ago (Gya) \\citep{ter74}. The nature of lunar bombardment before that time is controversial, as is the source of the impactors at 3.85 Gya. The terms \"Lunar cataclysm\" (LC) and \"Late Heavy Bombardment\" (LHB) have been introduced to describe this bombardment, and the two terms are now used almost interchangeably. A non-primordial source of the lunar cataclysm at 3.85 Gya is based on two lines of argument. One argument is that the decay of lunar bombardment at the time of the large Imbrium impact at about 3.85 Gya is too rapid to be consistent with a slowly decaying primordial population \\citet{ccg07}. Independently, dynamical calculations show that the primordial small bodies should have been all but extinct by 3.85 Gya \\citep{bot07}. This leads to a conclusion that the lunar cataclysm impactor population must have become unstable after a period of stability. The only plausible mechanism for a delayed instability of a numerous small body population would be a late rearrangements of the planets. This could be a local event, involving an additional terrestrial planet \\citep{cha07}, or global, affecting the giant planets and, indirectly, the rest of the system \\citep{lev01, tsi05, gom05}. While there is independent evidence for an unstable episode in the Solar System's history \\citep{mor05, nes07, mor09, mor10}, there is no independent evidence about the timing of this instability; it is most often assumed to coincide with the lunar cataclysm \\citep{gom05}. In \\citet{cuk10} we argue that the record of the cataclysm does not match the predictions of the planetary migration-based models. Youngest basins (Imbrium and Orientale) seem to have formed much later than the older basins (see section 3), and were produced by a non-asteroidal impactor population \\citep{cuk10}, while planetary migration predicts the late impactors to be derived from main-belt asteroids \\citep{gom05}. More generally, the necessity of a connection between the planetary instability and bombardment at 3.85 Gya hinges on the lack of other plausible explanations for the late delivery of lunar impactors. In this paper, we will show that there should still be a significant number of primordial Mars-crossers present at 3.85 Gya, so that this population can indirectly trigger a relatively short-lived lunar cataclysm, in full agreement with the available evidence. ", "conclusions": "Here we will detail a new scenario of the chronology and sources of early lunar bombardment. This scenario is tentative and will need further improvements if it is to explain the full complexity of the lunar record. However, we believe that by proposing a number of testable predictions we can help advance lunar science. On the basis of magnetic and cratering data, we conclude that there were two populations of ancient lunar impactors, a pre-Imbrian one and an Imbrian one. These groups are largely equivalent to Populations I and II from \\citet{str05}, although there is some disagreement about the nature of Population II. Here we find that the most likely source of Population I were the primordial Mars crossers, which would have decayed with the half-life of 80 Myr and may have been able in producing $D > 900$~km basins until 4.0~Gya. The size-distribution of the Population I impactors was very similar to present-day main belt asteroids (MBAs), but there is no reason to think that other (now extinct) inner solar system small body populations did not have the same SFD. It is reasonable to expect that MBAs, Mars-crossers, the \"E-belt\" \\citep{cuk08, bot10} and the leftovers from Earth's formation \\citep{bot07} would all have had similar size-distributions. The above discussion implies that the size-distribution and composition of Population I impactors are unlikely to be specific to their source region. However, chronology of their impacts should be indicative of their dynamical evolution. If the Nectaris basin is older than about 4.2 Gyr, it could plausibly be the product of a primordial impactor population, as proposed here. If, on the other hand, it is younger than 4.1 Gyr ago, it is likely to be a part of a cataclysm-type event \\citep{bot10}. The absolute age of Nectaris basin may be impossible to determine from the existing Apollo 16 samples, which appear dominated by Imbrium ejecta \\cite{kor87, has98, war03, nor10}. The 3.9~Gya Apollo 17 ages attributed to Serenitatis may actually reflect the formation of Imbrium, with Serenitatis being much older \\citep{has98, spu11}. Unfortunately, the factors that complicate dating of the nearside Nectarian basins may similarly affect future samples from other locations on the Moon. Moscoviense may be the only lower Nectarian basin sufficiently distant from both Imbrium and Orientale impacts that it was likely not contaminated much by their ejecta, and may be (in theory) the best candidate for sample return among Nectarian basins. Dating of the largest and oldest South Pole-Aitken (SPA) basin would put a upper limit on the age of lunar surface features and confirm or rule out the total resurfacing scenario \\citep{ter74}, but would not give us any information about the chronology of much later impactors. An age of about 4.4~Gyr for SPA would allow both \"medium\" \\citep{bot10} and minimal (as proposed here) versions of the lunar cataclysm, so at least one Nectarian basin would need to be dated in order to distinguish between these two tentative chronologies. In contrast, the timing of (lower Imbrian) Population II impactors is well established, as the Imbrium basin formed at 3.85-3.87 Gya, and the most Imbrian craters formed shortly afterwards. Since the size distribution of Imbrian impactors appears to be different from that of Nectarian ones \\citep{wil78, cuk10}, and different from asteroids, their source is unknown. We propose a collisional disruption of a Vesta-sized Mars-crosser as their source \\citep{wet75} on the basis of theoretical considerations, tentative composition of the Imbrium impactor \\citep{puc08}, and possible connection to mesosiderites. A minimal cataclysm from a Mars-crossing collisional disruption also affected Mars, but the number of superposed craters indicate that the largest impact features on Mars like Argyre and Hellas are likely to predate the lunar cataclysm \\citep{tan86, wer08}. In any case, amount of mass in 200 km bodies necessary to form these basins is not consistent with the breakup of one Vesta-sized body. Our hypothesis predicts that none of the largest basins formed at Mars at 3.85 Gya, and, in general, implies much older ages for Noachian features than the models that assume a larger-scale bombardment at 3.85 Gya \\citep{wer08}. Despite the limited extent of the spike at 3.9 Gya, Mars would have experienced substantial bombardment prior to 3.6 Gya due to Mars-crossers and this bombardment was the likely the source of shocking found in the ancient martian meteorite ALH84001 \\citep{ash96}. Shock ages of HED meteorites and H-chondrites are also sometimes used to constrain the early Solar System bombardment \\citep{bog03, swi09}. Interestingly, both HED meteorites and H-chondrites record a much longer bombardment episode than the Moon, making it hard to explain both records with the same impactor population \\citep{har03}. We think that it is significant that the HED and H-chondrite parent bodies \\citep[Vesta, and, most likely, Hebe ]{mcs10, gaf98, bot10b} are both located in the inner main belt and would be likely to suffer significant number of collisions with late surviving Mars-crossers/proto-Hungarias. The shock resetting of meteorite Ar-Ar ages appears to be dependent both on the size of the body, parameters of impacts and the family history \\citep{nes09}. Therefore, while the number of impacts by Mars-crossers onto Vesta and Hebe from 3.5 to 4.1 Gya did not dwarf their collisions with other asteroids since then (H-chondrites also show more recent Ar-Ar ages, while HEDs do not), it is likely that the shock ages of HED meteorites and H-chondrites do have a connection to Mars-crossers. Even more than meteorite shock ages, impact events recorded in lunar meteorites \\citep{coh00, coh05} seem at odds the lunar impact history inferred from Apollo samples \\citep{har03, har07}. Events postdating 3 Gya are recorded in lunar meteorites, long after the heavy bombardment ended. Our hypothesis does not offer an explanation for this discrepancy, and we suspect that the solution lies in the physics of lunar regolith rather than the dynamics of impactors. How can our hypotheses be tested? First, new crater counts (down to about $D=8$~km) on all non-magnetized basins (Imbrium, Orientale, Schr\\\" odinger and Hertzsprung) would confirm or falsify the existence of a non-asteroidal Imbrian size distribution, without having to resort to stratigraphic or morphological identification of craters. Otherwise, the study of mesosiderites and Hungarias could probably test their relationship with the lunar cataclysm faster than new lunar samples can be collected. Our hypothesis predicts that the mesosiderite parent body (or bodies) are likely located among Hungarias (although innermost main belt and Mars Trojans are also plausible refuges). Remote sensing of asteroid composition has advanced considerably in the last few decades, and it is possible that we may be able to identify the intermediate mesosiderite parent body, if it is big and bright enough (small Hungarias are brighter than small MBAs). Radar observations may also in principle be able to break the degeneracy among X-type asteroids among Hungarias \\citep{ock10}. Additionally, if the Dawn spacecraft identifies a mesosiderite unit exposed on Vesta (or if mesosiderites are traced to any other large asteroid deep in the main belt), our hypothesis about the connection between the mesosiderites and the cataclysm would be falsified. While the collisional disruption hypothesis can be formulated independently from mesosiderites \\citep{wet75}, the case becomes weaker once mesosiderites are excluded from consideration." }, "1112/1112.4386_arXiv.txt": { "abstract": "We comment on arXiv:1112.1320 and point out that baryonic oscillations of the matter power spectrum, while predicted by theories that do not incorporate collisionless cold dark matter, are strongly suppressed by the statistical window function that is used to process finite-sized galaxy samples. We assert that with present-day data sets, the slope of the matter power spectrum is a much stronger indicator of a theory's validity. We also argue that MOND should not be used as a strawman theory as it is not in general representative of modified gravity theories; some theories, notably our scalar-vector-tensor MOdified Gravity (MOG), offer much more successful predictions of cosmological observations. ", "introduction": " ", "conclusions": "" }, "1112/1112.1106_arXiv.txt": { "abstract": "Despite the success of Maxwell's electromagnetism in the description of the electromagnetic interactions on small scales, we know very little about the behaviour of electromagnetic fields on cosmological distances. Thus, it has been suggested recently that the problems of dark energy and the origin of cosmic magnetic fields could be pointing to a modification of Maxwell's theory on large scales. Here, we review such a proposal in which the scalar state which is usually eliminated be means of the Lorenz condition is allowed to propagate. On super-Hubble scales, the new mode is essentially given by the temporal component of the electromagnetic potential and contributes as an effective cosmological constant to the energy-momentum tensor. The new state can be generated from quantum fluctuations during inflation and it is shown that the predicted value for the cosmological constant agrees with observations provided inflation took place at the electroweak scale. We also consider more general theories including non-minimal couplings to the space-time curvature in the presence of the temporal electromagnetic background. We show that both in the minimal and non-minimal cases, the modified Maxwell's equations include new effective current terms which can generate magnetic fields from sub-galactic scales up to the present Hubble horizon. The corresponding amplitudes could be enough to seed a galactic dynamo or even to account for observations just by collapse and differential rotation in the protogalactic cloud. ", "introduction": "Out of the four fundamental interactions, gravity and electromagnetism are the only long-range forces in nature. The standard theories for these two interactions, namely General Relativity (GR) and Maxwell's electromagnetism have received a vast experimental support in a huge range of scales. GR is able to explain the gravitational phenomena from sub-millimeters distances up to Solar System scales and standard electromagnetism has been probed from the tiny scales involved in high-energy colliders up to distances of order 1.3 A.U. corresponding to the coherence lengths of the magnetic fields dragged by the solar wind{\\cite{limit}. However, despite the enormous success of these two theories, both of them present some unresolved problems when large scales are involved. For GR, cosmological observations require the presence of exotic components in the universe, i.e. dark matter and dark energy. Indeed, some attempts to account for such observations without invoking dark components include infrared modifications of GR. On the other hand, the presence of cosmic magnetic fields found in galaxies, clusters\\cite{galactic1,galactic2,galactic3} and, very recently\\cite{extragalactic1,extragalactic2,extragalactic3,extragalactic4}, also in the voids cannot be accommodated within Maxwell's theory. This could also be signaling that a more careful analysis of the behavior of electromagnetic fields in cosmological contexts is needed. A particularly interesting aspect is the quantization of gauge theories in non-trivial spacetimes. The usual quantization procedures of the electromagnetic field in flat spacetime rely on some type of subsidiary condition on the physical states to get rid of the unphysical gauge modes. Although all the approaches are equivalent in flat spacetime, such an equivalence has not been proven in curved background. In fact, the Gupta-Bleuler formalism for the covariant quantization has been shown to present difficulties in a time-dependent spacetime \\cite{Parker,EM2}. The BRST method also exhibits similar pathologies in certain spacetimes\\cite{ghosts1,ghosts2}. Here, we shall review a recent proposal of an extended theory of electromagnetism in which we can avoid the aforementioned difficulties. This extension is based on allowing the propagation of the state that is usually eliminated from the physical Hilbert space by means of the subsidiary condition. We shall show how this theory can be consistently quantized in an expanding universe with {\\it three} physical states comprising the two polarizations of the usual photons plus one additional scalar state. The remarkable feature of the resulting theory is that the additional state can be generated from quantum fluctuations during inflation and give rise to and effective cosmological constant on large scales, whereas on sub-Hubble scales it leads to the generation of cosmic magnetic fields. ", "conclusions": "We have reviewed the extended electromagnetic theory in which a gauge-fixing term is promoted into a physical contribution in the fundamental action. The quantization of the free theory can be performed without having to impose any subsidiary condition at the price of introducing an additional degree of freedom. This new mode can be gravitationally produced from quantum fluctuations during inflation and its amplitude will be determined precisely by the scale at which inflation takes place. In the subsequent cosmological evolution, this new mode has two effects. On super-Hubble scales, it behaves as an effective cosmological constant whose observed value can be explained if inflation occurred at the electroweak scale. On sub-Hubble scales, the additional mode gives rise to a stochastic background of longitudinal electric waves that generate magnetic fields from sub-galactic scales up to the present Hubble radius. On the other hand, we have also considered the effects of non-minimal couplings in the presence of the temporal electromagnetic mode and shown that in such a case, neutral massive objects can act as sources of electromagnetic fields thus implementing the old conjecture by Schuster, Einstein and Blackett of gravitational magnetism. The theory studied in this work shows how a modification of electromagnetism which does not require the introduction of new fields, dimensional parameters or potential terms could provide a simple explanation for the tiny value of the cosmological constant and, at the same time, a mechanism for the generation of magnetic fields on cosmological scales. Some open questions still remain to be studied such as the inclusion of electromagnetic interactions in the theory, since all the analysis performed so far are limited to the free theory. Also, it would be interesting to know the behaviour of the new mode in more general background space-times in order to determine the viability of the model from the theoretical and phenomenological point of views. Finally, the possibility of detecting the longitudinal electric wave background generated during inflation could provide a clear signal of the modification of electromagnetism on cosmological scales. \\vspace{0.2cm} {\\em Acknowledgments:} This work has been supported by MICINN (Spain) project numbers FIS 2008-01323 and FPA 2008-00592, CAM/UCM 910309 and MICINN Consolider-Ingenio MULTIDARK CSD2009-00064. JBJ is also supported by the Ministerio de Educaci\\'on under the postdoctoral contract EX2009-0305. \\vspace{0.2cm}" }, "1112/1112.2513_arXiv.txt": { "abstract": "{The quiet Sun magnetic fields produce ubiquitous bright points (BPs) that cover a significant fraction of the solar surface. Their contribution to the total solar irradiance (TSI) is so-far unknown.} {To measure the center-to-limb variation (CLV) of the fraction of solar surface covered by quiet Sun magnetic bright points. The fraction is referred to as {\\em fraction of covered surface}, or FCS. } {Counting of the area covered by BPs in $G$-band images obtained at various heliocentric angles with the 1-m Swedish Solar Telescope on La Palma. Through restoration, the images are close to the diffraction limit of the instrument ($\\sim 0\\farcs 1$). } { The FCS is largest at disk center ($\\simeq$ 1\\,\\%), and then drops down to become $\\simeq$ 0.2\\,\\% at $\\mu\\simeq 0.3$ (with $\\mu$ the cosine of the heliocentric angle). The relationship has large scatter, which we evaluate comparing different subfields within our FOVs. We work out a toy-model to describe the observed CLV, which considers the BPs to be depressions in the mean solar photosphere characterized by a depth, a width, and a spread of inclinations. Although the model is poorly constrained by observations, it shows the BPs to be shallow structures (depth$\\,<\\,$width) with a large range of inclinations. We also estimate how different parts of the solar disk may contribute to TSI variations, finding that 90\\,\\% is contributed by BPs having $\\mu > 0.5$, and half of it is due to BPs with $\\mu > 0.8$. } {} ", "introduction": "Our understanding of the quiet Sun magnetic fields has drastically improved during the last decade \\citep[for recent reviews, see, e.g.,][]{2009SSRv..144..275D,2011ASPC..437..451S}. We have gone from magnetic signals present only at the network boundaries \\citep[e.g.,][]{bec77b,sol93}, to ubiquitous polarization signals created through Hanle effect \\citep[e.g.,][]{fau93,tru04} and Zeeman effect \\citep[e.g.,][]{lin99,2000ApJ...532.1215S,2003ApJ...582L..55D,har07,lit08}. The wealth of quiet Sun magnetic structures makes them potentially important to understand the global magnetic properties of the Sun \\citep[][]{2003ApJ...585..536S,tru04}, and also makes it unlikely that the quiet Sun magnetism results from the decay of active regions \\citep[e.g.,][]{san03b}. Theoretical arguments, corroborated by numerical experiments, favor a different production mechanism \\citep{pet93,cat99b,vog07,pie10,2011ApJ...736...36M}. An efficient turbulent dynamo transforms into magnetic fields part of the kinetic energy of the granular convection. It generates a complex magnetic field which evolves in short time scales (a few min) and has small characteristic length-scales ($<\\,1\\,$Mm). In agreement with the turbulent dynamo scenario, quiet Sun magnetic fields come with strengths in the full range covering from almost zero to 2\\,kG \\citep[][]{2000ApJ...532.1215S,dom06,2008A&A...477..953M, 2009A&A...506.1415B,2011A&A...526A..60V}. Even if they only fill a small fraction of the quiet photosphere, the part having strong kG fields may be particularly important for a number of reasons. Firstly, the magnetic flux and energy increase with field strength, therefore, the energy and flux provided by kGs may surpass the contribution of the more common but weaker fields \\citep[][]{san04}. The need to consider kGs is illustrated by the numerical experiments set up by \\citet{2011arXiv1108.1155C}, where realistic granular convection redistribute initial hG fields so that daG, hG and kG field strengths have the same energy despite their very different area covering. Magnetic concentrations with kG fields may also be important because buoyancy makes them vertical \\citep[e.g.,][]{sch86} and so, they naturally provide a mechanical connection between the photosphere and the upper atmosphere \\citep[e.g.,][]{bal98,sch03b,goo04,jen06,san08}. They can function as guides that sustain magneto-acustic wave propagation, or be physical channels connecting plasmas of different atmospheric layers. Finally, kG magnetic concentrations are expected to be particularly bright due to the so-called hot-wall effect\\footnote{The magnetic pressure suffices to maintain kG structures in mechanical balance within the photosphere, therefore, they are evacuated and transparent, allowing us to look through into the sub-photosphere, which is generally hotter and so brighter.}\\citep[]{spr77,car04,kel04}. They produce bright points (BPs) which, depending on the variation across the solar disk, and during the solar cycle, may even contribute to the Total Solar Irradiance (TSI) variations as network and plage magnetic fields do \\citep[e.g.,][]{1997ARA&A..35...33L, 2004A&ARv..12..273F, 2011SSRv..tmp..133F}. The finding of BPs in the quiet Sun was immediately identified as the kG magnetic concentrations inferred from polarization measurements \\citep{2004ApJ...609L..91S}. Their basic properties and their ubiquitous presence have been confirmed by a number of researchers \\citep[][]{dew05,dew08,bov08,san08, 2009ApJ...700L.145V,2010ApJ...723..787V, 2010ApJ...714L..31G}. They vary on time scales of minutes similar to that of granulation, and many BPs are at the resolution limit of the current instrumentation (some 0.1~Mm). They are extremely common, filling at least 1\\% of the solar surface and outnumbering granules \\citep{2010ApJ...715L..26S}. \\citet{2011A&A...532A.136S} have recently estimated the excess of brightness produced by the quiet magnetic fields at the disk center, turning out to be of the order of 0.15\\,\\%. Their contribution is larger than the 0.08\\,\\%~TSI variations associated with cycle, however, determining the impact of these kG fields on TSI demands knowing their center-to-limb variation (CLV), as well as the variation of the quiet Sun magnetic fields with the solar cycle. These two properties, central to assess the role of quiet Sun fields on TSI, are poorly known. As far as the variation along the cycle is concerned, the claims in the literature are for little variation if any \\citep{san03d,shc03,har10}. Unfortunately, the uncertainty of such claims can be as large as a factor two \\citep{fau01}. There are several works on the CLV % signals associated with the quiet Sun magnetic fields \\citep[e.g.,][]{2008A&A...477..953M,lit08,% 2011ApJ...737...52L} however, to the best of our knowledge, nothing is known on the CLV % of the quiet Sun magnetic BPs. This is precisely the subject of our work, i.e., providing a first observational description of the CLV % of the quiet Sun BPs. To be more exact, we evaluate how the area covered by quiet Sun BPs varies with the position on the disk. Two main difficulties hinder the analysis. First, one has to use images with enough spatial resolution and of uniform quality, since the number of BPs depends critically on the spatial resolution of the observations \\citep{1996ApJ...463..797T,2010ApJ...715L..26S}. This issue is sorted out using data from the 1-m Swedish Solar Telescope \\citep[SST,][]{2003SPIE.4853..341S} obtained in a single day during moments of excellent seeing, and then restored to get images with uniform resolution close to the diffraction limit of the instrument \\citep{2005SoPh..228..191V}. Second, the BPs can be misidentified with granule borders and other structures \\citep[e.g.,][]{2007SoPh..243..121B}, a problem particularly severe in near-limb images. This second problem is addressed resorting to the cumbersome method of eye-ball identification which, however, has been proven to be reliable for our purpose \\cite[][and references therein]{2010ApJ...715L..26S}. The work is organized as follows; the observations are presented in \\S~\\ref{observation}. The actual CLV measurements are described in \\S~\\ref{sec_ctl}. In principle, such measurements has to be interpreted using realistic models of magneto-convection (such as those used for plages by \\citeauthor{car04}~\\citeyear{car04} or \\citeauthor{kel04}~\\citeyear{kel04}), where the 3-dimensional geometry of the photosphere is considered consistently. It would require repeating the analyses carried out on the observed images using synthetic images from the simulations. Then the CLV coming from a comprehensive battery of numerical simulations should be compared with the observations. This detailed realistic approach clearly exceeds the scope of the work. However, we attempt a toy-modeling which, despite its simplicity, considers the key ingredients that within the hot-wall paradigm determine the CLV (\\S~\\ref{interpretation}). The model is compared with the observed CLV in \\S~\\ref{results}, which constraints some of the properties of the magnetic structures. The implications for TSI variations are analyzed in \\S~\\ref{tsi_var}. These results are discussed and put into context in \\S~\\ref{conclusions}. ", "conclusions": "We have measured the center-to-limb variation (CLV) of the area covered by $G$-band bright points (BPs) in the quiet Sun (fraction of covered surface or FCS). It is a parameter difficult to determine since the detection of BPs critically depends on the angular resolution of the observation (\\S~\\ref{intro}). We employ several time series taken in two hours during moments of excellent seeing with the SST. The images were post processed using MOMFBD (see \\S~\\ref{observation}) which provides a homogeneous set of images adequate for these subtle measurements. They were restored to provide an angular resolution close to the diffraction limit of the instrument at the working wavelength (some 0\\farcs 1 at the $G$-band). We find the FCS to be largest at disk center ($\\simeq$ 1\\,\\%), and then it drops down to become $\\simeq$ 0.2\\,\\% at $\\mu\\simeq 0.3$. The relationship has large scatter, which we managed to estimate comparing different subfields within our FOVs (see the error bars in Fig.~\\ref{curvas8ptos} and Table~\\ref{ff_results}). The value obtained at the disk center agrees with previous estimates based on data of similar quality. We work out a toy model to describe the observed CLV. It assumes the magnetic bright points to be depressions in the mean solar photosphere, characterized by a depth, a width, and with a spread of inclinations. It is only an exploratory modeling which, however, includes the physical ingredients that seems to be responsible for a kG magnetic concentration to show up as a bright feature on the solar disk. The solutions offered by our toy-model are poorly constrained, but they seem to show the BPs to be shallow structures\\footnote{This result is not inconsistent with the magnetic concentrations producing BPs being modeled as a compactly-packed ensemble of narrow magnetic concentrations arranged in a micro-structured magnetic atmosphere MISMA fashion \\citep[][]{1996ApJ...466..537S, 2000ApJ...544.1135S,2001ApJ...555..978S}. The CLV of the FCS is sensitive to the global scale of the ensemble, whereas the MISMA accounts for the smallest scales responsible, among others, for the asymmetries of the spectral lines formed in the atmosphere. } (ratio depth to width $\\simeq 0.7\\pm 0.2$) with a large spread of inclinations ($\\simeq \\pm 70^\\circ$). Among others, the FCS is of interest because it determines the impact that quiet Sun magnetic fields may have on TSI variations, an influence so far unknown. Since the measured FCS is so peaked towards disk center, any role that quiet Sun magnetic fields may have to play on TSI will be due to the center of the disk. According to our estimate, 90\\,\\% of the TSI is contributed by BPs having $\\mu > 0.5$, and half of it is due to BPs with $\\mu > 0.8$ (\\S~\\ref{tsi_var}). This estimate is based on assuming the CLV of the BP (wavelength integrated) flux to scale as the observed $G$-band intensity. It is an ad-hoc assumption adopted for lacking of a better one which, however, has allowed us to provide a first constrain on the effect of quiet Sun BPs on TSI. In this sense, we have to stress that the FCS has been measured in the $G$-band, because the magnetic concentrations are particularly conspicuous at this wavelengths.Then we implicitly assume throughout the work that the FCS is the same at all wavelengths, i.e., that the area covered by magnetic concentration does not change with wavelength. (The $G$-band BPs are brighter but not bigger.) This assumption remains to be proven however, for the time being, it seems to be a reasonable working hypothesis. Our toy-model suggests the BPs to be shallow structures with varied inclinations. One could test these predictions observing polarization signals of quiet Sun BPs at disk center. On the one hand, the polarization signals provide magnetic field inclinations through Stokes inversion \\citep[e.g.,][]{2008ApJ...674..596S}. On the other hand, assuming the magnetic concentrations to be in mechanical balance, one can infer their depths \\citep[e.g.,][]{2000ApJ...532.1215S}. These inferences require low-noise spectro-polarimetry with an angular resolution similar to that of our $G-$band images. They represent a technical challenge, but such observations seem to be doable in the near future \\citep[see, e.g.,][]{2010ApJ...723L.164L}. As we mentioned in the introduction, the existing numerical simulations of magneto-convection predict the BPs to be depressed with respect to the mean photosphere \\citep[e.g.,][]{vog05,vog07}. Whether the faint BPs observed in quiet Sun are predicted to be superficial or deep remains to be worked out. However, the existing simulations of plage magnetic concentrations suggest the continuum intensity to be formed in a rather shallow region \\citep{kel04,car04}. -------------------------------" }, "1112/1112.5105_arXiv.txt": { "abstract": "We discuss a fast cross-Wigner transform based technique for detecting gravitational wave bursts, and estimating the direction of arrival, using a network of (three) non co-located interferometric detectors. The performances of the detector as a function of signal strength and source location, and the accuracy of the direction of arrival estimation are investigated by numerical simulations. The robustness of the method against instrumental glitches is illustrated. ", "introduction": "\\label{sect:intro} The next generation of interferometric detectors, of gravitational waves (henceforth GW) including AdLIGO \\cite{AdLIGO}, AdVirgo \\cite{AdVirgo} and GEO-HF \\cite{superGEO}, hopefully to be followed soon by LCGT \\cite{LCGT} and ACIGA \\cite{ACIGA}, and eventually by ET \\cite{ET}, is expected to observe tens of events per year, opening the way to gravitational wave astronomy \\cite{GWBs}. Identifying the direction of arrival (henceforth DOA) of the signals, and retrieving their shapes, will be a primary task in reconstructing the physics of the sources and their environments.\\\\ The possibility of retrieving the DOA from {\\it independent} estimates of the signal arrival time at {\\it each} detector was first suggested in \\cite{Boulanger}, and further discussed in Saulson seminal book \\cite{Saulson_book}. It was shown that three-interferometers are sufficient to retrieve the DOA up to a mirror-image ambiguity which can be solved in principle from knowledge of the detectors' directional responses. This method, often referred to as {\\it triangulation} was further elaborated by Sylvestre \\cite{Sylvestre_03}, Cavalier et al. \\cite{Cavalier_06}, and Merkovitz et al. \\cite{Zanolin_08}. In \\cite{Cavalier_06} a Gaussian distribution was assumed for the (independent) arrival time estimation errors, and a $\\chi^2$ minimization algorithm was accordingly proposed for retrieving the DOA, in the maximum likelihood spirit. In \\cite{Zanolin_08} it was shown that this method is affected by a systematic bias in the estimated DOA, a possible technique for removing the bias was discussed, and amplitude consistency tests for removing the mirror-image ambiguity were suggested. Fairhurst developed a similar analysis of the effect of arrival time estimation errors on the DOA estimation accuracy, for the special case of chirping signals, including waveform and calibration errors \\cite{Fairh_09}, \\cite{Fairh_11}.\\\\ DOA estimation algorithms are already implemented in the {\\it coherent} LIGO-Virgo pipelines for GW burst (henceforth GWB) detection \\cite{Xpipe}, \\cite{Klimenko_11}. DOA estimation in {\\it coherent} network data analysis, was studied first by Krolak and Jaranowski \\cite{Kro_Jar_94}, and then by Pai et al. \\cite{Pai_Dhur}, as part of the waveform parameter estimation problem, with specific reference to chirping waveforms from coalescing binaries, in a Gaussian noise background. The conceptual foundations of coherent data analysis for unmodeled waveforms were laid out by Flanagan and Hughes \\cite{Flan_Hug}, and further developed by Klimenko et al. \\cite{Klim1}-\\cite{Rakh}. G\\\"{u}rsel and Tinto \\cite{GurTin} first suggested the possibility of retrieving the DOA for un modeled signals using null-streams. This concept was analyzed in depth by Schutz and Wen \\cite{Wen_Schutz_05}, and further exploited by Chatterji et al. \\cite{Shurov}. A Fisher-matrix based analysis of arrival time estimation error in coherent network detection of modeled as well as {\\it unmodeled} signals was made in Wen et al. \\cite{Wen_Fan_Chen}.\\\\ In this paper we capitalize on the time-shift and localization properties of the cross-Wigner-Ville (henceforth XWV) transform to introduce a new and conceptually simple GWB detection and DOA reconstruction algorithm, using a network of non co-located interferometric detectors.\\\\ The Wigner-Ville transform is a well known powerful tool for the analysis of non-stationary signals \\cite{XWV1}, whose potential in GW data analysis, has been highlighted by several Authors, under different perspectives \\cite{Feo_etal}-\\cite{Chass}. Here we suggest its possible use as an effective tool for detecting GWBs, and estimating their DOA, which offers nice features in terms of performance, robustness against spurious instrumental/environmental transients (glitches).\\\\ Instead of using {\\it independent} estimates of the arrival times at {\\it each} detector, our DOA estimator uses data from (all) detector {\\it pairs} to estimate the needed propagation delays. In addition, it also provides an effective detection statistic, combining the data from {\\it all} detectors in the network, at a remarkably light computational cost.\\\\ DOA reconstruction from arrival-time delay estimation in a network of sensors is a well known problem in the technical Literature on Acoustics and Radar (see, e.g., \\cite{Berdugo} for a broad review). The standard method for time-delay estimation in Gaussian noise is (generalized) cross-correlation \\cite{Knapp}, which is known to perform reasonably well for relatively large signal to noise ratios \\cite{Fert_Sjo}. Remarkably, the correlation-based estimator offers worse performances compared to the XWV in the present context, as shown in Sect. 5.2\\\\ This paper is organized as follows. In Sect. 2 we introduce the XWV transform, and recall its time-shift properties, which are illustrated for the simplest case of sine-Gaussian (henceforth SG) GWBs. In the same section we recall the relationship between arrival time delays and DOA. In Sect. 3, we illustrate the proposed XWV transform based DOA reconstruction algorithm. In Sect. 4 we discuss the effect of noise in the data, and the related DOA reconstruction uncertainties. In Sect. 5 we present the results of extensive numerical simulations, aimed at characterizing the performance of our XWV based algorithm both as a detector and as a DOA estimator. The simulations are based on SG-GWBs, but the case of more realistic waveforms (including Dimmelmeier and binary merger waveforms) is also discussed. In Sect. 6 we include a short discussion of the robustness of the proposed algorithm against instrumental/environmental transients (glitches). Conclusions follow under Sect. 7. ", "conclusions": "We presented a simple, computationally light and fast algorithm for detecting short unmodeled GWB in a network of three interferometric GW detectors, and estimating the related DOA, based on XWV spectra. The algorithm is reasonably performant, and nicely robust against spurious transients (glitches) of instrumental origin corrupting the (otherwise Gaussian) detectors noise floor.\\\\ It does not provide waveform reconstruction; this latter, however can be accomplished in principle off-line, once the DOA has been estimated.\\\\ Generalization to larger networks, and other potentially interesting waveforms (e.g., chirps) is relatively straightforward. Such extensions will be explored in a forthcoming paper.\\\\ Based on the above preliminary results, we suggest that the proposed algorithm may be used as a quick-and-(not-so)-dirty on-line data sieving tool.\\\\ A quantitative comparison with existing GWB detection/DOA estimation algorithms in terms of efficiency and computational burden will be the subject of future investigation." }, "1112/1112.0516_arXiv.txt": { "abstract": "{Massive stars have a profound effect on the surrounding interstellar medium. They ionize and heat the neutral gas, and due to their strong winds, they swept the gas up forming large \\hi\\ shells. In this way, they generate a dense shell where the physical conditions for the formation of new stars are given.} {The aim of this study is to analyze the origin and evolution of the large \\hi\\ shell \\gs\\, and its role in triggering star forming processes.}{To characterize the shell and its environs, we carry out a multi-wavelength study. We analyze the \\hi\\ 21 cm line, the radio continuum, and infrared emission distributions. }{The analysis of the \\hi\\ data shows an expanding shell structure centred at ($l, b$) = (100\\fdg6, --2\\fdg04) in the velocity range from --29 to --51.7 \\kms. Taking into account non circular motions, we infer for \\gs\\, a kinematical distance of 2.8 $\\pm$ 0.6 kpc. Several massive stars belonging to Cep\\,OB1 are located in projection within the large \\hi\\, shell boundaries. The analysis of the radio continuum and infrared data reveal that there is no continuum counterpart of the \\hi\\ shell. On the other hand, three slightly extended radio continuum sources are observed in projection onto the dense \\hi\\ shell. From their flux density determinations we infer that they are thermal in nature. An analysis of the \\hi\\ emission distribution in the environs of these sources shows, for each of them, a region of low emissivity having a good morphological correlation with the ionized gas in a velocity range similar to the one where \\gs\\, is detected.} {Based on an energetic analysis, we conclude that the origin of \\gs\\, could have been mainly due to the action of the Cep\\,OB1 massive stars located inside the \\hi\\ shell. The obtained age difference between the \\hi\\ shell and the \\hii\\ regions, together with their relative location, led us to conclude that the ionizing stars could have been created as a consequence of the shell evolution. } ", "introduction": "The presence in the interstellar medium (ISM) of the Milky Way, when viewed in the $\\lambda\\sim$ 21 cm line emission of the neutral hydrogen (\\hi), of giant structures having linear dimensions of a few hundred parsecs in diameter is a widely known and well observed phenomena first noticed by \\citet{hei79}. These structures are usually detected as huge shells or arc-like features of enhanced \\hi\\, emission surrounding regions of low \\hi\\, emissivity, receiving the generic name of \\hi\\, supershells. These features may even be the dominant structure in the interstellar medium, taking up a large fraction of the volume of the galactic disk. The \\hi\\, structures could be also observed at infrared wavelengths. Based on the 60 and 100 $\\mu$m IRAS databases, \\citet{kon07} have performed an all-sky survey of loop- and arc-like structures. Similar structures have also been observed in nearby spiral galaxies \\citep{sta07,cha11}. In the Milky Way, these structures were initially catalogued by \\citet{hei79,hei84}. Though a large number of \\hi\\, features likely to be classified as either large \\hi\\, shells or \\hi\\, supershells have been catalogued in the outer (90$^\\circ \\leq l \\leq$ 270$^\\circ$) part of the Galaxy \\citep{ehl05}, only a small number of them have been studied in some detail. The later \\citep{jun96, sti01, uya02, mcc02, caz03, arn07, cic11} have galactocentric distances ranging from 9.7 to 16.6 kpc, diameters from 120 to $\\sim$ 840 pc, expansion velocities between $\\sim$10 and $\\sim$ 20 \\kms, and kinetic energies from $\\sim$ 1 $\\times$ 10$^{50}$ up to $\\sim$ 6 $\\times$ 10$^{51}$ erg. Among them only two (GS\\,305+01-24 and the feature studied by \\citet{caz03}) have an OB-association as their likely powering source, and other three (GSH\\,91.5+2--114, GS\\,234--02 and GS\\,263--02+45) show evidence of having induced the formation of new generation of stars. The general consensus is that those structures whose kinetic energy is of the order of, or less than, a few times 10$^{51}$ erg, very likely may have been created by the joint action of stellar winds and supernova explosions. A large number of examples are reported in the literature \\citep[e.g.][]{uya02,arn07,cic11}. On the other hand, for expanding \\hi\\, structures having kinetic energies in excess $\\sim 10^{52}$ erg, termed {\\it supershells}, the above mechanism may not be adequate to create them because one would need a stellar grouping (either an open cluster or an OB-association) with many more stars than the average found in the Milky Way. In these cases alternative mechanisms like the infalling of high velocity clouds \\citep{ten81} or gamma-ray bursts \\citep{per00} may be at work. Along its expansion, these structures (either a shell or a supershell) may became gravitationally unstable, forming clouds that later on may lead to the formation of stars along the periphery of these \\hi\\, structures, or else the expanding structure may hit and compress pre-existing {\\rm ISM} molecular clouds from one side. During this process a high density perturbation may move into the molecular clouds, which may eventually collapse into denser cores in which star formation may occur. A thorough review of observations and theory related to trigger star formation is given by \\citet{elm98}. A new large scale study aim at detecting in the outer part of the galaxy structures likely to be either large \\hi\\, shells or supershells is being carried out by one of the authors (L. A. Suad) as part of her PhD Thesis. In this paper we analyze a new large \\hi\\, shell observed at (\\textit{l, b}) $\\sim$ ($100^\\circ$, $-2^\\circ$), with the purpose of elucidating both its origin and its interaction with the surrounding ISM. We also look for signs of recent star formation activity likely to be related to this shell. ", "conclusions": "The large \\hi\\, shell \\gs\\, has been analyzed to study the interaction of massive stars with the interstellar medium and, in particular, the process of triggered star formation. From this analysis we conclude the following: \\begin{enumerate} \\item \\gs\\, is a large shell of a radius of about 102 pc located at a distance of 2.8 $\\pm\\,$ 0.6 kpc. The swept up mass in the shell is ($1.5 \\pm\\, 0.7$) $ \\times\\, 10^5$ M$_\\odot$ and the shell density $n_{sh} = 2.5 \\pm\\, 0.4$ cm$^{-3}$. The shell is expanding at a velocity of 11 $\\pm\\, 2$ \\kms\\, and its kinetic energy is ($1.8 \\pm\\, 0.8$) $\\times\\, 10^{50}$ erg. \\item Several evolved massive stars members of Cep\\,OB1 are projected inside the large shell. The distance to the OB association is compatible with the kinematical distance of \\gs\\, when non-circular motions are considered. An energetic analysis suggests that the wind energy provided during the main sequence phase of the stars could explain the origin of the shell. However, taking into account the SN rate in OB associations, the energy contribution of a SN explosion as well as of its massive progenitor can not be discarded. \\item From the 2695 MHz radio continuum and 60 $\\mu$m infrared images, we found three slightly extended sources, labelled G98, G100, and G103 projected onto the borders of \\gs. From the radio flux densities estimated at different wavelengths, the thermal nature of the sources was confirmed by the estimation of their spectral indexes. In addition, dust temperatures were estimated and found to be typical of \\hii\\, regions. \\item An inspection of the 1\\am\\, CGPS \\hi\\, data reveals \\hi\\, minima having a good morphological correlation with the \\hii\\, regions at velocity ranges compatible with the velocity spanning by \\gs. This leads to the conclusion that G98, G100, and G103 are located at the same distance than \\gs. \\item From O and OB star catalogues, the massive star candidates to be responsible for the ionized gas were identified. \\item The obtained age difference among the \\hii\\, regions and the shell, together with their relative location leads us to the conclusion that G98, G100, and G103 may have been created as a consequence of the action of a strong shock produced by the expansion of \\gs\\, into the surrounding gas. \\end{enumerate}" }, "1112/1112.0666_arXiv.txt": { "abstract": "The measurement of the anisotropies of cosmic ray arrival direction provides important informations on the propagation mechanisms and on the identification of their sources. In this paper we report the observation of anisotropy regions at different angular scales. In particular, the observation of a possible anisotropy on scales between $\\sim$10$^{\\circ}$ and $\\sim$30$^{\\circ}$ suggests the presence of unknown features of the magnetic fields the charged cosmic rays propagate through, as well as potential contributions of nearby sources to the total flux of cosmic rays. Evidence of new weaker few-degree excesses throughout the sky region $195^{\\circ}\\leq$ R.A. $\\leq 315^{\\circ}$ is reported for the first time. ", "introduction": "As cosmic rays (CRs) are mostly charged nuclei, their arrival direction is deflected and highly isotropized by the action of galactic magnetic field (GMF) they propagate through before reaching the Earth atmosphere. The GMF is the superposition of regular field lines and chaotic contributions. Altough the strength of the non-regular component is still under debate, the local total intensity is supposed to be $B=2\\div 4\\textrm{ $\\mu$G}$. In such a field, the gyroradius of CRs is given by $r_{a.u.}=100\\,R_{\\textrm{\\scriptsize{TV}}}$, where $r_{a.u.}$ is in astronomic units and $R_{\\textrm{\\scriptsize{TV}}}$ is in TeraVolt. However, different experiments \\cite{nagashima,kam07,tibet06,milagro09,eastop09,icecube11} observed an energy-dependent \\emph{\"large scale\"} anisotropy in the sidereal time frame with an amplitude of about 10$^{-4}$ - 10$^{-3}$, suggesting the existence of two distint broad regions, one showing an excess of CRs (called \"tail-in\"), distributed around 40$^{\\circ}$ to 90$^{\\circ}$ in Right Ascension (R.A.). The other a deficit (the \"loss cone\"), distributed around 150$^{\\circ}$ to 240$^{\\circ}$ in R.A.. The origin of these anisotropies is still unknown. Some authors claim that it can be explained within the diffusion approximation taking into account the role of the few most nearby and recent sources \\cite{blasi,erlykin06}. Other studies suggest that the observations may be due to a combined effect of the regular and turbolent GMF \\cite{battaner09}, or to local uni- and bi-dimensional inflows \\cite{amenomori10}. In the last years the Tibet AS$\\gamma$ \\cite{amenomori07} and Milagro \\cite{milagro08} Collaborations reported evidence of the existence of a medium angular scale anisotropy contained in the tail-in region. The observation of similar small scale anisotropies has been recently claimed by the Icecube experiment in the Southern hemisphere \\cite{icecube11}. So far, no theory of CRs in the Galaxy exists which is able to explain few degrees anisotropies in the rigidity region 1-10 TV leaving the standard model of CRs and that of the local GMF unchanged at the same time. More beamed the anisotropies and lower their energy, more difficult to fit the standard model of CRs and GMF to experimental results. The observation of anisotropy effects at a level of 10$^{-4}$ with an air shower array is a tricky job. A wrong estimation of the exposure may affect the significance and the relative intensity sky maps, even create artifacts (i.e. fake excesses or deficit regions). In fact, drifts in detector response and atmospheric effects on air shower development are quite hard to be modeled to sufficient accuracy. The envisaged solution is to use the data to estimate the detector exposure. But data contain either signal and background events, so that some distortions could be present in the results. The shape and the size of possible artifacts depend on the characteristic angle and time scale over which all the aspects of the data acquisition vary less than the effect to be observed. Therefore, if anisotropies of the order 10$^{-4}$ are looked for, operating conditions must be kept (or made up) stable down to this level, all across the field of view and during all the acquisition time. In addition, as it is well known, the partial sky coverage carried out by a given experiment will bias the observations of structure on largest scales. Finally, the observation of a possible small angular scale anisotropy region contained inside a larger one rely on the capability for suppressing the anisotropic structures at larger scales without, simultaneuosly, introducing effects of the analysis on smaller scales. The ARGO-YBJ experiment, located at the YangBaJing Cosmic Ray Laboratory (Tibet, P.R. China, 4300 m a.s.l., 606 g/cm$^2$), is an air shower array able to detect the cosmic radiation at an energy threshold of a few hundred GeV. The full detector is in stable data taking since November 2007 with a duty cycle greater than 85\\%. The trigger rate at the threshold is 3.6 kHz. The detector characteristics and performance are described in \\cite{moon11}, the main results are reported in \\cite{demitri11}. In this paper the observation of CR anisotropy at different angular scales with ARGO-YBJ is reported as a function of the primary energy. ", "conclusions": "In this paper the observation of CR anisotropy at different angular scales with ARGO-YBJ is reported as a function of the primary energy. The large scale CR anisotropy has been clearly observed up to about 25 TeV. Evidence of existence of different few-degree excesses in the Northern sky (the strongest ones positionally coincident with the regions detected by Milagro in 2008) is reported. The study of the CR anisotropy with air shower arrays is challenging, therefore deeper analysis is under way to investigate possible artifacts produced in the background calculations. However, a joint analysis of concurrent data recorded by different experiments in both hemispheres, as well as a correlation with other observables like the interstellar energetic neutral atoms distribution \\cite{ibex09,ibex11}, should be a high priority to clarify the observations." }, "1112/1112.2349_arXiv.txt": { "abstract": "{Dust particles are observed at mm sizes in outer regions of the disk, although theoretically, radial drift does not allow dust particles to form pebbles.} {In order to explain grain growth to mm sized particles and their retention in outer regions of protoplanetary disks, as it is observed at sub-mm and mm wavelengths, we investigate if strong inhomogeneities in the gas density profiles can slow down excessive radial drift and can help dust particles to grow.} {We use coagulation/fragmentation and disk-structure models, to simulate the evolution of dust in a bumpy surface density profile which we mimic with a sinusoidal disturbance. For different values of the amplitude and length scale of the bumps, we investigate the ability of this model to produce and retain large particles on million years time scales. In addition, we introduced a comparison between the pressure inhomogeneities considered in this work and the pressure profiles that come from magnetorotational instability. Using the Common Astronomy Software Applications ALMA simulator, we study if there are observational signatures of these pressure inhomogeneities that can be seen with ALMA.} {We present the favorable conditions to trap dust particles and the corresponding calculations predicting the spectral slope in the mm-wavelength range, to compare with current observations. Finally we present simulated images using different antenna configurations of ALMA at different frequencies, to show that the ring structures will be detectable at the distances of the Taurus Auriga or Ophiucus star forming regions.} {} ", "introduction": "\\label{sec1} The study of planet formation is an important field in astronomy with an increasing research since the middle of the twentieth century, however there are still countless unanswered questions. One of these questions is the observed grain growth to mm sized particles in the outer disk regions (\\cite{1991ApJ...381..250B}; \\cite{2000ApJ...534L.101W, 2005ApJ...626L.109W}; \\cite{2001ApJ...554.1087T, 2003A&A...403..323T}; \\cite{2005ApJ...631.1134A}; \\cite{2006A&A...446..211R}; \\cite{2007prpl.conf..767N}; \\cite{2009ApJ...701..260I}; \\cite{2009A&A...495..869L}; \\cite{2010A&A...512A..15R, 2011A&A...525A..81R}; \\cite{2011A&A...529A.105G}) that suggests a mechanism operating that prevents the rapid inward drift (\\cite{1997Icar..128..213K}, \\cite{2007A&A...469.1169B}, \\cite{2007Natur.448.1022J}). Different efforts are aimed to explain theoretically the growth from small dust particles to planetesimals, which have led to the development of different numerical models, e.g. \\cite{1981Icar...45..517N}, \\cite{2005A&A...434..971D}, \\cite{2008A&A...480..859B}, \\cite{2008A&A...489..931Z}, \\cite{2009ApJ...698.1122O}, \\cite{2010A&A...513A..79B}. Due to the fact that circumstellar disks exhibit a wide range of temperatures, they radiate from micron wavelengths to millimeter wavelengths, which is why they can be observed with infrared and radio telescopes. With the construction of different kinds of these telescopes, e.g. Spitzer, Herschel, SMA, EVLA or ALMA, astronomers can observe with more details the material inside accretion disks around young stars. The parallel development of theory and observations have allowed astrophysicists to study the different stages of planet formation, making this topic one of the most active fields in astronomy today. In the first stage of planet formation, the growth from sub-micron sized particles to larger objects is a complex process that contains many physical challenges. In the case of smooth disk with a radial gas pressure profile that is monotonically decreasing, the dust particles drift inwards owing to the fact that the gas moves with sub-keplerian velocity due to the gas pressure gradient. Before a large object can be formed, the radial drift causes dust pebbles to move towards the star. Moreover, the high relative velocities due to turbulence and radial drift cause the solid particles to reach velocities that lead to fragmentation collisions which do not allow dust particles to form larger bodies \\citep{1977MNRAS.180...57W, 2008A&A...480..859B, 2010A&A...516L..14B}. The combination of these two problems is called ``meter-size barrier'' because on timescales as short as 100 years, a one meter sized object at 1 AU moves towards the star due to the radial drift, preventing that any larger object could be formed. The observations in the inner regions of the disk, where planets like Earth should be formed, are very difficult because these regions are so small on the sky that few telescopes can spatially resolve them. Also, these regions are optically thick. However, what amounts to the meter-size barrier in the inner few AU is a ``millimeter-size barrier'' in the outer regions of the disk. These outer regions ($\\gtrsim 50$ AU) are much easier to spatially resolve and are optically thin. Moreover, one can use millimeter observations, which probe precisely the grain size range of the millimeter-size. Therefore, the study of dust growth in the outer disk regions may teach us something about the formation of planets in the inner disk regions. Observations of protoplanetary disks at sub-millimeter and mm wavelengths show that the disks remain dust-rich during several million years with large particles in the outer regions \\citep{2007prpl.conf..767N, 2010A&A...512A..15R}. However, it is still unclear how to prevent the inward drift and how to explain theoretically that mm-sized particles are observed in the outer regions of the disk. Different mechanisms of planetesimal formation have been proposed to resist the rapid inwards drift like: gravitational instabilities \\citep{2002ApJ...580..494Y}, the presence of zonal flows\\citep{2009ApJ...697.1269J, 2011A&A...529A..62J, 2011ApJ...736...85U} or dead zones of viscously accumulated gas which form vortices \\citep{2006A&A...446L..13V}. With the model presented here, we want to imitate mechanisms that allow to have long-lived pressure inhomogeneities in protoplanetary disks, by artificially adding pressure bumps onto a smooth density profile. \\begin{table} \\caption{Parameters of the model} \\label{table1} \\centering \\begin{tabular}{c c } \\hline\\hline Parameter & Values \\\\ \\hline $A$ & $\\{0.1; 0.3; 0.5; 0.7\\}$ \\\\ $f$ & $\\{0.3; 0.7; 1.0; 3.0\\}$ \\\\ $\\alpha$ & $10^{-3}$ \\\\ $R_{\\star} [R_\\odot]$ & $2.5$ \\\\ $T_{\\star} [K]$ & $4300$ \\\\ $M_{disk} [M_\\odot]$ & $0.05$ \\\\ $\\rho_s[g/cm^3]$& $1.2$ \\\\ $v_f[m/s]$& $10$ \\\\ \\hline \\end{tabular} \\end{table} To confront the millimeter-size barrier, it is necessary to stop the radial drift considering a radial gas pressure profile that is not monotonically decreasing with radius. Instead, we take a pressure profile with local maxima adding a sinusoidal perturbation of the density profile. These perturbations influence directly the pressure, following a simple equation of state for the pressure in the disk. Depending on the size of the particle, the dust grains will be nearly perfectly trapped in the pressure peaks, because a positive pressure gradient can cause those dust particles move outwards. On the other hand, turbulence can mix part of the dust particles out of the bumps, so that overall there may still be some net radial inward drift. More importantly, dust fragmentation may convert part of the large particles into micron size dust particles, which are less easily trapped and thus drift more readily inward. In the work of \\cite{2010A&A...516L..14B}, they compared the observed fluxes and mm spectral indices from Taurus \\citep{2010A&A...512A..15R} and Ophiucus \\citep{2010A&A...521A..66R} star-forming regions with predicted fluxes and spectral indices at mm wavelengths. They neglected the radial drift, forcing the dust particles to stay in the outer disk regions. They aimed to keep the spectral index at low values, which implies that the dust particles could acquire millimeter sizes \\citep {1991ApJ...381..250B}. However, they found over-predictions of the fluxes. As an extension of their work, the purpose of this paper is to model the combination of three processes: the radial drift, the radial turbulent mixing and the dust coagulation/fragmentation cycle in a bumpy surface density profile. Our principal aim is to find out how the presence of pressure bumps can help explain the retainment of dust pebbles in the outer regions of protoplanetary disks, while still allowing for moderate drift and thus obtaning a better match with the observed fluxes and mm spectral indeces. In addition, we show simulated images using different antenna configurations of the complete stage of ALMA, to study if it is possible to detect these kind of inhomogeneities with future ALMA observations. This paper is ordered as follows: Sect. \\ref{sec2} will describe the coagulation/fragmentation model and the sinusoidal perturbation that we take for the initial condition of the gas surface density. Section \\ref{sec3} will describe the results of these simulations, the comparison between existing mm observations of young forming disks and the results from our model. We discuss if the type of structures generated by our model can be detectable with future ALMA observations. In Sect. \\ref{sec4}, we explore the relation of our model with predictions of current simulations of the magnetorotational instability (MRI) \\citep{1991ApJ...376..214B}. Finally, Sect. \\ref{sec5} will summarize our results and the conclusions of this work. \\begin{figure*} \\centering \\includegraphics[width=18cm]{Figure1.pdf} \\caption{Comparison between: The gas surface density (left plot) taken in this work (Eq. \\ref{eq1}) for two different values of the amplitude and constant width (dashed and dot-dashed lines). The Rossby wave instability \\citep{2011arXiv1109.6177R}, and the presence of zonal flows due to MHD instabilities \\citep{2011ApJ...736...85U}. Right plot shows the pressure gradient for each of the gas surface density profiles.} \\label{comparacion} \\end{figure*} ", "conclusions": "\\label{sec5} Theoretical models of dust evolution in protoplanetary disks show that the growth from sub-micron sized particles to larger objects is prevented basically by two phenomena: radial drift and fragmentation. Nevertheless, infrared and radio observations show that millimeter sized particles can survive under those circumstances in the outer regions of disks. Therefore, various theoretical efforts have been focused on explaining the survival of those bodies. Taking into account strong inhomogeneities expected to be in the gas density profile e.g. zonal flows, and using the coagulation/fragmentation and disk-structure models by \\cite{2010A&A...513A..79B}, we have investigated how the presence of pressure bumps can cause the reduction of radial drift, allowing the existence of millimeter sized grains in agreement with observations. In this work, we assumed a sinusoidal function for the gas surface density to simulate pressure bumps. The amplitude and wavelength disturbances are chosen considering the necessary conditions to have outward angular momentum transport in an $\\alpha$-turbulent type disk, outward radial drift of dust and reasonable values compare to predictions from the recent work of zonal flows \\citep{2011ApJ...736...85U}. The results presented here suggest that the presence of pressure bumps with a width of the order of the disk scale-height and an amplitude of $30\\%$ of the gas surface density of the disk, provide the necessary physical conditions for the survival of larger grains in a disk with properties summarized in Table ~\\ref{table1}. Comparisons between the observed fluxes of the Taurus, Ophiucus and Orion Nebula Cluster star forming regions with the results of the models ratify that the effect of the radial drift is reduced allowing particles to grow. Figure \\ref{Fig3} shows how models with these kind of disturbances reproduce much better mm-observations than models with full or without radial drift. In addition, we presented a comparison between the bumpy density profile assumed in this work and 3D MHD models of zonal flows that can cause long lived bumps in protoplanetary disks. We showed that the pressure bumps cause by zonal flows of \\citep{2011ApJ...736...85U} are in agreement with the amplitudes and wavelengths used in this work. Therefore, taking those bumps, the survival of dust particles is possible in the outer regions after some Myr. The simulated images using CASA ALMA simulator (version 3.2.0) show that, with different antenna configuration of the final ALMA stage, the ring structures, due to the presence of the pressure bumps, should be detectable. Future ALMA observations will have an important impact for understanding the first stages of planet formation and it will be very important to investigate if the grain growth and retetion can be explained with the presence of these kind of inhomogeneities in the gas density profile." }, "1112/1112.2692.txt": { "abstract": "% Context: {Recent work identified a growth barrier for dust coagulation that originates in the electric repulsion between colliding particles. Depending on its charge state, dust material may have the potential to control key processes towards planet formation such as MHD (magnetohydrodynamic) turbulence and grain growth which are coupled in a two-way process.} %Aims: {We quantify the grain charging at different stages of disc evolution and differentiate between two very extreme cases: compact spherical grains and aggregates with fractal dimension $D_f = 2$.} %Methods: {Applying a simple chemical network that accounts for collisional charging of grains, we provide a semi-analytical solution. This allowed us to calculate the equilibrium population of grain charges and the ionisation fraction efficiently. The grain charging was evaluated for different dynamical environments ranging from static to non-stationary disc configurations.} %Results {The results show that the adsorption/desorption of neutral gas-phase heavy metals, such as magnesium, effects the charging state of grains. The greater the difference between the thermal velocities of the metal and the dominant molecular ion, the greater the change in the mean grain charge. Agglomerates have more negative excess charge on average than compact spherical particles of the same mass. The rise in the mean grain charge is proportional to $N^{1/6}$ in the ion-dust limit. We find that grain charging in a non-stationary disc environment is expected to lead to similar results.} % Conclusions {The results indicate that the dust growth and settling in regions where the dust growth is limited by the so-called ''electro-static barrier'' do not prevent the dust material from remaining the dominant charge carrier.} ", "introduction": "\\label{sec1} Small dust particles are regarded as the key ingredient to control the planet formation process in protoplanetary discs. At early stages of planet formation, submicron-sized compact particles are thought to grow towards larger but fluffy agglomerates. At mm to cm sizes these agglomerates are supposed to be compacted, whereby they loose their fractal structure. Because of the rise in the mass-to-surface ratio associated with the compaction, agglomerates tend to be more concentrated locally. However, there are a number of potential processes that may control grain growth, one of which is disc turbulence. The magnetorotational instability - MRI - (Balbus \\& Hawley 1991; Hawley \\& Balbus 1991) has been shown to be robust in generating turbulence in Keplerian discs. MHD turbulence provides the internal stress required for mass accretion, the rate of which is constrained by observations. Observations of young stars indicate that discs usually show signatures of active gas accretion onto the central star with mass flow rates of about $10^{-8\\pm1} \\ \\rm M_{\\odot} yr^{-1}$ (e.g., Sicilia-Aguilar et al. 2004). Recent observations from discs around young stars also set constraints on the turbulent linewidth and provide support for subsonic turbulence in discs (Hughes et al. 2011). However, numerical studies (e.g. Fleming \\& Stone 2003) on MRI-driven MHD turbulence have identified locations within the planet forming regions, the so-called ''dead zones'', where the coupling between the gas and the magnetic field is not sufficient to maintain MHD turbulence. Fleming \\& Stone also showed that a low Reynolds stress can be maintained in the dead zone, such that low levels of accretion are sustained there. That is why dead zones have been considered to advance the planet formation process. Dead zones provide a safe environment for grain growth and for planetesimals (Gressel et al. 2011). However, the presence of small grains in the weakly ionized dead zones leads to significant changes in the abundances of the charge carriers in the gas phase and therefore affects the MRI. In other words, MHD turbulence and grain growth are coupled in a two-way process.\\\\ \\indent Grain charging may also effect the dust growth: Okuzumi (2009) pointed out that the so-called ''electro-static barrier'' may inhibit dust coagulation in planet forming regions. In a subsequent study, Okuzumi et al. (2011a) extended the analysis including the dust size distribution. They found that under certain conditions the dust undergoes bimodal growth. While the small aggregates sweep up free electrons, the large aggregates continue to grow. However, the growing stalls if the electrostatic repulsion becomes strong and the collisions are driven by Brownian motion. The results of Okuzumi et al. (2011b) indicate that under minimum mass solar nebula conditions the dust growth halts at aggregate sizes beyond $[4 \\cdot 10^{-5}, 3\\cdot 10^{-2}] \\ \\rm cm$ depending on the radial position. Conventional chemical models studying the dust-grain chemistry in protoplanetary discs account for up to $|Z| = 3$ grain charges (e.g. Sano et al. 2000). More recently, the range of grain charges was extended considerably for various purposes. Focusing on surface layers of T-Tauri and transitional discs, Perez-Becker \\& Chiang (2011) examined the charge distribution on grains covering grain charges $[Z_{\\rm min}, Z_{\\rm max}] = [-200, 200]$. Okuzumi (2009) calculated the charge distribution on dust aggregates. He demonstrated that aggregates made of $N=10^{10}$ monomers can carry excess charges of about $Z_{\\rm min} \\sim - 10^5$. To account for higher grain charges we therefore improved our simple chemical network introduced in Ilgner \\& Nelson (2006a) allowing $[Z_{\\rm min}, Z_{\\rm max}] $ to become free of choice and appropriate to the conditions applied.\\\\ % \\indent In this paper we examine the grain charging for various stages of disc evolution ranging from static to non-stationary disc environment. Linking the grain charging with the gas-dust dynamics provides a natural environment to mimic the variation of the dust-to-gas ratio $\\Sigma_{\\rm d}/\\Sigma_{\\rm g}$. The grain charging is assumed to originate in collisional charging processes between grains, electrons, and gas-phase ions. We include X-ray ionisation from the central star as the primary source for ionisation and consider, to some extent, the contributions from cosmic ray particles. We furthermore assume an equilibrium distribution of dust material in vertical direction balancing turbulent stirring and sedimentation. In particular, we apply the model of Takeuchi \\& Lin (2005) to simulate the dynamics of the gas-dust disc. One of our primary goals is to compare the grain charging obtained for compact spheres and dust agglomerates and the mean grain charge in particular.\\\\ % \\indent We have investigated the effect that thermal adsorption/desorption of metals has on grain charging using two different chemical networks. In comparison with the values obtained by switching off the thermal adsorption of metals, the modified Oppenheimer-Dalgarno model produces lower, the Umebayashi-Nakano model higher values for the mean grain charge. This originates in differences in the thermal velocities of the dominant gas-phase ion: $v_{\\rm HCO^+} < v_{\\rm Mg^+} < v_{\\rm H_2^+}$. In line with expectations, we find that dust agglomerates have a higher charge-to-mass ratio carrying more charges than the corresponding compact spherical particles. Considering stationary disc configurations, we observe that grain charging on dust agglomerates is always associated with the so-called \"ion-dust regime\", where the charge balance is mainly maintained by negatively charged grains and positively charged gas-phase ions. (The ion-electron regime is characterised by the balance between gas-phase ions and electrons.) In particular, we show that the fractional abundances of the charge carriers are independent of the number $N$ of constituent monomers. Concerning compact spherical dust particles, we identified regions $z/h_{\\rm g} > 0$ associated with the ion-electron regime. Another conclusion of our work is that - for the parameter range considered (i.e., compact radii $a_{\\ast} \\le 10^{-3} \\ \\rm cm$ and agglomerates with characteristic radii $a_{\\rm c} \\le 10^{-2} \\ \\rm cm$ with $N \\le 10^6$, respectively\\footnote{The characteristic and compact radius is defined in Eqs. (\\ref{eq16:sec2}) and (\\ref{eq18:sec2}), respectively.}) - grain charging in a non-stationary disc environment is expected to lead to similar results. We also present semi-analytical solutions for both agglomerates and compact spheres that allow us to determine the equilibrium distribution of $<\\!\\!Z\\!\\!>$, $\\sqrt{<\\!\\!\\Delta Z^2\\!\\!>}$, and the fractional abundances of the charge carriers a priori.\\\\ % \\indent The paper is organised as follows. In Sect.2 we discuss the disc model we apply. In Sect. 3 we introduce the chemical network that we used to compare our results with the reaction network Okuzumi (2009) applied. Key problems concerning the ionisation rates are discussed in Sect. 4 while the numerical methods applied are described in Sect. 5. In Sect. 6 we present the results of our model, and finally in Sect. 7 we summarise the key findings of our study. % ================================================================= % New chapter: % ================================================================= ", "conclusions": "\\label{sec7} We have presented calculations of the grain charging under conditions that mimic different stages of protoplanetary disc evolution. Instead of parametrising the dust-to-gas ratio, we inferred the value for $\\Sigma_{\\rm d}(R)/\\Sigma_{\\rm g}(R)$ from the evolution of the gas-dust disc. In particular, we applied the disc model of Takeuchi \\& Lin (2002, 2005) and content ourself with order-of-magnitude estimates. We considered collisional charging as the dominant process to determine the charge state of grains. For that purpose, we generalised the modified Oppenheimer-Dalgarno model introduced in Ilgner \\& Nelson (2006a) to account for higher grain charges. Based on that simple chemical network, we examined the grain charging for two different types of grain topology: compact spherical grains and fractal agglomerates of $D_{\\rm f} = 2$. Our main conclusions are:\\\\ \\begin{enumerate} \\item The effect that thermal adsorption/desorption of metals has on grain charging depends on the mass of the dominant molecular gas-phase ion. If its mass is heavier than that of metal, the inclusion of the thermal adsorption of metals results in high negative excess charges on grains on average. Less negative charge excess on average is observed for molecular ions lighter than metals. \\item We extended the semi-analytical method proposed by Okuzumi (2009), which allowed us to determine steady-state solutions for the mean grain charge, electron and ion abundances associated with the modified Oppenheimer-Dalgarno model. The semi-analytical solutions were derived for both grain topologies: compact grains and fractal aggregates. \\item The results obtained confirm that dust agglomerates have a higher charge-to-mass ratio than the corresponding compact spheres. \\item We found that reducing the number $N$ of constituent monomers causes a drop in the mean value $<\\!\\!Z\\!\\!>$ of the grain charge and the standard deviation $<\\!\\!\\sqrt{\\Delta Z^2}\\!\\!>$. Under conditions valid for the ion-dust plasma limit (i.e., $x_{\\rm i} \\gg x_{\\rm e}$) the profiles for $<\\!\\!Z\\!\\!>x_{\\rm d}$, $x_{\\rm e}$, and $x_{\\rm i}$ remain unchanged with varying $N$. This is because the cumulative projected surface area of all BCCA agglomerates is independent of $N$. \\item The results obtained by switching from one type of grain topology (fractal agglomerates) to another (compact spheres) reveal that for compact spherical grains $<\\!\\!Z\\!\\!>$ is shifted towards more negative values while $<\\!\\!\\sqrt{\\Delta Z^2}\\!\\!>$ decreases. \\item The results obtained for agglomerate sizes $a_{\\rm c} = [10^{-4}, 10^{-2}] \\ \\rm cm$ indicate that the grain charging of BCCA agglomerates is governed by the ion-dust plasma limit. Transitions from the ion-dust to the ion-electron plasma limit are observed for compact spheres depending on altitude $z/h_{\\rm g}$. \\item Another conclusion is that grain charging in a non-stationary disc environment is expected to lead to similar results as long as the effective ionisation rate and the temperature of the gas match the stationary values. \\end{enumerate} We regard the drastically simplified description of the temperatures of the gas and the dust mentioned above as a potentially serious omission of our model. Working out more realistic conditions may result in significant changes in the local disc structure and might be a potential problem for future investigations. \\begin{acknowledgement} \\noindent I appreciate the discussions with Hiroshi Kobayashi, Satoshi Okuzumi, and Taku Takeuchi very much indeed. \\end{acknowledgement} % ================================================================= % References % =================================================================" }, "1112/1112.6040_arXiv.txt": { "abstract": "One possibility for explaining the apparent accelerating expansion of the universe is that we live in the center of a spherically inhomogeneous universe. Although current observations cannot fully distinguish $\\Lambda$CDM and these inhomogeneous models, direct measurement of the acceleration of the universe can be a powerful tool in probing them. We have shown that, if $\\Lambda$CDM is the correct model, DECIGO/BBO would be able to detect the positive redshift drift (which is the time evolution of the source redshift $z$) in 3--5 year gravitational wave (GW) observations from neutron-star binaries, which enables us to rule out any Lema\\^itre-Tolman-Bondi (LTB) void model with monotonically increasing density profile. We may even be able to rule out any LTB model unless we allow unrealistically steep density profile at $z\\sim 0$. This test can be performed with GW observations alone, without any reference to electromagnetic observations, and is more powerful than the redshift drift measurement using Lyman $\\alpha$ forest. ", "introduction": " ", "conclusions": "" }, "1112/1112.3948_arXiv.txt": { "abstract": "It is well known that the long-term evolution of the photospheric magnetic field plays an important role in building up free energy to power solar eruptions. Observations, despite being controversial, have also revealed a rapid and permanent variation of the photospheric magnetic field in response to the coronal magnetic field restructuring during the eruption. The \\hmi\\ instrument (HMI) on board the newly launched \\Sdo\\ (\\sdo) produces seeing-free full-disk vector magnetograms at consistently high resolution and high cadence, which finally makes possible an unambiguous and comprehensive study of this important back-reaction process. In this study, we present a near disk-center, \\goes-class X2.2 flare, which occurred in NOAA AR 11158 on 2011 February 15. Using the magnetic field measurements made by HMI, we obtained the first solid evidence of a rapid (in about 30 minutes) and irreversible enhancement in the horizontal magnetic field at the flaring magnetic polarity inversion line (PIL) by a magnitude of $\\sim$30\\%. It is also shown that the photospheric field becomes more sheared and more inclined. This field evolution is unequivocally associated with the flare occurrence in this sigmoidal active region, with the enhancement area located in between the two chromospheric flare ribbons and the initial conjugate hard X-ray footpoints. These results strongly corroborate our previous conjecture that the photospheric magnetic field near the PIL must become more horizontal after eruptions, which could be related to the newly formed low-lying fields resulted from the tether-cutting reconnection. ", "introduction": "Almost two decades ago, we discovered rapid and permanent changes of vector magnetic fields associated with flares \\citep{wang92,wang94}. Specifically, the transverse field near the flaring magnetic polarity inversion line (PIL) is found to enhance substantially and irreversibly across the time duration of the flare, which is also often accompanied by an increase of magnetic shear. Similar trend indicating a more horizontal orientation of the photospheric magnetic field after flares and coronal mass ejections (CMEs) has continued to be observed later on in many observations \\citep{wang02b,wang04,wang+liu05,liu05,wang07a,jing08,li09,chang11}, and shows some agreement with recent model predictions \\citep{li10}. Nevertheless, a majority of such studies are unavoidably hampered by the obvious limitations, ground-based observations (e.g., seeing variation and the limited number of observing spectral positions), probably because of which mixed results were also reported \\citep{ambastha93,hagyard99,chen94,li00a,li00b}. On the other hand, flare-related variations in the line-of-sight (LOS) component of photospheric magnetic field have been clearly recognized \\citep[e.g.,][]{wang02b,spirock02,yurchyshyn04,sudol05,wang06,wang10,petrie10}. In particular, the feature of unbalanced flux evolution of the opposite polarities could provide an indirect evidence for the more horizontal orientation of photospheric fields after flares/CMEs \\citep{wang10}. However, it is noted that the changes of the LOS field alone cannot provide complete understanding of the field restructuring \\citep{hudson11}. It is notable that vector magnetic field data has been made available with the \\hmi\\ (HMI) instrument \\citep{schou11} on board the newly launched \\Sdo\\ (\\sdo). Its unprecedented observing capabilities give a favorable opportunity to finally resolve any uncertainties regarding the evolution of photospheric magnetic field in relation to flares/CMEs. In this study, we investigate a near disk-center X2.2 flare on 2011 February 15, which provides the first solid evidence of the enhancement in the horizontal field at the flaring PIL using the seeing-free HMI data. We will discuss the implications of such a change in the context of magnetic reconnection model for flares. ", "conclusions": "We have used the unprecedented SDO/HMI vector field observations to analyze the changes of the photospheric magnetic field associated with the first X-class flare in the solar cycle 24, with the aid of images of flare emissions in multiple wavelengths. Main results are as follows. \\begin{enumerate} \\item A compact region R along the flaring PIL shows a rapid and permanent enhancement of $\\langle B_h \\rangle$ by 400~G (\\sm30\\% of the preflare magnitude) within about 30~minutes, which has a close temporal relationship with the flare HXR emission. Meanwhile, the nonpotentiality represented by magnetic shear also exhibits a pronounced increase near the surface. \\item The initial HXR sources FP1 and FP2 as well as the double J-shaped flare ribbons are at the two ends of the region R lying at the central of this sigmoidal active region. Two additional flare footpoints FP3 and FP4 are clearly seen in the hot 94~\\AA\\ channel, located at the far ends of the sigmoid. We suggest that the tether-cutting reconnection \\citep{moore01} between the loops FP3--FP2 and FP4--FP1 produces the short and low-lying loops FP1--FP2, which could explain the enhanced $B_h$ as well as $\\tilde{S}$ and $\\mathring{S}$ at the region R \\citep{melrose97}. The detected enhancement of nonpotentiality on the surface could also be due to the newly emerging, sheared fields \\citep{jing08}. \\end{enumerate} In summary, the HMI observations presented in this study constitute the first solid evidence of flare-induced changes of the photospheric magnetic field, which strongly endorses our previous results using ground-based vector magnetograms subject to seeing variation \\citep[][and references therein]{wang10}. The unambiguously observed enhancement of horizontal field on the surface strongly suggests that the photospheric magnetic field could respond to the coronal field restructuring by tilting toward the surface (i.e., toward a more horizontal state) near the PIL, and that this development may be due to the tether-cutting reconnection producing the flare. This view is also well in line with the recent theoretical development \\citep{hudson08,fisher10}, where the back reaction on the solar surface resulting from the coronal field evolution as required by the energy release is quantitatively assessed. Further systematic studies of flare-related magnetic field change, especially when aided with extrapolation models, are promising to provide further insight into the relationship between the surface field change and coronal magnetic reconnection \\citep[e.g.,][]{sun11,chang11}." }, "1112/1112.3386_arXiv.txt": { "abstract": "We present 15-20 \\mum\\, spectral maps towards the reflection nebula NGC\\,2023 obtained with the Infrared Spectrograph in short-wavelength, high-resolution mode on board the Spitzer Space Telescope. These spectra reveal emission from PAHs, C$_{60}$, and H$_2$ superposed on a dust continuum. These emission components exhibit distinct spatial distributions: with increasing distance from the illuminating star, we observe the PAH emission followed by the dust continuum emission and the H$_2$ emission. The C$_{60}$ emission is located closest to the illuminating star in the south while in the north, it seems to be associated with the H/H$_2$ transition. Emission from PAHs and PAH-related species produce features at 15.8, 16.4, 17.4, and 17.8 \\mum\\, and the 15-18 \\mum\\, plateau. These different PAH features show distinct spatial distributions. The 15.8 \\mum\\, band and 15-18 \\mum\\, plateau correlate with the 11.2 \\mum\\, PAH band and with each other, and are attributed to large, neutral PAHs. Conversely, the 16.4 \\mum\\, feature correlates with the 12.7 \\mum\\, PAH band, suggesting that both arise from species that are favored by the same conditions that favor PAH cations. The PAH contribution to the 17.4 \\mum\\, band is displaced towards the illuminating star with respect to the 11.2 and 12.7 \\mum\\, emission and is assigned to doubly ionized PAHs and/or a subset of cationic PAHs. The spatial distribution of the 17.8 \\mum\\, band suggests it arises from both neutral and cationic PAHs. In contrast to their intensities, the profiles of the PAH bands and the 15-18 \\mum\\, plateau do not vary spatially. Consequently, we conclude that the carrier of the 15-18 \\mum\\, plateau is distinct from that of the PAH bands. ", "introduction": "\\label{intro} Strong emission bands at 3.3, 6.2, 7.7 and 11.3 \\mum\\,- the so-called unidentified infrared (UIR) bands - dominate the mid-IR spectra of almost all objects, including reflection nebulae, planetary nebulae, the interstellar medium and \\HII\\, regions. These bands are generally attributed to IR fluorescence of a family of Polycyclic Aromatic Hydrocarbon molecules (PAHs) pumped by the UV radiation field. A key result in the observational studies of PAHs is that their mid-IR bands show clear variations in peak positions, shapes and (relative) intensities, not only between sources, but also spatially within extended sources \\citep[e.g.][]{Hony:oops:01, Peeters:prof6:02, Brandl:06, SmithJD:07, Galliano:08}. The variability in intensity ratio of the main PAH bands has been interpreted in terms of the charge state and excitation levels of the PAHs \\citep[e.g.][]{Allamandola:modelobs:99, Galliano:08} while the variation in the profiles of the main PAH bands \\citep[classified as class A, B and C profiles;][]{Peeters:prof6:02, vanDiedenhoven:chvscc:03} is attributed to the environment \\citep[CSM versus ISM;][]{Peeters:prof6:02, VanKerckhoven:phd:02, Boersma:08} and the degree of UV processing \\citep{Sloan:07, Tielens:08, Boersma:08}. In addition to the well-studied main PAH bands, a plethora of weaker PAH bands are present, for example in the 15-20 \\mum\\, region. PAH emission in this region was first observed with the Infrared Space Observatory (ISO): new features were reported at 15.8, 16.4, 17.4 and 17.9 \\mum\\, with the 16.4 and then 17.4 \\mum\\, PAH bands being the most prominent \\citep{Beintema:pahs:96, Moutou:16.4:00, Sturm:swsgal:00, VanKerckhoven:plat:00}. In addition, \\citet{VanKerckhoven:plat:00} presented evidence for a variable, broad plateau from 15 to 20 \\mum\\, which seems to be present solely in \\HII\\, regions \\citep{VanKerckhoven:plat:00, Peeters:plat:04, Peeters:spitzer:04}. Based upon observations with the Spitzer Space Telescope, additional broad components are reported at 16.6, 17.0 and 17.2 \\mum\\, \\citep{SmithJD:04, SmithJD:07, Sellgren:07}. But it is really the high sensitivity of Spitzer that revealed the omnipresence of these bands within the Milky Way and in other galaxies \\citep[e.g.][]{Werner:04, SmithJD:04, Brandl:04, Sellgren:07, SmithJD:07, Tappe:06}, allowing - for the first time - systematic investigations of these weaker bands. The 15-20 \\mum\\, emission is attributed to C-C-C bending vibrations \\citep[e.g.][]{ATB, Moutou:firempahs:96, VanKerckhoven:plat:00, Peeters:plat:04, Mattioda:09, Boersma:10, Ricca:10}. As a consequence, the bands in the 15-20 \\mum\\, region are expected to reveal more about the overall molecular structure of the carriers than do the major bands below 15 \\mum. An intriguing result of the Spitzer observations is the remarkable similarity (to first order) of the 15-20 \\mum\\, PAH emission spectra from regions spanning a large range of physical conditions. This suggests that the astronomical PAH population may be dominated by a handful of stable, molecular structures (cf. survival of the fittest) rather than comprised of a large number of PAHs with widely varying structures \\citep{Boersma:10}. Hence, a systematic study of a large sample of spectra showing these weaker PAH bands will provide a complementary view on the characteristics of the emitting PAH population. This paper reports such a study. Here we analyze Spitzer-IRS spectral maps of NGC\\,2023 in the 15-20 \\mum\\, region. In Section \\ref{source}, we describe the reflection nebula NGC\\,2023, while the observations and data reduction are discussed in Section \\ref{data}. The data analysis is presented in Section \\ref{analysis} and discussed in Section \\ref{discussion}. We end with a short summary in Section \\ref{conclusion}. \\begin{figure}[t!] \\centering \\resizebox{14cm}{!}{% \\includegraphics{fig1.eps}} \\caption{The IRAC [8.0] image of NGC\\,2023 with the SH FOV shown (white) for both the north and south positions studied here. The star HD37903 is indicated by a black circle, sources A, C and D are from \\citet{Sellgren:83}, and the white circles indicate 2MASS point sources located inside the SH apertures. S refers to the south ridge, SSE to the south-southeastern ridge, N to the north ridge, and NW to the northwestern ridge. Maps shown in this paper use the orientation denoted by the vectors (X, Y).} \\label{fov} \\end{figure} ", "conclusions": "\\label{conclusion} We studied the emission in the 15-20 \\mum\\, region towards the reflection nebula NGC\\,2023 by analysing two spectral maps obtained with Spitzer's Infrared Spectrograph (IRS), short-wavelength high-resolution mode. We observed PAH emission bands at 15.8, 16.4, 17.4, and 17.8 \\mum, a broad PAH plateau between 15-18 \\mum, C$_{60}$ emission at 17.4 and 18.9 \\mum, and H$_2$ emission at 17.0 \\mum\\, superposed on a dust continuum. \\\\ We found distinct spatial distributions for these emission components and found tight intensity correlations between some emission components. Based on these results, we collected the emission components in seven groups and discussed their specific characteristics. {\\it Group 1: the 11.2, 15.8 \\mum\\, PAH bands and the 15-18 \\mum\\, PAH plateau.} We attributed this group to large, neutral PAHs. {\\it Group 2: the 12.7 and 16.4 \\mum\\, PAH bands.} Compared to Group 1, these bands are displaced towards the illuminating star. We concluded that they must arise from species that are co-spatial with the cationic portion of the emitting PAH population. {\\it Group 3: the 17.4 \\mum\\, diffuse component.} We estimated the PAH contribution to the 17.4 \\mum\\, band and found that its spatial distribution is displaced towards the illuminating star compared to Groups 1 and 2. We assigned this PAH band to doubly ionized PAHs and/or a subset of the cationic PAH population, such as for example dehydrogenated PAHs. {\\it Group 4: the 17.8 \\mum\\, PAH band.} This band has a spatial distribution in between that of Groups 1 and 2. We suggested this band arises from both neutral and ionised PAHs. {\\it Group 5: the 18.9 and (some of) the 17.4 \\mum\\, features.} The C$_{60}$ emission shows a spatial distribution distinct from that of the PAH emission, consistent with \\citet{Sellgren:10}. Moreover, the spatial behavior of C$_{60}$ towards NGC2023 is different in the south and north. Specifically, the C$_{60}$ emission is located closer to the illuminating star compared to the PAH and H$_2$ emission in the south while in the north, it seems to be associated to the H/H$_2$ transition. {\\it Group 6: the underlying continuum producing the emission measured here at 14.9 and 19.2 \\mum.} The continuum emission is strongest deeper into the PDR compared to the PAH emission. This is consistent with the VSG component attributed to PAH clusters by, for example, \\citet{Rapacioli:05} and \\citet{Berne:07}. {\\it Group 7: the H$_2$ emission.} The H$_2$ emission is located furthest into the PDR.\\\\ We also investigated the profiles of the PAH bands and 15-18 \\mum\\, PAH plateau. All PAH bands have a symmetric band profile except for the 16.4 \\mum\\, band, which is slightly asymmetric. We did not observe any changes in the band profiles spatially across the maps. Likewise, we found that the profile of the 15-18 \\mum\\, PAH plateau does not vary. We concluded that the carrier of the underlying 15-18 \\mum\\, plateau is distinct from the individual PAH bands located on top of it." }, "1112/1112.4790_arXiv.txt": { "abstract": "In a power spectrum of the Sun's acoustic waves, the $p$ modes have distinctly skewed frequency profiles. Furthermore, the asymmetry is observed to have the opposite sign in power spectra made from line-of-sight velocity and continuum intensity. The asymmetry and its reversal in sign has previously been explained using a combination of mechanisms that involve the acoustic source. A localized acoustic source within an acoustic cavity naturally generates asymmetric profiles through wave interference; however, the sign of the asymmetry due to this mechanism is identical for all observables. The reversal of the asymmetry between velocity and intensity observations has been attributed to the visibility of the source itself (i.e., ``correlated noise\"). In this paper, I will show that asymmetry generated by a localized source can be interpreted as either a wave interference effect in physical space, or a mode interference effect in spectral space. I advocate a new mode-fitting procedure based on this new interpretation, whereby the complex phases of all the modes determine the mode asymmetries. Further, I suggest that information about the acoustic source function is encapsulated in the amplitude of each mode, and present a scheme by which the source function can be obtained from measured mode amplitudes by standard helioseismic inversion techniques. ", "introduction": "\\label{sec:introduction} \\setcounter{equation}{0} In power spectra of solar acoustic oscillations the frequency profiles of the $p$ modes are skewed and markedly non-Lorentzian. This property was first discovered using Ca II K observations made from the geographic South Pole \\citep{Duvall:1993} and were later confirmed \\citep{Nigam:1998a} with modern spacecraft-based telescopes using data from the Michelson Doppler Imager (MDI) onboard the Solar and Heliospheric Observatory (SOHO). In both sets of observations, a puzzling fact was revealed: for modes of the same frequency and harmonic degree, the asymmetry or skewness of the line profile has opposite signs for observations made in the continuum intensity and line-of-sight Doppler velocity. For both types of observations, the lowest-frequency modes are the most severely skewed and the asymmetry appears to be only a weak function of harmonic degree; however, for modes of high harmonic degree ($l \\gtrsim 150$) this statement is difficult to verify because the mode peaks are highly blended with the spatial leaks from modes of nearby degree, forming a broad ridge of power. Figures~\\ref{fig:GONGpower_l=190} and \\ref{fig:GONGpower_l=20} illustrate the asymmetry for both high and low degree. From the figures it is evident that the frequency profile of each individual mode is asymmetric and, at high degree, the blended ridge is also asymmetric. The physical mechanism for this skewness depends somewhat on whether we are considering the asymmetry of each mode profile or the asymmetry of a blended ridge. In the former case, the asymmetry of each mode is now thought to arise from a combination of effects due to the nature and location of the acoustic wave source. At high harmonic degree, the asymmetry of the $p$ mode ridge is the result of the skewness of each component mode profile as well as an uneven distribution of power amongst the spatial leaks. The wave source can introduce skewness in a variety of ways. The first mechanism that I discuss here relies in detail on the location of the acoustic source. The primary source of acoustic waves in the Sun is granular convection. Short-lived convective events of relatively high Mach number prolifically emit sound waves. Since granulation is confined to a narrow band just below the photosphere, the acoustic source is highly localized in depth. Waves generated in this layer propagate both upwards and downwards. Those that propagate downwards are eventually refracted back upwards at the bottom of the acoustic cavity by the sound speed gradient. When this initially downward propagating wave returns to the excitation layer, it can destructively interfere with the wave that was initially upward propagating. At particular frequencies, complete destructive interference occurs and zero wave amplitude is observed above the excitation layer (which is where the spectral lines used in actual observations are formed). The frequencies at which the interference is complete are asymmetrically placed between the mode ridges, thereby leading to asymmetric mode profiles \\citep{Gabriel:1992, Gabriel:1995, Roxburgh:1995, Abrams:1996, Rast:1998}. The second mechanism due to the source arises because the convective events that generate sound waves are themselves regions of intense temperature fluctuation and high velocity. The sources therefore provide a contribution to the power in either intensity or velocity observations. Since the acoustic waves are correlated with the source that generates them, the visibility of the source can lead to power that varies significantly over the line profile and cause skewness. This mechanism has been dubbed ``correlated noise\" and has been suggested as the reason why the asymmetry is opposite for intensity and velocity spectra \\citep{Roxburgh:1997, Nigam:1998a, Nigam:1998b}. The current understanding is that in order to reproduce the asymmetry of both intensity and velocity power spectra and to explain the rapid variation with frequency of the phase of the cross-spectra between intensity and velocity \\citep{Oliviero:1999}, correlated noise must provide significant power to both types of spectra \\citep[e.g.,][]{Nigam:1999, Skartlien:2000, Jefferies:2003}. Numerical simulations of granulation have provided a counter explanation \\citep{Georgobiani:2003}. In these simulations, the velocity and intensity fluctuations have power spectra with asymmetric mode profiles. However, if the two quantities are sampled at the same geometric height the sign of the asymmetries are identical. Only after artificial Dopplergrams and intensity images have been constructed using radiative transfer do the two types of power spectra exhibit opposite asymmetries. Therefore, these simulations suggest that details of the radiative transfer, such as the different heights at which the two observables are formed and the nonlinear interaction between wave fluctuations and the radiation field, are responsible for the opposite skewnesses. Unfortunately, many mode-fitting schemes used to measure $p$ mode frequencies assume that the line profile is symmetric, a Lorentzian in fact. \\cite{Duvall:1993} estimate that failing to account for the asymmetry in the mode's frequency profile can lead to mismeasurement of mode frequencies by as much as one part in $10^4$, a significant error for the purposes of helioseismic inversions. This is probably the source of the systematic offset ($\\sim 0.1~\\mu$Hz) that was observed between frequencies obtained from intensity and velocity spectra \\citep{Toutain:1997}. To account for such systematic shifts, researchers have suggested a variety of asymmetric model profiles with which to fit the observations \\citep[e.g.,][]{Duvall:1993, Nigam:1998b, Rosenthal:1998}. Due to its practicality, the formulation of \\cite{Nigam:1998b} is probably the most widely used, having been adapted for both the measurement of global mode frequencies \\citep{Toutain:1998, Reiter:2002, Korzennik:2005} and the fitting of ring-analysis spectra \\citep{Basu:1999, Basu:2001, Tripathy:2009}. In essence, the profile of \\cite{Nigam:1998b} is a Lorentzian times a polynomial in frequency, where the polynomial has been calculated by performing a low order expansion about the mode frequency $\\omega_0$, \\begin{eqnarray} P(\\omega) &=& \\frac{A}{1+x^2} \\, \\left[s^2 + (1 + sx)^2\\right] \\; , \\label{eqn:Nigam} \\\\ x &\\equiv& (\\omega-\\omega_0)/\\gamma \\; . \\end{eqnarray} \\noindent In this formula, $\\gamma$ is the linewidth, $A$ is the power amplitude, and $s$ is an asymmetry parameter. This expression is formally valid under fairly limited conditions. For example, it has been derived as a small argument expansion in terms of $sx$, which requires that asymmetry is weak and more restrictively that we confine our attention to the core of the line and avoid the wings. One clear sign of this restriction is that the profile isn't integrable, and does not contain a finite amount of energy. Second, it ignores the presence of nearby modes which may contribute to the power. This is a particularly onerous assumption for high degree modes for which the line widths are sufficiently large that the mode and the neighboring spatial leaks blend into a ridge. Figure~\\ref{fig:GONGpower_l=190} illustrates this later point quite well. Each ridge is a blending of mode peaks and the ridges are wide enough that they have begun to overlap. Our goal here is to generate a fitting function that takes the skewness of the $p$-mode frequency profiles into account while maintaining its validity over a wide range of conditions. I do so by using a modal expansion of the wavefield to represent the power spectra using a small number of free parameters. This expansion is accomplished in a general manner that doesn't depend on details of the atmosphere or the mechanisms for wave damping and excitation. In the next section, Section~\\ref{sec:driven}, I discuss the driving and damping of acoustic $p$ mode oscillations in broad terms. In Section~\\ref{sec:ModeDecomp}, I derive the modal expansion of the wave field and compute the resulting power spectrum. In Section~\\ref{sec:ModeFitting}, I promulgate a fitting profile of the spectrum that treats the asymmetry as a mode interference phenomenon. Finally, in Section~\\ref{sec:discussion}, I discuss my findings and suggests ways in which properties of the acoustic source could be deduced through inversion of the fitted complex mode amplitudes. ", "conclusions": "\\label{sec:discussion} \\setcounter{equation}{0} I have suggested an alternate formula---Equation~\\eqnref{eqn:FitFunc}---for fitting $p$-mode frequency profiles that is based on an eigenfunction expansion of the Green's function. This formulation takes into account the skewness of the line profiles that is inherently produced by a localized acoustic source by including the effects of mode interference in spectral space. In addition, the fitting formula has the salutory features that it is valid for all frequencies across the mode profile (not just in the core of a line) and each mode contains a finite amount of energy. The suggested scheme can be applied to the fitting of global modes of relatively low harmonic degree ($l \\lesssim 200$) directly as envisioned. The asymmetry in a mode's profile is primarily the result of interference with nearby spatial leaks, which are most likely from the same mode order or ridge (except at very low harmonic degree $l \\lesssim 10$). On the other hand, the proposed fitting model is not immediately relevant to the fitting of ring-analysis spectra. The asymmetry observed in the $p$-mode ridges at large harmonic degree ($l \\gtrsim 200$) are not produced by the same mechanisms that generate the asymmetry for global modes. Instead, the asymmetry is largely a result of the spatial window function (the ring-analysis version of the blending of a mode and the nearby leaks). Ring-analysis spectra are manufactured by Fourier transforming the wavefield observed on a small patch on the surface of the Sun. The small spatial domain translates into a broad spread of power in horizontal wavenumber. This widens the ridges and produces an effective line width that is much larger than one would expect based just on the mode lifetime. Furthermore, the resulting line profile depends on the window function applied during the Fourier analysis, the slope of the $p$-mode ridge in the dispersion diagram ($d\\omega_n/dk$), and the variation in power along the ridge. Unlike in the global mode analyses where the effects of leakage have been considered in detail for over a decade, leakage has not been treated self-consistently in any ring-analysis studies. While the fitting formula presented here could be used to model the resulting ring-analysis spectra quite well with the asymmetry being produced by the interference between the ridges themselves, the underlying physical mechanism would be wrong. I leave it to a subsequent paper to include the effects of the leakage function on mode profiles. \\subsection{Measuring the Acoustic Source Function} \\label{subsec:SourceFunc} One novel property of the fitting function given by Equation~\\eqnref{eqn:FitFunc} is that the effect of the source is directly represented by the complex mode amplitudes. Therefore, the additional parameters have a physical significance that can be used to ferret out information about acoustic sources and the convection that engenders them. This can be seen most easily if I relax the assumption that the source is proportional to a delta function in height. If instead the source is distributed, as usual the wavefield is obtained through a convolution of the Green's function with the source, \\begin{eqnarray} \\psi(z;\\bvec{k},\\omega) &=& \\sum_{n \\neq 0} \\frac{a_n(\\bvec{k}, \\omega)}{\\omega-\\hat{\\omega}_n(k)} \\; , \\\\ &=& \\sum_{n \\neq 0} \\intall dz' \\; \\frac{\\psi_n(z';k) \\, \\psi_n(z;k)}{2\\hat{\\omega}_n(k)}\\; \\frac{S(z';\\bvec{k},\\omega)}{\\omega-\\omega_n(k)} \\; . \\end{eqnarray} \\noindent Assuming that the source function is isotropic with a frequency variation that is slow compared to the damping rate, the complex mode amplitude is therefore given by an integral over the source, \\begin{eqnarray} a_n(k) &=& \\intall dz' \\; {\\cal K}_n(z';k) \\, S(z';k,\\omega_n) \\; , \\label{eqn:Aintegrals} \\\\ {\\cal K}_n(z';k) &\\equiv& \\frac{\\psi_n(z;k) \\, \\psi_n(z';k)}{2\\hat{\\omega}_n(k)} \\; , \\end{eqnarray} \\noindent where the set of ${\\cal K}_n$ are sensitivity kernels. Given a set of measured amplitudes, the coupled integral equations~\\eqnref{eqn:Aintegrals} could be inverted using standard helioseismic inversion procedures to obtain the source as a function of height in the atmosphere. \\subsection{Effects of Correlated Noise} \\label{subsec:CorrNoise} The theory and procedures that have been developed in this paper have so far ignored the existence of correlated noise. Previous studies have shown that correlated noise is a necessary ingredient to explain the reversal of asymmetry between Doppler-velocity spectra and continuum-intensity spectra \\citep{Roxburgh:1997, Nigam:1998a, Nigam:1998b}, as well as the rapid phase variation in velocity--intensity phase difference spectra \\citep{Skartlien:2000}. The modal expansion presented here can easily be adapted to include the effects of correlated noise \\citep[e.g.,][]{Skartlien:2000}; one simply adds a complex correlated noise component, $\\psi_{\\rm corr}$, to the wavefield before computing the power, \\begin{equation} P_{\\rm model}(\\omega) = B(\\omega) + \\left| \\frac{a_j}{\\omega-\\hat{\\omega}_j + i\\Gamma_j} + \\sum_{j'} c_{jj'} \\, \\frac{a_{j'}}{\\omega-\\hat{\\omega}_{j'} + i\\Gamma_{j'}} + \\psi_{\\rm corr} \\right|^2 \\; , \\label{eqn:PowerCorrNoise} \\end{equation} \\noindent Assuming that the correlated noise is due to the visibility of the source itself, one would expect that the phase and amplitude of the correlated noise $\\psi_{\\rm corr}$ would be a slowly varying function of frequency. Therefore, if one is fitting the power in a narrow frequency band, it can safely be assumed that the noise is a complex constant. Thus, in any fit to data the correlated noise adds only two additional parameters that must be fit. If one were fitting a broad range of frequencies, it might be necessary to parameterize the frequency dependence of the correlated noise. Since, the frequency variation should be rather weak, almost any smooth function with a only a few free parameters (polynomial, power law, gaussian, etc.) would probably work." }, "1112/1112.4926_arXiv.txt": { "abstract": "Nanoflares, the basic unit of impulsive energy release may produce much of the solar background emission. Extrapolation of the energy frequency distribution of observed microflares, which follows a power law to lower energies can give an estimation of the importance of nanoflares for heating the solar corona. If the power law index is greater than 2, then the nanoflare contribution is dominant. We model time series of extreme ultraviolet emission radiance, as random flares with a power law exponent of the flare event distribution. The model is based on three key parameters, the flare rate, the flare duration and the power law exponent of the flare intensity frequency distribution. We use this model to simulate emission line radiance detected in 171 \\AA, observed by STEREO/EUVI and SDO/AIA. The Observed light curves are matched with simulated light curves using an Artificial Neural Network and parameter values are determined across regions of active region, quiet sun, and coronal hole. The damping rate of nanoflares is compared with radiative losses cooling time. The effect of background emission, data cadence, and network sensitivity on the key parameters of model is studied. Most of the observed light curves have a power law exponent, $\\alpha$, greater than the critical value 2. At these sites nanoflare heating could be significant. ", "introduction": "The mechanism of forming nanoflares is the dissipation of current sheets arisen from tangential discontinuities in the continuously evolving corona (Levine 1974, Parker 1988). Parker presumes that the change in the magnetic field across the current sheet, $\\Delta B$, is critical for the onset. When the strength $|\\Delta{B}|$ of the discontinuity exceeds some threshold, there is a runaway dynamical instability leading to an explosive reconnection phase. This is similar to the sand pile model that has been used to explain the comparison is also made between results of different methods. Avalanches of magnetic reconnection (Lu \\& Hamilton 1991), pointed out that the magnetic field of the corona is in a state of self organized criticality. To determine whether nanoflares are the main source of heat input to the corona or not, it is necessary to measure the energy frequency distribution of the smallest observed flares. Hudson (1991) pointed out that the flare occurrence follows a power law distribution such as, $dN\\sim E^{-\\alpha}dE$, in which $dN$ is the number of flares per energy interval $E$ and $E+dE$ (Lu \\& Hamilton (1991) obtained this distribution for complex systems which are in a self organized critical state). The power law index, $\\alpha$, is a critical value for determining whether more weight is given to small scale events (nanoflares) or larger ones (flares). Determination of the power law index of the flare frequency distribution is a scientific challenge. Many authors have attempted to find this index. Benz \\& Krucker (1998), using Yohkoh and SoHO, shown that the power law index at the microflare frequency distribution is 2.5. Parnell \\& Jupp (2000), used TRACE observations and concluded that the energy of nanoflares is insufficient to heat the quiet sun corona. Aschwanden \\& Parnell (2002) also get some other power law distribution by combining scaling law and fractal geometry for different observations. Aschwanden \\& Charbonneau (2002), gathered the statistics of solar flares, microflares, and nanoflares and concluded that, the power law index falls bellow the critical value. An extended review on the simulation and observational results is given by ( Klimchuk 2006 and Klimchuk et al. 2009). {To determine the contribution of small scale events (nanoflares) on the solar coronal heating, some applicable models has been investigated (e. g., Vekstein \\& Katsukawa 2000, Sakamoto et al. 2009, Terzo et al. 2011). Here,} we use a model to simulate the observed Extreme Ultra Violet (EUV) emitted radiation from STEREO and SDO which has been applied successfully to the UV radiance fluctuation in the quiet Sun (Pauluhn \\& Solanki 2007) and later to SUMER observations of the corona in an active region (Bazarghan et al. 2008). This paper is organized as follow: STEREO/EUVI and SDO/AIA data analysis are described in Section \\ref{data}. The nanoflare model is treated in Section \\ref{model}. A kind of Artificial Neural Networks (Probabilistic Neural Network) is briefly discussed in Section 4. In sections \\ref{method} and \\ref{result} the method and results are presented. The conclusions are given in Section \\ref{conc}. ", "conclusions": "\\label{conc} The basic aim of the model, as mentioned earlier, is to determine the coronal heat input by numerous randomly distributed small scale events specially nanoflares by finding the power law index. This will decides whether the small scale event contribution is of importance or not. The problem of the too power law index indicated by researchers may have been just a bias due to neglecting the overlapping nanoflares. Here, a simple nanoflare model based on three key parameters (the flare rate, the flare decay time, and the power law exponent of the flare energy frequency distribution) is used to simulate emission line radiances from the STEREO/EUVI and SDO/AIA in the corona. The simulation code ran to generate more than 22000 light curves (train set) for each combination of $\\alpha$, $\\tau_d$, and $p_f$. For each of the marked regions on the full solar disk images, more than 15000 light curves were generated for average intensities. This large number of perfect light curves which enable us in statistical description was focus of the present-paper. Light curve pattern recognition by a Probabilistic Neural Network (PNN) was employed to determine values of the key parameters. We found that more than 85\\% of the observed light curves have a power law index greater than 2. Since the network's sensitivity depends on the training set in which the network must see all the possible patterns during the training session, we may have some errors in recognizing the correct patterns. Empirically, the network is sensitive to steps $\\Delta \\tau_d\\approx1$, $\\Delta p_f\\approx0.1$, and $\\Delta\\alpha\\approx0.1$ for three key parameters. This means that, for shorter steps the light curves are too like, so that the network is not able to classify. The results can be summarized as follows: \\begin{description} \\item[-] A physical picture of how the model's parameters affect the simulated light curves is discussed. Decreases in both average and variance of the light curves are the function of increasing power law index (greater $\\alpha$ value corresponds to greater number of small events). The higher moments (skewness and kurtosis) values of the time series are accompanied with increasing $\\alpha$, $\\tau_d$, and $p_f$ values (because of the lognormal shape of the distributions). Both the skewness and kurtosis values are positive numbers. {The distributions of both simulated and observed light curves are asymmetric (Terzo et al. 2011). } \\item[-] The average $\\alpha$ and $ p_f$ did not change with changes in data cadences but the average $\\tau_d$ is a sensitive function of decrease in cadence. \\item[-] With regard to average dimensionless range of $\\tau_d$=40-50s and by multiplying it by cadence of $2.5$ min (for STEREO/EUVI, 13 June 2007) and 1.5 min (for SDO/AIA, 22 August 2010) we obtain values of $100-125$ min and $100-75$ min, respectively. Assuming that plasma cooling through the narrow band filter is dominated by radiative cooling we find that the ranges are consistent with previous results. \\item[-] The effect of the background emission on the flare rate, $p_f$ is studied. In the coronal hole regions with less background majority of events has low flare rates. \\end{description} The next logical step is to determine the actual flare energies and the total energy input to the corona which is still a problem for researchers to be solved in the future." }, "1112/1112.3447_arXiv.txt": { "abstract": "We study gaseous outflows from disk galaxies driven by the combined effects of ram pressure on cold gas clouds and radiation pressure on dust grains. Taking into account the gravity due to disk, bulge and dark matter halo, and assuming continuous star formation in the disk, we show that radiation or ram pressure alone is not sufficient to drive escaping winds from disk galaxies, and that both processes contribute. We show that in the parameter space of star formation rate (SFR) and rotation speed of galaxies, the wind speed in galaxies with rotation speed $v_c\\le 200$ km s$^{-1}$ and SFR $\\le 100$ M$_{\\odot}$ yr$^{-1}$, has a larger contribution from ram pressure, and that in high mass galaxies with large SFR, radiation from the disk has a greater role in driving galactic winds. The ratio of wind speed to circular speed can be approximated as ${v_w \\over v_c} \\sim 10^{0.7} \\, \\left[{\\rm SFR\\over 50 \\, {\\rm M}_{\\odot} \\, {\\rm yr}^{-1}}\\right] ^{0.4} \\ \\left[{v_c\\over 120\\, km/s}\\right]^ {-1.25}$. We show that this conclusion is borne out by observations of galactic winds at low and high redshift and also of circumgalactic gas. We also estimate the mass loading factors under the combined effect of ram and radiation pressure, and show that the ratio of mass loss rate to SFR scales roughly as $v_c^{-1} \\Sigma_g^{-1}$, where $\\Sigma_g$ is the gas column density in the disk. ", "introduction": "Galactic winds have been observed at different wavelengths in galaxies of various masses and in a range of redshifts. Galaxies, especially with star formation rates ($\\Sigma_{SFR}$) $\\ge 10^{-1}$ M$_{\\odot}$ yr$^{-1}$ kpc$^{-2}$, often show large outflow of hot gas that emits X-rays and in which cold clouds are found to be embedded, which are observed with H$\\alpha$ or NaD lines \\citep{hec00,mar05}. The speed of the clouds in the wind range from a few tens to several hundred km s$^{-1}$, and the total mass loss rate can be several times the star formation rate \\citep{vei05}. These outflows play a crucial role in the evolution of galaxies by expunging gas, and thereby suppressing the star formation. The attempts to understand galactic evolution in the cosmological context have since long encountered the so-called 'cooling catastrophe' problem, since left to its own device the baryonic gas would cool and form stars more rapidly than observed. It is generally believed that a feedback loop inhibits this, and that the process of star formation excites an outflow and quenches itself. The observed mass-metallicity relation in galaxies also indicate that galactic outflows play a major role in the chemical evolution in galaxies. Furthermore, these outflows enrich the intergalactic medium with metals. The standard model to understand galactic outflows involves a heated interstellar medium (ISM) under the influence of supernovae (SN), and the hot gas being driven by thermal pressure \\citep{che85, hec02}. The expansion speed of this hot gas can be large enough to eject it out of the galaxy \\citep{lar74,sai79,dek86}. The observations of cold gas in these outflows \\citep{hec00} led to the proposal that the cold gas entrained in the hot gas moved due to ram pressure. The wind speed was however not found to correlate with galaxy mass \\citep{hec00,mar99}, and it was argued that the supernovae rate increased with SFR and hence the wind velocity might correlate with SFR. Simulations also supported this scenario \\citep{suc94,str00}. However, there is a limiting cloud speed implicit in this process since ram pressure acts on the cold gas until the cold gas velocity becomes equal to that of hot gas. This scenario, however, has met with problems from new observations of cold component which show that the terminal outflow speed depends on galactic properties like rotation speed \\citep{mar05,rup05}.% It has been proposed that these observations can be explained by radiation pressure driving the outflow \\cite{mqt05, mar05, sha11}. It has also been pointed out that a natural course of events leading from a starburst would be a radiation pressure driven wind in the beginning, and ram pressure acting on it after a period of $\\sim 3\\hbox{--}5$ Myr, the life time of massive stars \\citep{nath08, mur11}. This scenario also naturally explains the puzzling fact that cold clouds are observed at large distances although their survival time-scales in the hot gas would have inhibited them from being pushed out to such distances. In the face of two processes leading to outflows, one wonders if both processes contribute equally, or if there are regimes in which one of these two processes dominate over the other. In this paper we present an analytical calculation for the dynamics of cold clouds taking into account both ram and radiation pressure and all sources of gravity, and compare our results with observations. ", "conclusions": "The most important result of our calculation is that galactic outflows require both ram and radiation pressure, especially for high mass and high SFR cases. Our calculation has a number of ingredients from stellar physics and disk and halo parameters, and apart from the value of the hot wind speed $v_h$, there is no free parameter in this calculation. It is therefore interesting to note that our theoretical results are consistent with most data of outflows when studied in the parameter space of $v_c$ and SFR. It is also interesting that a recent simulation with ram and radiation pressure driven outflows has concluded that these two processes are important in different mass regimes, although it is not clear where the dividing line between the two regimes lies \\citep{hopkins11,sch11}. Cold cloud outflows from galaxies on the left of the contours in figure \\ref{figcie} are unlikely to escape into the IGM and likely get trapped in the circumgalactic region as observed by \\cite{tum11} (data shown by blue cross) or fall back \\citep{opp08}. Although strictly speaking our calculation refers to cold clouds being driven out along the pole of the disk galaxies, and we cannot infer the mass loss rate without doing a 2-D calculation, but we can speculate on the scaling of the mass loss rate with galactic mass by making some simple assumption. Let us assume that the dynamics of cold clouds beyond the polar regions are similar to that along the pole. Assuming a one-dimensional mass flow, the mass loss rate from the disk is approximately $\\dot{M}_w\\propto v_w [\\Sigma_g \\pi r_d^2]$, where $\\Sigma_g$ is the gas column density and $r_d$ is the scale length of the disk. We note that in the prescription of MMW98, one has $v_c \\propto r_d$. We therefore have, $\\dot{M}_w \\propto v_c^{2-0.25} \\dot{M}_\\ast ^{0.4} \\Sigma_g$, where we have used eqn \\ref{eq:fit}, after multiplying both sides by $v_c$. The ratio of mass outflow rate to the SFR is therefore $\\dot{M}_w/\\dot{M}_\\ast \\propto v_c^{1.75} \\Sigma_g \\dot{M}_\\ast^{-1.4}$. Using Kennicutt's law of star formation, which gives $\\dot{M}_\\ast \\propto \\Sigma_g^{1.4} r_d^2 \\propto \\Sigma_g ^{1.4} v_c^2$, we have finally, $ {\\dot{M}_w \\over \\dot{M}_\\ast} \\propto v_c^{1.05} \\Sigma_g^{-0.96} $. We can therefore conclude that roughly, \\begin{equation} {\\dot{M}_w \\over \\dot{M}_\\ast} \\propto v_c^{-1} \\Sigma_g ^{-1} \\,. \\end{equation} Interestingly, similar power law dependence has also been found in simulations \\citep{hopkins11}. We note that our results assumed a value of $v_h\\sim 800$ km s$^{-1}$, and a column density of cold clouds of $\\sim 10^{21}$ cm$^{-2}$. If we assume a larger value of $v_h$ ($\\sim 1000$ km s$^{-1}$), then the contour for only ram pressure will be able to explain the winds in ULIGs with large SFR and high mass. A similar result will follow from larger values of $\\kappa$ for the radiation pressure case. % It is interesting to note that the contour for only radiation pressure can explain the ULIG region of Figure \\ref{figcie} (top right corner). Extending to larger SFR, our results indicate that radiation pressure will also be important for HLIGs (Hyperluminous Infra-red galaxies) \\citep{row00}. Lastly, although it may appear that the role of radiation pressure in galaxies other than ULIGs is less dominant than ram pressure as far as energetics is concerned, radiation pressure may still play an important role in lifting the clouds to a large height before it is embedded in the hot wind to help it survive long \\citep{nath08, mur11}." }, "1112/1112.1168_arXiv.txt": { "abstract": " ", "introduction": "The Australian SKA Pathfinder (ASKAP) \\citep{Johnston08, Deboer09} is a new radio telescope being built on the Australian candidate Square Kilometre Array (SKA) site in Western Australia. ASKAP will consist of 36 12-metre antennas spread over a region 6 km in diameter. Although the array of antennas is no larger than many existing radio telescopes, each antenna will be equipped with a phased-array feed of 96 dual-polarisation pixels, giving it a 30 deg$^2$ field of view and a very fast survey speed. The Evolutionary Map of the Universe (EMU) project \\citep{Norris11} is a wide-field radio continuum survey planned for ASKAP. The primary goal of EMU is to make a deep (rms $\\sim$ 10$\\mu$Jy/bm) radio continuum survey of the entire Southern Sky at 1.3 GHz, extending as far North as $+30^{\\circ}$ declination, with a 10 arcsec resolution. EMU is expected to detect and catalogue about 70 million galaxies, including typical star-forming galaxies up to z=1, powerful starbursts to even greater redshifts, and AGNs to the edge of the visible Universe. The amount of data involved with ASKAP ($\\sim$2.5 GB/s, or 100 PB/year) requires that the source detection and measurement is fast, robust and highly automated. Source detection and measurement is a problem common to all astronomical imaging surveys and projects, and numerous software tools have been developed in order perform this initial step in the analysis of imaging data. With the advent of large area surveys the automation of this process is clearly crucial. This has led to a variety of survey-specific source-finders being developed, each optimised to address the specific issues associated with the imaging technology of each survey, and its corresponding image properties and artifacts. What has not developed in parallel is an analysis of the common steps in source-identification and measurement in order to assess the optimal approaches or algorithms that should be used to maximise the robustness and scientific utility of the resulting source catalogues. This is partially a consequence of images obtained using different telescope or imaging technologies (UV/optical/NIR/FIR, compared to radio, or X-ray, or gamma-ray) having very different characteristics. Consequently, the assumptions being used in source-finders developed for images at one wavelength or technology are not typically applicable for others. The analysis presented here looks at the first steps taken in the source-finding process, those of background estimation and thresholding. This is done explicitly in the context of radio interferometric imaging, although the expectation is that the conclusions should be more broadly applicable. Furthermore, our investigation has a focus on identifying an optimal approach for the source-identification and measurement to be implemented in the ASKAP image analysis software pipeline. This analysis complements that of Hancock et al.\\ 2011 (in prep.), which presents a detailed exploration of existing source-finding tools, their underlying algorithms, and how they perform on simulated radio interferometer images. Together, these analyses comprise the first stage of a thorough investigation of each stage of the source-identification and measurement process that is being pursued as part of the Design Study for EMU. Several radio source identification and measurement tools are in common use. These include the {\\sc Miriad}/AIPS Gaussian fitting routines IMSAD, SAD and VSAD, {\\sc sfind} \\citep{hopkins2002}, and Duchamp (Whiting 2008) as well as SExtractor \\citep{bertin1996}. There are also a variety of survey-specific tools, such as HAPPY (a modified version of SAD used in the FIRST survey, \\citealp{white1997}), a machine-learning back-end to VSAD used to construct the SUMSS catalogue \\citep{mauch2003}, BDSM (used for the LOFAR source-finding, N. Mohan, in prep), and the floodfill algorithm being used in the Australia Telescope Large Area Survey (ATLAS) Data Release 2 (\\citealp{murphy2007}; Hales et al.\\ 2011 in prep.). As standalone tools, none of these is adequate for EMU, due to limitations evident in the treatment of background estimation, approaches to treating a varying noise level across an image, and importantly in the numbers of artifacts incorrectly identified as sources. Here we aim to test the approaches to both background estimation and thresholding, in order to identify an optimal approach for the EMU survey. All existing tools have implemented the complete sequence of steps, from background and noise level estimation, through threshold setting, and ``source-pixel\" identification, to source-measurement. It is thus challenging to extract robust information about each independent step in the source-identification process. We do this here for a selection of tools through a judicious choice of parameters in the tasks we investigate, and interpret our results cautiously as a consequence. We emphasise that we are focussing here on two-dimensional data (radio continuum images), with our results expected to be applicable generically to two-dimensional image source-identification. We present details of the simulated data used in our analysis in \\S\\,\\ref{simdata}, and the algorithms being tested in \\S\\,\\ref{algorithms}. The background estimation and local rms noise estimates are discussed in \\S\\,\\ref{bkg}, with the reliability and completeness statistics being used as a metric to compare the different approaches in \\S\\,\\ref{relcom}. Our results are summarised in \\S\\,\\ref{conc}. ", "conclusions": "\\label{conc} We have tested SExtractor, Selavy and {\\sc sfind} to explore the effects of background and noise estimation, along with two approaches to thresholding, a simple $n\\sigma$ level compared to the false-discovery rate method, on source extraction. The tests were performed on two sets of simulations, the ASKAP simulations which are based on SKADS input source catalogue and include instrumental artefacts, and the Hancock et al.\\ simulation which has only Gaussian noise. The Hancock et al.\\ simulation is an idealised case that is useful for testing the algorithms in ``perfect\" conditions. The first step in source extraction is background subtraction and noise estimation. We have confirmed the results from the Planck team (Planck Collaboration 2011) and previous deep continuum radio surveys \\citep{huynh2005,schinnerer2007,schinnerer2010} that mesh-sizes of 10 to 20 PSFs or beamwidths produce satisfactory background and noise images. We find that SExtractor background mesh-sizes of 10 and 20 beamwidths produce similar results in terms of reliability. A visual inspection shows the background and noise images are still affected by local bright sources for a mesh-size of 10 beamwidths, but combining a mesh-size of 10 beamwidths with smoothing of 3 meshes produces the highest quality background and noise images. The reliability of the catalogues resulting from 10 and 20 beamwidth mesh-sizes, however, does not differ significantly. The fact that the background estimation step, so crucial in all the subsequent stages of source-identification and measurement, still needs to be manually tuned in most existing source-detection software, is a major concern. The optimum mesh-size for background subtraction and noise estimation is likely to be image specific, and to vary perhaps substantially depending on the distribution of sources within the image. Developing an automated process for setting the mesh-size when implementing the background and noise properties is clearly a priority, and will need to be developed as part of an automated pipeline for radio telescopes of the future such as the Australian Square Kilometre Array Pathfinder (ASKAP) and the Square Kilometre Array (SKA). One possible method could be to use a tree-based approach, identifying rms noise levels for the whole image and for progressively smaller regions, so that each pixel can be associated with a ``tree\" of rms noise values on each scale. Identifying the scale for which the rms noise plateaus for each pixel could be a suitable approach, and will be tested as part of the EMU Design Study. The thresholding comparison was limited for the ASKAP simulation due to a ceiling in the completeness values, resulting from bright input sources with small separations that were not deblended well by any of the algorithms, and by the low peak flux density values for extended sources. Nevertheless, in this simulation we find that {\\sc sfind} with $\\alpha$ values of 1 to 5 results in similar completeness to SExtractor run with thresholds of 4 to 5$\\sigma$. The reliability of the {\\sc sfind} sources is higher however, $\\sim$92\\% for {\\sc sfind} compared to 80 -- 88\\% for the SExtractor. Selavy results in higher completeness than SExtractor or {\\sc sfind} for the ASKAP simulations, but at the cost of lower reliability. In regions with significant artefacts such as sidelobes from bright sources {\\sc sfind} and Selavy perform much better than SExtractor in rejecting spurious detections. In the case of the Hancock et al.\\ simulation, where noise is Gaussian and the sources more well-separated, SExtractor is 98\\% reliable with thresholds of $4\\,\\sigma$ or greater. For the Hancock et al.\\ simulation we find that $n\\,\\sigma$ threshold approach of SExtractor and the FDR approach of {\\sc sfind} perform similarly, although the FDR thresholding of {\\sc sfind} seems to give somewhat better reliability at thresholds that produce comparable levels of completeness. In this idealised simulation SExtractor and Selavy gives similar completeness but the Selavy sources are 15\\% to 30\\% less reliable. There is a trade-off between completeness and reliability in source extraction algorithms: parameters which give high completeness result in lower reliability. Overall, the false-discovery rate method, as tested with {\\sc sfind}, results in more reliable sources than SExtractor or Selavy, for parameters that give similar completeness levels. While more fine-tuning of Selavy, the prototype source finder for EMU, is required, our analysis suggests that the FDR approach is worthwhile pursuing and we recommend this be implemented in Selavy. This analysis demonstrates that existing approaches to background and noise estimation seem to be limited not by the specific algorithms but rather the requirement to select appropriate mesh-sizes over which to calculate a ``local\" background and noise estimate. Future work will require development of an automated background mesh-size estimation process for the ASKAP software pipeline, in particular for the EMU images. In addition, a complementary analysis of the source-fitting and parameter measurement approaches is also underway (Hancock et al.\\ in prep.) which will establish the optimum approaches to these latter stages of source extraction." }, "1112/1112.0028_arXiv.txt": { "abstract": "Recently published {\\it Spitzer Space Telescope} observations of the classical Cepheid archetype $\\delta$~Cephei revealed an extended dusty nebula surrounding this star and its hot companion HD~213307. At far infrared wavelengths, the emission resembles a bow shock aligned with the direction of space motion of the star, indicating that $\\delta$~Cephei is undergoing mass-loss through a stellar wind. Here we report \\HI\\ 21-cm line observations with the Very Large Array (VLA) to search for neutral atomic hydrogen associated with this wind. Our VLA data reveal a spatially extended \\HI\\ nebula ($\\sim13'$ or 1~pc across) surrounding the position of \\DC. The nebula has a head-tail morphology, consistent with circumstellar ejecta shaped by the interaction between a stellar wind and the interstellar medium (ISM). We directly measure a mass of circumstellar atomic hydrogen $M_{\\rm HI}\\approx0.07~M_{\\odot}$, although the total \\HI\\ mass may be larger, depending on the fraction of circumstellar material that is hidden by Galactic contamination within our band or that is present on angular scales too large to be detected by the VLA. It appears that the bulk of the circumstellar gas has originated directly from the star, although it may be augmented by material swept from the surrounding ISM. The \\HI\\ data are consistent with a stellar wind with an outflow velocity $V_{\\rm o}=35.6\\pm1.2$~\\kms\\ and a mass-loss rate of ${\\dot M}\\approx(1.0\\pm0.8)\\times 10^{-6}~M_{\\odot}$ yr$^{-1}$. We have computed theoretical evolutionary tracks that include mass loss across the instability strip and show that a mass-loss rate of this magnitude, sustained over the preceding Cepheid lifetime of \\DC, could be sufficient to resolve a significant fraction of the discrepancy between the pulsation and evolutionary masses for this star. ", "introduction": "} Owing to the tight coupling between the pulsation properties of Cepheid variables and their fundamental stellar parameters, these stars have long provided crucial tests for stellar evolution models (Gautschy \\& Saio 1996) and play a key role in the calibration of the extragalactic distance scale (Feast \\& Walker 1987; Freedman et al. 2001). Nonetheless, for more than 40 years, a puzzle has persisted concerning the lingering discrepancy between stellar masses derived for Cepheids using different methods (e.g., Cox 1980). In spite of improved radiative opacity calculations (Iglesias et al. 1990; Seaton et al. 1994), advances in evolutionary modeling (Bono et al. 2002), and the consideration of metallicity effects (Keller \\& Wood 2006), masses derived from stellar evolutionary models are found to be systematically $\\sim$10-15\\% higher than those derived from stellar pulsation models (Caputo et al. 2005) or orbital dynamics, when available (Evans 2009). As described by Neilson et al. (2011), one of the most promising solutions to the so-called ``Cepheid mass discrepancy'' is likely to be convective core overshooting (Chiosi et al. 1992) coupled with {\\em mass-loss} during the Cepheid phase of evolution (see also Bono et al. 2006). Mass-loss from Cepheids has long been postulated, not only as a means of resolving the Cepheid mass discrepancy, but as a natural consequence of stellar evolution and pulsation models (e.g., Iben 1974; Willson \\& Bowen 1984; Neilson \\& Lester 2008). The observational confirmation of mass loss from Cepheids would impact not only our understanding of Cepheid evolution, but would also have implications for the use of Cepheids as distance indicators. For example, mass-loss can non-negligibly affect the structure and scatter in the infrared period-luminosity relation (Neilson et al. 2009), and the presence of circumstellar material can impact Cepheid distance determinations made using the interferometric Baade-Wesselink method (M\\'erand et al. 2007). A number of previous authors have attempted to identify direct evidence of past or ongoing mass-loss from Cepheids using observations ranging from ultraviolet to radio wavelengths (e.g., Deasy \\& Butler 1986; Deasy 1988; Welch \\& Duric 1988; Bohm-Vitense \\& Love 1994; M\\'erand et al. 2007; Kervella et al. 2006, 2009; Neilson et al. 2009). However, the results have been inconclusive, as the mass-loss rates (or upper limits) derived from these studies (${\\dot M}\\approx 10^{-12}$ to $10^{-5}~M_{\\odot}$~yr$^{-1}$) span many orders of magnitude and are typically rather uncertain owing to their reliance on trace atomic and molecular species (which may not be optimal tracers of past and ongoing mass-loss), their inability to sample very extended spatial scales, and finally, the strong dependencies of derived mass-loss rates on a variety of underlying assumptions (e.g., wind structure, gas-to-dust ratios, ionization fractions). With the goal of obtaining new empirical constraints on Cepheid mass-loss, Marengo et al. (2010a) recently used the {\\it Spitzer Space Telescope} to survey a sample of 29 nearby Cepheids. While IR excesses were not directly detected, ruling out the presence of a large amount of warm ($\\sim$500~K) dust in close proximity to the stars, the {\\it Spitzer} images revealed extended emission around several targets (Marengo et al. 2010a; Barmby et al. 2011). Perhaps the most intriguing was the discovery of a prominent nebula around the archetype Cepheid variable, \\DC\\ (Marengo et al. 2010b). Some basic stellar properties of \\DC\\ are summarized in Table~1. \\DC\\ is part of a wide binary (Benedict et al. 2002), with a hot companion HD~213307 ($T_{\\rm eff}=$8800~K; Cenarro et al. 2007) separated by a projected distance of $40''$. The IR emission surrounding \\DC\\ (and HD~213307) is visible in multiple {\\it Spitzer} bands and shows a roughly parabolic structure with an extent of $\\sim5'$ ($\\approx2\\times10^{4}$~AU). The symmetry axis of the parabola is aligned with the direction of the star's motion through the interstellar medium (ISM; Marengo et al. 2010b), where $V_{\\rm space}\\approx 10.3$~\\kms\\ and P.A.=\\ad{58}{3}.\\footnote{For the present work, we have recomputed the space motion vector for \\DC\\ using the updated solar constants from Sch\\\"onrich et al. 2010.} The IR emission therefore appears to trace a bow shock structure, as arises when a moving, mass-losing star interacts with the local ISM (e.g., Wilkin 1996). Bow shocks have been seen previously in FIR images of a number of mass-losing stars, including the supergiants $\\alpha$~Cam (Van Buren \\& McCray 1988) and $\\alpha$~Ori (Noriega-Crespo et al. 1997; Ueta et al. 2008) and the asymptotic giant branch (AGB) stars R~Hya (Ueta et al. 2006; Wareing et al. 2006), Mira (Ueta 2008), R~Cas (Ueta et al. 2010), IRC+10216 (Ladjal et al. 2010), TX~Psc, and X~Her (Jorissen et al. 2011). The emission mechanism responsible for the FIR emission from these circumstellar bow shocks is uncertain, but is likely to be mainly thermal emission from dust, with possible contributions from low-excitation atomic emission lines (Ueta et al. 2008; Marengo et al. 2010b). While the companion of \\DC, HD~213307, cannot strictly be excluded as the source of the IR emitting material seen by {\\it Spitzer}, this appears quite unlikely based on the B7-B8~III-V spectral type of this star (see Marengo et al. 2010b for discussion). Thus the {\\it Spitzer} observations of Marengo et al. (2010b) provide strong evidence that \\DC\\ is undergoing mass loss through a stellar wind. To further characterize the nature of this wind and its interaction with the ISM, we have now used the Very Large Array (VLA) to search for \\HI\\ 21-cm line emission in the circumstellar environment of \\DC. \\HI\\ observations have been used previously to trace extended circumstellar emission surrounding a number of mass-losing AGB stars (e.g., G\\'erard \\& Le~Bertre 2006; Matthews \\& Reid 2007; Libert et al. 2008; Matthews et al. 2008, 2011). Frequently, the \\HI\\ emission associated with AGB stars is highly extended (up to $\\sim$1~pc) and shows signatures of interaction with the ISM in the form of trailing \\HI\\ wakes, velocity gradients caused by ram pressure effects, and/or density enhancements that demarcate the interstellar-circumstellar interaction zone. Furthermore, several of the evolved stars detected in \\HI\\ are among those with FIR (and/or FUV) emitting bow shocks, including Mira, R~Cas, and IRC+10216, underscoring the complementarity of these tracers for probing the chemistry and kinematics of stellar outflows and their interaction with their environments. This paper represents the first extension of this approach to the study of Cepheids. \\begin{deluxetable}{lcc} \\tabletypesize{\\scriptsize} \\tablewidth{0pc} \\tablenum{1} \\tablecaption{Coordinates and Stellar Properties of $\\delta$~Cephei} \\tablehead{\\colhead{Parameter} & \\colhead{Value} & \\colhead{Ref.}} \\startdata $\\alpha$ (J2000.0) & 22 29 10.2 & 1\\\\ $\\delta$ (J2000.0) & +58 24 54.7 & 1 \\\\ $l$ & \\ad{105}{19} & 1 \\\\ $b$ & \\ad{+0}{53} & 1\\\\ Distance (pc) & 273$\\pm$0.011 & 2\\\\ Spectral Type & F5Ib-G1Ib & 3\\\\ Pulsation period (days) & 5.366341 & 3\\\\ Mean $T_{\\rm eff}$ (K) & 5910 & 4\\\\ Mass (pulsation)$^{b}$ ($M_{\\odot}$) & 4.5$\\pm$0.3 & 5\\\\ Mass (evolutionary)$^{c}$ ($M_{\\odot}$) & 5.7$\\pm$0.5 & 5\\\\ Mean $M_{V}$ & $-$3.47$\\pm$0.10 & 2 \\\\ Mean Luminosity ($L_{\\odot}$)& $\\sim$2000 & ...\\\\ Mean radius ($R_{\\odot}$) & 44.5$^{a}$ & 6\\\\ $V_{\\rm LSR}^{d}$ & $-4.7$~\\kms\\ & 7 \\\\ \\enddata \\tablenotetext{a}{Based on the mean, limb darkened angular diameter $\\phi_{\\rm LD}$=1.520~mas and the distance adopted here.} \\tablenotetext{b}{Based on $V-K$ colors.} \\tablenotetext{c}{Assuming a canonical model with no overshoot; see \\S~\\ref{resolution}).} \\tablenotetext{d}{Derived from the heliocentric radial velocity $V_{\\rm h}=-16.8$~\\kms.} \\tablecomments{Units of right ascension are hours, minutes, and seconds, and units of declination are degrees, arcminutes, and arcseconds. All quantities have been scaled to the distance adopted in this paper.} \\tablerefs{(1) SIMBAD database; (2) Benedict et al. 2002; (3) Samus et al. 2011; (4) Andrievsky et al. 2005; (5) Caputo et al. 2005; (6) Armstrong et al. 2001; (7) Wilson 1953 } \\end{deluxetable} ", "conclusions": "We have used the VLA to search for \\HI\\ 21-cm line emission in the circumstellar environment of the archetype of Cepheid variables, \\DC. We have detected an extended ($13'$, or $\\sim$1~pc) nebula at the position of the star. The nebula exhibits a head-tail morphology, with the head of the structure aligning closely with the infrared-emitting nebula and bow shock previously detected by Marengo et al. (2010b), while the tail appears to be a turbulent structure that trails the motion of the star through the ISM. We measure an \\HI\\ mass for the nebula of $M_{\\rm HI}\\approx0.07~M_{\\odot}$, although its total \\HI\\ mass could be $\\sim$2-3 times larger, depending on the fraction of emission that is hidden by the strong Galactic emission that contaminates a portion of our observing band. Additional material may also be present on large angular scales ($\\gsim15'$) to which the VLA is insensitive. We interpret the bulk of the \\HI\\ nebula surrounding \\DC\\ as arising from a stellar wind with a mass-loss rate of ${\\dot M}\\approx(1.0\\pm0.8)\\times10^{-6}~M_{\\odot}$~yr$^{-1}$ and an outflow velocity $V_{\\rm o}\\approx35$~\\kms. By computing evolutionary models that include mass loss across the instability strip, we show that a mass-loss rate of this magnitude, sustained over the preceding Cepheid lifetime of \\DC, could be sufficient to resolve a significant fraction of the longstanding discrepancy between the masses of the star derived from stellar pulsation versus stellar evolutionary models." }, "1112/1112.2507_arXiv.txt": { "abstract": "The newly discovered Be/X-ray binary in the Small Magellanic Cloud, \\src, provides the first example of a robust association with a supernova remnant (SNR). The short age estimated for the SNR qualifies \\src\\ as the youngest known source in its class, $\\tau\\approx 10^4\\ \\textrm{yr}$. As such, it allows to test current models of magneto-rotational evolution of neutron stars in a still unexplored regime. Here we discuss possible evolutionary scenarios for \\src\\ in the attempt to reconcile its long spin period, $P=1062\\ \\textrm{s}$, and short age. Although several options can be considered, like an anomalously long initial period or the presence of a fossil disc, our results indicate that \\src\\ may host a neutron star born with a large initial magnetic field, typically in excess of $\\sim 10^{14}\\, {\\rm G}$, which then decayed to $\\sim 10^{13}\\, {\\rm G}$. ", "introduction": "\\label{intro} Be/X-ray binaries (or BeXBs for short) form a subclass of the high-mass X-ray binaries (HMXBs) in which the neutron star (NS) companion is a Be star, a spectral class B giant/subgiant with emission lines and large IR flux. These peculiar properties are explained in terms of an equatorial disc, formed by matter lost by the rapidly rotating Be star. X-ray emission is believed to be powered by accretion of material in the equatorial disc onto the NS \\cite[see e.g.][for a recent review]{reig11}. BeXBs are both transient and persistent X-ray sources. Transient systems are characterized by type I-II outbursts during which their flux increases by a factor 10--$10^4$ over the quiescent level. They typically contain a not-too-slow NS ($P\\la 100\\, {\\rm s}$) on a moderately eccentric orbit, $\\po\\la 100\\, {\\rm d}$, $e\\ga 0.2$. On the other hand, persistent BeXBs exhibit a rather flat lightcurve, lower X-ray luminosity ($L\\approx 10^{34}$--$10^{35}\\, \\lum$), longer spin and orbital periods \\cite[$P\\ga 200\\, {\\rm s}$, $\\po\\ga 200\\, {\\rm d}$; see again][]{reig11}. There are presently about 30 well-established BeXBs in the Galaxy, plus $\\sim 40$ candidates. In addition, $\\sim 50$ sources (plus $\\sim 20$ candidates) are known in the Small Magellanic Cloud (SMC)\\footnote{Figures from \\cite{reig11} and the BeXB online catalogue at http://xray.sai.msu.ru/\\textasciitilde raguzova/BeXcat; see also \\cite{ragpo05}.}. Very recently \\cite{henbru11} reported the discovery of a new BeXB in the Wing of the SMC. The new source (\\src) has the typical properties of a persistent BeXB: a B0-0.5(III)e+ donor, $L\\sim 6\\times 10^{35}\\, \\lum$, $P\\sim 1062\\, {\\rm s}$, and an orbital period $\\po\\sim 300\\, {\\rm d}$, as estimated from the Corbet diagram \\cite[][]{corb09}. What makes \\src\\ to stand out amongst its kinship is its (likely) association with a supernova remnant (SNR). BeXB-SNR associations have been already reported (again in the SMC) but in all previous cases they appear uncertain \\cite[][]{hugsm94,coe00}. In the case of \\src\\ the association looks robust and allows for the first time to estimate the NS age in a BeXB from that of the parent SNR, $\\tau\\sim 2$--$4\\times 10^4\\, {\\rm yr}$ \\cite[][]{henbru11}. The suggested association of \\src\\ with a SNR has been further strengthened by a reanalysis of the same \\xmm\\ datasets, supplemented with optical and radio observations, by \\cite{hab11}. Their estimate for the SNR age, $1.6\\times 10^4$ yr, is even shorter than, albeit fully compatible with, that of \\cite{henbru11}. \\cite{hab11} were also able to measure the source period evolution, which results in a positive (i.e. spin-down) rate $\\dot P\\sim 3\\times 10^{-6}\\, {\\rm s\\,s}^{-1}\\sim 95\\, {\\rm s\\,yr}^{-1}$. The long spin periods ($P \\ga 1000\\, {\\rm s}$) of some persistent BeXBs have been for a long time a major issue. According to the standard picture, there are four stages in the spin evolution of a neutron star embedded in a medium: ejector, propeller, accretor and georotator\\footnote{The last one is of no concern here, since it occurs under specific conditions, hardly met in HMXBs.} \\cite[e.g.][]{lip92}. Once the NS entered the accretor stage after a short propeller phase, its spin period quickly settles at an equilibrium value, $P_{eq}$. In the conventional model, based on \\cite{davpri81} results, the star dipole field should be $B\\ga 10^{14}\\, {\\rm G}$ to have $P_{eq}>10^3 \\, {\\rm s}$, unless the accretion rate is orders of magnitude below what needed to account for the observed X-ray luminosity. Since observations support the presence of a NS with standard magnetic field (at least in some BeXBs), the subsonic propeller stage, which can delay the onset of accretion until a much longer period is reached \\cite[][]{iks07}, has been invoked to explain ultra-slow BeXBs \\cite[see][]{reig11}. More recent investigations of wind-fed accretion onto magnetized NSs indicate, however, that the equilibrium period can be as high as $\\sim 1000\\, {\\rm s}$ even for $B\\approx 10^{12}$--$10^{13}\\, {\\rm G}$ and $\\dot M\\approx 10^{16}\\, {\\rm g\\,s}^{-1}$, as expected in BeXBs \\cite[][]{shak11}. Whatever the details of the braking torques, the previous argument implicitly relies on the assumption that the present age of the source is long enough for the NS to have entered the propeller stage. Unless the accretion rate is way above typical, a NS in a BeXB with $B\\ga 10^{12}\\, {\\rm G}$ starts its evolution in the ejector (or pulsar) phase. Its duration can be roughly estimated as $\\tau_{ej}\\ga 10^6 (B/10^{12}\\, {\\rm G})^{-1}(\\dot M/10^{15}\\, {\\rm g\\,s}^{-1})^{-1/2}\\, {\\rm yr}$ (see Section~\\ref{spindown}). This is comfortably below (by a factor $\\approx 10$) the lifetime of the Be companion, so there is ample room for the binary to start an accretion-powered X-ray stage. In the case of \\src, however, it would be impossible for the NS to enter the propeller stage (and hence to become an accretor) in a time as short as a few $\\times 10^4\\, {\\rm yr}$, the estimated SNR age, for typical values of $B$ and $\\dot M$. The accretion rate in \\src\\ is $\\dot M=L/\\eta c^2\\sim 6\\times 10^{15} \\, {\\rm g\\,s}^{-1}$ for an efficiency $\\eta=0.1$, so this points to a highly magnetized NS, with an initial magnetic field substantially above $10^{12}\\, {\\rm G}$. ", "conclusions": "The quite short age implied by the association of the newly discovered BeXB in the SMC \\src\\ with a SNR \\cite[][]{henbru11, hab11} raises a number of questions about the properties of the neutron star and its evolutionary status. Normally, one expects that accreting X-ray pulsars spin close to their equilibrium period. However, for a such a young system this appears far from granted. \\cite{hab11} reported a large spin-down rate for \\src, which may suggest that $P_{eq}$ has not been reached yet. According to the standard evolutionary scenario \\cite[][]{lip92}, the (maximum) spin-down rate in the accretor stage is $\\dot P\\sim 2\\pi B^2R_{NS}^6/(GMI)$ which implies $B\\approx 3\\times 10^{14}\\, {\\rm G}$ for $\\dot P\\approx 100\\, {\\rm s\\,yr}^{-1}$. On the other hand, if Eq. (\\ref{shaktorque}) is used to estimate the magnetic field, a much lower value is obtained, $B\\approx 10^{13}\\, {\\rm G}$, very close to what is predicted assuming that the source spins at the equilibrium period (see Eq. [\\ref{bshak}]). This supports a picture in which the NS actually rotates close to $P_{eq}$. A further argument in favor of this is the very short duration of the spin-down phase in the accretor stage, $P/\\dot P\\la 100\\, {\\rm yr}$, which makes it very unlikely to catch the source in this state. Our conclusion is that both the young age and the large spin-down rate of \\src\\ argue in favor of an initially highly magnetic NS which experienced field decay. Details of evolution past the ejector stage are uncertain, so a fine tuning of the model discussed in Sec. \\ref{spindown} is entirely premature. Still, some general considerations can be made. The main one is that the timescale required for reaching $P_{eq}$ after the NS left the ejector stage is very short. A further point to be stressed is that the zero, or negative, period derivative at $P=P_{eq}$ (see Figure \\ref{perplot}) does not account for variations in $\\dot M$ due to orbital motion in the BeXB and irregularities in the wind. This can result in quite large values of $\\dot P$ and rapid changes in the period derivative which are indeed measured in wind-fed X-ray binaries \\cite[see e.g.][]{bild97}. In this respect the spin-down rate in \\src\\ is large but not exceptionally so. For example, GX~301-2 is known to have a larger value of $\\dot P$ \\cite[][]{koh97}. \\cite{hab11} measured the average period derivative over a time-span less than a month, $\\sim$ one tenth of the orbital period. Fluctuations in $\\dot P$ can occur on a timescale of days, so a more complete dataset is definitely needed to assess the real nature of the period variation. Alternative scenarios to explain the long period and short age of \\src\\ can be envisaged. For instance, \\cite{hab11} suggested that the NS could have been born with an initial period much longer than $\\sim 0.01\\, {\\rm s}$. The value of $P_0$ can be evaluated by requiring that the transition period given by Eq. (\\ref{pejec}) is reached in less than the source age, and it turns out to be $\\sim 1\\, {\\rm s}$ for the $B$-field corresponding to $P_{eq}$ [Eq. (\\ref{bshak})]. If this is the case no field decay is required. Although possible, no compelling observational evidence for such long initial periods in NSs exists. There are hints that some central compact objects (CCOs) in SNRs may have a current period very close to the initial one, $P_0\\approx 0.1\\, {\\rm s}$. These sources, however, are suspected to host weakly magnetized NSs \\cite[$B\\approx 10^{11}\\, {\\rm G}$; e.g.][]{gotthalp10}. Whether the low field and the long period are related, depending on the conditions under which the NS formed, is still unclear. Still, the estimated value of $B$ in \\src\\ is much higher and the initial period required in the present case is more than one order of magnitude longer. Very recently \\cite{knig11} presented evidence that the population of BeBXs is bimodal, with the two sub-populations having distinct typical spin and orbital periods, and eccentricities. They suggest that the NSs in the two groups are produced through different channels, iron-core-collapse and electron capture supernovae, with the long spin period sources associated to the former channel. Whether this may result in a population of long $P_0$ NSs is a still open question. Another possibility is that the NS in \\src\\ could, at least in the early stages following its formation, have been surrounded by a debris (or fossil) disc, left by the parent supernova explosion. The existence of such a disc could also lead to rapid spin-down and large period. This, in addition to a super-strong field, was suggested to explain the ultra-long period ($\\sim 6.67\\, {\\rm hr}$) in the enigmatic source RCW~103 \\cite[][]{deluca06,li07}. Although this remains a possibility worth of further investigations, preliminary calculations, based on the model by \\citet[][see also \\citealt{esp11}]{li07}, indicate that a large initial field ($B\\ga 10^{14}\\ {\\rm G}$) is still required, unless the disc is quite massive. For $M_{disc}(0)\\sim 10^{-2}\\ M_\\odot$, the NS can enter the propeller stage earlier than $\\sim 10^4\\ {\\rm yr}$ also for $B\\sim 10^{13}\\ {\\rm G}$. If \\src\\ indeed contains an initially strongly magnetized neutron star, then studies of this system can shed light on the origin of magnetars. In the standard scenario \\cite[][]{dt92} super-strong fields are generated via dynamo processes. This requires rapid rotation of the proto neutron star. Primary components in high-mass binary systems are not expected to form rapidly rotating cores at the end of their lives \\cite[][]{pp06, bp09}. Up to now no strongly magnetized compact objects have been identified in binary systems with certainty. \\cite{chashkina11} have recently derived estimates of the $B$-field in HMXBs using \\cite{shak11} formula and no evidence of ultra-high fields was found. It has been suggested that supergiant fast X-ray transients \\cite[SFXTs, see e.g.][for a review]{sid11} may host a magnetar \\cite[][]{boz08}. More recently, \\cite{tor11} reported magnetar-like behaviour from the peculiar binary LS I +61 303. However, no definite confirmation has been given yet, and, in this respect, \\src\\ may be a unique example. The existence of a (evolved) magnetar in a high-mass binary system will pose new challenges on the origin of such neutron stars." }, "1112/1112.5111_arXiv.txt": { "abstract": "A complete sample of bright {\\it Swift} Gamma--ray Bursts (GRBs) has been recently selected by Salvaterra et al. (2011). The sample has a high level of completeness in redshift ({\\bf 90}$\\%$). We derive here the intrinsic absorbing X--ray column densities of these GRBs making use of the {\\it Swift} X--ray Telescope data. This distribution has a mean value of $\\log(N_H/{\\rm cm^{-2}})=21.7\\pm0.5$. This value is consistent with the distribution of the column densities derived from the total sample of GRBs with redshift. We find a mild increase of the intrinsic column density with redshift. This can be interpreted as due to the contribution of intervening systems along the line of sight. Making use of the spectral index connecting optical and X--ray fluxes at 11 hr ($\\beta_{OX}$), we investigate the relation of the intrinsic column density and the GRB `darkness'. We find that there is a very tight correlation between dark GRBs and high X--ray column densities. This clearly indicates that the dark GRBs are formed in a metal-rich environment where dust must be present. ", "introduction": "Long duration Gamma--ray bursts (GRBs) are thought to originate from the collapse of massive stars. Several lines of evidence points toward this conclusion, ranging from the association to type Ib/c supernovae (Woosley \\& Bloom 2006 and references therein), to the occurrence of GRBs in the most luminous part of their host galaxies (Svensson et al. 2010). The ambient medium in which GRBs explode is expected to be denser than the interstellar medium and typical of star forming regions. The values of the absorbing column densities as measured in X--rays are high (Galama \\& Wijers 2001; Stratta et al. 2004; Campana et al. 2006; Watson et al. 2007). An analysis of the intrinsic column densities of all {\\it Swift} GRBs observed within 1,000 s and with a known redshift has been carried out by Campana et al. (2010). The selected sample consisting of 93 GRBs was biased. The distribution of the intrinsic X--ray absorption column density is consistent with a lognormal distribution with mean $\\log N_H(z)=21.9\\pm0.1$ ($90\\%$ confidence level). This distribution is in agreement with the expected distribution for GRBs occurring randomly in giant molecular clouds similar to those within the Milky Way (Campana et al. 2006; Reichart \\& Price 2002). Looking at the distribution of X--ray column densities vs. redshift, there is a lack of non-absorbed GRBs at high redshift and a lack of heavily absorbed GRBs at low redshift. This might be the outcome of biases present in the sample. Looking at the distribution of X--ray column densities versus redshift a lack of non-absorbed GRBs at high redshift and a lack of heavily absorbed GRBs at low redshift were found in previous studies (Campana et al. 2010). The former might be explained in terms of more compact and dense star formation regions in the young Universe (or to a sizable contribution from intervening systems). The latter might be interpreted as due to a change in the dust properties with redshift, with GRBs at redshift $z\\lsim 2-3$ having a higher dust to gas ratio for the same X--ray column density (e.g. different grain size or composition). This will naturally provide a lack of heavily (X--ray) absorbed GRBs at small redshifts. In the optical the presence of a large amount of absorbing material is much less clear, since a large number of GRB afterglows are not affected by absorption (Kann, Klose \\& Zeh 2006; Schady et al. 2010; Zafar et al. 2011; but see Greiner et al. 2011 and Covino et al. 2011, in preparation). In this respect photoionisation of dust grains can play an important role (Lazzati, Perna \\& Ghisellini 2001; Lazzati, Covino \\& Ghisellini; Draine \\& Hao 2002; Campana et al. 2007). Moreover, the absorbing column densities measured in the optical based on damped Lyman-$\\alpha$ absorption are a factor of $\\sim 10$ lower than those measured in the X--ray band (Campana et al. 2010; Fynbo et al. 2009). This has been interpreted as due to photoionization of the surrounding medium by GRB photons (Campana et al. 2006, 2007; Watson et al. 2007; Campana et al. 2010; Schady et al. 2011). The presence of a large amount of material is also testified by the existence of `dark' GRBs. There are several definitions of dark GRBs. The easiest is that they do not show an optical counterpart (Fynbo et al. 2001). Since this definition is clearly related to the sensitivity (and availability) of the instruments used for the follow-up a more general definition is needed. Based on the predictions of the fireball model (M\\'esz\\'aros \\& Rees 1997) one can require that the optical to X--ray spectral index $\\beta_{OX}$ (i.e. the slope between the fluxes in the $R$-band and at 3 keV at 11 hr after the burst) should be lower than 0.5 (Jakobsson et al. 2004). This will individuate optically sub-luminous bursts, i.e. fainter than expected from the fireball model. Alternatively, with the advent of {\\it Swift}, X--ray spectral slopes were more easily available and a somewhat different definition was put forward by van der Horst et al. (2009) for which $\\beta_{OX}$ is shallower than $\\beta_X - 0.5$. The darkness of a GRB can have different causes: it can be due to intrinsically optically faint GRBs, it can be due to absorption by intervening material within the host galaxy or it can be due to high redshift GRBs, thus being absorbed by the intergalactic medium. Several works have addressed this topic in the {\\it Swift} era when a number of facilities allowed a quick follow-up of the afterglows. The fraction of dark bursts has been estimated to be $\\sim 25-50\\%$ according to Jakobsson's definition (Melandri et al. 2008; Roming et al. 2009; Cenko et al. 2009; Greiner et al. 2011; Melandri et al. 2011). It is now believed that the faint optical afterglow emission of dark bursts might be due to a moderate intrinsic extinction at moderate redshifts. Greiner et al. (2011) estimated a $\\sim 20\\%$ contribution from high redshift ($z\\gsim 4-5$) GRBs to the dark population only. Salvaterra et al. (2011, see also Nava et al. 2011) selected a complete sample of bright GRBs based on optical observability (Jakobsson et al. 2006) and {\\it Swift} BAT peak flux $P\\ge 2.6$ ph s$^{-1}$ cm$^{-2}$. The sample consists of 58 GRBs and it is complete in spectroscopic redshift at $90\\%$ (with $95\\%$ of GRBs having some constraints on the redshift). This sample offers the opportunity to study in an unbiased way the distribution of the X--ray column densities and its relation to GRB darkness. The paper is organised as follows. In section 2 we derive the X--ray absorbing column densities for the Salvaterra's sample and briefly describe how the slope $\\beta_{OX}$ has been computed for each GRB of the sample. In section 3 we discuss our findings and in section 4 we draw our conclusions. ", "conclusions": "Salvaterra et al. (2011) selected a complete sample of bright GRBs with a high degree of completeness in redshift. In a series of papers we investigate the impact of this sample on GRB studies. Here we focus on the properties of the sample with respect to the intrinsic X--ray absorption. We found that the intrinsic column density distribution of our complete sample is consistent with the total distribution of column densities presented in Campana et al. (2010). The mean of the two distributions are in fact $21.7\\pm0.5$ and $21.9\\pm0.1$, respectively. This likely indicates that the GRB brightness, as well as any other bias present in the total sample of GRBs with redshift (e.g. dust), does not heavily influence the total distribution of intrinsic column densities. At variance with the total distribution presented in Campana et al. (2010), we see in the complete sample presented here that the region at high column densities and low redshift is now more populated by GRBs. This clearly reveals a bias present in the non-complete sample, where this region is heavily underpopulated due to the lack of a redshift determination of dark bursts. Even if not statistically compelling there is an increase of the intrinsic column density with redshift (this is more apparent in the full sample of GRBs with redshift, Campana et al. 2010). We evaluate the mean contribution to $N_H(z)$ due to the intervening systems along the GRB line of sight. We find that, if we take into account the larger number of observed systems affecting the line of sight of GRBs with respect to the quasar one (Vergani et al. 2009), the population of sub-Damped Lyman-$\\alpha$ and Damped Lyman-$\\alpha$ systems can account for the increase with redshift of $N_H(z)$. It would be interesting to confirm this directly through the study of high-$z$ GRB lines of sight. Unfortunately this effect plays a significant role at very high redshift, where the number of GRB afterglow spectra is very low. It is indeed difficult to measure absorption from Lyman-$\\alpha$. However, the column density of neutral gas can still be traced by weakly ionised metal lines (e.g. Zn II, Si II), which in fact is a more logical method of comparing absorption in X--rays and the optical, given that the X--rays are absorbed by metals and not neutral hydrogen (e.g. Schady et al. 2011). The X-shooter instrument mounted at the ESO/VLT offers the best opportunities for these studies. Making use of the $\\beta_{OX}$ computed by Melandri et al. (2011), we found a strong correlation between GRB darkness and X--ray absorbing column densities. Since metals are a key ingredient for dust production (Draine 2003), our findings are consistent with a picture in which the darkness of a GRB is in most cases due to absorption by circumburst material." }, "1112/1112.6218_arXiv.txt": { "abstract": "{Solar cycles vary in their amplitude and shape. There are several empirical relations between various parameters that link cycle's shape and amplitude, the foremost the Waldmeier relations.} {The solar cycle is believed to be a result of the solar dynamo action, therefore these relations require an explanation in the framework of this theory, which we aim to present here.} {We related the cycle-to-cycle variability of solar activity to fluctuations of solar dynamo drivers and primarily to fluctuations in the parameter responsible for the recovery of the poloidal magnetic field from the toroidal one. To be specific, we developed a model in the framework of the mean-field dynamo based on the differential rotation and $\\alpha$-effect.} {We demonstrate that the mean-field dynamo model, which is based on a realistic rotation profile and on nonlinearity that is associated with the magnetic helicity balance, reproduces both qualitatively and quantitatively the Waldmeier relations observed in sunspot data since 1750. The model also reproduces more or less successfully other relations between the parameters under discussion, in particular, the link between odd and even cycles (Gnevyshev-Ohl rule).} {We conclude that the contemporary solar dynamo theory provides a way to explain the cycle-to-cycle variability of solar activity as recorded in sunspots. { We discuss the importance of the model for stellar activity cycles which, as known from the data of the Mount Wilson HK project, which measures the Ca H and K line index for other stars, demonstrate the cycle-to-cycle variability similar to solar cycles.}} ", "introduction": "} Solar activity has a periodic nature, but the cycle amplitude and shape vary from one cycle to the other. { This challenges the prognostic abilities of solar activity models. The sunspot activity can be quantified by using various tracers derived from observations. These tracers show an interrelation among each other. The indices characterizing the tracers can be employed to predict the future evolution of solar activity.} Waldmeier (\\citeyear{w35}) first suggested this option (an inverse correlation between the length of the ascending phase of a cycle and the peak sunspot number of that cycle) and applied it (\\citealp{w36}) to predict the subsequent cycle. Later, other relations of this kind were proposed and called Waldmeier relations (see for a review \\citealp{vetal86} and \\citealp{hathaw02}). The nature of physical processes underlying the Waldmeier relations is not clear \\citep[see discussion, e.g., in][]{camschu07,detal08,karak11}. We note, that these statistical properties of the magnetic activity also exist for other tracers related to the sunspot activity (e.g., sunspot group number and area, see \\citealp{vetal86,hathaw02,karak11}), and even for the other kind of solar and stellar activity indices, e.g., for the Ca II index \\citep{soon94}. The Waldmeier relations are considered as a valuable test for dynamo models \\citep{karak11,pk11}. Clarifying the physics underlying the Waldmeier relations is particularly attractive to support the prognostic abilities concerning solar cycle. It is believed that the cyclic solar activity is driven by a dynamo, i.e. a mechanism that transforms the kinetic energy of hydrodynamic motions into a magnetic one. Many modern solar dynamo models \\citep[see, e.g.,][]{stix:02} assume that the toroidal magnetic field that emerges on the surface and forms sunspots is generated near the bottom of the convection zone, in the tachocline or just beneath it in a convection overshoot layer. This kind of dynamo can be approximated by the Parker dynamo waves \\citep{park93}. The direction of the dynamo wave propagation in the framework of the $\\alpha\\Omega$-dynamo is defined by the Parker-Yoshimura rule \\citep{par55,yosh1975}, according to which the wave propagates along iso-surfaces of the angular velocity. The propagation can be affected by the turbulent transport (associated with the mean drift of magnetic activity in the turbulent media by means of the turbulent mechanisms), by the anisotropic turbulent diffusivity \\citep{k02}, and by meridional circulation \\citep{yosh1975,choud95}. An alternative to the Parker's surface dynamo waves is the distributed dynamo with subsurface shear \\citep[e.g.,][]{b05}, where the dynamo wave propagates along the radius in the main part of the solar convection zone \\citep{k02}. Near-surface activity is determined by the subsurface shear. Another popular option is the flux-transport dynamo \\citep[e.g.,][]{choud95,dc99}. In the context of dynamo theory, the Waldmeier relations can be explained by invoking physical mechanisms of the solar magnetic field generation and a mechanism that drives variations of the amplitude and shape of the activity cycle. For example, \\cite{pk11}, hereafter PK11, showed that variations of the $\\alpha$-effect amplitude may explain the correlation between the cycle rise rate and the cycle amplitude and other types of the Waldmeier relations as well. It was suggested \\citep{choud92,h93} that the fluctuations of the $\\alpha$-effect (associated with kinetic helicity fluctuations) are likely to be one of the natural sources of the cycle parameter variations. In addition to the statistical relations between the cycle parameters within a separate cycle there are correlations relating the parameters in subsequent cycles, for example, the odd-even cycle and the last cycle period-amplitude effects. These effects are closely related to the memory of the dynamo processes and to the strength of the saturation processes, which damp deviations of the cycle parameters from the reference state characterizing the cycle attractor \\citep{oss-h96a,oss-h96b}. It was argued \\citep{choud92,h93}, that a dozen percent is a reasonable estimate for the noise component of the $\\alpha$-effect. Previous calculations (see the above cited papers) showed that a straightforward application of the idea with the vortex turnover time and the vortex size as the correlation time and length for the $\\alpha$-fluctuations needs fluctuations much stronger than the mean $\\alpha$. On the other hand, the results of direct numerical simulations \\citep[e.g,][]{bs02} and results of current helicity (related to $\\alpha$) observations in solar active regions \\citep[e.g.,][]{zetal10} {suggest that the correlation time for $\\alpha$-fluctuations can be comparable to the cycle length and the correlation length comparable to the extent of the latitudinal belts}. Using these results, \\citet{moss-sok08} and \\citet{uetal09} showed that an $\\alpha$-noise on the order of few dozen percents is sufficient to explain the Grand minima of solar activity. The aim of this paper is to examine the result of $\\alpha$-fluctuations in the statistical properties of the solar cycle including the Waldmeier relations and the odd-even cycle effect. We chose a particular model for the solar cycle in which $\\alpha$-fluctuations are introduced. Of course, it is impractical to try all available models to learn which one is better to obtain the relations under discussion, but we select below the model among a relative wide choice of the models that gave better results in the preliminary simulations \\citep{ps11}. ", "conclusions": "We have studied the impact of low-amplitude Gaussian fluctuations of the $\\alpha$-effect on the statistical properties of the magnetic dynamo cycle, such as the Waldmeier relations and the Gnevyshev-Ohl rules. The dynamo model includes long-term fluctuations of the $\\alpha$-effect and employs two types of a nonlinear feedback of the mean-field on the $\\alpha$-effect, including algebraic quenching and dynamic quenching due to the magnetic helicity generation. The general properties of the dynamo, such as the direction of the toroidal magnetic field drift, the polar magnetic field sign reversal at the maximum of a cycle, etc., are consistent with observations. Our model does not include the meridional circulation effect, which is advocated by the Babcock-Leighton and the flux-transport type dynamo models \\citep[e.g.,][]{choud95,dc99,detal08}. It was shown that a part of the Waldmeier relations can be possibly explained by a specially tuned flux-transport model that considers fluctuations of the meridional circulation speed \\citep{karak11}. Still, observational constraints on the distribution of the meridional circulation inside the convection zone are not very strong because we have measurements for the surface. The angular momentum balance in mean-field models supports for the circulation pattern, which has a deep circulation stagnation point and a strong concentration of the velocity speed towards the bottom and the top boundaries of the solar convection zone (e.g., \\citealp{kit11}). Yet, most of the flux-transport models (including \\citealp{karak11}) use a very different circulation pattern. Following to this argumentation, we postpone a more complete study of the effects of the meridional circulation fluctuations to the future. We showed, confirming the previous findings of \\cite{pk11} and \\cite{ps11}, that variations of the $\\alpha$-effect amplitude result in variations of the cycle amplitude and period. Taking into account random fluctuations of the $\\alpha$-effect, we calculated statistical properties relating the cycle amplitude, the cycle shape, the rise time, etc., on the basis of the simulated SN data set covering period of more than 10000 years. Our results agree well with observations for the Waldmeier relations and the Gnevyshev-Ohl rules. From the qualitative point of view these results were anticipated from the earlier analysis of the helicity fluctuation effect in the dynamo given by \\cite{choud92} and \\cite{h93} (see, also \\citealp{oss-h96a,oss-h96b,moss-sok08,uetal09}). Our results presented in Figure~\\ref{minsurf} about the correlation of the polar dipole field and the cycle amplitude and as the results for the Gnevyshev-Ohl rules suggest that the Waldmeier relations can be understood by considering the general properties of the magnetic field generation processes, which are involved in the dynamo. Our model shows a good correlation (with low variance) between the strength of the polar dipole magnetic field in the cycle minimum and the amplitude of the subsequent cycle. This results from the deterministic process of the toroidal magnetic field generation by the differential rotation from the large-scale poloidal magnetic field. This correlation is often used for the cycle prediction \\citep{hath09} by Babcock-Leighton type dynamo models and it is for the first time demonstrated in the mean-field dynamo. The rise rate of the sunspot cycle depends on the differential rotation and the amplitude of the poloidal field. Therefore, the correlation between the rise rate and amplitude of the cycle is a derivative property and is a consequence of the link between the polar dipole magnetic field in the cycle minimum and the strength of the toroidal field in the subsequent cycle. Furthermore, following the general idea of \\cite{GM78} (cf, \\citealp{h93,chetal07}), we can interpret the Gnevyshev-Ohl rule as evidence that the solar cycle is a nonlinear self-excited oscillation that tends to preserve the property of the attractor under random perturbations. The amplitude and phase of the subsequent cycles are related by the so-called Zaslavsky map. The strength of the link between the parameters of the subsequent cycles is controlled by the fluctuation amplitude and by the perturbation's decrement. The latter strongly depends on the nonlinear mechanisms involved in the dynamo. To examine this idea we made additional simulations with a lower helicity dissipation rate (high $R_{\\chi}$) and found that the correlation coefficients between the parameters of the subsequent cycle increase with increasing $R_{\\chi}$. Therefore the link between the odd and even cycles, and the period to amplitude correlation in subsequent cycles can be considered as evidence for the fluctuation impact on the dynamo and evidence for nonlinear damping of these perturbations in the dynamo. This conclusion needs to be investigated in more detail especially by comparing the results of the $\\alpha$-effect and meridional circulation fluctuations. Long-term variations of the magnetic cycle in the dynamo can be induced in different ways. Two main mechanisms can be identified: nonlinear deterministic chaos and an effect of fluctuations of the turbulent parameters involved in the dynamo process. Generally, we anticipate that statistical properties of long-term cycle variations are depended on the force that drives the long-term variations. We examined weakly nonlinear models with the amplitude of the $\\alpha$-effect close to the threshold. In our models, the typical ratio between the energy of the large-scale toroidal field and the kinetic energy of convective flows does not exceed $0.3$. As a result, the chaotic regime in the model is not as evident as the impact of the $\\alpha$-effect fluctuations. Figure~\\ref{prob} illustrates the difficulty to obtain extended episodes of low magnetic activity in this case, while these episodes are common in the solar dynamo \\citep{setal04}. To amplify the chaotic regime in the model, we have tried additional possibilities and included an angular momentum balance into the dynamo problem to take into account the nonlinear feedback of the magnetic field on the differential rotation. The model was described earlier by \\cite{p99,p04}. Figure~\\ref{ndifr} shows the simulated SN for the model involving the nonlinear effect of the magnetic field on the angular momentum balance in the solar convection zone. This model shows higher intermittency in the cycle variations than that in Figure~\\ref{waldr}, and indeed it has a similar probability of the occurrence of lowactivity episodes as that in the reconstruction data set. \\begin{figure} \\begin{centering} \\includegraphics[width=0.95\\columnwidth]{fig9} \\par\\end{centering} \\caption{Time series of the simulated sunspot number for the extended 2D1$\\alpha$ model involving the fluctuations of the alpha effect and the magnetic feedback on the differential rotation.}\\label{ndifr} \\end{figure} It is natural to expect that at least stellar magnetic cycles of solar-like stars should demonstrate a variability similar to the solar one, including relations comparable with the Waldmeier relations. Available observations of stellar activity provide some hints that support this expectation. { Stellar activity data of the Mount Wilson HK project measuring the Ca H and K line index for other stars \\citep{bal} are available for two activity cycles.} The wavelet analysis of the data for several stars \\citep{balfrick} demonstrated that the subsequent cycles for a given star can differ in their cycle amplitudes. We note that a monitoring of stellar activity of solar-like stars to obtain relations similar to the Waldmeier ones could establish our prognostic abilities of solar activity based on these relations much better. Summarizing the results of the paper, we conclude that the mean-field solar dynamo theory provides a way to explain the cycle-to-cycle variability of solar activity as recorded in sunspots. The results given in the literature and the results obtained in the paper suggest that the Waldmeier relations can be explained invoking very different kinds of dynamo models. More work is necessary to study the relations between the statistical properties of the dynamo cycle, and the dynamo mechanisms involved in the magnetic activity will help to obtain more insight into the processes operating in the stellar and solar dynamo." }, "1112/1112.1438_arXiv.txt": { "abstract": "We present the results of a systematic Green Bank Telescope (GBT) and Giant Metrewave Radio Telescope (GMRT) survey for 21-cm absorption in a sample of 10 damped Lyman-$\\alpha$ (DLA) systems at $2\\le$ \\zabs$\\le 3.4$. Analysis of L-band Very Long Baseline Array (VLBA) images of the background QSOs are also presented. We detect 21-cm absorption in only one DLA (at \\zabs\\ = 3.1745 towards J1337+3152). Thus the detection rate of 21-cm absorption is $\\sim 10\\%$ when no limit on the integrated optical depth ($\\int \\tau(v) dv$) is imposed and $\\sim13$\\% for a 3$\\sigma$ limit of 0.4~\\kms. Combining our data with the data from the literature (a sample of 28 DLAs) and assuming the measured core fraction at milliarcsecond scale to represent the gas covering factor, we find that the H~{\\sc i} gas in DLAs at $z\\ge 2$ is predominantly constituted by warm neutral medium. The detection rate of 21-cm absorption seems to be higher for systems with higher $N$(H~{\\sc i}) or metallicity. However, no clear correlation is found between the integrated 21-cm optical depth (or the spin-temperature, $T_{\\rm S}$) and either $N$(H~{\\sc i}), metallicity or velocity spread of the low ionization species. There are 13 DLAs in our sample for which high resolution optical spectra covering the expected wavelength range of \\h2 absorption are available. We report the detection of \\h2 molecules in the \\zabs\\ ~=~3.3871 21-cm absorber towards J0203+1134 (PKS 0201+113). In 8 cases, neither \\h2 (with molecular fraction $f$($H_2$)$ \\le 10^{-6}$) nor 21-cm absorption (with $T_{\\rm S}/f_{\\rm c} \\ge$~700~K) are detected. The lack of 21-cm and H$_2$ absorption in these systems can be explained if most of the H~{\\sc i} in these DLAs originate from low density high temperature gas. {In one case we have a DLA with 21-cm absorption not showing \\h2 absorption.} In two cases, both species are detected but do not originate from the same velocity component. In the remaining 2 cases 21-cm absorption is not detected despite the presence of \\h2 with evidence for the presence of cold gas. All this is consistent with the idea that the \\h2 components seen in DLAs are compact (with sizes of $\\le$~15~pc) and contain only a small fraction (i.e typically $\\le 10\\%$) of the total $N$(H~{\\sc i}) measured in the DLAs. This implies that the molecular fractions $f$(\\h2) reported from the \\h2 surveys should be considered as conservative lower limits for the \\h2 components. ", "introduction": "The Galactic interstellar medium (ISM) has a multiphase structure with neutral hydrogen being distributed between the cold neutral (CNM), warm neutral (WNM) and warm ionized (WIM) media. A large fraction of the gas is also found in diffuse, translucent and dense molecular clouds. Newly formed stars are associated with these dense molecular clouds and strongly influence the physical state of the rest of the gas in different forms through radiative and mechanical inputs. The physical conditions in the multiphase ISM depend on the UV background radiation field, metallicities, dust content and the density of cosmic rays \\citep[see Figs.~5, 6 and 7 in][]{Wolfire95}. In addition, the filling factor of the different phases depends sensitively on the supernova rate \\citep{deavillez2004}. Therefore, detecting and studying the multiphase ISM in external galaxies has great importance for our understanding of galaxy evolution. Damped Lyman-$\\alpha$ systems (DLAs) are the highest H~{\\sc i} column density absorbers seen in QSO spectra, with $N$(H~{\\sc i})$\\ge 2\\times 10^{20}$~cm$^{-2}$. These absorbers trace the bulk of the neutral hydrogen at $2\\le z\\le 3$ \\citep [][]{Prochaska05,Noterdaeme09dla} and have long been identified as revealing the interstellar medium of the high-redshift precursors of present day galaxies \\citep[for a review see,][]{wolfe05}. The typical dust-to-gas ratio of DLAs, is less than one tenth of that observed in the local ISM, and only a small fraction ($<\u223c10$\\%) of DLAs show detectable amounts of molecular hydrogen \\citep{Petitjean00,Ledoux03,Noterdaeme08} with the detection rate being correlated to the dust content of the gas \\citep{Petitjean06}. The estimated temperature and molecular fraction in these systems are consistent with them originating from the CNM \\citep{Srianand05}. It has been shown recently that strong C~{\\sc i} absorbers detected in low-resolution Sloan Digital Sky Survey (SDSS) spectra are good candidates for H$_2$ bearing systems. Indeed these absorbers have yielded the first detections of CO molecules in high-$z$ DLAs \\citep{Srianand08,Noterdaeme09co,Noterdaeme10co,Noterdaeme11}. The properties of these absorbers are similar to those of translucent molecular clouds. The fact that no DLA is found to be associated with a dense molecular cloud, a fundamental ingredient of star-formation, is most certainly related to the large extinction that these clouds are expected to produce and/or the small size of such regions \\citep{Zwaan06} making detections difficult. Thus, most DLAs detected in optical spectroscopic surveys seem to probe the diffuse H~{\\sc i} gas \\citep{Petitjean00}. However, about 50\\% of the DLAs show detectable C~{\\sc ii}$^*$ absorption \\citep{Wolfe08}, and \\citet{Wolfe03b} argued that a considerable fraction of the C~{\\sc ii}$^*$ absorption in DLAs originates from CNM gas \\citep[see however][]{Srianand05}. { Detection of 21-cm absorption is the best way to estimate the CNM fraction of DLAs as it is sensitive to both $N$(H~{\\sc i}) and thermal state of the gas \\citep{Kulkarni88}.} \\begin{figure*} \\centerline{ \\vbox{ \\hbox{ \\includegraphics[trim= 25.0mm 50.0mm 90.0 40.0mm, clip, scale=0.33,angle=90.0]{J0733+2721c.ps} \\includegraphics[trim= 25.0mm 50.0mm 90.0 50.0mm, clip, scale=0.33,angle=90.0]{J0801+4725c.ps} \\includegraphics[trim= 25.0mm 50.0mm 90.0 50.0mm, clip, scale=0.33,angle=90.0]{J0816+4823c.ps} } \\hbox{ \\includegraphics[trim= 25.0mm 50.0mm 70.0 40.0mm, clip, scale=0.33,angle=90.0]{J0839+2002c.ps} \\includegraphics[trim= 25.0mm 50.0mm 70.0 50.0mm, clip, scale=0.33,angle=90.0]{J0852+2431c.ps} \\includegraphics[trim= 25.0mm 50.0mm 70.0 50.0mm, clip, scale=0.33,angle=90.0]{J1017+6116c.ps} } \\hbox{ \\includegraphics[trim= 25.0mm 50.0mm 70.0 40.0mm, clip, scale=0.33,angle=90.0]{J1223+5037c.ps} \\includegraphics[trim= 25.0mm 50.0mm 70.0 50.0mm, clip, scale=0.33,angle=90.0]{J1237+4708c.ps} \\includegraphics[trim= 25.0mm 50.0mm 70.0 50.0mm, clip, scale=0.33,angle=90.0]{J1242+3720c.ps} } \\hbox{ \\includegraphics[trim= 25.0mm 50.0mm 70.0 40.0mm, clip, scale=0.33,angle=90.0]{J1406+3433c.ps} \\includegraphics[trim= 25.0mm 50.0mm 90.0 50.0mm, clip, scale=0.33,angle=90.0]{J1413+4505c.ps} \\includegraphics[trim= 25.0mm 50.0mm 90.0 50.0mm, clip, scale=0.33,angle=90.0]{J1435+5434c.ps} } }} \\caption[]{ SDSS spectra showing the \\lya lines for 12 DLAs in our sample. The best fitted Voigt profiles (solid curves) together with the associated 1$\\sigma$ errors (shaded regions) are over-plotted. The dotted curve gives the best fitted continuum (in some cases the continuum fit includes the emission lines also). We have used VLT UVES spectra to get $N$(H~{\\sc i}) in the case of \\zabs = 3.1745 system towards J1337+3152 \\citep[see][]{Srianand10} and \\zabs = 2.595 and 2.622 systems towards J0407$-$4410 \\citep[CTS 247, see][]{Ledoux03}. } \\label{dlafits} \\end{figure*} This is why it is important to search for 21-cm absorption in DLAs over a wide redshift range. While a good fraction of DLAs/sub-DLAs preselected through Mg~{\\sc ii} absorption seems to show 21-cm absorption at $z\\sim 1.3$ \\citep[see for example][]{Gupta09, Kanekar09mg2}, searches for 21-cm absorption in DLAs at $z_{\\rm abs}\\ge2$ have mostly resulted in null detections \\citep[see][]{Kanekar03,Curran10} with only four detections reported till now \\citep[see][]{Wolfe85,Kanekar06,Kanekar07,York07}. { The low detection rate of 21-cm absorption in high-$z$ DLAs can be related to either the gas being warm (i.e high spin temperature, $T_{\\rm S}$, as suggested by \\citet{Kanekar03}) and/or the low value of covering factor ($f_c$) through high-z geometric effects \\citep{Curran06}. The best way to address the covering factor issue is to perform milliarcsecond scale spectroscopy in the redshifted 21-cm line using very long baseline interferometry (VLBI) to measure the extent of absorbing gas \\citep{Lane00}. Unfortunately due to limited frequency coverage and sensitivity of the receivers available with VLBI such studies cannot be extended to high redshift DLAs. Alternatively, the core fraction measured in the milliarcsecond scale images can be used to get an estimate of the covering factor \\citep[see][] {Briggs89, Kanekar09vlba}. Here one assumes that the absorbing gas completely covers at least the emission from the milliarcsec scale core. Therefore, to address this issue, one needs, not only to increase the number of systems searched for 21-cm absorption but also to perform milliarcsecond scale imaging of the background radio sources. We report here the results of a search for 21-cm absorption in 10 DLAs at $z>2$ we have carried out using GBT and GMRT, complemented by L-band VLBA images of the background QSOs. \\begin{table} \\caption{Log of GBT and GMRT observations to search for 21-cm absorption} \\begin{center} \\begin{tabular}{lccccc} \\hline \\hline \\multicolumn{1}{c}{Source}& Tele- & Date & Time & BW & Ch. \\\\ \\multicolumn{1}{c}{name} & scope & & & & \\\\ & & yy-mm-dd & (hr) & (MHz) & \\\\ \\multicolumn{1}{c}{(1)} & (2) & (3) & (4) & (5) & (6) \\\\ \\hline J0407$-$4410 & GBT & 2006-10-20 & 4.7 & 0.625 & 512 \\\\ (CTS247) & & 2007-01-05 & & & \\\\ & & 2007-01-06 & & & \\\\ & & 2007-01-08 & & & \\\\ J0733+2721 & GBT & 2007-12-05 & 10.7 & 1.25 & 512 \\\\ & & 2007-12-06 & & & \\\\ J0801+4725 & GMRT & 2006-12-22 & 10.8 & 1 & 128 \\\\ & & 2006-11-23 & & & \\\\ J0852+2431 & GBT & 2009-08-07 & 5.6 & 1.25 & 1024 \\\\ & & 2009-08-08 & & & \\\\ & & 2009-09-06 & & & \\\\ J1017+6116 & GBT & 2008-10-15 & 4.5 & 1.25 & 1024 \\\\ & & 2008-10-16 & & & \\\\ & & 2008-10-18 & & & \\\\ J1242+3720 & GMRT & 2007-06-08 & 6.1 & 0.5 & 128 \\\\ & & 2007-06-10 & & & \\\\ J1337+3152 & GMRT & 2009-01-13 & 6.2 & 1 & 128 \\\\ & & 2009-03-17 & 7.8 & 0.25 & 128 \\\\ J1406+3433 & GBT & 2009-03-05 & 8.0 & 1.25 & 1024 \\\\ & & 2009-03-06 & & & \\\\ & & 2009-03-07 & & & \\\\ & & 2009-03-08 & & & \\\\ & & 2009-04-07 & & & \\\\ J1435+5435 & GMRT & 2007-06-10 & 6.7 & 1 & 128 \\\\ & & & & & \\\\ \\hline \\end{tabular} \\end{center} \\begin{flushleft} Column 1: Source name. Column 2: Telescope used for 21-cm absorption search. Column 3: Date of observations. Column 4: Total time on source i.e. after excluding telescope set-up time and calibration overheads. Columns 5 and 6: Spectral setup for the observations i.e. bandwidth (BW) and number of spectral channels respectively. \\end{flushleft} \\label{obslog} \\end{table} This survey has resulted in the detection of 21-cm absorption in the \\zabs = 3.1745 DLA towards J1337+3152. A detailed analysis of this system and two sub-DLAs close to this system are presented in \\citet{Srianand10}. Section~\\ref{samp} presents the details of our sample. In Section 3 we present the details of GBT and GMRT spectroscopic observations, VLBA continuum observations, and data reduction. The detection rate of 21-cm absorption in DLAs is discussed in Section~\\ref{detect}. In Section~\\ref{gencor} we study the correlations between the parameters derived from 21-cm observations, $N$(H~{\\sc i}), metallicity and redshift. In Section~\\ref{mole} we study the relation between 21-cm and \\h2 absorption. The results are summarized in Section~\\ref{results}. In this work we assume a flat Universe with $H_0$~=~71~\\kms~Mpc$^{-1}$, $\\Omega_{\\rm m}$~=~0.27 and $\\Omega_\\Lambda$~=~0.73. ", "conclusions": "\\label{results} We have carried out a systematic search for 21-cm absorption in 10 DLAs at \\zabs$>2$ using GMRT and GBT. We detect 21-cm absorption in only one of them. From our sample we find the 21-cm detection rate is 13\\% for a $\\int \\tau dv$ limit of 0.4 km/s (the detection limit reached in the case of J1337+3152). We also obtained 1420~MHz VLBI images for the sources in our sample. The 21-cm detection at $z\\ge 2$ seems to favour systems with high metallicity and/or high $N$(H~{\\sc i}) \\citep[see also][]{Kanekar09ts, Curran10}. This basically means that the probability of detecting cold components that can produce detectable 21-cm absorption is higher in systems with high values of $N$(H~{\\sc i}) and Z. However, we do not find any correlation between the integrated optical depth (or T$_{\\rm S}$/$f_{\\rm c}$) and $N$(H~{\\sc i}) or metallicity. It is important to address the covering factor issue before drawing any conclusions on $T_S$. Ideally one should do high spatial resolution VLBA spectroscopy for this purpose \\citep[see for example][]{Lane00}. However, this is not possible at present specially for $z\\ge2$ absorbers. Therefore, we proceed by assuming that the core fraction found in the VLBA images as the covering factor of the absorbing gas \\citep[as in the case of][]{Kanekar09vlba}. We find that more than 50\\% of DLAs have weighted mean spin temperature ($T_{\\rm S}$) in excess of 700 K. For the assumed temperature of the CNM gas $T_{\\rm S}^C = 200$ K (as seen in \\h2 components in high-z DLAs) we find that more than 73\\% of H~{\\sc i} in such systems is originating from WNM. The median value CNM fraction (i.e $f$(CNM)) obtained for the detections and the median value of upper limits in the case of non-detections are in the range 0.2 to 0.25. We study the connection between 21-cm and \\h2 absorption in a sub-sample of 13 DLAs where both these species can be searched for. We report the detection and detailed analysis of \\h2 molecules in the \\zabs=3.3871 DLA system towards J0203+1134 where 21-cm absorption is also detected. For a $b$ parameter in the range 1-5~\\kms\\ we find 14.57$\\le$log~$N$(\\h2)$\\le$16.03. The inferred kinetic temperature is in the range 48-108~K based on $T_{01}$ of H$_2$. However no 21-cm absorption is detected at the very position of this \\h2\\ component. This suggests that the H~{\\sc i} column density associated with this component is $\\le$ 10$^{19}$~cm$^{-2}$. However, the lack of proper coincidence between 21-cm and any of the strong UV absorption components may also mean that the radio and optical sight lines probe different volumes of the gas. In the case of 8 DLAs, neither 21-cm nor H$_2$ are detected. Typical upper limits on the molecular fraction ($f_{\\rm H_2}$) in these systems are $\\le 10^{-6}$. The lack of \\h2 in DLAs can be explained if the H~{\\sc i} gas originates from low density regions photoionized by the metagalactic UV \\citep[see for example,][]{Petitjean92,Petitjean00,Hirashita05}. This also indicates that the volume filling factor of \\h2 in DLAs is small \\citep{Zwaan06}. Typical limits obtained for $T_{\\rm S}$ in these systems are consistent with only a small fraction of the H~{\\sc i} gas originating from the CNM phase as suggested by the lack of \\h2 absorption. In two cases strong \\h2 absorption is detected and kinetic temperatures are in the range 100-200~K, but 21-cm absorption is not detected. Even in two cases where both the species are detected they do not originate from the same velocity component. The lack of 21-cm absorption directly associated with \\h2\\ indicates that only a small fraction (typically $\\le$ 10\\%) of the neutral hydrogen seen in the DLA is associated with the \\h2 components \\citep[see also][]{Noterdaeme10co}. This implies that the molecular fractions $f$(\\h2) reported from the \\h2 surveys should be considered as conservative lower limits for the \\h2 components. For two of the \\h2-bearing DLAs with density measurements based on C~{\\sc i} fine-structure excitation we derive an upper limit on the line of sight thickness of $\\le 15$~pc. { This is consistent with the size estimate for the H$_2$-bearing gas in \\zabs = 2.2377 DLA towards Q1232+082 based on partial coverage \\citep{Balashev11}.} In principle, the presence of \\h2 and 21-cm absorptions in a single component provides a unique combination to simultaneously constrain the variation of the fine-structure constant ($\\alpha$), the electron-to-proton mass ratio ($\\mu$) and the proton G-factor. As shown here, DLAs with 21-cm and \\h2 detections are rare. Even in these cases the presence of multiphase structure at parsec scale is evident, introducing velocity shifts between the different absorption components that will affect the constraints on the variation of constants." }, "1112/1112.3860_arXiv.txt": { "abstract": "{The clumpy density structure of photon-dominated regions is well established, but the physical properties of the clumps and of the surrounding interclump medium are only approximately known.} {The aim of this paper is to constrain the physical and chemical conditions in the Orion Bar, a prototypical nearby photon-dominated region.} {We present observations of the HF $J$=1--0 line, which appears in emission toward the Orion Bar, and compare the brightness of the line to non-LTE radiative transfer calculations. } {The large width of the HF line suggests an origin of the emission in the interclump gas, but collisional excitation by \\hh\\ in the interclump gas underpredicts the observed line intensity by factors of 3--5. In contrast, an origin of the line in the dense clumps requires a density of $\\sim$10$^9$\\ccm, 10--100 times higher than previous estimates, which is unlikely. However, electron impact excitation reproduces our observations for $T$=100\\,K and $n_e$=10\\,\\ccm, as expected for the interclump gas.} {We conclude that HF emission is a signpost of molecular gas with a high electron density. Similar conditions may apply to active galactic nuclei where HF also appears in emission.} ", "introduction": "\\label{s:intro} Photon-dominated regions (PDRs) are the surface regions of molecular clouds, where ultraviolet radiation with photon energies between a few and 13.6\\,eV drives the thermal and chemical balance of the gas \\citep{hollenbach1999}. This situation occurs in regions of high-mass star formation, but also in protoplanetary disks and in the nuclei of active galaxies. Studying the structure of PDRs therefore has a wide astrophysical application. In PDRs, gas heating proceeds by photo-electric emission from dust grains, while the main cooling channels are the fine structure lines of C$^+$ and O and the rotational lines of CO \\citep{kaufman1999}. Absorption of the impinging ultraviolet radiation by dust and gas in the PDR creates a layered structure, where chemical transitions such as H$^+$ $\\to$ H $\\to$ \\hh\\ and C$^+$ $\\to$ C $\\to$ CO occur. The Orion Bar is a prototypical PDR, located between the Orion molecular cloud and the Orion Nebula, the H$^+$ region surrounding the Trapezium stars, at a distance of 414\\,pc \\citep{menten2007}. Observations at infrared and submillimeter wavelengths indicate a geometry for the Bar where the PDR is wrapped around the Orion Nebula, and changes from a face-on to an edge-on view where the molecular emission peaks (\\citealt{hogerheijde1995}; \\citealt{walmsley2000}). The mean temperature of the molecular gas in the Bar is 85\\,K, while the temperature rises to several 100\\,K toward the ionization front, where the emission from PAH particles and vibrationally excited \\hh\\ peaks. While the temperature structure of the Orion Bar is reasonably well understood, the same cannot be said about the density structure. The mean density of the molecular gas is 10$^5$\\,\\ccm, but single-dish observations already indicate the presence of small-scale density variations without apparent pattern, usually called ''clumps'' \\citep{hogerheijde1995}, that are also seen toward other PDRs \\citep{stutzki1988,wang1993}. While interferometric observations have confirmed the presence of clumps \\citep{youngowl2000}, the densities of both the clumps and the interclump medium are somewhat uncertain. The interclump medium probably has a density between a few 10$^4$ and \\pow{2}{5}\\,\\ccm\\ \\citep{simon1997}, while estimates of the clump density range from \\pow{1.5}{6} to \\pow{6}{6}\\,\\ccm\\ \\citep{lis2003}. This Letter presents observations of the HF $J$=1--0 line, which appears in emission toward the Orion Bar. The HF molecule is expected to be the dominant carrier of gas-phase fluorine, because the reaction F + \\hh\\ $\\to$ HF + H is exothermic. For diffuse clouds where the effect of depletion on grains should be unimportant, models by \\citet{neufeld2009} predict an HF abundance of $\\sim$\\pow{3.6}{-8} relative to \\hh. Recent observations of the HF 1--0 line confirm this prediction: the line is seen in absorption toward several background sources, indicating abundances of $\\sim$\\pow{1--2}{-8} \\citep{neufeld2010}. Toward dense clouds, the abundance is measured to be $\\sim$100 times lower \\citep{phillips2010}, suggesting that depletion of F on grain surfaces or excitation effects play a role. Like CO, HF is a linear rotor with a regular line spectrum, where $\\nu = 2B(J+1)$ and $A_{ij} \\propto \\nu^3$. Unlike CO, HF has a small reduced mass and a large dipole moment, so that the lines have high frequencies and radiative decay is rapid. In particular, the HF 1--0 line has a frequency of 1232\\,GHz and an Einstein~A coefficient of \\pow{2.422}{-2}\\,s$^{-1}$. Thermal excitation of the $J$=1 level of HF thus requires extremely high gas densities, which is why the line usually appears in absorption \\citep{sonnentrucker2010,monje2011}. ", "conclusions": "\\label{s:concl} We have presented observations of the HF $J$=1--0 line toward the Orion Bar, which is the first time that this line is seen in emission from the Galactic interstellar medium. We present calculations and arguments why this emission cannot be caused by collisional excitation of HF by \\hh, either from dense clumps or from the inter-clump medium. However, collisional excitation by electrons in the interclump gas explains our observations. The appearance of HF in emission therefore seems to be a signpost of molecular gas with a high electron density, which in the case of the Orion Bar is the combined effect of the high gas density and the strong ultraviolet radiation field. The physical conditions of the Orion Bar may be similar to those in the nucleus of the active galaxy Mrk 231, where the radiation field is strong and where the HF line also appears in emission \\citep{vdwerf2010}. Other active galactic nuclei where HF appears in absorption such as Arp 220 \\citep{rangwala2011} and the Cloverleaf \\citep{monje2011clover} probably have lower electron densities, which may indicate a softer radiation field. Support for an origin of the HF emission in the interclump medium comes from the detection of CF$^+$ toward the Orion Bar \\citep{neufeld2006}. Additional support for this model would come from observations of H$_2$F$^+$, for which line frequencies are known \\citep{fujimori2011}. However, our data do not show any features at the predicted line frequencies down to an rms of $\\approx$50\\,mK in $T_A^*$, which is a factor of $\\sim$10 above the expected signal. In the future, maps of HF emission with Herschel-HIFI will help to clarify the spatial distribution of dense gas clumps in the Orion Bar and other PDRs. Another useful test would be the observation of HF $J$=2--1 or higher-$J$ lines in the $v$=0 or $v$=1 states. Unfortunately, the $J$=2--1 line at 2463\\,GHz lies very close (0.13\\,\\mic) to the N$^+$ $^3P_2$--$^3P_1$ fine-structure line at 121.8\\,\\mic, so that the lines are blended at the resolution of the PACS instrument onboard Herschel. Heterodyne observations in this frequency range may become possible with the future STO telescope, the SOFIA airborne observatory, or the Millimetron mission." }, "1112/1112.4995_arXiv.txt": { "abstract": "The universe evolution during the radiation-dominated epoch in the $R^2$--extended gravity theory is considered. The equations of motion for $R$ and $H$ are solved analytically and numerically. The particle production rate by the oscillating curvature is calculated in one-loop approximation and the back-reaction of particle production on the evolution of $R$ is taken into account. Possible implications of the model for cosmological creation of non-thermal dark matter is discussed. ", "introduction": "During the last two decades the accumulated astronomical data have unambiguously proved that the universe expansion is accelerated. The data include the observation of the large scale structure of the universe, the measurements of the angular fluctuations of the cosmic microwave background radiation, the determination of the universe age (for a review see~\\cite{cosm-prmtr}), and especially the discovery of the dimming of distant Supernovae~\\cite{Nobel_2011}. The driving force behind this accelerated expansion in unknown. Among possible explanations, the most popular is probably the assumption of a new (unknown) form of cosmological energy density with large negative pressure, $ P < -\\rho/3$, the so-called dark energy, for a review see e.g.~\\cite{DE_Peebles_Ratra}. % A competing mechanism to describe the accelerated expansion is represented by gravity modifications at small curvature, the so-called $f(R)$-gravity theories, as suggested in ref.~\\cite{grav-mdf}. In these theories the standard Einstein-Hilbert Lagrangian density, proportional to the scalar curvature $R$, is replaced by a function $f(R)$, so the usual action of General Relativity acquires an additional term: \\be S = -\\frac{m_{Pl}^2}{16\\pi} \\int d^4 x \\sqrt{-g}\\,f(R)+S_m= -\\frac{m_{Pl}^2}{16\\pi} \\int d^4 x \\sqrt{-g}\\,\\left[R+F(R)\\right]+S_m\\, , \\label{A1} \\ee where $m_{Pl}= 1. 2 2\\cdot 10^{19}$ GeV is the Planck mass and $S_m$ is the matter action. The original version of such models suffers from a strong instability in presence of gravitating bodies~\\cite{DolgKaw} and because of that more complicated functions $ F(R) $ have been proposed~\\cite{Starob, HuSaw, ApplBatt,Noj-Odin-2007}, which are free from the mentioned exponential instability. It was claimed in an earlier paper~\\cite{odin-instab} that the instability could be eliminated by a mere addition of terms of the type $R^q/m_q^{2(q-1)}$ to the action, where $m_q$ is a constant parameter with dimension of mass. To some extent such terms may suppress the instability for systems with relatively high mass density, $\\rho > 1 {\\rm g/cm^{3}}$, but for less dense systems e.g. for those with $\\rho \\sim 10^{-24} {\\rm g/cm{^3}}$ this mechanism demands too large coefficients in front of $R^q$ (small $m_q$), which are at odds with big bang nucleosynthesis. Though free of instability~\\cite{DolgKaw}, the models proposed in~\\cite{Starob, HuSaw, ApplBatt} possess another troublesome feature, namely in a cosmological situation they should evolve from a singular state in the past~\\cite{appl-bat-08}. Moreover, it was found in refs.~\\cite{frolov, Arb_Dolgov} that in presence of matter, a singularity may arise in the future if the matter density rises with time; such future singularity is unavoidable, regardless of the initial conditions, and is reached in a time which is much shorter than the cosmological one. In the standard Friedmann cosmology, in which the energy density decreases with time, a future singularity may appear in several scenarios (phantom cosmology, quintessence models, \\dots), but not in models~\\cite{Starob, HuSaw,ApplBatt}, as shown in ref.~\\cite{Arb_Dolgov}. This statement is in straight disagreement with earlier papers~\\cite{odin-future-sing}. The subject of modified gravity includes about $10^3$ papers at the present time and it is impossible to quote many relevant works. For a review one may look into references~\\cite{App-Bat-Star,NojOd,clifton}. The aforementioned problems can be cured by adding to the action an $R^2$-term, which prevents from the singular behavior both in the past and in the future. In the present work we study the cosmological evolution of the Universe in a theory with only an additional $R^2$ term in the action, neglecting other terms which have been introduced to generate the accelerated expansion in the contemporary universe. The impact of such terms is negligible in the limit of sufficiently large curvature, $|R|\\gg |R_0|$, where $R_0$ is the cosmological curvature at the present time. In other words, we study here the cosmological evolution of the early and not so early universe in the model with the action: \\be S = -\\frac{m_{Pl}^2}{16\\pi} \\int d^4 x \\sqrt{-g} \\left(R-\\frac{R^2}{6m^2}\\right)+S_m\\,, \\label{A-R2} \\ee where $m$ is a constant parameter with dimension of mass.\\\\ Cosmological models with an action quadratic in the curvature tensors were pioneered in ref.~\\cite{Gur-Star}. Such higher-order terms appear as a result of radiative corrections to the usual Einstein-Hilbert action after taking the expectation value of the energy-momentum tensor of matter in a curved background. In such models, like for instance the Starobinsky model~\\cite{Starobinsky_1980}, the universe may have experienced an exponential (inflationary) expansion without invoking phase transitions in the very early universe. This model has a graceful exit to matter-dominated stage which is induced by the new scalar degree of freedom, the \\textit{scalaron} (curvature scalar), which becomes a dynamical field in $R^2$-theory. The reheating process, due to gravitational particle production from scalaron (curvature scalar) oscillations, leads to a transition to a Friedmann-like universe. These features of the model are thoroughly discussed, for instance, in refs.~\\cite{Vilenkin_1985,Mijic_Morris_Suen_1986,Suen_Anderson}. Cosmological dynamics of fourth-order gravity were investigated in several works, see e.g.\\cite{Carloni_Dunsby_Troisi_2009} and references therein. A somewhat similar study was performed in ref.~\\cite{davidson} where a version of massive Brans-Dicke (BD) theory without kinetic term (i.e. with BD parameter $\\omega =0$) was considered. The Hubble parameter and curvature demonstrate oscillating behavior which resembles the one found in our paper (and earlier in many others), but quantitative features are very much different. Beside $R^2$-terms, terms containing the Ricci tensor squared $ R_{\\mu\\nu} R^{\\mu\\nu}$ are induced by radiative corrections as well, and with similar magnitude. However, the natural magnitude of such radiatively induced terms is quite small. The characteristic mass parameter, in fact, is of the order of the Planck mass in both cases, which makes this situation non-interesting for applications discussed below. On the other hand, $R^2$ cosmology (without $ R_{\\mu\\nu} R^{\\mu\\nu} $) has been considered in the literature with much larger magnitude of $R^2$ than the natural value from radiative corrections. The assumption of large $R^2$ terms is made \\textit{ad hoc} to formulate a model which could, for instance, cure singularities. We follow the spirit of those works. However, it could be worthwhile to study the consequences of more complicate models with both $R^2$ and $ R_{\\mu\\nu} R^{\\mu\\nu} $ terms. It may be a subject for future investigation. The paper is organized as follows. In sec.~\\ref{sec:eqs_of_motion} the cosmological equations modified by the presence of $R^2$-term are presented and solved both analytically and numerically in the case of a radiation-dominated (RD) universe. We study two equivalent sets of equations, for the time-time component of the (modified) Einstein equations and for their trace. In sec.~\\ref{s-part-prod} particle production by the external oscillating gravitational field is studied. First, we derive the equation of motion for the evolution of $R$ with the account of the back-reaction from particle production. This leads to an exponential damping of the oscillating part of $R$, while the non-oscillating \"Friedmann\" part remains practically undisturbed. The particle production influx into the cosmological plasma is estimated in the case of a massless, minimally-coupled scalar field. In the conclusions we discuss possible implications of the scenario, in particular for heavy supersymmetric dark matter. ", "conclusions": "The characteristic decay time of the oscillating curvature is \\be \\tau_R = \\frac{1}{2\\Gamma_R} = \\frac{24 m^2_{Pl}}{m^3} \\simeq 2 \\left(\\frac{10^5\\text{ GeV}}{m}\\right)^3 \\text{ seconds}\\,. \\label{tau-R} \\ee The contribution of the produced particles into the total cosmological energy density reaches its maximum value at approximately this time. The ratio of the energy density of the newly produced energetic particles and that of those already existing in the plasma, according to eq.~(\\ref{rho-long-time}), is: \\be \\frac{\\rho_{hi}}{\\rho_{therm}} = \\frac{8\\beta^2 N_{eff}}{\\kappa(2\\alpha_1-1)}\\,\\frac{1-(2\\Gamma_Rt_{in})^{2\\alpha_1-1}}{(2\\Gamma_R t_{in})^{2\\alpha_1-2}}\\,. \\label{ratio} \\ee If we take $t_{in} \\simeq 1/m$, then $t_{in} \\Gamma_R \\simeq m^2/m_{Pl}^2 \\ll 1$ and the effects of non-thermalized matter may be negligible. However, for sufficiently large $\\beta$ and possibly small $\\kappa$ the non-thermal particles may play a significant role in the cosmological history.\\\\ The influx of energetic protons and antiprotons could have an impact on BBN. Thus this would either allow to obtain some bounds on $m$ or even to improve the agreement between the theoretical predictions for BBN and the measurements of primordial light nuclei abundances. The oscillating curvature might also be a source of dark matter in the form of heavy supersymmetric (SUSY) particles. Since the expected light SUSY particles have not yet been discovered at LHC, to some people supersymmetry somewhat lost its attractiveness. The contribution of the stable lightest SUSY particle into the cosmological energy is proportional to \\be \\Omega \\sim{ m^2_{SUSY} }/m_{Pl} \\label{Omega-SUSY} \\ee and for $m_{SUSY} $ in the range $100-1000$ GeV the cosmological fraction of these particles would be of order unity. It is exactly what is necessary for dark matter. However, it excludes thermally produced LSP's if they are much heavier. If LSP's came from the decay of $R$ and their mass is larger than the ``mass'' of $R$, i.e. $m$, the LSP production could be sufficiently suppressed to make a reasonable contribution to dark matter. These and other manifestations of the considered scenario will be discussed elsewhere." }, "1112/1112.2057_arXiv.txt": { "abstract": "We consider the construction of generic spherically symmetric thin-shell traversable wormhole spacetimes in standard general relativity. By using the cut-and-paste procedure, we comprehensively analyze the stability of arbitrary spherically symmetric thin-shell wormholes to linearized spherically symmetric perturbations around static solutions. While a number of special cases have previously been dealt with in scattered parts of the literature, herein we take considerable effort to make the analysis as general and unified as practicable. We demonstrate in full generality that stability of the wormhole is equivalent to choosing suitable properties for the exotic material residing on the wormhole throat. ", "introduction": "Traversable wormholes are hypothetical tunnels in spacetime, through which observers may freely travel~\\cite{Morris, 269623}. These geometries are supported by ``exotic matter'', involving a stress-energy tensor violating the null energy condition (NEC). That is, there exists at least one null vector $k^\\mu$ such that $T_{\\mu\\nu}k^{\\mu}k^{\\nu}<0$ on or in the immediate vicinity of the wormhole throat (see, in particular,~\\cite{Visser, gr-qc/9704082, gr-qc/9710001}). In fact, wormhole geometries violate all the standard pointwise energy conditions, and all the standard averaged energy conditions~\\cite{Visser}. Although (most) classical forms of matter are believed to obey (most of) the standard energy conditions~\\cite{hawkingellis}, it is a well-known fact that they are violated by certain quantum effects, amongst which we may refer to the Casimir effect, Hawking evaporation, and vacuum polarization. (See~\\cite{Visser}, and more recently~\\cite{gr-qc/0205066}, for a review. For some technical details see~\\cite{gvp}.) It is interesting to note that the known violations of the pointwise energy conditions led researchers to consider the possibility of averaging of the energy conditions over timelike or null geodesics~\\cite{Tipler-AEC}. For instance, the averaged weak energy condition (AWEC) states that the integral of the energy density measured by a geodesic observer is non-negative. (That is, $\\int T_{\\mu\\nu}U^\\mu U^\\nu \\,d\\tau \\geq 0$, where $\\tau$ is the observer's proper time.) Thus, the averaged energy conditions are weaker than the pointwise energy conditions, they permit localized violations of the energy conditions, as long as they hold when suitably averaged along a null or timelike geodesic~\\cite{Tipler-AEC}. As the theme of exotic matter is a problematic issue, it is useful to minimize its usage. In fact, it is important to emphasize that the theorems which guarantee the energy condition violation are remarkably silent when it comes to making quantitative statements regarding the ``total amount'' of energy condition violating matter in the spacetime. In this context, a suitable measure for quantifying this notion was developed in \\cite{VKD}, see also~\\cite{gr-qc/0405103}, where it was shown that wormhole geometries may in principle be supported by arbitrarily small quantities of exotic matter (an interesting application of the quantification of the total amount of energy condition violating matter in warp drive spacetimes was considered in \\cite{gr-qc/0406083}). In the context of minimizing the usage of exotic matter, it was also found that for specific models of stationary and axially symmetric traversable wormholes the exotic matter is confined to certain regions around the wormhole throat, so that certain classes of geodesics traversing the wormhole need not encounter any energy condition violating matter \\cite{Teo:1998dp}. For dynamic wormholes the null energy condition, more precisely the averaged null energy condition, can be avoided in certain regions \\cite{hochvisserPRL98, hochvisserPRD98, gr-qc/9901020}. Evolving wormhole geometries were also found which exhibit ``flashes'' of weak energy condition (WEC) violation, where the matter threading the wormhole violates the energy conditions for small intervals of time \\cite{Kar}. In the context of nonlinear electrodynamics, it was found that certain dynamic wormhole solutions obey (suitably defined versions of) the WEC \\cite{Arellano:2006ex}. It is interesting to note that in modified theories of gravity, more specifically in $f(R)$ gravity, the matter threading the wormhole throat can be forced to obey (suitably defined versions of) all the energy conditions, and it is the higher-order curvature terms that are responsible for supporting these wormhole geometries \\cite{modgravity}. (Related issues arise when scalar fields are conformally coupled to gravity~\\cite{gr-qc/0003025, gr-qc/9908029}.) In braneworlds, it was found that it is a combination of the local high-energy bulk effects, and the nonlocal corrections from the Weyl curvature in the bulk, that may induce an effective NEC violating signature on the brane, while the ``physical''/``observable'' stress-energy tensor confined to the brane, threading the wormhole throat, nevertheless satisfies the energy conditions~\\cite{Lobo:2007qi}. An interesting and efficient manner to minimize the violation of the null energy condition, extensively analyzed in the literature, is to construct thin-shell wormholes using the thin-shell formalism \\cite{Visser,thinform,Israel} and the cut-and-paste procedure as described in~\\cite{Visser,VisserPRD,VisserNPB,VisserPLB,Poisson}. Motivated in minimizing the usage of exotic matter, the thin-shell construction was generalized to nonspherically symmetric cases~\\cite{Visser,VisserPRD}, and in particular, it was found that a traveler may traverse through such a wormhole without encountering regions of exotic matter. In the context of a (limited) stability analysis, in \\cite{VisserNPB}, two Schwarzschild spacetimes were surgically grafted together in such a way that no event horizon is permitted to form. This surgery concentrates a nonzero stress energy on the boundary layer between the two asymptotically flat regions and a dynamical stability analysis {(with respect to spherically symmetric perturbations)} was explored. In the latter stability analysis, constraints were found on the equation of state of the exotic matter that comprises the throat of the wormhole. Indeed, the stability of the latter thin-shell wormholes was considered for certain specially chosen equations of state \\cite{Visser,VisserNPB}, where the analysis addressed the issue of stability in the sense of proving bounded motion for the wormhole throat. This dynamical analysis was generalized to the stability of spherically symmetric thin-shell wormholes by considering linearized radial perturbations around some assumed static solution of the Einstein field equations, without the need to specify an equation of state \\cite{Poisson}. This linearized stability analysis around a static solution was soon generalized to the presence of charge \\cite{Eiroa}, and of a cosmological constant \\cite{LoboCraw}, and was subsequently extended to a plethora of individual scenarios \\cite{thinshellwh,Ishak}, some of them rather ad hoc. The key point of the present paper is to develop an extremely general, flexible, and robust framework that can quickly be adapted to general spherically symmetric traversable wormholes in 3+1 dimensions. We shall consider standard general relativity, with traversable wormholes that are spherically symmetric, with all of the exotic material confined to a thin shell. The bulk spacetimes on either side of the wormhole throat will be spherically symmetric and static but otherwise arbitrary (so the formalism is simultaneously capable of dealing with wormholes embedded in Schwarzschild, Reissner--Nordstr\\\"om, Kottler, or de~Sitter spacetimes, or even ``stringy'' black hole spacetimes). The thin shell (wormhole throat), while constrained by spherical symmetry, will otherwise be permitted to move freely in the bulk spacetimes, permitting a fully dynamic analysis. This will then allow us to perform a general stability analysis against spherically symmetric perturbations, where wormhole stability is related to the properties of the exotic matter residing on the wormhole throat. We particularly emphasize that our analysis can deal with geometrically imposed ``external forces'', (to be more fully explained below), a feature that has so far been missing from the published literature. Additionally we emphasize the derivation of rather explicit and very general rules relating the internal structure of the wormhole throat to a ``potential'' that drives the motion of the throat. This paper is organized in the following manner: In Sec. \\ref{secII} we outline in detail the general formalism of generic dynamic spherically symmetric thin-shell wormholes, and provide a novel approach to the linearized stability analysis around a static solution. In Sec. \\ref{secIII}, we provide specific examples by applying the generic linearized stability formalism outlined in the previous section. In Sec. \\ref{conclusion}, we draw some general conclusions. ", "conclusions": "In this work, we have developed an extremely general, flexible and robust framework, leading to the linearized stability analysis of spherically symmetric thin shells. The analysis is well-adapted to general spherically symmetric thin shell traversable wormholes and, in this context, the construction confines the exotic material to the thin-shell. The latter, while constrained by spherical symmetry is allowed to move freely within the bulk spacetimes, which permits a fully dynamic analysis. To this effect, we have considered in great detail the presence of a flux term, which has been widely ignored in the literature. This flux term corresponds to the net discontinuity in the conservation law of the surface stresses of the bulk momentum flux, and is physically interpreted as the work done by external forces on the thin shell. Relative to the linearized stability analysis, we have reversed the logic flow typically considered in the literature, and introduced a novel approach to the analysis. We recall that the standard procedure extensively used in the literature is to define a parametrization of the stability of equilibrium, so as not to specify an equation of state on the boundary surface \\cite{Poisson,Eiroa,LoboCraw}. More specifically, the parameter $\\eta(\\sigma)= d{\\cal P}/d\\sigma$ is usually defined, and the standard physical interpretation of $\\eta$ is that of the speed of sound. In this work, rather than adopt the latter approach, we considered that the stability of the wormhole is fundamentally linked to the behaviour of the surface mass $m_s(a)$ of the thin shell of exotic matter, residing on the wormhole throat, via a pair of stability inequalities. More specifically, we have considered the surface mass as a function of the potential. This novel procedure implicitly makes demands on the equation of state of the matter residing on the transition layer, and demonstrates in full generality that the stability of thin shell wormholes is equivalent to choosing suitable properties for the material residing on the thin shell. We have applied the latter stability formalism to a number of specific examples of particular importance: some presented to emphasize the features specific to possible asymmetry between the two universes used in traversable wormhole construction; some to emphasize the importance of NEC non-violation in the bulk; and some to assess the simplifications due to symmetry between the two asymptotic regions. In particular, we have considered the case of borderline NEC non-violation in the bulk. This is motivated by the knowledge that, on extremely general grounds, the NEC must be violated somewhere in the spacetime of a generic traversable wormhole, and if this were to happen in the bulk region, this would be equivalent to imposing $\\Phi'_\\pm(r)< 0$ in the bulk. Thus to minimize NEC violations for thin-shell wormholes, we have considered the specific case of $\\Phi_\\pm=0$, which is particularly interesting for its mathematical simplicity, and for its physical interest as it corresponds to the constraint that the bulk regions on either side of the wormhole throat be on the verge of violating the NEC. We have also considered the simplification when the two bulk regions are identical, and analysed the stability regions of asymmetric thin-shell Schwarzschild wormholes and thin-shell Reissner-Nordstr\\\"{o}m wormholes in great detail. It was instructive to consider explicit cases which violate some of the energy conditions in the bulk. For instance, we considered thin-shell variants of the Ellis wormhole and two new toy models, and explored the linearized stability analysis in the presence of zero momentum flux and non-zero external forces. Finally, we analyzed thin-shell dilatonic wormholes, where the exterior spacetime solutions corresponded to a combined gravitational-electromagnetic-dilaton system. In conclusion, by considering the matching of two generic static spherically symmetric spacetimes using the cut-and-paste procedure, we have analyzed the stability of spherically symmetric dynamic thin-shell traversable wormholes --- stability to linearized spherically symmetric perturbations around static solutions. The analysis provides a \\emph{general and unified framework} for simultaneously addressing a large number of wormhole models scattered throughout the literature. As such we hope it will serve to bring some cohesion and focus to what is otherwise a rather disorganized and disparate collection of results. A key feature of the current analysis is that we have been able to include ``external forces'' in the form of non-zero values for the metric functions $\\Phi_\\pm(r)$. (This feature is absent in much of the extant literature.) Another key aspect of the current analysis is the focus on $m_s(a)$, the ``mass'' of the thin shell of exotic matter residing on the wormhole throat --- and the realization that stability of the wormhole is fundamentally linked to the behaviour of this exotic matter via a pair of simple and relatively tractable inequalities." }, "1112/1112.0863_arXiv.txt": { "abstract": "Peculiar A stars are so named because they exhibit abundance peculiarities in their atmospheres. It is believed that these arise as a result of differentiation of chemical species in large magnetic spots in which convective mixing is inhibited: there might be just two antipodal spots, whose axis is inclined to the axis of rotation. Many of the Ap stars that are rotating slowly also pulsate, with periods substantially shorter than the period of the fundamental radial mode. The pulsations appear to be nonradial, but axisymmetric, with their common axis usually aligned with the axis of the spots. In this lecture I shall first discuss the magnetic suppression of convection in the spots, and then I shall try to explain the pulsation phenomenon, reviewing some of the suggestions that have been made to explain the alignment and the excitation mechanism, and finally raising some issues that need to be addressed. ", "introduction": "It was my privilege to be the first research student to be registered by the University of Cambridge under Roger Tayler's supervision. That did not make me Roger's first student in practice, however, because Roger was a very kindly man and would supervise other astrophysics students in the department who needed more help than their registered supervisors provided; and so there were already several students under Roger's wing when I arrived. Consequently Roger was at that time intellectually active in a variety of fields, embracing stellar structure and evolution, the physics of thermonuclear reactions, magnetohydrodynamics -- particularly stability theory, with which Roger had a great deal of experience -- and stellar pulsation. It is therefore appropriate that I devote this lecture to a subject that combines many of these fields. I have chosen to address some of the issues that arise from trying to understand the phenomenon of rapid oscillations in peculiar A stars. Roger did not himself work in this area, but, recognizing his interests, one can easily imagine that he might well have done so had his attention not been drawn more strongly to other matters. My own interest in rapidly oscillating Ap (roAp) stars was triggered by Donald Lynden-Bell, when in 1981, if I recall correctly, he told me on his return from a visit to the South African Astronomical Observatory of Don Kurtz's recent, and yet unpublished, discovery. What intrigued me was that the stars were spotted and that the oscillations were apparently dipolar, with axes aligned with the spots. There were only a few examples, but nevertheless they must surely have at least hinted very strongly that the oscillations are always (almost) aligned with the spots, and therefore do not wander far off under the influence of Coriolis precession. Soon after that time, N\\\"{o}el Dolez visited me from Paris, and we decided to carry out some simple calculations that might address the most obvious questions: why are only dipolar oscillations excited; why are their axes aligned with the spots; and why are only rapid oscillations observed, with frequencies much higher than that of the fundamental dipole mode? I address unashamedly in this lecture the picture that is emerging at least in my mind from that early work, and which has been substantially elaborated upon in recent years; I cannot here also review in detail the alternative suggestions that have been propounded. And I emphasize here that the line of argument that I follow is strongly influenced by what I learnt at Roger's feet when I was a student. ", "conclusions": "It is my opinion that a theory of the roAp-star phenomenon is beginning to emerge. The stars are slow rotators. They are spotted as a result of there being a large-scale predominantly dipole magnetic field pervading the star, inclined from the axis of rotation, which suppresses subsurface convection to form two antipodal spots where the field is almost vertical. This permits the establishment of chemical abundance anomalies brought about a combination of radiative levitation and gravitational settling against diffusion. There might also be a stellar wind from the spot, where the magnetic field lines are effectively open. This would modify the chemical abundances, as also might transport by fingering in regions where the mean molecular mass increases upwards, although this process, unfortunately, is not well understood. The oscillations are low-degree $\\rm{p}$ models of high order, which are probably axisymmetric with axes more-or-less aligned with the spots. They are excited by the $\\kappa$ mechanism in the spots, and damped by convection elsewhere. Only a few high-order modes are overstable, in accord with observation. Some calculations have found overstability also in some relatively-low-order modes, with much lower growth rates; perhaps such oscillations are present too, but at amplitudes too low yet to have been observed. An important matter that has never been seriously addressed is an explanation of the amplitudes of the modes. If the modes are overstable, their growth must be limited by some nonlinear process. Is that process similar to that which limits the growth of the oscillations of Cepheids and RR Lyrae stars? If so, why does it operate effectively at so low and amplitude? Only when the spots are sufficiently large, and the rotation of the star is sufficiently small, can axisymmetric oscillation modes be dynamically locked with their axes close enough to the spots for excitation to dominate over damping, causing overstability; nonaxisymmetric modes precess, due to Coriolis effects, and cannot be firmly locked. Because the oscillation amplitude is greatest on the axis of symmetry, overstability is most likely to result when the modes are aligned with the spots. Of course, the larger the spots, the less precise need be the alignment, and in the fullness of time a robust theoretical criterion for overstability might be found: a relationship between spot size and inclination of the oscillation axis from the spots. Nonaxisymmetric modes have lower growth rates, and are overstable only if the spots are very large; that they appear not to be observed would set an upper limit to the size of the spots, once a robust theory is available. In some circumstances the stellar rotation might be so large that the principal axis of pulsation is more-or-less aligned with the rotation axis, and not with the spots; in that case, it might be that other modes of the same degree precess slowly across the spot and become overstable only when they are appropriately orientated, giving the impression of dynamical alignment. The theoretical models are at present extremely primitive. The influence of the spots on the pulsations has been accommodated in a piecemeal way, and must in future be properly incorporated into the pulsation dynamics to produce a plausible quantitative model. There has been substantial progress in studying the influence of the Lorentz forces on the pulsations, but only in stellar models without spots. Lateral inhomogeneity of the background state, convection-pulsation interactions, Lorentz forces and nonadiabaticity must all be combined in a consistent way. And our understanding of the dynamics of the oscillations in the atmosphere, where nonlinearity in the dynamics can be important, and where the processes of amplitude limitation are probably the most effectual, must be improved. Only then can one reliably use future observations of mode frequencies and amplitudes, intensity-velocity phase relations and nonlinearities in the light-curves to draw reliable inferences about the structure of the stars. We have many (but not all) of the ingredients in hand. Therefore a preliminary theory might be almost in sight." }, "1112/1112.2782_arXiv.txt": { "abstract": "We demonstrate that the integrated gravitational wave signal of Type Ia supernovae (SNe Ia) in the single-degenerate channel out to cosmological distances gives rise to a continuous background to spaceborne gravitational wave detectors, including the Big Bang Observer (BBO) and Deci-Hertz Interferometer Gravitational wave Observatory (DECIGO) planned missions. This gravitational wave background from SNe Ia acts as a noise background in the frequency range 0.1 - 10 Hz, which heretofore was thought to be relatively free from astrophysical sources apart from neutron star binaries, and therefore a key window in which to study primordial gravitational waves generated by inflation. While inflationary energy scales of $\\gtrsim 10^{16}$ GeV yield inflationary gravitational wave backgrounds in excess of our range of predicted backgrounds, for lower energy scales of $\\sim10^{15}$ GeV, the inflationary gravitational wave background becomes comparable to the noise background from SNe Ia. ", "introduction": "Type Ia supernovae (SNe Ia) are powerful thermonuclear explosions of a white dwarf star that release a kinetic energy of $\\sim10^{51}$ ergs, and serve as standard candles for cosmology \\cite {phillips93, riessetal98, perlmutteretal99}. While the precise nature of the SNe Ia progenitor remains a topic of active investigation, current theoretical models \\cite {townsleyetal07, jordanetal09, hoflichetal10} of the detonation of a white dwarf star brought to a mass near the Chandrasekhar limit by a companion red giant or main sequence star are consistent with both the energetics and nucleosynthetic yield of the explosion, as well as its observed asymmetry \\cite {maedaetal10}. We refer to this particular mechanism of SNe Ia as the single degenerate channel (SD)\\citep{whelanetal73,nomotoetal84}, as opposed to the double degenerate (DD) channel, which consists of the merger and subsequent detonation of two degenerate white dwarfs \\citep {webbink84, ibentutukov84}. Two models for detonations in the SD channel have been studied extensively -- the deflagration-to-detonation transition (DDT) \\citep {khokhlov91, blondinetal11}, and the gravitationally-confined detonation (GCD). Utilizing detailed three-dimensional (3D) hydrodynamical simulations of the GCD mechanism of the SD channel of SNe Ia, \\cite{faltaetal11} demonstrated that the intrinsic asymmetry of the explosion gives rise to a gravitational wave (GW) signature. Based on the GCD model prediction, a nearby individual SNe Ia event may possibly be detected in the Milky Way and nearby galaxies by third-generation GW detectors in the $\\sim 2$ Hz frequency range. While the focus of the earlier paper \\citep {faltaetal11} was upon the GCD model for detonation in the SD channel, both the DDT and DD models are likely to yield off-centered detonations \\citep {ropkeetal07, vankerkwijketal10}. Recent work has demonstrated that the explosion energies and nucleosynthetic yields are similar in off-centered detonations in which only a small fraction of the total energy release occurs prior to detonation \\citep {chamulaketal11}. Since the great majority of the GW energy is radiated during the detonation itself, we expect both the DDT and DD models are likely to yield off-centered detonations whose gravitational wave signature is qualitatively similar to the GCD model prediction. Previous work explored the stochastic gravitational wave background due to the distribution of unresolved ``chirping'' binary white dwarfs, and demonstrated that this background obscures the inflationary signal between frequencies of 10$^{-5}$ Hz and 0.1 Hz \\cite {farmerandphinney03}. The frequency range between 0.1 - 10 Hz was previously thought to be relatively free from gravitational wave emission from astrophysical sources, with the primary contribution believed to originate from neutron star binaries \\cite {schneider01}, and possibly also core-collapse supernovae at cosmological distances, and Pop III (e.g., first-generation) stars \\cite {buonannoetal05}. In this paper, we consider an additional stochastic GW background source : SNe Ia at cosmological distances. If sufficiently strong, the SNe Ia stochastic background may be a source of noise to future detectors in the 0.1 - 10 Hz frequency range. Thus, the presence of a stochastic gravitational wave background due to SNe Ia may further impose additional challenges in extracting a possible inflationary GW background in this frequency range. The paper is organized as follows. In section II, we present the formalism underlying the computation of the cosmological gravitational wave background of stochastic sources. In section III, we discuss the cosmic rate of SNe Ia. In section IV, we review the gravitational wave signature of a single SNe Ia event predicted by the GCD mechanism. Finally, in sections V and VI, we present our results and conclude. ", "conclusions": "A population of purely SD GCD SNe Ia at cosmological distances will generate a continuous GW stochastic background within the 1 - 2 Hz frequency range that could be a source of noise to third-generation detectors seeking the inflationary GW signal. The SNe Ia GW stochastic signal strength is relatively insensitive to the DTD, with the prompt DTD producing a slightly larger signal than the $\\beta = -1.1$ power law. The SNe Ia background could be a significant source of noise for the ultimate DECIGO detector if the inflationary signal is at a level less than $\\Omega_{GW} h_0 \\simeq 10^{-19}$. However assuming a slow roll inflation that can be constrained by the CMB/LSS, the inflationary signal could be as strong as $\\Omega_{GW} h_0 \\lesssim 6\\times 10^{-15}$, in which case the SNe Ia background would be too weak to be a source of noise. If the SNe Ia GW signal strength is significantly stronger, such that the background is closer to the upper bound estimate, it could also be a source of noise to the correlated BBO. Previous work on the primordial inflationary gravitational wave spectrum has primarily focused upon neutron-star binaries as a source of noise in the frequency range of 0.1 - 10 Hz \\cite {schneider01}. The gravitational wave signal generated by these binaries is relatively clean. Therefore it seems that it may be possible to subtract the neutron-star binary spectrum to reveal the underlying primordial gravitational wave spectrum. In contrast, the physics underlying the SNe Ia is far more complex, involving both hydrodynamics and nuclear combustion. Furthermore, though supernova rates are relatively well-constrained from optical observations at low redshift, they are poorly constrained at the high-redshifts responsible for the background. In this respect, the numerous uncertainties surrounding the stochastic background of SNe Ia are similar to those of core-collapse SNe \\cite {buonannoetal05}. If the inflationary energy scale does lie close to $10^{15}$ GeV, the GW emission of SNe Ia will likely prove to be an additional source of noise to the primordial gravitational wave signal -- one which will be very challenging to extract. The authors acknowledge useful discussions with Gaurav Khanna. DF acknowledges support from NSF grant numbers PHY-0902026. RF acknowledges research support from NSF grant CNS-0959382 and AFOSR DURIP grant FA9550-10-1-0354. The software used in this work was in part developed by the DOE-supported ASC / Alliance Flash Center at the University of Chicago. This research was supported in part by the National Science Foundation through TeraGrid resources provided by the Louisiana Optical Network Initiative under grant number TG-AST100038. \\newcommand{\\aj}{Astronomical Journal} \\newcommand{\\apjl}{Astrophysical Journal Letters} \\newcommand{\\apjs}{Astrophysical Journal Supplement} \\newcommand{\\aap}{Astronomy and Astrophysics} \\newcommand{\\araa}{Annual Review of Astronomy and Astrophysics} \\newcommand{\\mnras}{Monthly Notices of the Royal Astronomical Society} \\newcommand{\\jcap}{Journal of Cosmology and Astroparticle Physics}" }, "1112/1112.0114_arXiv.txt": { "abstract": "We present multi-epoch Very Long Baseline Array (VLBA) observations of V773~Tau~A, the 51-day binary subsystem in the multiple young stellar system V773~Tau. Combined with previous interferometric and radial velocity measurements, these new data enable us to improve the characterization of the physical orbit of the A subsystem. In particular, we infer updated dynamical masses for the primary and the secondary components of 1.55 $\\pm$ 0.11 \\Msun, and 1.293 $\\pm$ 0.068 \\Msun, respectively, and an updated {\\it orbital parallax} distance to the system of 135.7 $\\pm$ 3.2 pc, all consistent with previous estimates. Using the improved orbit, we can calculate the absolute coordinates of the barycenter of V773~Tau~A at each epoch of our VLBA observations, and fit for its {\\it trigonometric parallax} and proper motion. This provides a direct measurement of the distance to the system almost entirely independent of the orbit modeling. The best fit yields a distance of 129.9 $\\pm$ 3.2 pc, in good agreement (i.e.\\ within 1$\\sigma$) with the distance estimate based on the orbital fit. Taking the mean value of the {\\it orbital} and trigonometric parallaxes, we conclude that V773~Tau is located at $d$ = 132.8 $\\pm$ 2.3 pc. The accuracy of this determination is nearly one order of magnitude better than that of previous estimates. In projection, V773~Tau and two other young stars (Hubble~4 and HDE~283572) recently observed with the VLBA are located toward the dark cloud Lynds 1495, in the central region of Taurus. These three stars appear to have similar trigonometric parallaxes, radial velocities, and proper motions, and we argue that the weighted mean and dispersion of their distances ($d$ = 131.4 pc and $\\sigma_d$ = 2.4 pc) provide a good estimate of the distance to and depth of Lynds~1495 and its associated stellar population. The radio emission from the two sources in V773~Tau~A is largely of gyrosynchrotron origin. Interestingly, both sources are observed to become typically five times brighter near periastron than near apastron (presumably because of increased flaring activity), and the separation between the radio sources near periastron appears to be systematically smaller than the separation between the stars. While this clearly indicates some interaction between the individual magnetospheres, the exact mechanisms at play are unclear because even at periastron, the separation between the stars ($\\sim$ 30 $R_*$) remain much larger than the radius of the magnetospheres around these low-mass young stars ($\\sim$ 6 $R_*$). ", "introduction": "In spite of significant progress in recent years, the formation and early evolution of stars are still not fully understood (e.g.\\ Hillenbrand \\& White 2004, Mathieu et al.\\ 2007). One way to foster progress is to measure as accurately as possible the intrinsic characteristics of individual young stars (luminosity, effective temperature, mass, disk properties, etc.), and compare them with the predictions of detailed theoretical models. Young binary systems are particularly interesting in this respect, because tracking their orbital motions provides a direct means to estimate their dynamical mass. In particular, if astrometric and spectroscopic data are combined, the physical orbit and the individual masses of the system members can be determined. A recurrent obstacle to the accurate determination of the intrinsic properties of young stars have been fairly large uncertainties (typically 20 to 50\\%) in distance estimates to even the nearest star-forming regions. Significant progress has been possible in recent years thanks to direct trigonometric parallax measurements obtained using Very Long Baseline Interferometry (VLBI) multi-epoch observations. Such results have been reported, in particular, in the previous papers in this series (Loinard et al.\\ 2005, 2007, 2008; Torres et al.\\ 2007, 2009; Dzib et al.\\ 2010, 2011). They provide uncertainties of a few percent or better, that typically surpass the accuracy of previous determinations by one order of magnitude. V773~Tau (HD~283447, HBC~367) is a young stellar system located toward the dark cloud Lynds~1495 in Taurus. V773~Tau was first identified as a T~Tauri star by Rydgren et al.\\ (1976), and was established as a visual double (components designated A and B) with an apparent separation of roughly 150 mas in high-angular resolution studies independently by Ghez et al.\\ (1993) and Leinert et al.\\ (1993). Roughly contemporaneously, Martin et al.\\ (1994) and Welty (1995) established the brighter (A) visual component as a short-period (51 days) double-lined spectroscopic binary. Duch\\^ene et al.\\ (2003; hereafter D2003) and Woitas (2003) independently identified a third visual component (designated C; note that D2003 use an alternate component notation) in the system, indicating that V773~Tau is (at least) a quadruple system.\\footnote{It should be noted that multiple nomenclatures have been used to describe the components of V773~Tau. D2003 designate the spectroscopic binary V773~Tau~A/B, and the optical and infrared companions V773~Tau~C and V773~Tau~D, respectively. Alternately White \\& Ghez (2001), Woitas (2003), and Massi et al.\\ (2008) among others, designate the spectroscopic binary as V773~Tau~A (containing stars Aa and Ab), and the two companions as B and C, respectively. A third nomenclature is used in the Double Star Catalog of Mason et al.\\ (2001), which designates the spectroscopic binary Aa, containing stars Aa1 and Aa2. In this work we will follow the notation established by White \\& Ghez (2001), reflecting the hierarchical nature of the V773~Tau architecture.} V773~Tau~A has long been known to be a strong radio source (Kutner et al.\\ 1986).\\footnote{The two companions V773~Tau~B and C, on the other hand, are not detected at radio wavelengths at the level of sensitivity of existing observations.} Indeed, it was the strongest source in the 5 GHz VLA survey of WTTS in the Taurus-Auriga molecular cloud complex by O'Neal et al.\\ (1990). From detailed multi-frequency observations, Feigelson et al.\\ (1994) concluded that the radiation was most likely of non-thermal origin. This was confirmed by Phillips et al.\\ (1996; hereafter P1996) who obtained VLBI observations, and resolved the radio emission into a clear double source, most likely corresponding to the two components of the spectroscopic binary. More recently, Massi et al.\\ (2002, 2006) showed that the radio emission exhibits periodic variations with a period corresponding to the 51 day orbital period of the spectroscopic binary. This variability is due to an increase in the flaring activity near periastron and might reflect interactions between extended magnetic structures associated with the two stars when they get close to one another. Finally, Boden et al.\\ (2007; hereafter B2007) and Massi et al.\\ (2008) also resolved the radio emission from V773~Tau~A into two components, which they associate with the two stars in the spectroscopic binary. The dynamics in V773~Tau has been studied by a number of authors. Relative orbital motion among the A, B, C components has been monitored by D2003, B2007 and Boden et al.\\ (2011; a companion paper to this one -- hereafter B2011). Following Welty (1995) and P1996, B2007 used radio and near-IR interferometry and spectroscopic radial velocity (RV) datasets to estimate the A subsystem physical (three-dimensional) orbit. They obtain dynamical mass estimates of 1.54 and 1.33 \\Msun\\ for the Aa and Ab components, respectively. Further, B2007 estimated the A subsystem distance by means of ``orbital parallax'' (comparing the angular and physical orbit size), yielding 136.2 $\\pm$ 3.7 pc. There is also a direct trigonometric parallax measurement based on multi-epoch VLBI observations for this source (Lestrade et al.\\ 1999). This VLBI-based distance measurement ($d$ = 148.4$^{+5.7}_{-5.3}$ pc) is roughly consistent (at the 2--3 $\\sigma$ level) with the value obtained from the A subsystem orbit modeling. In this paper we present new VLBI observations of V773~Tau~A which resolve the A subsystem, and yield additional insights on its structure. These data are described in \\S 2, and modeled jointly with earlier observations from B2007 in \\S3 to update the A subsystem physical orbit and resulting physical parameters (component dynamical masses, orbital parallax). Then the updated A orbit model is used in conjunction with the VLBA global astrometry to compute a new trigonometric parallax to V773~Tau~A. The relevance of these new results for the distance to the dark cloud Lynds~1495 is discussed in \\S~3.3. Several interesting features of the V773~Tau~A radio emission are apparent in these new observations; the insights that they yield on the magnetospheric physics of the A subsystem components are discussed in \\S~4.2. {\\scriptsize \\begin{deluxetable}{clcr} \\tablecaption{Project code and date for each observation.\\label{table_observations}} \\tablehead{ \\multicolumn{1}{c}{Epoch}& \\multicolumn{1}{c}{Project Code}& \\multicolumn{1}{c}{Mean UT date}& \\multicolumn{1}{c}{Julian Day}\\\\ {}& {}& {[yyyy.mm.dd ~ hh:mm]}& {}}% \\startdata 01 & BM 198 A & 2004.03.11 ~ 20:12 & 2453076.3417 \\\\ 02 & BM 198 B & 2004.03.12 ~ 20:12 & 2453077.3417 \\\\ 03 & BM 198 C & 2004.03.13 ~ 20:12 & 2453078.3417 \\\\ 04 & BM 198 D & 2004.03.14 ~ 20:12 & 2453079.3417 \\\\ 05 & BM 198 E & 2004.03.15 ~ 20:11 & 2453080.3414 \\\\ 06 & BM 198 F & 2004.03.16 ~ 20:11 & 2453081.3414 \\\\ 07 & BM 198 G & 2004.03.17 ~ 20:11 & 2453082.3414 \\\\ 08 & BL 128 AA & 2005.09.08 ~ 12:01 & 2453622.0013 \\\\ 09 & BL 128 AB & 2005.11.15 ~ 07:31 & 2453689.8136 \\\\ 10 & BL 128 AC & 2006.01.21 ~ 03:11 & 2453756.6327 \\\\ 11 & BL 128 AD & 2006.04.01 ~ 22:31 & 2453827.4386 \\\\ 12 & BL 128 AE & 2006.06.12 ~ 17:48 & 2453899.2423 \\\\ 13 & BL 128 AF & 2006.09.05 ~ 12:14 & 2453984.0102 \\\\ 14 & BL 146 B & 2007.08.23 ~ 13:06 & 2454336.0461 \\\\ 15 & BL 146 C & 2007.08.29 ~ 12:42 & 2454342.0295 \\\\ 16 & BL 146 D & 2007.09.05 ~ 12:15 & 2454349.0106 \\\\ 17 & BL 146 E & 2007.09.11 ~ 11:51 & 2454354.9943 \\\\ 18 & BL 146 F & 2007.09.16 ~ 11:32 & 2454359.9806 \\\\ 19 & BL 146 G & 2007.09.21 ~ 11:12 & 2454364.9669 \\\\ 20 & BL 146 H & 2007.09.27 ~ 10:48 & 2454370.9504 \\\\ 21 & BL 146 I & 2007.10.03 ~ 10:25 & 2454376.9342 \\\\ 22 & BL 146 J & 2007.10.09 ~ 10:01 & 2454382.9177 \\\\ 23 & BL 146 K & 2007.10.17 ~ 09:30 & 2454390.8960 \\\\ 24 & BL 146 L & 2007.10.23 ~ 09:06 & 2454396.8794 \\\\ 25 & BL 146 M & 2007.10.27 ~ 08:50 & 2454400.8684 \\\\ 26 & BL 146 N & 2007.11.17 ~ 07:28 & 2454421.8114 \\\\ 27 & BM 306 & 2009.09.27 ~ 08:07 & 2455101.8386 \\\\ \\enddata \\end{deluxetable} } ", "conclusions": "" }, "1112/1112.0004_arXiv.txt": { "abstract": "Microwave Kinetic Inductance Detectors, or MKIDs, have proven to be a powerful cryogenic detector technology due to their sensitivity and the ease with which they can be multiplexed into large arrays. A MKID is an energy sensor based on a photon-variable superconducting inductance in a lithographed microresonator, and is capable of functioning as a photon detector across the electromagnetic spectrum as well as a particle detector. Here we describe the first successful effort to create a photon-counting, energy-resolving ultraviolet, optical, and near infrared MKID focal plane array. These new Optical Lumped Element (OLE) MKID arrays have significant advantages over semiconductor detectors like charge coupled devices (CCDs). They can count individual photons with essentially no false counts and determine the energy and arrival time of every photon with good quantum efficiency. Their physical pixel size and maximum count rate is well matched with large telescopes. These capabilities enable powerful new astrophysical instruments usable from the ground and space. MKIDs could eventually supplant semiconductor detectors for most astronomical instrumentation, and will be useful for other disciplines such as quantum optics and biological imaging. ", "introduction": "Cryogenic detectors are currently the preferred technology for astronomical observations over most of the electromagnetic spectrum, notably in the far infrared through millimeter (0.1--3~mm)~\\cite{Bintley:2010p4108,Niemack:2008p4173,Carlstrom:2011p4239}, X-ray~\\cite{Kelley:2009p4337}, and gamma-ray~\\cite{Doriese:2007p4284} wavelength ranges. In the important ultraviolet, optical, and near infrared (0.1--5~$\\mu$m) wavelength range a variety of detector technologies based on semiconductors, backed by large investment from both consumer and military customers, has resulted in detectors for astronomy with large formats, high quantum efficiency, and low readout noise. However, these detectors are fundamentally limited by the band gap of the semiconductor (1.1 eV for silicon) and thermal noise sources from their high ($\\sim$100~K) operating temperatures~\\cite{Eisaman:2011p6692}. Cryogenic detectors, with operating temperatures on the order of 100~mK, allow the use of superconductors with gap parameters over 1000 times lower than typical semiconductors. This difference allows new capabilities. A superconducting detector can count single photons with no false counts while determining the energy (to several percent or better) and arrival time (to a microsecond) of the photon. It can also have much broader wavelength coverage since the photon energy is always much greater than the gap energy. While a CCD is limited to about 0.3--1~$\\mu$m, the new arrays described here are sensitive from 0.1~$\\mu$m in the UV to greater than 5~$\\mu$m in the mid-IR, enabling observations at infrared wavelengths vital to understanding the high redshift universe. This approach has been pursued in the past with two technologies, Superconducting Tunnel Junctions (STJs)~\\cite{Martin:2006p4412,Hijmering:2008p4501} and Transition Edge Sensors (TESs)~\\cite{Romani:2001p1716,Burney:2006p4521}. While both of these technologies produced functional detectors, they are limited to single pixels or small arrays due to the lack of a credible strategy for wiring and multiplexing large numbers of detectors, although recently there have been proposals for larger TES multiplexers~\\cite{Niemack:2010p4657}. Microwave Kinetic Inductance Detectors, or MKIDs\\cite{Day03}, are an alternative cryogenic detector technology that has proven important for millimeter wave astrophysics\\cite{Schlaerth:2010p3957,Roesch:2010p4655} due to their sensitivity and the ease with which they can be multiplexed into large arrays. MKIDs use frequency domain multiplexing~\\cite{Mazin:2006p5} that allows thousands of pixels to be read out over a single microwave cable. While the largest STJ array is 120 pixels~\\cite{Verhoeve:2006p3383} and the largest optical TES array is 36 pixels~\\cite{Burney:2006p4521}, the MKID arrays described below are 1024 pixels, with a clear path to Megapixel arrays. The ability to easily reach large formats is the primary advantage of MKID arrays. In this paper we describe the first photon-counting, energy-resolving ultraviolet, optical, and near infrared MKID focal plane array. These Optical Lumped Element (OLE) MKID arrays have significant advantages over semiconductor detectors like charge coupled devices (CCDs)~\\cite{Smith:2011p4039}. They can count individual photons with essentially no false counts and determine the energy and arrival time of every photon with good quantum efficiency. Their physical pixel size and maximum count rate is well matched with large telescopes. These capabilities enable powerful new astrophysical instruments usable from the ground and space. ", "conclusions": "" }, "1112/1112.2237_arXiv.txt": { "abstract": "We have discovered two compact sources of shocked H$_2$ 2.12-$\\mu$m emission coincident with \\object{Mol 160} (\\object{IRAS 23385+6053}), a massive star-forming core thought to be a precursor to an ultracompact \\ion{H}{2} region. The 2.12-$\\mu$m sources lie within 2$^{\\prime\\prime}$ (0.05 pc) of a millimeter-wavelength continuum peak where the column density is $\\ge$ 10$^{24}$ cm$^{-2}$. We estimate that the ratio of molecular hydrogen luminosity to bolometric luminosity is $>$ 0.2\\%, indicating a high ratio of mechanical to radiant luminosity. CS J=2$\\rightarrow$1 and HCO$^+$ J=1$\\rightarrow$0 observations with CARMA indicate that the protostellar molecular core has a peculiar velocity of $\\sim$ 2 km s$^{-1}$ with respect to its parent molecular cloud. We also observed 95 GHz CH$_3$OH J=8$\\rightarrow$7 Class I maser emission from several locations within the core. Comparison with previous observations of 44-GHz CH$_3$OH maser emission shows the maser sources have a high mean ratio of 95-GHz to 44-GHz intensity. Our observations strengthen the case that \\object{Mol 160} (\\object{IRAS 23385+6053}) is a rapidly accreting massive protostellar system in a very early phase of its evolution. ", "introduction": "Studying the earliest evolutionary phases of massive protostars is difficult, because they are rare and their formation times are brief. Mol 160/IRAS 23385+6053 (hereafter Mol 160) is one of a number of sources identified by Molinari et al. (1996, 1998a, 2000, 2002) as possible precursors to ultracompact (UC) \\ion{H}{2} regions (Wood \\& Churchwell 1989). They were selected on the basis of their cool IRAS colors, high infrared luminosities, absence of centimeter-wavelength continuum emission, H$_2$O maser activity, ammonia line emission, millimeter continuum observations, and presence of energetic molecular outflows. Based on follow-up observations with the Owens Valley Radio Observatory (OVRO) and the Infrared Satellite Observatory (ISO), Molinari et al. (1998b, hereafter M98) suggested that Mol 160 might be the first bona fide example of a massive Class 0 protostar, an object still accreting matter at a high rate from a massive envelope. It is generally accepted that outflows from low-mass young stellar objects (YSOs) are driven by magnetic stresses in accretion disks that eject some of the inflowing matter and carry off angular momentum (Pudritz \\& Norman 1983, 1986; Lovelace et al. 1987; K\\\"onigl 1989; Pelletier \\& Pudritz 1992; Wardle \\& K\\\"onigl 1993; Safier 1993; Paatz \\& Camenzind 1996; Shu et al. 2000). Optical and near-infrared observations often reveal jets and Herbig-Haro objects that show in detail the interaction of the underlying winds with the ambient medium. If massive stars are built up through similar accretion processes, one might expect them to have outflows that are scaled-up versions of their low-mass counterparts. However, observationally linking jets and outflows with their sources is challenging, even in regions forming intermediate-mass stars (e.g., Wolf-Chase et al. 2003). The problems are compounded for massive stars, which often form in tight clusters, lie in heavily obscured regions at distances greater than several kiloparsecs, and may be evolving on dynamical time scales less than a few thousand years. Observations of molecular hydrogen lines in the near infrared are particularly valuable, since they directly measure kinetic energy deposited by high-speed shocks and can be observed with large ground-based telescopes at angular resolutions similar to those that can be achieved with millimeter wavelength interferometry. In this paper, we present narrowband, 0.7$^{\\prime\\prime}$-resolution, near-infrared images of Mol 160 made with the Apache Point Observatory (APO) 3.5-m telescope and 2$^{\\prime\\prime}$-resolution 3-mm continuum and line observations with the Combined Array for Research in Millimeter-wave Astronomy (CARMA). The near-infrared measurements were made as part of an emission-line survey for shocked molecular hydrogen gas associated with massive protostars (Wolf-Chase et al. 2011, in preparation). In the Mol 160 region, we detected a number of compact molecular hydrogen emission-line sources. Surprisingly, two of the brightest lie within 2$^{\\prime\\prime}$ of a millimeter-wavelength continuum peak where the hydrogen column density is $\\ge$ 10$^{24}$ cm$^{-2}$. Our CARMA measurements include CS, a tracer of dense gas, HCO$^+$, a tracer of cloud kinematics, and a Class I maser transition of CH$_3$OH, a tracer of shock-excited gas. Our observations support the hypothesis that Mol 160 is a massive protostellar system in a very early phase of evolution, but call into question some of the specific conclusions drawn from earlier outflow studies of this object (see \\S 4.3.1). Prior work on Mol 160 includes observations covering a wide range of wavelengths and angular resolutions, from the arcminute scales of the IRAS measurements to the arcsecond scales of interferometric observations. This can potentially lead to ambiguities in the association of source names with specific physical systems. In this paper, we follow the nomenclature generally used in the literature (e.g., Wang et al. 2011 and references therein). We use ``clumps'' to refer to parsec-scale molecular cloud structures and``cores'' to the smaller structures ($\\sim$ 0.1 pc) that may further collapse and coalesce into stars or groups of stars. We will use the term ``Mol 160'' to narrowly denote the protostar (or protostellar cluster) and the dense molecular core from which it is forming (diameter $\\le$ 6$^{\\prime\\prime}$) and ``Mol 160 region'' to refer to the arcminute-scale area that also encompasses potentially related structures including larger molecular clouds and two \\ion{H}{2} regions. \\section {OBSERVATIONS AND DATA REDUCTION} \\subsection{Apache Point Observatory: NICFPS} We obtained narrowband, near-infrared images using the Near-Infrared Camera and Fabry-Perot Spectrometer (NICFPS) on the Astrophysical Research Consortium (ARC) 3.5-m telescope at APO in Sunspot, NM. NICFPS uses a 1024 $\\times$ 1024 Rockwell Hawaii HgCdTe detector with 18-$\\micron$ pixels. On the APO 3.5 m, this yields a pixel scale of 0.273$\\arcsec$ pixel$^{-1}$ and gives a field of view of 4.6$\\arcmin \\times 4.6\\arcmin$ (Vincent et al. 2003). We used the following filters: K$_s$ (500 s), H$_2$ 2.12-$\\micron$ (1800 s), H$_2$ 2.25-$\\micron$ (1200 s), and a narrow-band H$_2$-continuum filter centered on 2.13 $\\micron$ (1200 s). Because of weather and scheduling constraints, the H$_2$ 2.12-$\\micron$ and K$_s$ data were recorded on 2005 Jun 17 and the remaining data were obtained on 2007 Nov 21. Seeing conditions were good the first night ($\\sim0.7\\arcsec$ at 2.2 $\\micron$) and average the second night ($ \\sim1.3\\arcsec$ at 2.2 $\\micron$). All images were processed using IRAF and standard reduction techniques (e.g. Joyce 1992). We dithered the telescope between each exposure in a five-point pattern, and used the median of the resulting images to subtract the sky background from each image. To produce final images in each filter, we used stars to register the reduced images and then took the median of the aligned frames. We produced a continuum-subtracted H$_2$ 2.25-$\\micron$ image by taking the difference between H$_2$ 2.25-$\\micron$ and H$_2$-continuum images. Ideally, we would have done the same for the H$_2$ 2.12-$\\micron$ data, but it was not possible to obtain H$_2$-continuum filter data during our 2005 run. Additionally, the instrument electronics were upgraded between the two runs, which resulted in different read noise and gain levels on the chip that complicated the comparison of 2005 and 2007 data. For these reasons, we chose to use K$_s$ data taken on the same night and thus with the same electronics and seeing conditions as the H$_2$ 2.12-$\\micron$ data, scaling the K$_s$ image so that the stars disappeared when we subtracted the K$_s$ image from the H$_2$ 2.12-$\\micron$ image. Both continuum subtraction techniques worked well on all but the brightest stars, which show small residual artifacts. We performed aperture photometry on compact sources using the DAOPHOT package in IRAF. To avoid source confusion while still maintaining good signal-to-noise, we chose an aperture radius of 8 pixels (2.2$\\arcsec$), which is roughly 3 times the FWHM of the stars in the H$_2$ 2.12-$\\micron$ images. The data were calibrated by comparing the instrumental magnitudes of several dozen stars to their K$_s$ magnitudes in the 2MASS point source catalog. For each filter, we calculated a zero point correction between our instrumental magnitudes and the 2MASS K$_s$ magnitudes, then computed fluxes using: $$ F_{Jy} = F_0 \\times 10^{-Kmag/2.5} \\eqno{(1)} $$ $$ F_{line} = F_{Jy} \\times 10^{-26} \\times d\\nu \\eqno{(2)} $$ \\noindent where F$_0 =$ 666.7 Jy and $d\\nu = c d\\lambda/\\lambda^2$. For the H$_2$ 2.12-$\\micron$ filter, $d\\lambda = 6.93$ nm and for the H$_2$ 2.25-$\\micron$ filter, $d\\lambda = 7.6$ nm (Apache Point Observatory 2010). WCS coordinates were calibrated using 2MASS point sources. The 2MASS positions are accurate to within 0.1$^{\\prime\\prime}$ for K$_s$ $<$ 14 (Skrutskie et al. 2006). Given the seeing conditions ($\\sim$ 0.7$^{\\prime\\prime}$) during acquisition of the NICFPS data and pixel size of 0.27$^{\\prime\\prime}$, we estimate that our positions are accurate to within $\\pm$ 0.3$^{\\prime\\prime}$. \\subsection{CARMA} We observed a field around the Mol 160 core approximately 1.5$^{\\prime}$ across with CARMA on 2010 Mar 23 and 24 during three tracks. The interferometer was in the C configuration with baselines ranging from 26-370 meters. We simultaneously observed the continuum and 3 spectral lines. At the time, CARMA was configured to observe in both upper and lower sidebands of three spectral bands, and a fourth was being integrated into the system. For the first track, two bands were configured for maximum continuum bandwidth, covering 468 MHz each, for a total of 1.872 GHz (including both sidebands). A third observed CH$_3$OH J=8$\\rightarrow$7 at 95.169 GHz in the upper side band with a 31-MHz bandwidth and 63 channels, resulting in a velocity resolution of 1.538 km s$^{-1}$. The fourth band observed HCO$^+$ J=1$\\rightarrow$0 at 89.189 GHz in the lower side band with 62-MHz bandwidth and 383 channels, giving a resolution of 0.547 km s$^{-1}$. There were calibration problems with this band; the upper side band data were corrupted, so the fourth band was not used during the other tracks. For the second and third tracks, we used only three bands. One observed continuum, for a total of 936-MHz bandwidth. The band observing CH$_3$OH was unchanged. The third band observed the HCO$^+$ line in the lower side band and CS J=2$\\rightarrow$1 in the upper side band at 97.981 GHz. It was set to a 62-MHz bandwidth with 63 channels, yielding 3.283 km s$^{-1}$ resolution for HCO$^+$ and 2.988 km s$^{-1}$ resolution for CS. The passband calibrator was 3C84 for the first track, 1927$+$730 on the second track, and 3C454.3 on the third. The flux calibrator was Uranus for the first and second track and MWC349 on the third. The phase calibrator was 0102$+$584 for all tracks. In our analysis, we combined CS and HCO$^+$ data from the second and third tracks and CH$_3$OH and continuum data from all three tracks. The total integration time on source was 2.35 hours for the second and third tracks and 4.75 hours for all three tracks. The size (FWHM) of the synthesized beam was 1.982$^{\\prime\\prime}\\times$1.811$^{\\prime\\prime}$. We reduced and calibrated the data using the MIRIAD data reduction package (Sault et al. 1995). We used an intermediate Brigg's visibility factor to balance sensitivity to compact structure in the Mol 160 core with suppression of sidelobes that might otherwise distort clump-scale emission. Coverage of the u-v plane over the multiple tracks contains numerous shorter baselines, so we expect good qualitative rendering of structures $\\le$30$^{\\prime\\prime}$ in size, though some fraction of the flux will be missing. \\section {RESULTS} \\subsection{H$_2$ Emission in the Mol 160 Region} Our H$_2$ 2.12-$\\micron$ and continuum-subtracted H$_2$ 2.12-$\\micron$ images are shown in Figures 1 \\& 2, respectively. Positions and designations of the 24 $\\micron$ point sources (crosses) presented in Table 1 of Molinari et al. (2008a, hereafter M08) are plotted in Figure 2. Source ``A'' lies close to the Mol 160 molecular core. Some of the pure line emission in Figure 2 arises from fluorescently excited gas in photo-dissociation regions (PDRs) associated with two \\ion{H}{2} regions (indicated by the large ovals in Figure 1) and some from shock-excited gas. In a catalog hosted by the Joint Astronomy Centre in Hawaii, Davis et al. (2010) have established a numbering scheme for ``molecular hydrogen emission-line objects'' or ``MHOs'', compact emission-line sources thought to be associated with outflows. We have identified 10 candidate MHOs and numbered them according to this scheme. Their positions are plotted with open diamonds in Figure 2 and identified by the last two digits of the MHO numbers. Table 1 lists the MHO designations (column 1), their positions (columns 2 \\& 3), and their 2.12-$\\micron$ line fluxes (column 4). None of the MHOs were detected in the H$_2$ 2.25-$\\micron$ filter. Using equations (1) \\& (2) to calibrate the faintest stars in the H$_2$ 2.25-$\\micron$ image against stars also present in the 2MASS catalog, we derive a 3$\\sigma$ upper limit for H$_2$ 2.25-$\\micron$ line intensity of $9.49 \\times 10^{-19}$ W m$^{-2}$. Column 5 lists the ratios of the 2.12-$\\micron$ fluxes to the 2.25-$\\micron$ upper limit. The difference in extinction at 2.12 $\\micron$ and 2.25 $\\micron$ amounts to only a tenth of a magnitude (assuming A$_{\\lambda} \\sim {\\lambda}^{-1.7}$), with a corresponding brightness ratio of $\\sim$1.1, which would change the numbers in column 5 only slightly. The expected 2.12-$\\micron$/2.25-$\\micron$ line ratio for UV excitation (as in PDRs) is $\\sim$ 1.9, and $\\sim$ 7.7 for shocks (Black \\& van Dishoeck 1987; Gredel \\& Dalgarno 1995). Shock excitation is consistent with the observed ratios for all the MHOs and is very likely for the brightest. Star formation is clearly occurring throughout the Mol 160 region. MHOs 2921 \\& 2922 are of particular interest because of their proximity to the Mol 160 3.2-mm continuum peak. At the present time, there is no evidence linking any of the other MHOs to the Mol 160 core itself. MHO 2922 is extremely compact, appearing almost stellar in the 2.12-$\\micron$ image (see Figure 3a). Comparison with stellar images of comparable peak intensity yields a deconvolved source size of 0.7$^{\\prime\\prime} \\times 0.6^{\\prime\\prime}$. Assuming a distance of 4.9 kpc to Mol 160 (M98)\\footnote{Zhang et al. (2005) used a distance of 6.9 kpc, and M08 suggest the distance may be closer to 8 kpc.}, this corresponds to a physical size of $\\sim$ 3430 AU. MHO 2921 is slightly larger and more irregular in appearance, particularly at low intensity levels. It is displaced to the south of the peak position of the 3.2-mm continuum source, and there is a very dark region symmetrically displaced to the north, with a low-intensity arc of H$_2$ 2.12-$\\micron$ emission at its northern edge, as can be seen in the continuum-subtracted 2.12-$\\micron$ image (Figure 3b). Fitting a two-dimensional Gaussian to the bright central region of MHO 2921 yields a deconvolved size of 1.4$^{\\prime\\prime} \\times 0.7^{\\prime\\prime}$. Subtracting the Gaussian fits to MHO 2921 and MHO 2922 from the continuum-subtracted H$_2$ image of Figure 3b results in the image shown in Figure 3c. The residual emission has the form of a ring of approximately the same diameter as the outer contours of the -53.7 km s$^{-1}$ CS and -52.5 km s$^{-1}$ HCO$^+$ peaks seen in Figure 4. The ring also coincides with a number of CH$_3$OH maser sources. The only feature of comparable intensity within the ring is a patch of emission near the positions of Masers 3 and 4 (see \\S 3.2.2). In Figure 3, the relative intensities at the peaks along the ring and the peaks at MHO 2921 \\& MHO 2922 are roughly 1:4:10. Note that the intensities of the dark area and arc north of the main continuum source are the same in Figures 3b and 3c. That is, they are not artifacts of the subtraction. \\subsection{CARMA Results} \\subsubsection{HCO$^+$, CS, and 3.2-mm Continuum Emission} The 3.2-mm (93.6 GHz) continuum emission peaks at the position 23$^h$40$^m$54.5$^s$, +61$^{\\circ}$10$^{\\prime}$28.1$^{\\prime\\prime}$. The beam size was 2.0$^{\\prime\\prime}$ $\\times$ 1.8$^{\\prime\\prime}$, with a position angle of 40$^{\\circ}$. A two-dimensional Gaussian fit to our continuum data yields a size of 3.3$^{\\prime\\prime}$ $\\times$ 2.6$^{\\prime\\prime}$, with a position angle of 40$^{\\circ}$. The deconvolved source size is 2.7$^{\\prime\\prime}$ $\\times$ 1.9$^{\\prime\\prime}$ with a position angle of 35$^{\\circ}$. We measured a peak intensity of 10.8 mJy beam$^{-1}$. Integrating under the Gaussian fit yields a total flux of 21.6 mJy. Our position lies within 0.4$^{\\prime\\prime}$ of the positions measured by M98 and Fontani et al. (2004, hereafter F04) at 3.4mm (88 GHz) and 3 mm ($\\sim$100 GHz), respectively. M98 measured a peak intensity of 9 mJy beam$^{-1}$ in a 4.1$^{\\prime\\prime}$ $\\times$ 3.5$^{\\prime\\prime}$ beam, a total flux of 19 mJy, and derived a deconvolved source size of 4.5$^{\\prime\\prime}$ $\\times$ 3.6$^{\\prime\\prime}$. F04 measured a peak flux of 8 mJy beam$^{-1}$ in a 2.3$^{\\prime\\prime}$ $\\times$ 1.9$^{\\prime\\prime}$ beam, a total flux of 12.4 mJy (integrated to the 3$\\sigma$ contour), and a deconvolved source size of 1.5$^{\\prime\\prime}$. Our measured flux at 3.2 mm is consistent with the SED for Mol 160 presented in Fig. 4 of M08, and does not change their conclusions regarding envelope properties or integrated bolometric luminosity (3170 L$_{\\odot}$, assuming a distance of 4.9 kpc). Maps of the three CS and three HCO$^+$ channels containing significant signal are shown in Figure 4 and Figure 5. The bandwidths of each channel are $\\sim$ 3 km s$^{-1}$ (3.283 km s$^{-1}$ for HCO$^+$ and 2.988 km s$^{-1}$ for CS). The central frequencies of the HCO$^+$ channels lie near the boundaries of the CS channels. Figure 4 shows the data at full angular resolution. In Figure 5, the data have been smoothed using a Gaussian kernel with a radius of 4$^{\\prime\\prime}$, to emphasize structure with scale lengths between those of the full-resolution interferometer data and measurements made with single-dish telescopes. Contours of velocity-integrated CS emission are plotted on the H$_2$ image in Figure 1. Although the interferometer data do not recover all the flux at larger scales, the data have been smoothed on scales to which CARMA is still sensitive, so the structures seen in the smoothed maps are either real emission, or real emission contaminated by side-lobe artifacts. To test for the latter, we compared the smoothed, velocity-integrated CS emission with the CARMA PSF smoothed to the same radius. We found no correlation of CS emission features with side-lobe features, and therefore conclude that the emission features seen in the smoothed maps are real. Qualitatively, Figure 5 is more sensitive to parsec-scale ``clumps'' and Figure 4 to sub-parsec-scale ``cores''. Furthermore, we note that the CS clumps in the smoothed maps correlate well with depressions in the extended H$_2$ emission (Figure 1), as would be expected for dense clumps in the foreground of a bright PDR background. In the smoothed maps, clump-scale emission is most evident in the CS channel centered at -50.7 km s$^{-1}$ and the HCO$^+$ channels centered at -49.2 and 52.5 km s$^{-1}$. The most prominent clump is the one containing Mol 160. Near the position of the 3.2-mm continuum peak, the velocity distribution is broader and ``bluer'' than at other positions in the clump, and the compact nature of the Mol 160 core is most evident in the CS channel centered at -53.7 km s$^{-1}$ and the HCO$^+$ channel centered at -52.5 km s$^{-1}$. \\subsubsection{CH$_3$OH Masers} Kurtz, Hofner, \\& \\'Alvarez (2004, hereafter KHA) detected 44-GHz CH$_3$OH Class I maser emission at four positions near the Mol 160 core with a synthesized beam of $2.01^{\\prime\\prime} \\times 1.25^{\\prime\\prime}$ with the VLA in its D configuration. We found 95.169-GHz CH$_3$OH J=8$\\rightarrow$7 emission at all the positions they reported and also at a position south of the Mol 160 core continuum peak. We obtained a copy of the 44-GHz data (S. Kurtz, private communication), and compared it with our CARMA observations. Examination of the individual channel maps revealed 44-GHz emission at the position of our southern 95-GHz source and at slightly lower but statistically significant intensities at several other positions. The channel maps also show that the KHA Source 3 position lies between two partially resolved sources, one to the north and one to the east of the Mol 160 core continuum peak. At the 0.166 km s$^{-1}$ resolution of the 44-GHz data, many of the sources shift in position from channel to channel, particularly Source 2, which extends in velocity across 6 channels and traces a path several arcseconds long. In Table 2 we list the peak positions and velocities of all the 44-GHz sources. We have extended the nomenclature used by KHA to include the three additional masers with intensities below the cut of their survey data and to reflect the resolution of their Source 3 into two spatial components. To more directly compare the intensities of the sources at 95 GHz and 44 GHz, we velocity-binned the KHA data to the channel resolution of the 95-GHz data and smoothed the higher angular resolution VLA data to match the CARMA beam. The resulting maps are shown in Figure 6. Table 2 compares the properties of the 95-GHz and 44-GHz sources. The positions listed in Table 2 correspond to positions of peak single-channel intensity of the 44-GHz observations, which display more structure than the 95-GHz observations because of their higher angular resolution. The large diamonds in the panels identify the four KHA positions. Small diamonds indicate the positions of the sources with intensities lying below the KHA cutoff (Sources 5, 6, and 7) and the two partially-resolved components of Source 3 (3N and 3E). The crosses in the panels show the positions of peak single-channel 95-GHz intensity. Column 7 in Table 2 lists the peak single-channel 44-GHz intensities. Columns 8 and 9 comprise, respectively, the 44-GHz intensities integrated over a velocity range corresponding to the 95-GHz channel bandwidth and the 95-GHz peak intensities at the positions of Column 1 and 2. Column 10 lists the ratios of the 95-GHz intensities to the smoothed, velocity-integrated 44-GHz intensities. ", "conclusions": "In their review of high mass star formation, Zinnecker and Yorke (2007) suggest four evolutionary phases. (1) Compression: the formation of dense, cold, gravitationally bound cores with masses $\\ge$ 100 M$_{\\odot}$ from a larger molecular cloud by gravo-turbulent cloud fragmentation. (2) Collapse: gravitational collapse of portions of the core into one or more quasi-hydrostatic protostellar embryos, accretion disks, and infall envelopes. (3) Accretion: accretion of matter onto protostellar objects, eventually leading to one or more massive stars. (4) Disruption: disruption of the natal core by stellar winds, outflows, ionizing radiation, or supernovae. In practice, linking details of these proposed evolutionary categories to observations of massive star-forming regions is complicated by many factors. The total radiant luminosity of a massive protostellar core may arise from a combination of Kelvin-Helmholtz emission from quasi-static protostellar embryos, protostellar and disk accretion shocks, and shocks associated with outflows or relative motions of core components. During the earliest evolutionary phases, there could also be significant contributions from external heating sources. The relative importance of each could vary dramatically and perhaps erratically with time, depending on the detailed dynamics of what may be a complex and chaotically evolving system. The expected conversion efficiency of gas to stars is sufficiently small that the mass of a core may not change dramatically until it enters the disruption phase (see, e.g., the theoretical tracks of core mass vs. luminosity in Molinari et al. 2008b). However, its temperature, density structure, chemistry, and kinematics may undergo significant alterations throughout the compression, assembly, and accretion phases. The ratio of mechanical luminosity from outflows to total radiant luminosity may also vary in a complicated way. In a general sense, high accretion rates probably correlate with high outflow rates, but kinetic energy can be stored and released asynchronously. Additionally, outflows in massive star-forming regions may be produced through different mechanisms. Two distinctly different types of outflows have been observed in some of the nearest massive star-forming regions. Whereas the outflow from IRAS 20126$+$4104 appears to be a scaled-up version of the disk-mediated accretion outflow scenario associated with low-mass YSOs (Caratti o Garatti et al. 2008), outflow from the Orion BN/KL complex has been attributed to the explosive disintegration of a massive star cluster about 500 years ago (e.g., Rosenthal et al. 2000; Zapata et al. 2009; Bally et al. 2011). Since most massive star-forming regions (including Mol 160) are too distant to fully resolve outflow structure with current instruments, the frequency of occurrence of these fundamentally different types of outflow events, and their effects on evolving protoclusters, remains to be seen. For example, Mol 160 is approximately ten times farther away than OMC-1. At this distance, the entire BN/KL complex, including both the ``18 km s$^{-1}$'' SiO outflow and the higher velocity gas associated with the shocked H$_2$ ``bullets'' or ``fingers'' (Allen and Burton 1993; Rosenthal et al. 2000; Colgan et al. 2007), would span $\\le$6$^{\\prime\\prime}$, similar to the size of the Mol 160 core. Theoretical models can offer some insights but are poorly constrained by actual observations. The observational problem at hand is how detailed studies of Mol 160 and other young objects can illuminate the phenomenology of massive star formation and place constraints on possible models. \\subsection{Environment of Mol 160} The molecular cloud containing Mol 160 was studied by F04. They observed C$^{18}$O J=1$\\rightarrow$0, C$^{17}$O J=1$\\rightarrow$0, and C$^{17}$O J=2$\\rightarrow$1 with the IRAM 30-m telescope with 23$^{\\prime\\prime}$, 22$^{\\prime\\prime}$, and 11$^{\\prime\\prime}$ beams, respectively. At the position of the 3-mm continuum peak, they found there were two distinct kinematic components, centered at -50.5 and -47.8 km s$^{-1}$. Maps over a 40$^{\\prime\\prime} \\times 40^{\\prime\\prime}$ area showed that the -47.8 km s$^{-1}$ emission peaked south of the position of the Mol 160 continuum peak and that the peak of the -50.5 km s$^{-1}$ emission was extended in a N-S direction and displaced about 4$^{\\prime\\prime}$ to the west. They suggested that the lack of a prominent feature at the position of the Mol 160 core might be explained by depletion of molecular gas on grains in the cool, high-density core. They also mapped NH$_3$ (1,1) over the same region with the VLA, using a 4.3$^{\\prime\\prime} \\times 3.7^{\\prime\\prime}$ beam. They found the emission was extended in a N-S ridge parallel to the CO peak, but displaced only 1$^{\\prime\\prime}$-2$^{\\prime\\prime}$ to the west of Mol 160. At the declination of Mol 160, there was actually a depression in the ridge, with the peak intensity occurring $\\sim$ 5$^{\\prime\\prime}$ to the south, again suggesting depletion in the core. In NH$_3$, they saw only the -50.5 km s$^{-1}$ component, and they concluded that the interferometer was resolving out the -47.8 km s$^{-1}$ component seen in single-dish observations with the Effelsberg 100-m telescope. Their CO maps also showed that the -47.8 km s$^{-1}$ emission was more extended E-W than the -50.5 km s$^{-1}$ component. Henceforth, we will refer to these components as the -50.5 and -47.8 km s$^{-1}$ clouds. On a larger scale, Molinari et al. (2002) observed 3.6-cm continuum emission from two extended \\ion{H}{2} regions lying to the west and east of Mol 160. They named them VLA 1 and VLA 2, respectively. Their ionization fronts almost intersect at a position just to the east of the Mol 160 core (see the large ovals on Figure 1). They appear to be bounded along portions of their ionization fronts by dense molecular clouds. In such cases, the high optical depths and densities toward the cloud result in high-surface-brightness H$_2$, polycyclic aromatic hydrocarbon (PAH), and dust emission from layers parallel to the ionization front, leading to particularly intense emission along tangent lines of sight. For the purposes of this discussion, we will broaden the conventional definition of the term PDR to include not only the zone just outside the ionization front but also the high-density zone inside the front where grains flowing into the \\ion{H}{2} region from the molecular cloud are exposed to high radiation densities and emit strongly at mid-infrared wavelengths. Contours of velocity-integrated CS emission are plotted on the H$_2$ image in Figure 1. We smoothed the CS data using a Gaussian kernel with a 4$^{\\prime\\prime}$ radius in order to highlight emission from parsec-scale structures generally referred to as ``clumps'' (Wang et al. 2011 and references therein). Most of the flux seen in the integrated map is also present in the single velocity channel centered at -50.7 km s$^{-1}$. In the H$_2$ images, one can see dark nebulosity overlying a brighter background of fluorescent emission from PDRs excited by the stars ionizing the \\ion{H}{2} regions. A similar morphology can be seen in the 15-$\\micron$ ISO image of M98. The molecular clump containing the Mol 160 core lies in a particularly dark band whose E-W extent is comparable to the width of the -50.5 km s$^{-1}$ CO cloud. It seems likely that the -50.5 km s$^{-1}$ cloud and the Mol 160 core are associated with this dark nebulosity and that they lie on the near side of the \\ion{H}{2} regions and their associated PDRs. The CS clumps lie along the eastern side of the dark band, to the east of the CO ridge and just to the west of the intense arc of PDR emission at the western boundary of VLA 2. M08 observed Mol 160 with MIPS on Spitzer at 24 $\\micron$ and 70 $\\micron$. The prominent arcs of emission in their 24-$\\micron$ map all have analogs in our 2.12-$\\micron$ image. The extended 24-$\\micron$ and fluorescent H$_2$ radiation probably arise from contiguous regions in the PDR, within the first few optical depths at FUV wavelengths. Our arcsecond-resolution H$_2$ data suggest that much of the flux that they attribute to embedded 24-$\\micron$ point sources (the positions for which are shown in Figure 2) could arise from knots of intense PDR emission. Separating truly point-like mid-infrared sources will require significantly better angular resolution than possible with Spitzer. The exciting stars of VLA 1 and VLA 2 may as yet be undetected. It is likely that much of the 70-$\\micron$ emission also arises from the PDRs. The overall morphologies of the Spitzer 24-$\\micron$ and 70-$\\micron$ maps made by M08 are similar. One can plausibly ascribe most of the differences to the fact that the hotter grains that dominate the 24-$\\micron$ emission lie close to or within the \\ion{H}{2} regions, where the stellar radiation is most intense. The principal exception is the far infrared peak near the position of Mol 160. Even in that case, the relative importance of internal and external heating is uncertain. The 70-$\\micron$ flux peaks $\\sim$ 5$^{\\prime\\prime}$ northeast of the Mol 160 continuum source. The discrepancy could result from positional uncertainties, as they suggest, but it is perhaps as likely that the shift is real and a significant contribution to the peak intensity in the 22$^{\\prime\\prime}$ Spitzer beam comes from nearby PDR emission, as is the case in arcminute-scale observations of the Trapezium/BNKL region in Orion (Harper 1974). Because the column density of the Mol 160 core is sufficiently high to absorb mid-infrared flux from the PDR, there could also be a significant external contribution to the heating of the core. For these reasons, the value M08 derive for the radiant luminosity of the Mol 160 core should be considered an upper limit. \\subsection{The Mol 160 Core} F04 have summarized estimates of the mass, column density, volume density, and temperature of the Mol 160 core derived from both line and continuum data taken with single dish telescopes and interferometers at angular resolutions between 1.5$^{\\prime\\prime}$ and 18$^{\\prime\\prime}$. For a region with diameter of $\\sim$4$^{\\prime\\prime}$-8$^{\\prime\\prime}$ and an assumed distance of 4.9 kpc, they find a kinetic temperature of 26 K, masses of $\\sim$100-400 M$_{\\odot}$, H$_2$ column densities of $\\sim$ 1.5-4 $\\times$10$^{24}$ cm$^{-2}$, and space densities of $\\sim$ 3.4-16 $\\times$10$^{6}$ cm$^{-3}$, with higher numbers from continuum data than from molecular lines. They note that there is evidence that molecules are depleted onto grains in the densest part of the core and suggest that the higher values should be preferred. For a region 1.3$^{\\prime\\prime}$-1.9$^{\\prime\\prime}$ in diameter ($\\sim$ 0.03 pc), they derive a kinetic temperature of 42 K, masses of 15-150 M$_{\\odot}$, H$_2$ column densities of $\\sim$3.3-15.0 $\\times$ 10$^{24}$ cm$^{-2}$, and space densities of $\\sim$ 1.7-6.0 $\\times$10$^{7}$ cm$^{-3}$, again with the highest numbers derived from continuum data. From the source SED, they estimate a temperature of 40 K. M08 derived a similar temperature of 37 K by using the automatic SED fitting tool provided by Robitaille et al. (2007), and a bolometric luminosity of 3170 L$_{\\odot}$. They showed that these values place Mol 160 firmly into the protostar category on the evolutionary diagram developed by Molinari et al. (2008b). We note that their estimate of the bolometric luminosity of Mol 160 should be considered an upper limit, due to emission from the nearby, overlapping PDRs. Taken together, these observations and models indicate Mol 160 is in an evolutionary stage that is a precursor to a hot core. In spite of the uncertainties, it seems reasonable to conclude that the mass of the core is sufficient to form one or more massive stars. Furthermore, the high column densities suggest that massive stars {\\it should} form once the core begins to collapse. Recent theoretical work predicts a mass column density threshold $\\ge$ 1 g cm$^{-2}$ for massive star formation (Krumholz \\& McKee 2008; Krumholz et al. 2010). Regions where intermediate mass stars are thought to be forming typically have mass column densities $\\sim$ 0.1 - 0.5 g cm$^{-2}$ (e.g., Arvidsson et al. 2010; Wolf-Chase et al. 2003; Wolf-Chase, Walker, \\& Lada 1995). In the Arvidsson et al. (2010) study, estimates were made from peak column densities derived from single-antenna millimeter-wave observations of regions at different distances, implying a `clump' rather than `core' scale for many of these regions; nevertheless, the derived values are significantly smaller than those obtained for UC \\ion{H}{2} regions at comparable distances. In the Wolf-Chase et al. (1995, 2003) studies, the 2264 S1 core in the Mon OB1 dark cloud has a mass column density of $\\sim$ 0.27 g cm$^{-2}$ on $\\sim$ 0.1 pc scale. Spitzer observations indicate 2264 S1 contains at least 10 low-mass protostars (Young et al. 2006; Teixeira et al. 2006). Taking 4.0 $\\times$ 10$^{24}$ cm$^{-2}$ as a lower limit for the column density of the compact Mol 160 core (see Tables 5 \\& 6 in F04), the mass column density is $\\ge$ 16 g cm$^{-2}$. M98 noted that the Mol 160 core was slightly elongated to the southeast and to the north at 3.4 mm. Using a $0.94^{\\prime\\prime} \\times 0.76^{\\prime\\prime}$ beam at a wavelength of 1.3 mm, F04 resolved the millimeter continuum source into two components. The principal component was centered at the position previously determined by M98. The flux density per beam and integrated flux density of the smaller source were, respectively, 50\\% and 8\\% of those of the primary. Our 3.2-mm continuum observations do not resolve these two components, but are slightly extended in the direction of the secondary peak (see contours on Figure 3c). Also, the position of the secondary source coincides with the position of MHO 2922. If the mass within a 1.5$^{\\prime\\prime}$ diameter region centered on the primary continuum peak is $\\sim$ 150 M$_{\\odot}$, as F04 suggest, and if the mass scales with the 1.3-mm continuum flux, the mass associated with the secondary peak would be $\\sim$ 12 M$_{\\odot}$. The Mol 160 region was not covered in the Spitzer GLIMPSE survey (Benjamin et al. 2003; Churchwell et al. 2009), but has been observed by WISE. Figure 7 is a 3-color WISE image (3.4, 4.6, \\& 12-$\\micron$ bands) of the Mol 160 region. As in the color scheme chosen for the GLIMPSE survey, objects that are unusually bright in the 4.6-$\\mu$m band appear to be green. In the GLIMPSE data, ``extended green objects'' or ``EGOs'' have been linked to shocked H$_2$ emission from massive outflow candidates (Cyganowski et al. 2008). MHOs 2921 \\& 2922 are the only MHOs listed in Table 1 that are detected as EGOs in the WISE image. Furthermore, Mol 160 is {\\it only} detected in the WISE 4.6-$\\mu$m band and was not seen in the continuum by ISO at 15 $\\mu$m (M98). Hence it is likely, as for the EGOs, that the 4.6-$\\mu$m emission is dominated by spectral lines from shocked molecular gas. \\subsection{Outflow from Mol 160} Observations of millimeter-wave, H$_2$, and CH$_3$OH maser emission independently trace outflowing gas from Mol 160. Taken together, these observations point to the extreme youth of Mol 160 and suggest new avenues to explore in establishing an evolutionary sequence for massive star formation. \\subsubsection{Millimeter-wave Spectroscopy} M98 and Molinari et al. (2002) presented evidence for energetic molecular outflows in Mol 160. The wings of their SiO spectrum at the position of Mol 160 form a broad plateau in velocity space. The wings have a relatively large amplitude compared to the intensity in the central frequency channel, and the intensity drops sharply at the edges of the plateau. The HCO$^+$ wings are much smaller, compared to the central peak. The blue wing tapers down smoothly from the systemic peak within $\\sim$ ~10 km s$^{-1}$. The red wing is broader, with a dip at -46 km s$^{-1}$ and a peak at -43 km s$^{-1}$. The blue and red HCO$^+$ outflow maps of M98 peak $\\sim$ 1$^{\\prime\\prime}$ to the east of the continuum peak, and the blue peak (integrated from -63 km s$^{-1}$ to -53 km s$^{-1}$) is much stronger and broader than the barely detected red peak (integrated from -47 km s$^{-1}$ to -37 km s$^{-1}$). The frequency range over which they integrated to make their ``blue'' outflow map overlaps a third of our -52.5 HCO$^+$ channel and nearly all of our 53.7 km s$^{-1}$ CS channel, so it is likely that their map is dominated not by outflow gas but by the blue wing of the systemic core gas. On the other hand, the morphologies of their maps of the SiO wings are similar to each other. The peak of the blue lobe lies $\\sim$ 0.5$^{\\prime\\prime}$ south of the continuum peak, and the red peak lies $\\sim$ 1$^{\\prime\\prime}$ to the north. Based on the small angular separation between the lobes, M98 argued that the outflow is essentially parallel to the line of sight, ruling out the possibility that their failure to detect 15-$\\mu$m emission with ISO was caused by geometrical effects such as an edge-on dust disk and thus implying that the flow is completely contained within the optically thick core. In our HCO$^+$ data, we do see a NE-SW velocity gradient across the core (cf. Figure 4 and the color image in Figure 8). The channel maps of the 1.5 km s$^{-1}$ data from Track 1 show the same trend, but the instrumental problems with the correlator prevent a more precise estimate of the velocity dispersion. One can set a rough upper limit of $\\sim$6 km s$^{-1}$ from the span between the central velocities of our -49.2 and -55.8 km s$^{-1}$ HCO$^+$ channels. This is much less than the $ >$20 km s$^{-1}$ extent of the SiO plateau and could plausibly result from systematic motions established during the compression/collapse phase of core evolution rather than protostellar outflows. We suggest that kinematics of the HCO$^+$ and CS gas can best be understood in terms of a model in which the compact, high-surface-brightness molecular core associated with the Mol 160 continuum source is blueshifted by $\\sim$2 km s$^{-1}$ from a parent clump with systemic velocity of -50.5 km s$^{-1}$. Most of the flux from the compact core lies within the -52.5 km s$^{-1}$ HCO$^+$ channel and -53.7 km s$^{-1}$ CS channel. The source brightness distributions in these two channels are similar and their outer contours are roughly congruent with the low-intensity 2.12-$\\micron$ halo around MHO 2921 and MHO 2922 seen in Figure 3. The observed blueshift of the core gas with respect to the -50.5 km s$^{-1}$ cloud would be consistent with compression by a background \\ion{H}{2} region. Our observations are consistent with the assertion by M98 that the extent of the SiO outflow is small compared to the size of the core but suggest an alternative possibility for the geometry. The symmetrical displacements of MHO 2921 to the south of the continuum peak position and an opposing dark area to the north is reminiscent of the simulated near-infrared images of wide-angle outflows constructed by Zhang \\& Tan (2011). See, for example, the example with an outflow inclination of 60$^{\\circ}$ in their Figure 9. Their models assumed the source of the near-infrared radiation from the arc was continuum radiation from the protostar but would also be consistent with an intense source of line emission located close to the center of the core. Such a scenario would imply that H$_2$ emission associated with the red SiO outflow lobe is hidden by dust in the waist of the accretion envelope (and/or a disk or pseudodisk). From their SiO observations, M98 deduced an outflow mass of $\\sim$ 20 M$_{\\odot}$, a dynamical timescale of $\\le 6 \\times 10^3$ yr, a mass loss rate of 3.5 $\\times 10^{-3}$ M$_{\\odot}$ yr$^{-1}$ and a mechanical luminosity of 22 L$_{\\odot}$. They based these estimates on the assumptions that the outflow angle was 30$^{\\circ}$ with respect to the line of sight and that the extent of the outflow was equal to the radius of the molecular core. In light of the small angular displacement of MHO 2921 from the continuum peak, the mechanical luminosity and mass loss rate could be larger and the time scale shorter, indicating an even younger, more powerful outflow. \\subsubsection{H$_2$ Luminosity} Until spectra of MHOs 2921 \\& 2922 are available, we can only estimate the effects of extinction. Assuming that the 2.12-$\\micron$ flux is $\\sim$ 5\\% of the total ro-vibrational H$_2$ emission (e.g., Caratti o Garatti 2006, 2008), the combined H$_2$ luminosity for the two MHOs is $3.4 \\times 10^{-3} \\times$ (D/4.9 kpc)$^2 \\times 10^{0.4 A_{2.12}}$ L$_{\\odot}$, where D is the distance in kpc and $A_{2.12}$ is the extinction at 2.12 $\\micron$. Because of extinction, the actual luminosities of the 2.12-$\\micron$ sources are probably much greater than the observed luminosities. The average column densities, N(H$_2$), of the large -50.5 km s$^{-1}$ cloud inferred from large-beam measurements with single-dish telescopes range from 1$\\times 10^{23}$ to 3.6$\\times 10^{23}$ cm$^{-2}$. The values derived from interferometric measurements at higher angular resolution are larger by an order of magnitude or more. Conservatively assuming a column density of N(H$_2$) = 5$\\times 10^{22}$ cm$^{-2}$, an amount equal to half the lowest value derived from large-beam measurements, leads to a 2.12-$\\micron$ extinction of 5.1 and a combined H$_2$ luminosity of 7.4 L$_{\\odot}$ for MHO 2921 and MHO 2922. This is 0.23\\% of the bolometric luminosity inferred from the SED of the continuum source (L$_{bol} = 3170$ L$_{\\odot}$, M08) and is comparable to the mechanical luminosity inferred by M98 from the high-velocity SiO outflow. Caratti o Garatti (2006) found a relationship between total H$_2$ luminosity and L$_{bol}$ for low-mass YSOs. Caratti o Garatti (2008) suggested that this relationship could tentatively be extended to high-mass YSOs based on their results for IRAS 20126$+$4104. They further suggested that this relationship might be taken as evidence that outflows from massive YSOs are similar in nature to those from lower-mass YSOs (i.e., produced by disk-mediated accretion). However, the Orion BN/KL complex, which has a bolometric luminosity of 10$^5$ L$_{\\odot}$, and a total H$_2$ luminosity of 120 L$_{\\odot}$ (Rosenthal et al. 2000), would also fit their relationship. In this kind of impulsive event there is no reason to expect strict proportionality between the bolometric luminosity of the source and either the kinetic energy of the outflow or its rate of dissipation in shocks. Therefore, caution must be exercised in attempting to draw conclusions about the nature of outflows from massive YSOs from the L$_{H_2}$/L$_{bol}$ relationship alone. We note that the L$_{H_2}$/L$_{bol}$ ratio for Mol 160 conservatively places Mol 160 a factor of two above the L$_{H_2}$/L$_{bol}$ relationship determined by Caratti o Garatti (2006, 2008) and the discrepancy could be larger if (a) the internal heating is smaller; (b) the extinction has been underestimated; or (c) the ratio of total H$_2$ to H$_{2.12}$ luminosity has been underestimated (e.g., if there is significant emission from cooler shocked hydrogen, cf. Caratti o Garatti 2006, 2008). As noted in \\S 4.3.1, MHO 2922 is coincident with a secondary continuum peak (F04). It is possible that MHO 2922 emission arises from a distinct outflow from the central source, a secondary accretion center, or an object ejected from the core. Present observations do not allow us to distinguish between the possibilities; however, we note that this does not change our conclusion that Mol 160 has a very high L$_{H_2}$/L$_{bol}$ ratio, since this ratio was computed based on the estimated bolometric luminosity for the entire core, and our computed ratio is a conservative lower limit. \\subsubsection{CH$_3$OH Masers} Class I CH$_3$OH masers are frequently and perhaps exclusively found near young sources. They are well-correlated with molecular outflows in massive star forming regions and are thought to be collisionally pumped (Cyganowski et al. 2009 and references therein; Fontani et al. 2010). Schnee \\& Carpenter (2009) found a strong correlation between the presence of compact 3-mm continuum emission and 95-GHz Class I CH$_3$OH maser emission. In contrast, they detected no 3-mm continuum emission toward UC \\ion{H}{2} regions lacking maser emission, suggesting that the masers are signposts of an early stage in the evolution of a massive protostar before an expanding UC \\ion{H}{2} region has destroyed the accretion disk. DeBuizer et al. (2009) note that the relatively low velocities of CH$_3$OH masers, combined with their occasional locations slightly offset from the outflow axis, suggest they arise in ``systemic'' gas in outflow cavity walls. The velocities of the Mol 160 masers vary from -49.0 to 52.2 km s$^{-1}$, and the most intense emission comes from sources nearest the continuum peaks and at velocities near $\\sim$ 52 km s$^{-1}$. The maser positions seem to lie along two principal axes (see Figure 8), one approximately north-south through the main continuum source and one roughly along a line between the main and secondary continuum sources. It is tempting, but probably premature, to associate them with the walls of a wide-angle outflow cavity. The source could be significantly more complex than a single central source and a single bipolar outflow. The maser sources are as yet unresolved, spectrally or spatially, but they are bright enough to permit observations at higher spectral and spatial resolution that could significantly improve our understanding of the structure, kinematics, and excitation of the outflows. In their survey of CH$_3$OH masers, Fontani et al. (2010) found that sources for which both 44-GHz and 95-GHz Class I masers were observed, had similar spectra, confirming a common physical origin. They also concluded that the 95-GHz line is intrinsically fainter, based on their detection rates. The mean value of the ratio of 95-GHz to 44-GHz line intensity in Table 2 is 4.4. Some of the individual sources studied by Fontani et al. (2010) also have intensity ratios this large, suggesting that their detection statistics must result not from the intrinsic line formation process but from the probability of finding sources with conditions favoring the 95-GHz transition. Perhaps the ratio will prove to be a useful diagnostic for the shortest-lived (e.g., very early) phases of massive star formation. \\subsection{Mol 160: Current Status \\& Future Explorations} Taken together, the core and outflow properties of Mol 160 strongly suggest that this object is in an early phase of the 3rd stage (accretion) outlined for massive star formation at the beginning of this section (Zinnecker \\& Yorke 2007). The temperature, mass, mass column density, and luminosity of Mol 160 are consistent with physical properties expected for a precursor to a hot core, before winds and radiation from embedded protostars have removed a significant amount of the accreting envelope, and while the protostar mass and luminosity are still increasing. The presence of a very compact outflow, which is independently confirmed via millimeter-wave spectroscopy, H$_2$ shocks, and CH$_3$OH masers, also points to the extreme youth of one or more embedded, accreting, massive objects. Because we expect the youngest massive protostars to be rare, more refined selection methods are highly desirable. A high value of L$_{H_2}$/L$_{bol}$ may be a useful discriminant, but the example of Mol 160 also highlights some of the difficulties. Near- and mid-infrared spectroscopy are needed for more accurate estimates of extinction and L$_{H_2}$. It is also likely that massive accreting objects will be located in or near regions containing other products of massive star formation such as bright PDRs. For these sources, bolometric luminosities can easily be overestimated. High-angular-resolution, multicolor, far-infrared observations with Herschel or the Stratospheric Observatory for Infrared Astronomy (SOFIA) will help reduce confusion in complex regions. Systematic searches in infrared dark clouds more clearly separated from PDRs (e.g., for compact Herschel sources and GLIMPSE ``green objects'' more pointlike than EGOs) may also be helpful. The Milky Way Project is producing a new catalog of PDRs in the galactic plane using citizen scientist identifications from Spitzer GLIMPSE/MIPSGAL images (Simpson et al. 2011) that should be useful in this regard. High-angular-resolution observations of CH$_3$OH Class I maser line ratios may also prove to be a useful diagnostic for physical conditions characteristic of very young sources, one that will become even more valuable with the deployment of advanced receivers on CARMA and as the Atacama Large Millimeter/submillimeter Array (ALMA) comes on line." }, "1112/1112.1658_arXiv.txt": { "abstract": "We present a new version of our code for modeling the atmospheric circulation on gaseous exoplanets, now employing a ``double-gray\" radiative transfer scheme, which self-consistently solves for fluxes and heating throughout the atmosphere, including the emerging (observable) infrared flux. We separate the radiation into infrared and optical components, each with its own absorption coefficient, and solve standard two-stream radiative transfer equations. We use a constant optical absorption coefficient, while the infrared coefficient can scale as a powerlaw with pressure. Here we describe our new code in detail and demonstrate its utility by presenting a generic hot Jupiter model. We discuss issues related to modeling the deepest pressures of the atmosphere and describe our use of the diffusion approximation for radiative fluxes at high optical depths. In addition, we present new models using a simple form for magnetic drag on the atmosphere. We calculate emitted thermal phase curves and find that our drag-free model has the brightest region of the atmosphere offset by $\\sim$12\\degrees~from the substellar point and a minimum flux that is 17\\% of the maximum, while the model with the strongest magnetic drag has an offset of only $\\sim$2\\degrees~and a ratio of 13\\%. Finally, we calculate rates of numerical loss of kinetic energy at $\\sim$15\\% for every model except for our strong-drag model, where there is no measurable loss; we speculate that this is due to the much decreased wind speeds in that model. ", "introduction": "It has been almost one decade since the first atmospheric measurement of a hot Jupiter \\citep{Charbonneau2002} and yet this class of exotic exoplanet still provides us with many mysteries waiting to be solved. These include the culprit responsible for the stratospheric temperature inversions inferred for many hot Jupiters \\citep{Hubeny2003,Fortney2008,Spiegel2009,Zahnle2009,Knutson2010}, the implication of a super-solar carbon-to-oxygen ratio in at least one hot Jupiter \\citep{Madhu2011}, the possible influence of magnetic drag on the atmospheric circulation \\citep{Perna2010a}, and a peculiar planet where dawn seems to be much hotter than noon \\citep{Crossfield2010}. In the midst of a growing collection of observations, some with unexpected or surprising results, is the continuing development of atmospheric models trying to interpret and understand the measurements. One set of atmospheric models are numerical ones that simulate the global circulation patterns on close-in gas giants. With masses and radii comparable to Jupiter, but subject to incident stellar fluxes 10,000 times stronger than what Jupiter receives from the Sun, and expected to be tidally locked into synchronous orbits, hot Jupiters exist in an atmospheric regime unlike anything in our solar system, and models of their atmospheric circulation are expanding into uncharted territory. In order to try to understand how atmospheres work in this new regime, the models that have been developed so far represent a range of complexities, various approaches, and the use of different sets of assumptions. Some of the simplest models solve the shallow water or equivalent barotropic equations and use various schemes to include the effect of radiative heating \\citep{Cho2003,Cho2008,Langton2007}. Others solve the primitive equations of meteorology, either using a Newtonian relaxation scheme for the radiative forcing \\citep{Showman2002,Cooper2005,Showman2008,MR09,RM10}, a dual-band radiative transfer scheme \\citep[][similar to the one we present here]{Heng2011}, or more complex non-gray radiative transfer \\citep{Showman2009}. Finally, there are also models that solve the full set of fluid equations, using dual-band flux-limited diffusion for the radiative transfer \\citep{DobbsDixon2008,DobbsDixon2010}. Here we present an updated version of our previous general circulation model that now includes a ``double-gray\" radiative transfer scheme. Fluxes throughout the atmosphere are separated into optical and infrared components, each with its own absorption coefficient, which is constant for the optical band and can scale as a powerlaw with pressure for the infrared band. We use standard two-stream radiative transfer equations to solve for the vertical fluxes and calculate heating rates from those. In addition to producing self-consistent radiative fluxes and heating rates, including the observable infrared flux emerging from the top boundary, this new code has the advantage of maintaining only a moderate level of complexity. This facilitates comparison between simulated results and analytic profiles \\citep[][see also Hansen 2008]{Guillot2010} and makes it easier to clearly identify the effects due to changes in opacity. Our new code is described in detail in Section~\\ref{sec:code}. We demonstrate the functionality of our code by presenting a model of a generic hot Jupiter (Section~\\ref{sec:models}) and we investigate issues related the atmosphere's behavior at deep pressures (Section~\\ref{sec:deep}). We then study the effect of magnetic drag on atmospheric circulation by applying a simplified drag scheme to our model, with drag strengths ranging from weak to strong (Section~\\ref{sec:drag}). In Section~\\ref{sec:conc} we summarize our results. ", "conclusions": "\\label{sec:conc} We have presented a new radiative transfer scheme for our atmospheric circulation code. It divides all radiation into optical and infrared wavelengths. The optical flux is incident on the top boundary and its attenuation is controlled by a constant optical absorption coefficient. The infrared flux is absorbed and emitted at each level, as governed by an infrared absorption coefficient that goes as a powerlaw with pressure (and can be constant). Below the infrared photosphere the infrared fluxes are calculated using the flux-limited diffusion approximation, which we found was necessary in order to reproduce analytic temperature profiles correctly. Using this new code, we present a fiducial model for a generic hot Jupiter and find that the temperature and wind structures agree well with previously published models. We also show vertical velocities, radiative fluxes, and heating/cooling rates as a function of pressure and region of the atmosphere. Our radiative scheme allows us to self-consistently predict the infrared flux emitted by the planet and we map this as a function of location on the globe, reproducing the standard shift of the brightest region eastward of the substellar point. We show temperature profiles for models that: 1) did not use the flux diffusion scheme, or 2) assumed zero flux from the interior. In both cases the deepest pressure levels are mostly isothermal, instead of increasing in temperature toward the inner convective zone, significantly changing the ability of the atmosphere to exchange energy and momentum with the interior. However, the upper atmosphere (including and above the infrared photosphere) are relatively insensitive to the behavior in the deepest levels and have similar observable properties to our fiducial model. We present models that use our new radiative transfer scheme in combination with a simplified form of magnetic drag, as an improvement on the models presented in \\citet{Perna2010a}. As in those earlier models, we find that as the drag strength is increased, the structure of the atmosphere changes; the strongest level of drag is able to significantly alter the atmosphere from its drag-free state. As a new result, we are able to measure the amount by which the brightest region of the atmosphere is offset from the substellar point, as a function of the strength of the magnetic drag used in each model. The weak- and medium-drag models have orbital thermal phase curves similar to the drag-free model. The winds are slowed enough in the strong-drag model that the brightest region of the atmosphere is closely aligned with the substellar point and, related to this effect, the day-night flux amplitude in the strong-drag model is slightly larger than for the drag-free model. We estimate the amount of ohmic heating that would be produced for each of our models and find that the medium-drag model should have a significantly inflated planetary radius, while the amount of heating in the strong-drag model is so high that it could perhaps lead to planet evaporation. A better analysis of magnetic effects on hot Jupiter atmospheres must wait for improvements in our code, however, using a more realistic form for the drag and explicitly including the effect of ohmic heating. Finally, we estimate the rate of numerical loss of kinetic energy in each of our models and find it to be at the level of $\\sim$15\\%, except for our model with strong magnetic drag, which we calculate to have losses consistent with zero at the percent level. Our strong-drag model has much slower winds than the other models presented here and we speculate that strong numerical loss is directly related to those high speed winds." }, "1112/1112.3970_arXiv.txt": { "abstract": "We use a sample of active galaxies from the Cosmic Evolution Survey to show that host galaxy morphology is tied to the accretion rate and X-ray obscuration of its active galactic nucleus (AGN). Unobscured and rapidly accreting broad-line AGNs are more likely to be in spheroid-dominated hosts than weak or obscured AGNs, and obscured AGNs are more likely to have disturbed host galaxies. Much of the disagreement in previous work on the AGN-merger connection is likely due to each study probing AGNs with different obscuration and accretion properties. Instead, only obscured AGNs seem to merger-driven, while weak AGNs are fed by stochastic processes in disks and rapidly-accreting broad-line AGNs require massive bulges. Our observed ``unified model'' for AGN hosts fits with theoretical models for merger-driven AGN evolution, but is also consistent with steady-state AGN activity. ", "introduction": "The most popular theoretical framework for coevolving active galactic nuclei (AGNs) and galaxies invokes major mergers to fuel both starbursts and quasars \\citep{san88,hop06}. But there also exist secular processes which can grow AGNs in disks, through stochastic gas accretion and stellar mass loss \\citep{hh06} or cold streams and disk instabilities \\citep{bou11}. The relative dominance of these fueling modes may evolve: perhaps secular processes are efficient in the gas-rich $z>0.5$ universe but mergers are required to funnel gas to galaxy nuclei at $z \\sim 0$, in a way analogous to star-forming galaxies \\citep[e.g.][]{gen10}. Observations of AGN host galaxy morphologies have the potential to distinguish secular processes from merger fueling. Figure \\ref{fig:mergerfrac} shows examples of host morphology studies from X-ray AGNs at three different redshifts: $z \\sim 0.03$ \\citep{koss10}, $z \\sim 0.75$ \\citep{cis11}, and $z \\sim 1.8$ \\citep{koc11}. In each survey, AGNs and mass-matched inactive galaxies are blindly classified by visual inspection. Naively comparing the three results suggests evolution in the importance of merger fueling for AGNs, but each study is biased to selecting different kinds of AGNs. \\begin{figure}[t] \\begin{center} \\scalebox{0.8} {\\plotone{jrtrump_fig1.eps}} \\end{center} \\caption{The fraction of AGNs and inactive galaxies which are disturbed, measured from X-ray selected samples at three different redshifts: $z \\sim 0.03$ \\citep{koss10}, $z \\sim 0.75$ \\citep{cis11}, and $z \\sim 1.8$ \\citep{koc11}. At first glance, it appears that the AGN-merger connection evolves, with mergers driving AGNs locally but not in the past. However the apparent trend is probably caused by differences in the AGN samples rather than true evolution. \\label{fig:mergerfrac}} \\end{figure} Contrary to the classical unified model \\citep{ant93}, recent work has shown that rapidly accreting quasars have very different fueling and feedback modes from weak AGNs \\citep{ho08,tru11,ant11}. In particular, \\citet{tru11} shows that broad-line quasars have strong radiative winds and luminous accretion disks, while weakly accreting narrow-line AGNs have radiatively inefficient accretion flows and radio-mode feedback. Understanding AGN-galaxy coevolution then requires studying host galaxy type across the two axes of AGN accretion rate and obscuration. Yet previous AGN host studies (including those shown in Figure \\ref{fig:mergerfrac}) have generally not distinguished between weakly accreting LINERs, Compton-thick AGNs, and powerful quasars. Here we study host galaxy morphologies for 70 X-ray selected AGNs from the Cosmic Evolution Survey \\citep[COSMOS,][]{sco07}. These AGNs span three orders of magnitude in each of accretion rate and column density, and their observed data demonstrate that host galaxy type is connected to AGN properties. ", "conclusions": "" }, "1112/1112.0138_arXiv.txt": { "abstract": "{The distance to an isolated dark globule is often unknown and yet crucial for understanding its properties, in particular its mass. A new approach to this problem is discussed} {The purpose of the present paper is to investigate how well the distances of more or less reddened field stars can be determined by using multi-colour imaging.} {We observed a test globule, B\\,335 in U, B, g, r, and I, and together with the 2MASS survey, this data set gives a well-defined spectral energy distribution (SED) of a large number of stars. The SED of each star depends on the interstellar extinction, the distance to the star, and its intrinsic SED. As we had good reasons to suspect that the wavelength dependence of the extinction (the reddening) changes from the outskirts of the globule to the central parts, we did not assume any specific reddening law. Instead, we use a scheme that allows independent determination of the extinction in each line of sight as determined by groups of adjacent stars. The method is based on the use of stellar atmospheric models to represent the intrinsic SEDs of the stars. Formally, it is then possible to determine the spectral class of each star and thereby its distance. For some of the stars we have optical spectra, allowing us to compare the photometric classification to the spectrometric. } {As expected, the main problem is that there are few stars found within each distance bin for the small field size defining a typical dark globule. However, the characterisation of the extinction and photometric classification give consistent results and we can identify one star at the front side of the globule. It has a photometric distance of 90\\,pc. The closest star behind the B\\,335 globule has a distance of only $\\approx 120$\\,pc and we therefore determine the distance to B\\,335 as 90--120\\,pc. Our deep U image shows a relatively bright south-western rim of the globule, and we investigate whether it might be due to a local enhancement of the radiation field. A candidate source, located 1.5 arcminutes outside our field, would be the field star, HD\\,184982. This star has an entry in the Hipparcos Catalogue and its distance is 140--200\\,pc. However, we come to the conclusion that the bright SW rim is more likely due to the wing of the point-spread-function (PSF) of this star.} {} ", "introduction": "It is in principle easy to determine the distance to a dark cloud: it is just a matter of finding a star just in front of the cloud and another closely behind it. Or, even better, a star close enough to the cloud to give rise to a reflection nebula. The closest cloud complexes, like that in Taurus/Aurigae, cover large regions on the sky and thanks to the Hipparcos mission (\\cite{Hipparcos}) there are a number of stars with known distances within the field. The remaining task is that of placing the stars in front of or behind the cloud. The traditional way is to look for the onset of interstellar reddening (see Lombardi et al. \\cite{Lombardi} for a recent example). As an attractive alternative, the onset of polarisation can be determined (Alves \\& Franco \\cite{Alves}). For small isolated clouds, however, there are too few (if any) Hipparcos stars within the field, and the distances to the stars must be determined by spectroscopic/photometric means. The reason why it is important to determine the distances to such clouds, including dark globules, is their role of hosting isolated star formation. Thus, important properties as the luminosity of the forming star and the mass and the radius of the cloud cannot be determined without knowing the distance. Maheswar \\& Bhatt (\\cite{Maheswar}) have estimated the distances to nine dark globules using V, R and I photometry in combination with the J, H and K photometry from the 2MASS catalogue (Skrutskie et al. \\cite{2MASS}). By assuming a standard interstellar extinction law (valid for \\emph {diffuse} clouds) and adjusting visual extinction, A$_V$, until the SED fits that of a \\emph{main sequence} (MS) star, these authors determine the extinction and the distance for each star. There are two uncertain assumptions in their approach, which may have strong impact on the estimates of the distances. The first is the assumption of a standard extinction law, while it is known that the $R_{V}$ value ($R_{V}$=A$_{V}$/E(B-V)) changes from typically $R_{V}$=3.1 in the diffuse ISM to $R_{V}\\approx 5$ in molecular clouds (see e g Ossenkopf \\& Henning \\cite{Ossenkopf}, Strafella et al \\cite{Strafella}). The second is the assumption that all the stars are MS stars. In the present paper we discuss these problems and present a more elaborate method which is based on imaging in eight filters. As a test case we use the well studied dark globule B\\,335. It hosts a protostar heavily obscured by a flattened core structure, including an accretion disk seen edge-on (G\\aa lfalk \\& Olofsson \\cite{Gaalfalk}) . In most of the numerous investigations of this source, a distance of 250\\,pc as given by Tomita et al. (\\cite{Tomita}) is assumed. Stutz et al. (\\cite{Stutz}) have recently revised the distance estimate by using archive photometry of stars within a radius of 20\\arcmin. Their revised distance is 150\\,pc (60--200\\,pc). The diameter of B\\,335 as seen in the optical is just $\\approx$ 4\\arcmin\\,. CO observations (Frerking et al. \\cite{Frerking}), however, show that it extends far to the NE, including stars within such a large radius as 20\\arcmin. \\begin{table} \\begin{minipage}[t]{\\columnwidth} \\caption{Observation log} \\label{obslog} \\renewcommand{\\footnoterule}{} % \\begin{tabular}{llllll} \\hline \\hline object &camera &filter ¢er &exp time&exp time\\\\ & & & wavel &per frame &total\\\\ & & &$ [nm]$&[s]&[s]\\\\ B\\,335 &EMMI-blue &U602 &353.96 &720 &26600\\\\ B\\,335 &EMMI-red &Bb605 &413.22 &720 &7900\\\\ B\\,335 &EMMI-red &g772 &508.86 &480 &4700\\\\ B\\,335 &EMMI-red &r773 &673.29 &300 &2100\\\\ B\\,335 &EMMI-red &I610 &798.50 &300 &2400\\\\ B\\,335 &NOTCAM\\footnote{co-added image provided by M. G{\\aa}lfalk\\\\} &Ks &2144. &300 &2800\\\\ B\\,335ref &EMMI-blue &U602 &353.96 &720 &2200\\\\ B\\,335ref &EMMI-red &Bb605 &413.22 &720 &800\\\\ B\\,335ref&EMMI-red &g772 &508.86 &480 &500\\\\ B\\,335ref &EMMI-red &r773 &673.29 &300 &300\\\\ B\\,335ref &EMMI-red &I610 &798.50 &300 &300\\\\ \\hline\\\\ \\end{tabular} \\end{minipage} \\end{table} ", "conclusions": "\\begin{enumerate} \\item The fundamental limitation when it comes to determining the distance to a small dark cloud is the space density of stars and as the cool MS stars are the most abundant, any method should be able to sort out these stars and reliably separate them from cool giant stars. \\item It turns out that multi-colour imaging in combination with synthetic colours can provide reliable spectral classification and extinction, and we find that the U band is particularly important, both for the spectral classification and for tracing the extent of the globule. \\item We have tested a new SED-fitting method which allows for different R$_{V}$-values across the field. For the test case, B\\,335, we get a distance of 90--120\\,pc. \\item Our deep U image of B\\,335 shows a bright halo due to scattered star light, and the south-western rim is much brighter. We find that a possible explanation is that the globule is illuminated by a star, HD\\,184982, located 1.5 arcminutes to the south-west of our field and a few tenths of a parcec behind the cloud. The distance to this star is 140--200\\,pc according to the Hipparcos Catalogue. However, the surface brightness can also be explained as due to the PSF wing of the same star, and we judge that this is in fact a more likely explanation. \\end{enumerate}" }, "1112/1112.0656.txt": { "abstract": "Observed in cool chromospheric lines such as H$\\alpha$ or \\ion{Ca}{2} H, coronal rain corresponds to cool and dense plasma falling from coronal heights. Considered rather as a peculiar sporadic phenomenon of active regions, it has not received much attention since its discovery more than 40 years ago. Yet, it has been shown recently that a close relationship exists between this phenomenon and the coronal heating mechanism. Indeed, numerical simulations have shown that this phenomenon is most likely due to a loss of thermal equilibrium ensuing from a heating mechanism acting mostly towards the footpoints of loops. We present here one of the first high resolution spectroscopic observations of coronal rain, performed with the CRISP instrument at the Swedish Solar Telescope. This work constitutes the first attempt to assess the importance of coronal rain in the understanding of the coronal magnetic field in active regions. With the present resolution, coronal rain is observed to literally invade the entire field of view. A large statistical set is obtained in which dynamics (total velocities and accelerations), shapes (lengths and widths), trajectories (angles of fall of the blobs) and thermodynamic properties (temperatures) of the condensations are derived. Specifically, we find that coronal rain is composed of small and dense chromospheric cores with average widths and lengths of $\\sim370$~km and $\\sim1500$~km respectively, average temperatures below 7000~K, displaying a broad distribution of falling speeds with an average of $\\sim70$~km~s$^{-1}$ and accelerations largely below the effective gravity along loops. Through estimates of the ion-neutral coupling in the blobs we show that coronal rain acts as a tracer of the coronal magnetic field, thus supporting the multi-strand loop scenario, and acts as a probe of the local thermodynamic conditions in loops. We further elucidate its potential in coronal heating. We find that the cooling in neighboring strands occurs simultaneously in general suggesting a similar thermodynamic evolution among strands, which can be explained by a common footpoint heating process. Constraints for coronal heating models of loops are thus provided. Estimates of the fraction of coronal volume with coronal rain give values between 10~\\% and 40~\\%. Estimates of the occurrence time of the phenomenon in loops set times between 4 and 15~hours, implying that coronal rain may be a common phenomenon, in agreement with the frequent observations of cool downflows in EUV lines. The coronal mass drain rate in the form of coronal rain is estimated to be on the order of $10^{10}$~g~s$^{-1}$, a significant quantity compared to the estimate of mass flux into the corona from spicules. ", "introduction": "Solar observations over the past 30 years have presented a very dynamic picture of the Sun. Loops, closed magnetic field entities encountered basically everywhere in the Sun at different spatial scales (of a few Mm and up to a few hundred Mm) are perfect examples of this dynamic character. Especially in active region coronae, studies of such entities reveal a large range of lifetimes, dynamics, and thermodynamic properties which have led to a main trifold classification as hot, warm and cool loops \\citep[see the extensive review by][]{Reale_2010LRSP....7....5R}. Several debates exist over the characteristics of the heating leading to such observed properties. For instance, it is not known whether the heating has mainly an impulsive or a uniform character in time. It is not known whether the heating is generally localized at specific atmospheric heights or is rather uniform over the loop. And, maybe most importantly, the main agent converting the magnetic energy into thermal energy is not known \\citep[generally known as AC or DC mechanism depending on its timescale,][]{Klimchuk_2006SoPh..234...41K}. Several observational studies have shown, however, that a significant fraction of loops in active regions are out of hydrostatic equilibrium \\citep{Aschwanden_2001ApJ...550.1036A, Aschwanden_2001ApJ...559L.171A, Winebarger_etal_2003ApJ...587..439W, Schmieder_etal_2004ApJ...601..530S, Aschwanden_2009ApJ...695...12A}. This family of loops shows in general temperatures below $2-3\\times10^{6}~$K and show significant changes over relatively short timescales as compared to hotter loops. Indeed, heating and cooling processes seem to occur continuously in these loops \\citep{Kjeldseth_Brekke_1998SoPh..182...73K, UgarteUrra_etal_2006ApJ...643.1245U, Warren_2007PASJ...59S.675W, Ugarte-Urra_etal_2009ApJ...695..642U, Landi_etal_2009ApJ...695..221L, Dudik_etal_2011AA...531A.115D}. A recurrent finding in these and other active region studies is the presence of cool downflows along the legs of the loops \\citep{Foukal_1976ApJ...210..575F, Foukal_1978ApJ...223.1046F, Schrijver_2001SoPh..198..325S, Oshea_etal_2007AA...475L..25O, Tripathi_etal_2009ApJ...694.1256T}. This non-equilibrium state is generally explained through footpoint heating, a scenario in which the loops have their main heating source located towards the footpoints. Footpoint heating in active region loops has received further observational support. Upflows often appear correlated with nonthermal velocities at the footpoints of these loops \\citep{Doschek_etal_2007ApJ...667L.109D, Hara_2008ApJ...678L..67H, Nishizuka_Hara_2011ApJ...737L..43N}. Recently, \\citet{Hansteen_2010ApJ...718.1070H} and \\citet{DePontieu_etal_2011Sci...331...55D} have shown that a considerable part of the hot coronal plasma could be heated at low spicular heights, thus explaining the fading character of the ubiquitous type II spicules \\citep{Rutten_2006ASPC..354..276R, DePontieu_etal_2007PASJ...59S.655D, Rouppe_etal_2009ApJ...705..272R}. But perhaps one of the clearest evidences for footpoint heating is put forward by the presence of cool structures in the active region coronae, such as filaments/prominences and coronal rain. Coronal rain is not a newly observed phenomenon. Although observed for the first time almost 40 years ago \\citep{Kawaguchi_1970PASJ...22..405K, Leroy_1972SoPh...25..413L} few observational work exists on the subject. Furthermore, the term `coronal rain' is often erroneously attributed to basically any cold material falling down from coronal heights, as has been the case for describing prominence material falling down following a prominence eruption. Coronal rain corresponds to cool and dense blob-like material forming in the hot coronal environment in a timescale of minutes, which subsequently falls down to the surface along loop-like paths. The temperatures of the plasma composing coronal rain seem to range from transition region to chromospheric, according to the lines in which it has been observed: in \\ion{Ne}{7} 465~\\AA\\ and \\ion{O}{6} 1032~\\AA\\ with {\\it Skylab} \\citep{Levine_Withbroe_1977SoPh...51...83L}, in \\ion{O}{5} 629~\\AA\\ with {\\it SoHO}/CDS \\citep{Kjeldseth_Brekke_1998SoPh..182...73K}, in the 1600~\\AA\\ channel of {\\it TRACE} \\citep{Schrijver_2001SoPh..198..325S}, in the 304~\\AA\\ channel of {\\it SoHO}/EIT \\citep{DeGroof_2004AA...415.1141D, DeGroof05, Stenborg_etal_2008ApJ...674.1201S}, in the 304~\\AA\\ channel of {\\it SDO}/AIA \\citep{Kamio_etal_2011AA...532A..96K}, in \\ion{Ca}{2}~H with {\\it Hinode}/SOT \\citep{Antolin_2010ApJ...716..154A, Antolin_Verwichte_2011ApJ...736..121A}, in H$\\alpha$ \\citep[at Okayama Astrophysical Observatory, the Observatoire du Pic du Midi, Big Bear Observatory, the Swedish Vacuum Solar Telescope (SVST):][respectively]{Kawaguchi_1970PASJ...22..405K, DeGroof_2004AA...415.1141D, Leroy_1972SoPh...25..413L, Muller_2005ESASP.596E..37M}. \\citet{Schrijver_2001SoPh..198..325S} followed the various stages of coronal rain formation in loops with the channels of {\\it TRACE}. The hot loops rapidly cool down through thermal conduction and radiation until becoming thermally unstable. These radiative instabilities occur locally in the corona and lead to the rapid formation of blob-like condensations. These first appear in the 1216~\\AA\\ channel and then in the 1600~\\AA\\ channel suggesting continuous cooling below transition region temperatures. \\citet{DeGroof05} analyzed coronal rain simultaneously with {\\it SOHO}/EIT in the 304~\\AA\\ channel and in H$\\alpha$ images from Big Bear Observatory. The blobs first appear in EIT before becoming bright in H$\\alpha$, supporting the progressive cooling scenario. Furthermore, \\citet{Schrijver_2001SoPh..198..325S} shows that the loops hosting the condensations bright in the 1216~\\AA\\ were still visible in the 171~\\AA\\ channel, suggesting a multi-temperature structure in the loops. Recently, \\citet{Kamio_etal_2011AA...532A..96K} in observations with {\\it SDO}/AIA, similarly report cool condensations bright in the 304~\\AA\\ channel that appear to be surrounded by diffuse hotter plasma in the 171~\\AA\\ channel. Numerical simulations have pinpointed the now generally accepted mechanism behind coronal rain. Termed thermal non-equilibrium (or catastrophic cooling) the mechanism is based on the fact that a coronal loop whose heating input is mostly concentrated towards the footpoints will undergo a thermal instability locally in the corona. Due to the footpoint heating source the loop gets rapidly hot and dense. Thermal conduction acts efficiently for slowing the temperature increase over the loop. The heating source together with chromospheric evaporation will gradually increase the density in the loop bringing the heating per unit mass down and with it the average coronal temperature. Eventually, in a time scale of an hour, the temperature is sufficiently low to allow recombination of atoms locally, decreasing the temperature and the pressure locally in a timescale of minutes. The pressure drop accretes mass from the surroundings and produces upflows from the footpoints leading to the condensations. The dense clumps cool rapidly through radiation until the plasma becomes optically thicker and the cooling slows down \\citep{Goldsmith_1971SoPh...19...86G, Hildner_1974SoPh...35..123H, Mok_etal_1990ApJ...359..228M, Antiochos_Klimchuk_1991ApJ...378..372A, Antiochos_1999ApJ...512..985A, Karpen_etal_2001ApJ...553L..85K, Muller_2003AA...411..605M, Muller_2004AA...424..289M, Muller_2005AA...436.1067M, Mendozabriceno_2005ApJ...624.1080M, Mok_etal_2008ApJ...679L.161M, Antolin_2010ApJ...716..154A, Xia_etal_2011ApJ...737...27X}. It is still not clear how cool can coronal rain become from this process (before falling down to the solar surface). Since they have so far been observed in the same spectrum lines, coronal rain is often associated with prominences, and probably for both the thermal non-equilibrium mechanism is equally important \\citep{Dahlburg_etal_1998ApJ...495..485D, Karpen_etal_2001ApJ...553L..85K, Karpen_etal_2005ApJ...635.1319K, Karpen_etal_2006ApJ...637..531K, Liu_etal_11SPD....42.2119L}. However, one fundamental difference seems to be the magnetic field structure in which they appear, most likely determining the time scale of the phenomenon. While a prominence can live for days or even weeks, coronal rain can only exist during the time it takes for the blobs to fall to the surface (on the order of half an hour or less). The thermodynamic properties of the plasma are very likely linked to the underlying magnetic field topology. Coronal rain is restrained to coronal loops, where the magnetic field is basically bipolar and magnetic field lines are not far from being potential. On the other hand, the magnetic field topology in prominences seems to present complicated geometries such as `dips' or `peaks' \\citep[see][and references therein]{Lin_2010SSRv..tmp..112L}. Helical magnetic fields conforming flux ropes have also been proposed, in which apart from offering support from the Lorentz force to balance gravity, thermal isolation from the surrounding hot corona is also achieved due to the cross-field thermal conduction inhibition \\citep{Kippenhahn_Schluter_1957ZA.....43...36K, Low_Hundhausen_1995ApJ...443..818L}. The longer timescales involved in prominences may produce different element population leading to different ionization fraction, and consequently mechanisms such as ambipolar diffusion may play a different role. Another type of prominence model has been suggested in which the material is continuously renewed through flows between the prominence and the chromosphere, giving the impression of long-lived material \\citep{Priest_Smith_1979SoPh...64..267P, Tandberg-Hansen_1995ASSL..199.....T, Zirker_etal_1998Natur.396..440Z, Liu_etal_11SPD....42.2119L, Karpen_etal_2001ApJ...553L..85K, Karpen_etal_2005ApJ...635.1319K}. Since this case does not involve long-term gravitational support the required magnetic field may be more simple and features such as `dips' are not necessary: conditions that are more alike to coronal rain formation. High resolution observations have revealed that prominences are often composed of a myriad of fine threads, outlining a fine-scale structure of the magnetic field and the presence of flows along the threads \\citep{Heinzel_Anzer_2006ApJ...643L..65H, Lin_2005SoPh..226..239L, Lin_2008ASPC..383..235L, Lin_2010SSRv..tmp..112L, Martin_2008SoPh..250...31M}. Observations of coronal rain with {\\it Hinode}/SOT have also revealed a thread-like character in coronal loops, due to the separation and elongation of the blobs down to very small sizes \\citep{Antolin_2010ApJ...716..154A, Antolin_Verwichte_2011ApJ...736..121A}. The observations with the {\\it SST} presented here allow a closer look into this thread-like character of loops, suggesting the actual tracing of strands by the blobs. Due to the strong footpoint heating and the subsequent very high densities that follow a flare, thermal non-equilibrium leading to the formation of condensations is also proposed as the mechanism behind the observed cool downflows in post-flare loops \\citep{Foukal_1978ApJ...223.1046F, Schmieder_etal_1995SoPh..156..337S, Shimojo_etal_2002SoPh..206..133S, Hara_etal_2006ApJ...648..712H}. Despite the very different thermodynamic conditions in post-flare loops with respect to warm loops, the coronal rain observed in both cases appears similar. Although a rigorous comparison awaits, this could be easily explained since radiative cooling increases with the square of the density, and therefore achieving the same low temperatures in a similar timescale. It is possible however that the element population of the obtained blobs in post-flare loops is different, leading to different properties at short length scales. It is thus important to make the distinction between flaring and non-flaring active regions for coronal rain studies. Coronal rain is often observed to follow loop-like paths, suggesting a strong coupling between neutrals coming from recombination and the local ion population. Since the thermodynamical state and ionization fraction of such clumps are still poorly known it is not clear how strong this coupling is and, consequently, how well coronal rain blobs can trace the coronal magnetic field. Assuming that a blob does not alter the magnetic field lines, by studying its dynamics, morphology changes and thermodynamics we can learn about the local conditions in coronal loops. This is an idea that is exploited in this work. Supported by theoretical estimates of coupling processes based on the obtained results we show that the very small sizes of blobs allow them to be magnetic field tracers and probes of the local environment. In all of the observational papers concerning coronal rain it has been shown that the falling speeds are smaller than free fall, leading to downward accelerations considerably lower than that of the solar gravity at the surface. The gas pressure gradient inside the loops has often been suggested as the main agent regulating the speeds of the blobs. This is also supported by numerical simulations of thermal non-equilibrium \\citep{Mackay_2001SoPh..198..289M, Muller_2003AA...411..605M}. Magnetic pressure exerted by transverse MHD waves have also been suggested as factors determining the kinematics of blobs. In \\ion{Ca}{2}~H observations with {\\it Hinode}/SOT, \\citet{Antolin_Verwichte_2011ApJ...736..121A} detected such waves through coronal rain tracking. It is shown that the pressure force from the waves can explain the low falling speeds. Further investigation on the effect of waves on blobs such as those of coronal rain is being carried out \\citep{Verwichte_etal_2011}. Whether blobs are capable of producing such waves is also being investigated in that work. Based on the {\\it TRACE} observations, \\cite{Schrijver_2001SoPh..198..325S} estimated the occurrence time of coronal rain in a coronal active region loop to be at most once every two days stressing the sporadic character of coronal rain in active regions, and thus suggesting that thermal non-equilibrium may not play a significant role in coronal heating. These estimates are constrained by the resolution of {\\it TRACE}. Here we show that in order to have a clear picture about the phenomenon a resolution at least twice that of {\\it TRACE} is needed. New estimates of the occurrence time of coronal rain and the coronal volume involved are presented, as well as the coronal mass drain rate. The importance of thermal non-equilibrium (catastrophic cooling) in coronal heating is a matter of significant debate. So far, observations have pointed out five characteristics of warm loops ($\\sim1~$MK) in active regions, which any coronal heating model must explain. Namely, an excess density, a flat temperature profile, a super-hydrostatic scale height, an unstructured intensity profile, and a $1000-5000~$s lifetime. \\cite{Klimchuk_2010ApJ...714.1239K} have analyzed through one-dimensional simulations the case of thermal non-equilibrium in this context, considering both monolithic and multi-stranded loops. The simulations failed in general to satisfactorily reproduce the observations. For instance, monolithic models end up with far too much intensity structure, while multi-strand models are either too structured or too long-lived. On the other hand, 3D-MHD models of an active region with a footpoint heating function leading to thermal non-equilibrium have provided satisfactory results \\citep{Mok_etal_2008ApJ...679L.161M}. The resulting evolution of the light-curves, the variation of temperature along the loops, the density profile, and the absence of small scale structures in the intensity profiles are all compatible with real loops, thus supporting the thermal non-equilibrium scenario \\citep{Lionello_etal_2010AGUFMSH31C1811L}. Impulsive versus steady heating of loops is a common debate in studies of coronal heating \\citep[see][and references therein]{Klimchuk_2006SoPh..234...41K}. The scenario in which loops are heated impulsively to hot temperatures and subsequently cool down to EUV or lower temperatures is often proposed in observational studies of the corona \\citep{Warren_etal_2002ApJ...579L..41W, Winebarger_etal_2003ApJ...593.1164W, Winebarger_etal_2003ApJ...593.1174W, Viall_Klimchuk_2011ApJ...738...24V} but neither this nor the steady heating scenario are fully compatible with some observations \\citep{UgarteUrra_etal_2006ApJ...643.1245U}. Indeed, in some cooling events the {\\it TRACE} or {\\it SDO}/AIA filters do not seem to brighten according to their temperature. A possible solution to this has been put forward by \\citet{Kamio_etal_2011AA...532A..96K}, where, apart from investigating the upward propagation of quasi-periodic brightness fluctuations observed with {\\it SDO}/AIA, they also analyze sporadic cool downflows that they interpret as coronal rain. The peculiar order of the brightening in the AIA channels in time seem to be due to these cool downflows, particularly, because of the contribution of cool lines to hot channels in the AIA temperature response functions. Throughout their lifetimes the condensations progressively cool and the densities increase (in order to keep the local pressure roughly constant), producing a considerable increase in the cool line emission. The authors thus favor the thermal non-equilibrium mechanism to explain the observational results. Being the observational signature of thermal non-equilibrium (and hence to footpoint heating), coronal rain is deeply linked with coronal heating. As shown by the numerical simulations of the works cited previously thermal non-equilibrium is very sensitive to parameters such as the heating scale height and the loop length. By studying coronal rain we can then learn about the heating mechanisms. \\citet{Antolin_2010ApJ...716..154A} showed that Alfv\\'en wave heating, a strong coronal heating candidate, is not a predominant heating mechanism in loops with coronal rain. When propagating from the photosphere into the corona, Alfv\\'en waves can nonlinearly convert to longitudinal modes due to density fluctuations, wave-to-wave interaction, and deformation of the wave shape during propagation \\citep{Vasheghani_2011AA...526A..80V}. These modes subsequently steepen into shocks and heat the plasma uniformly along the loop \\citep{Moriyasu_2004ApJ...601L.107M, Antolin_2010ApJ...712..494A}, thus avoiding the loss of thermal equilibrium in the corona. In this work we deepen the link between coronal rain and coronal heating by studying the occurrence character of thermal non-equilibrium in neighboring strands. This paper is organized as follows. In section~\\ref{observations} we present the {\\it SST}/CRISP observations. In section~\\ref{results} the results of the statistical analysis are presented. These comprise dynamics, shapes, trajectories and thermodynamic properties. Discussion is presented in section~\\ref{discussion}, where we try to assess the role of coronal rain in the understanding of the coronal magnetic field. Conclusions are presented in section~\\ref{conclusions}. ", "conclusions": "In this paper we have analyzed H$\\alpha$ observations of coronal rain with the CRISP spectropolarimeter of the {\\it SST}. The phenomenon is observed above an active region at the limb. Kinematics (total velocities and accelerations), shapes (widths and lengths), trajectories (angles of falling blobs) and thermodynamic properties (temperatures) were derived for the condensations. On-disc blobs dark in absorption were also analyzed and shown to have the same dynamical, morphological and thermodynamical properties as the off-limb blobs. We have further shown that the profile of off-limb blobs becomes an absorption profile when passing against a bright background, from which we have concluded that on-disc blobs correspond to the same phenomenon. Being one of the first spectroscopic studies on coronal rain we have been able to measure the total velocities and shown that the falling velocities are lower than free fall, as has been suggested in previous works on the subject \\citep{Schrijver_2001SoPh..198..325S, DeGroof_2004AA...415.1141D, Muller_2005ESASP.596E..37M}. The corresponding downward accelerations are lower than the effective gravity along elliptic loops, suggesting the presence of other forces than gravity, possibly gas pressure as suggested by most numerical simulations. Transverse magnetic waves could not be clearly detected due to seeing effects so we cannot discard their presence and influence on the dynamics of the blobs. The blobs could often be traced to chromospheric heights, close to what appears to be the footpoints of the loops that host them. Strong decelerations were observed in some cases, suggesting the increase of gas pressure at those heights expected from the higher local densities. Through orders of magnitude estimates we have shown that the coronal rain observed in H$\\alpha$ is expected to come from a neutral Hydrogen population strongly coupled to the protons. At least on length scales on the order of the blob sizes (a few hundred km) no diffusion across the magnetic field is expected. Coronal rain can thus be considered as a tracer of the coronal magnetic field. Combining the projected and Doppler velocities we have shown that it is possible to retrieve the angles of fall of the blobs, allowing a reconstruction of the coronal magnetic field. No evidence for twisting or braiding of strands in loops was found, but we do not discard their existence at lower length scales (under 100 km), where turbulence may also be important. The tracing of strands by the blobs further suggest a constant area cross-section in the corona for these loops with no significant expansion down to chromospheric heights. The expected flux tube expansion may happen at lower (photospheric?) heights. One of our main results concerns the sizes of the observed blobs. Lengths and widths on the order of 1.5~Mm and 0.37~Mm, respectively, were found, stressing the need of high spatial (and temporal) resolution to fully observe this phenomenon. The blobs further present average temperatures of 7000~K, and possibly even lower if turbulence and the Stark effect are important. Coronal rain thus present cool and compact chromospheric cores of a few 100~km in width. Some evidence of progressive cooling over time was found, as suggested from previous work. We further found that coronal rain occurs frequently in neighboring strands in a simultaneous way, forming groups of condensations which, if close enough together, are seen as large clumps. We have termed these sporadic events as `showers', and they can have widths up to a few Mm. This is probably what has been observed in the past with instruments of lower spatial resolution such as {\\it TRACE} and {\\it SOHO}/EIT. Since we did not observe here any evidence of cool blobs nor showers in the co-temporal {\\it Hinode}/EIS observation we suspect that loops with coronal rain are multi-thermal, as suggested previously by \\cite{Schrijver_2001SoPh..198..325S}. A scenario in which the cool coronal rain cores are surrounded by diffuse hotter plasma needs to be further investigated. Our observations support the multi-stranded loop scenario and suggest that a significant fraction of strands in a loop do not have an independent thermodynamic evolution. Indeed, neighboring strands often display a coherent cooling (in the form of coronal rain) of a significant number of strands in a loop, suggesting a heating mechanism which acts in roughly the same way on neighboring strands. A footpoint heating mechanism imparting a similar heating scale height over all or part of the strands in a loop is a simple scenario that can explain these observations. Since we expect an expansion of the loops at low (possibly photospheric) heights spatially small heating events may suffice for such purpose. A coherent evolution of strands in a loop throughout their lifetime (from heating to cooling, from being dense to being evacuated) is a puzzling scenario that needs further observational testing. In case of being a common scenario it may pose several constraints on coronal heating models, especially on one-dimensional models. For instance, one constraint may be that a loop cannot be modeled by a collection of independently evolving one-dimensional strands. This may explain the disagreement regarding the importance of thermal non-equilibrium in the corona between the 1D \\citep{Klimchuk_2010ApJ...714.1239K} and the 3D simulations \\citep{Klimchuk_2010ApJ...714.1239K, Lionello_etal_2010AGUFMSH31C1811L}. The present observations suggest that coronal rain may not be a sporadic phenomenon of active regions as previously thought, but a rather common phenomenon deeply linked to the heating mechanisms of coronal loops. We have estimated the fraction of coronal volume with coronal rain to be between 10~\\% and 40~\\%. This is however strongly dependent on the magnetic field filling factor and on the projection effects. We have further estimated the occurrence time of this phenomenon in a loop to be between 4 and 15~hours. Longer datasets with larger fields of view are needed to clearly answer the question on how common coronal rain is. The obtained mass drain rate in the form of coronal rain is roughly $10^{10}$~g~s$^{-1}$ taking a mass density of $10^{-13}$~g~cm$^{-3}$ for the condensations. This number is on the same order of magnitude as the estimated mass flux into the corona from spicules, reinforcing the idea that coronal rain is an important phenomenon. This suggests a scenario in which the hot spicular material injected into the corona falls back cool, `raining' down. Irrespectively of the occurrence time of coronal rain, we have shown in this paper the large potential that this phenomenon beholds in the understanding of the coronal magnetic field. In \\citet{Antolin_2010ApJ...716..154A} we have shown that coronal rain can act as a marker of the coronal heating mechanisms. Here, besides further elucidating its link to coronal heating, we have shown that it constitutes a tracer of the internal structure of loops and that it can be a probe of the local thermodynamic conditions in loops." }, "1112/1112.0886_arXiv.txt": { "abstract": "The Kilo Degree Survey (KiDS) is a 1500 square degree optical imaging survey with the recently commissioned OmegaCAM wide-field imager on the VLT Survey Telescope (VST). A suite of data products will be delivered to ESO and the community by the KiDS survey team. Spread over Europe, the KiDS team uses \\textsf{Astro-WISE} to collaborate efficiently and pool hardware resources. In \\textsf{Astro-WISE} the team shares, calibrates and archives all survey data. The data-centric architectural design realizes a dynamic 'live archive' in which new KiDS survey products of improved quality can be shared with the team and eventually the full astronomical community in a flexible and controllable manner. ", "introduction": "One of the radical advances that optical astronomy has seen in recent years is the advent of wide-field CCD-based surveys. On Paranal, ESO has recently started operating two dedicated survey telescopes: VISTA in the infra-red wavelength region and the VLT Survey Telescope (VST) in the optical. The lion's share of the observing time on both survey telescopes will be invested in a set of Public Surveys. The largest of the optical surveys is the Kilo-Degree Survey (KiDS), which will image 1500 square degrees in four filters ($u$,$g$,$r$,$i$) over a period of 3--4 years. Combined with one of the VISTA surveys, VIKING, which will observe the same area in ZYJHK, this will provide a sensitive, 9-band multi-colour survey. \\begin{figure}[ht] \\includegraphics[width=\\textwidth]{P157_fig1.ps} \\caption{Lay-out of the KiDS-North (top) and KIDS-South (bottom) fields, shown by the hatched areas. Also shown are the areas where 2DF spectra are available, indicated by the large dots, and the area covered by DR7 of the SDSS survey, indicated by the small dots. The CFHTLS-W2 field and the DS/COSMOS deep field are overplotted on the top panel.} \\label{fig:areas} \\end{figure} {\\bf Observational set-up.} KiDS will cover 1500 square degrees, which is approximately 7\\% of the extragalactic sky. It consists of two patches that ensure that observations can take place year-round. The Northern patch lies on the celestial equator, while the Southern area straddles the South Galactic Pole (Fig. \\ref{fig:areas}). These specific areas were chosen because they were the target of massive spectroscopic galaxy surveys already: the 2dF redshift survey \\citep{2dfgrs} covers almost the same area, and KiDS-North overlaps with the SDSS spectroscopic and imaging survey \\citep[SDSS, ][]{sdssdr8}. The exposure times for KiDS and VIKING have been chosen to yield a median galaxy redshift of 0.8, so that the evolution of the galaxy population and matter distribution over the last $\\sim$ half of the age of the universe can be studied. They are also well-matched to the natural exposure times for efficient VST and VISTA operations, and balanced over the astro-climate conditions on Paranal (seeing and moon phase) so that all bands can be observed at the same average rate. This strategy makes optimal use of the fact that all observations are queue-scheduled, allowing the best seeing time to be used for deep $r$-band exposures, for example, and the worst seeing for $u$. {\\bf Science drivers.} The main scientific objective of KiDS and VIKING is to map the matter distribution in the universe through weak gravitational lensing and photometric redshift measurements. The large numbers of galaxies that KiDS will detect, with accurate photometric redshifts up to $z\\simeq1.2$ will allow the Baryonic Accoustic Oscillations, an important cosmological standard candle, to be measured over a large redshift range, and thus unveil its evolution. Galaxy-galaxy lensing (GGL) studies into the structure of galaxy halos for various redshifts and galaxy types, will exploit the excellent image quality of the OmegaCAM wide-field camera and the VST on the one hand and the shear size of the KiDS data set on the other. The deep photometry and accurate photometric redshifts also will ensure that KiDS data will be a powerful tool to study the evolution of galaxies and clusters out to redshifts of $z\\simeq1.5$. Additionally, the extensive data set that KiDS will deliver, will be useful in a broad range of rsearch areas in astronomy, for example the study of stellar streams and the Galactic halo. {\\bf Survey data products.} Being a Public Survey, all KiDS data will be made publicly available. The KiDS catalogue will contain some 100,000 sources per square degree (150 million sources over the full survey area), and for each square degree there will be 10 GB of final image data, 15 TB for the whole survey. A set of basic data products will be made public, both through ESO and through the \\textsf{Astro-WISE} database: calibrated coadded images, weight maps, calibration images, single-band and multi-band catalogues. In the long-term, we intend to provide more advanced data products, for example images with gaussianized point-spread-functions, or morphological parameters of all detected sources. ", "conclusions": "" }, "1112/1112.0248_arXiv.txt": { "abstract": "In cosmic-ray physics, large field of view experiments are triggered by a number of signals laying on different angular scales: point-like and extended gamma-ray sources, diffuse emissions, as well as large and intermediate scale cosmic-ray anisotropies. The separation of all these contributions is crucial, mostly when they overlap with each other. Needlets are a form of spherical wavelets that have recently drawn a lot of attention in the cosmological literature, especially in connection with the analysis of CMB data. Needlets enjoy a number of important statistical and numerical properties which suggest that they can be very effective in handling cosmic-ray and gamma-ray data analysis. An application of needlets to astroparticle physics is shown here. In particular, light will be thrown on how useful they might be for estimating background and foreground contributions. Since such an estimation is expected to be optimal or nearly-optimal in a well-defined mathematical sense, needlets turn out to be a powerful method for unbiased point-source detections. In this paper needlets were applied to two distinct simulated datasets, for satellite and EAS array experiments, both large field of view telescopes. Results will be compared to those achievable with standard analysis tecniques in any of these cases. ", "introduction": "\\label{sec:needlets} Needlets enjoy quite a few important properties that make them very suitable for spherical data analysis. Indeed, needlets do not rely on any tangent plane approximation and they are perfectly adapted to standard packages; their main property is the capability of unfolding directional data in the harmonic space, so that every order $j$ of needlets contains power only from within a certain range in the $l$ space. In the real space for any fixed angular distance the tail of the needlets decay faster than any polynomial, i.e. quasi-exponentially as the frequency increases. It is worth noticing that for needlets the exact reconstruction property holds, i.e. back-transform exists giving the input signal within numerical accuracy. We do not give details about needlets construction here (on this purpose, see \\cite{npw1,bookmarinucci}). Nonetheless, we recall that the needlet idea has been extended by Geller and Mayeli \\cite{gm1} with the construction of so called Mexican needlets, see also \\cite{scodeller} for numerical analysis and implementation in a cosmological framework. Mexican needlets have localization properties in real space even better than standard needlets, that is why they were applied to get the results presented here. Numerical codes have been developed by the author to compute the needlet estimators $\\widehat{\\beta}_{jk}$ for any sky map containing density information. Some other codes are publicly available \\cite{needatool} and results reported here have been suitably cross-checked. The index $j$ refers to the needlet order, while $k$ runs along the pixels into which the sphere has been discretized (the Healpix package is used here \\cite{Healpix}, with the ``ring'' numbering scheme). When the needlet transform is numerically computed, three parameters are important, i.e. $j_0$ (the first order computed) and $n_j$ (how many needlet orders are computed) and $B>1$ which is a number used to pass from the $j$-th order to the $(j+1)$-th. A golden rule to be used is that the order $j$ contains power mostly from multipoles in the range $[B^{j-1},B^{j+1}]$\\footnote{That is, if $B=1.6$, then $j=6$ corresponds to $l\\sim1.6^{6}=17$, or $\\psi\\sim10^{\\circ}$.}. If mexican needlets are used, one more parameter is important, $p$, which describes how fast needlets go to zero in the harmonic space. We shall now consider the analysis of data from cosmic rays observatories, following classical approaches to wavelet-based density estimation. An idea that we shall pursue is to implement \\emph{thresholding} estimates, as discussed for instance by \\cite{bkmpAoSb}. More precisely, we can consider the nonlinear estimate $\\widehat{f}_{n}^{\\ast }(x)$ obtained by suitably weighting the needlet coefficients in the back-transform. The weight function $w_{jk}(\\widehat{\\beta }_{jk})$ is some nonlinear function that ``shrinks'' beta towards zero. Such estimates can be shown to be robust and nearly optimal over a wide class of density functions and different loss functions, i.e. figures of merit by which to measure when the estimate is \"close\" to the density to be estimated. Hereafter, we will name {\\em source-map} the sky-map as it comes from the experiment; {\\em beta-maps} those containing the coefficient estimators $\\widehat\\beta_{jk}$; {\\em reconstructed maps} or {\\em back-transformed maps}, those obtained after the application of the inverse needlet transform. ", "conclusions": "Recently, needlets drew the attention of the scientific community for their important applications in data-analysis of cosmological data as a new form of spherical wavelets. We presented here some applications of the needlet transform to high energy cosmic ray physics, showing the very good properties of localization.The needlet transform is sensitive in the whole harmonic domain, provided that enough orders of needlets are computed. In particular, the application to a simulation of the ARGO-YBJ and the Fermi-LAT data-sets found again the well-known intermediate scale anisotropy at low orders and the point gamma-ray sources at higher orders, thus showing the possiblity of decting sources directly in the needlet space. The significance of the beta-maps is carried out quite easily if the background distribution is known and the variance of the beta coefficients may be used to threshold the signal in the beta space, which turned out to be a very promising technique." }, "1112/1112.4141_arXiv.txt": { "abstract": "Rastall's theory is based on the non-conservation of the energy-momentum tensor. We show that, in this theory, if we introduce a two-fluid model, one component representing vacuum energy whereas the other pressure-less matter (e.g. baryons plus cold dark matter), the cosmological scenario is the same as for the $\\Lambda$CDM model, both at background and linear perturbation levels, except for one aspect: now dark energy may cluster. We speculate that this can lead to a possibility of distinguishing the models at the non-linear perturbation level. ", "introduction": "Since the formulation of General Relativity, about one hundred years ago, many alternative geometric theories have been proposed in order to explain gravitation phenomena (e.g. \\cite{Cartan1922, Cartan1923, Brans1961, Rosen1971, Rastall1973, Rastall1976, Moffat:1994hv, Bekenstein:2004ne}). Some of these touch one important aspect of General Relativity: the conservation law. To our knowledge, the first non-conservative theory of gravity was the steady-state model \\cite{Bondi1948, Hoyle1948}, following some ideas already presented in the end of the forties by Jordan \\cite{Jordan:1949zz}. In the beginning of the seventies, Rastall proposed one new version of a non-conservative theory of gravity, following the remark that the conservation law ${T^{\\mu\\nu}}_{;\\mu} = 0$ may not hold true in a curved space-time \\cite{Rastall1973, Rastall1976}. Hence, he argued that new gravitational equations can be obtained considering a modification of the conservation law such that \\begin{eqnarray} \\label{mot1} {T^{\\mu\\nu}}_{;\\mu} = \\kappa R^{;\\nu}, \\end{eqnarray} where $T^{\\mu\\nu}$ is the energy-momentum tensor, $\\kappa$ is a coupling constant and $R$ is the Ricci scalar curvature. Hence, in the weak field limit, the usual expressions are preserved. Since, generally, the Ricci scalar curvature is connected with the trace of the energy-momentum tensor, Eq.~(\\ref{mot1}) can be re-written as \\begin{eqnarray} {T^{\\mu\\nu}}_{;\\mu} = \\bar\\kappa T^{;\\nu}\\;, \\end{eqnarray} where $\\bar\\kappa$ is a new constant and $T$ is the trace of the energy-momentum tensor. It is curious to remark that the phenomenon of particle creation in cosmology \\cite{Gibbons:1977mu, Parker:1971pt, Ford:1986sy} also leads to a violation of the classical conservation laws and, in this sense, Rastall's idea may be viewed as a kind of classical formulation of that quantum phenomenon, since the violation of the energy-momentum conservation is connected with the curvature. More in detail, Rastall's modification to Einstein equations take the following form ($c = 1$ units): \\begin{eqnarray} \\label{eq1R} R_{\\mu\\nu} - \\frac{1}{2}g_{\\mu\\nu}R &=& 8\\pi G\\left(T_{\\mu\\nu} - \\frac{\\gamma- 1}{2}g_{\\mu\\nu}T\\right)\\;,\\\\ \\label{eq2R} {T^{\\mu\\nu}}_{;\\mu} &=& \\frac{\\gamma - 1}{2}T^{;\\nu}\\;, \\end{eqnarray} where $\\gamma$ is a parameter (the choice $\\gamma = 1$ restores General Relativity). Note that it seems possible to have a Lagrangian formulation from which the above equations are deduced \\cite{Smalley1984}. Since for a radiative fluid $T = 0$, implying $R = 0$, we can expect that the cosmological evolution during the radiative phase is the same as in the standard cosmological scenario. At same time, a single fluid inflationary model, described by a cosmological constant, is the same as it would be in the General Relativity case. Hence, Rastall cosmologies may have an important departure from the standard cosmological model from the beginning of the matter dominated phase on \\cite{Batista:2010nq, Fabris:2011rm, Fabris:2011wz, Capone:2009xm}. ", "conclusions": "We have shown that there is a subset of cosmological scenarios based on Rastall's energy-momentum tensor non-conservation that are equivalent to the $\\Lambda$CDM cosmology, except for one aspect: vacuum energy may agglomerate. Two effects could allow to discriminate between the two models: non-linear effects in the matter power spectrum and the transfer function for cosmological perturbations. Vacuum energy is negligible in the past, hence the impact of its fluctuation in the evolution of the perturbations in the primordial periods of the evolution of the universe may not be relevant. However, according to the scenario described above, dark energy must be present in virialized systems, like galaxies and cluster of galaxies, and the effect of agglomeration of dark energy must be relevant at this level. Therefore, Rastall's cosmology and the $\\Lambda$CDM model seem to be distinguishable only at the non-linear regime of the evolution of cosmic perturbations. Though we have given some hints of the latter possibility addressing the simple case of a spherical top-hat collapse, a deeper analysis shall be the subject of future investigation. \\subsection*{Acknowledgements} The authors thank the anonymous referee for his kind and useful suggestions. J. C. F. thanks CNPq (Brasil) for partial financial support." }, "1112/1112.1880_arXiv.txt": { "abstract": "The thermal evolution of isothermal neutron stars is studied with matter both in the hadronic phase as well as in the mixed phase of hadronic matter and strange quark matter. In our models, the dominant early-stage cooling process is neutrino emission via the direct Urca process. As a consequence, the cooling curves fall too fast compared to observations. However, when superfluidity is included, the cooling of the neutron stars is significantly slowed down. Furthermore, we find that the cooling curves are not very sensitive to the precise details of the mixing between the hadronic phase and the quark phase and also of the pairing that leads to superfluidity. ", "introduction": "Neutron stars are natural laboratories for physics under extreme conditions with their extremely high densities, powerful energy emission, large magnetic fields, and millisecond rotation periods. At the densities near the surface of such a star, atoms break apart into nuclei and electrons. At higher densities, the electrons neutralize with the protons in the nuclei to form neutrons. These stars thus consist of a large fraction of neutrons and are supported from gravitational collapse by the neutron degeneracy pressure, from which the neutron star derives its name. However, at high densities the existence of more exotic particles is expected. These particles are generated by processes which produce strangeness, such as \\begin{equation}\\label{lambda} n+n\\rightarrow n+\\Lambda^0+K^0, \\end{equation} where $n$ is the neutron, $\\Lambda^0$ the Lambda hyperon, and $K^0$ the strange meson. The strange meson can decay via various weak processes and the final products, usually photons and neutrinos, leak out of the star. Therefore, the reverse process is reduced and some net amount of strangeness survives in the dense core of the star \\cite{Glendenning:1985}. It is generally believed that these strangeness carrying particles, called hyperons, can exist in the center of neutron stars. They coexist with the nucleons as well as the leptons $e^-$ and $\\mu^-$. The interactions between them are dominated by the complicated nuclear force whose carriers are the mesons. We refer to such a system as the hadronic phase of matter. The system with only nucleons and leptons, i.e., without hyperons, is referred to as the nuclear phase of matter. At even higher densities, as a consequence of asymptotic freedom, quarks become deconfined from the hadrons. Therefore, strange quark matter, which consists of $u$, $d$ and $s$ quarks, may also exist in the neutron star core. There may also be a phase-separated mixture of strange quark matter and hadronic matter in a certain range of densities \\cite{Glendenning:1992}, which we call the mixed phase. At present, there is still a lot unknown about the deconfinement phase transition of quarks. The contemporary knowledge on matter at high densities and temperatures is shown by the schematic QCD phase diagram in Fig.~\\ref{QCD}. Since the baryon chemical potential $\\mu_B$ is a monotonously increasing function from the surface to the center of the neutron star, the $\\mu_B$ axis can be mapped to the stellar radius and the different phases in the neutron star are explicitly indicated. \\begin{figure} \\resizebox{1\\linewidth}{!}{\\includegraphics*{QCD.eps}} \\caption{The schematic QCD phase diagram as a function of temperature $T$ and baryon chemical potential $\\mu_B$. The solid lines denote first- and second-order phase transitions, whereas the dashed lines denote smooth crossovers. According to the typical temperature of a proto-neutron star of $T \\sim 10^{10}$ K $\\sim 1$ MeV, the state of matter in the neutron star should be close to the bottom line, as indicated by the grey thick line.}\\label{QCD} \\end{figure} Because of the limited knowledge on the state of matter at high densities and the complexity of the interactions between the particles, many effective models for matter inside neutron stars have been constructed. In general, these models supply an equation of state, which determines the particle composition of the neutron star. The observation of neutron stars, in turn, constrain these models. For example, if an equation of state is too soft it is incapable of supporting a very large stellar mass, such that some models become disfavored whenever the data on the heaviest neutron star is updated. The stellar mass is not the only constraint on the models. The cooling of neutron stars can be studied from their luminosity as a function of time. A proto-neutron star is born with a typical temperature larger than $10^{10}\\textrm{ K}$, after which it mainly loses its energy by two processes, namely by neutrino emission everywhere inside the star and by photon radiation at the surface. In the early stages of the thermal evolution of the star, neutrino emission is the dominant cooling effect after which photon radiation ultimately takes over \\cite{Yakovlev:2001}. Since neutrino emission occurs everywhere in the neutron star, it provides a probe for studying the state of matter inside the star. The most efficient neutrino-emission process is called the direct Urca (DUrca) process \\begin{equation}\\label{DUrca} n\\rightarrow p+e^-+\\bar{\\nu}_e,\\quad p+e^-\\rightarrow n+\\nu_e. \\end{equation} This process is only possible if the proton fraction is more than a certain threshold in order for it to satisfy energy and momentum conservation \\cite{Lattimer:1991}. Historically, the proton abundance in neutron stars was underestimated. In that scenario, it is reasonable to take into account also the modified Urca (MUrca) process, where a bystander helps the momentum conservation, for example \\begin{align}\\label{MUrca} n+n\\rightarrow p+n+e^-+\\bar{\\nu}_e,&\\ p+n+e^-\\rightarrow n+n+\\nu_e,\\nonumber\\\\ n+p\\rightarrow p+p+e^-+\\bar{\\nu}_e,&\\ p+p+e^-\\rightarrow n+p+\\nu_e. \\end{align} Some other processes can have a neutrino emissivity which is smaller or comparable to the modified Urca process, such as neutrino bremsstrahlung and plasmon decay \\cite{Yakovlev:2001}. However, all of them are negligible whenever the direct Urca channel is open. In the hadronic phase, the direct Urca processes are quite rich and can be summarized as \\begin{equation}\\label{DUrcabaryon} b_1\\rightarrow b_2+l^-+\\bar{\\nu}_l,\\quad b_2+l^-\\rightarrow b_1+\\nu_l, \\end{equation} where $b_1$ and $b_2$ denote two different baryons and $l^-$ represents one of the leptons, $e^-$ or $\\mu^-$. For strange quark matter, the direct Urca processes are \\begin{align}\\label{DUrcaquark} d\\rightarrow u+l^-+\\bar{\\nu}_l,\\quad u+l^-\\rightarrow d+\\nu_l,\\nonumber\\\\ s\\rightarrow u+l^-+\\bar{\\nu}_l,\\quad u+l^-\\rightarrow s+\\nu_l, \\end{align} which are simply the direct Urca processes of the baryons at the quark level. Inside the neutron star, there is no threshold for the direct Urca process in quark matter, since the Fermi momenta of the two quarks and lepton can always satisfy momentum conservation. As mentioned, the neutrino emissivity depends strongly on the type of matter contributing to the process, such that the thermal evolution of a star directly probes its composition. The main goal of this paper is to study the cooling of neutron stars with different types of equations of state and try to constrain them by comparing with observational data. A large amount of research on the cooling process of neutron stars has been carried out in the last several decades, with a strong focus on the nuclear phase of matter. To the best of our knowledge, a unifying model covering the low-density nuclear phase to the deconfined quark phase, consistent with the QCD phase diagram, has been less thoroughly explored. The strange quark star composed of strange matter, either with or without the nuclear crust, has been considered in some papers \\cite{Schaab:1996,Page:2002}. However, the cooling behavior of neutron stars containing also all hyperons and possibly also a mixed phase of strange quark matter and hadronic matter, has not been extensively studied. Two possible reasons for this can be given. First, the hadronic and mixed equations of state are rather soft and are continuously being challenged by new data on heavy neutron stars, such as PSR J1903+0327 which has $M=1.67\\pm0.01\\textrm{ M}_\\odot$ \\cite{Freire:2009} and PSR J1614$-$2230 which even has $M=1.97\\pm0.04\\textrm{ M}_\\odot$ \\cite{Demorest:2010}. Second, the existence of a mixed phase is being questioned in view of screening and surface effects \\cite{Voskresensky:dual}. With respect to the first concern, we note that most of the observed neutron star masses still lie below the maximum mass a hadronic or mixed equation of state can allow for. According to Ref.~\\cite{Timmers:1996}, most nearby young neutron stars \\cite{Popov:2003} have masses no bigger than $1.4\\textrm{ M}_{\\odot}$ \\cite{Page:2006}. Therefore, at least for the study of the thermal evolution of these stars, the hadronic or mixed phase can still be of great importance. In fact, because of the uncertainties in the interaction between particles at extremely high density, it is hard to exactly determine the equation of state at high density. Although the equations of state used in this paper cannot support neutron stars as massive as those reported above, we can still use them for a discussion on medium-mass neutron stars with more complicated compositions. As for the stability of the mixed phase, the arguments are still indecisive. For example, many details of the surface tension, which strongly influences the stability calculation, are still uncertain. Although it is expected that screening and surface effects diminish the mixed-phase regime, it is far from certain that its existence can be excluded \\cite{Endo:2011}. The mixed phase can in particular have significant effects on the cooling behavior, especially for the heat transport inside the star. However, we show below that its effect will be less important after the star has become isothermal, when the thermal evolution is determined by the heat capacity and neutrino emissivity integrated over the whole volume, and the inner thermal conductivity no longer plays a role. The temperature of a neutron star is generally much smaller than the typical Fermi energy as a consequence of the very high densities in the star. Therefore, superfluidity may play an important role. According to BCS theory, fermions can form Cooper pairs at low temperatures via an attractive interaction and thereby lower the energy of the system. The resulting pairs, which obey Bose statistics, can form a Bose-Einstein condensate and the system becomes superfluid. Pair formation changes the single-particle dispersion around the Fermi surface and consequently the heat capacity and neutrino emissivity will be influenced. We find that, without superfluidity, the cooling of neutron stars is too fast compared with observations. However, by including the effects of superfluidity we obtain a more realistic cooling behavior. The quark phase is usually referred to as an exotic phase of extremely dense matter, in contrast to the nuclear phase or the hadronic phase. Other exotic phases have also been proposed, among which the pion and kaon condensates have attracted much attention \\cite{Ohnishi:2009,Brown:2008}. However, the existence of such phases inside the neutron star is still an open question. On the one hand, with such condensates, the equation of state is further softened and thus the corresponding maximum star mass is reduced \\cite{Potekhin:2011}, which makes such phases less favored when compared to the observational data of massive stars, as mentioned above. On the other hand, for the cooling process, it was reported that a meson condensate can increase the neutrino emissivity over the typical modified Urca emissivity by several orders of magnitude \\cite{Yakovlev:2001}, but it is still much less efficient than the direct Urca process. Since in our calculation the direct Urca process is always present, such enhancement from the meson condensate has a negligible effect. Besides, the kaon condensation may even reduce the pressure and cause the star to collapse into a black hole \\cite{Brown:2008}. Therefore, considering all the above arguments, we do not include such meson condensates in this paper. The paper is organized as follows, we first introduce in Sec. II our theoretical framework which includes the relativistic stellar structure, the unified mean-field model describing the state of matter inside the star, and the thermal evolution equations. In Sec. III, we then present the numerical results for the cooling curves of neutron stars with the nuclear, hadronic, and mixed equations of state. Due to the efficiency of the direct Urca process the cooling is seen to be too fast compared to observations. Therefore superfluidity is included in Sec. IV and, as a consequence, the cooling is slowed down and we find a much better agreement with observations. ", "conclusions": "We have carefully compared the thermal evolution of different types of neutron stars, namely with the hadronic, the mixed phase of hadronic and strange quark matter, and the nuclear equation of state. We find that the direct Urca process is open in all of these cases and thus results in relatively fast cooling behavior. Although the details concerning the heat capacity and neutrino emissivity can be rather different in these cases, the cooling curves are quite similar after the stars become isothermal. However, the behavior in the early stages before the stars become isothermal could be significantly different, but such a study requires the knowledge of the thermal conductivity of these complex systems. The geometrical structure of the mixed phase is also expected to play an important role, although no decisive conclusion has yet been drawn. As we have seen, the fast cooling in the non-superfluid case did not agree with observations. In order to remedy this discrepancy, superfluidity was introduced, which significantly reduces the efficiency of the direct Urca process as well as the heat capacity. The resulting cooling curve is much closer to the observational data. We also found that the particular pairing type of the superfluid baryons is not very important to the thermal evolution of the star. The thermal evolution after the star becomes isothermal is not strongly dependent on the initial temperature of the star or on the critical temperature related to superfluidity, as long as they are within reasonable ranges. The robustness of the results with superfluidity is quite helpful in order to remove the uncertainties concerning baryon superfluidity at high density. Note that the cooling process is almost completely determined by the direct Urca process and photon radiation even after including the effects due to superfluidity. The cooling curves of the neutron stars with the three equations of state are still quite similar when superfluidity is included. We expect that a calculation including the magnetic field and rotation gives an even better agreement with the observational data. Besides, since our nuclear model is geared originally towards nuclear matter near the saturation point, a further improvement of it may play an important role in getting a better agreement with the observations. By incorporating the properties at higher densities, e.g., the coupling constants for hyperons, it is possible to get cooling curves covering most of the observational data, as in Ref. \\cite{Page:2006}." }, "1112/1112.4539_arXiv.txt": { "abstract": "The $k$d-tree is a fundamental tool in computer science. Among other applications, the application of $k$d-tree search (by the tree method) to the fast evaluation of particle interactions and neighbor search is highly important, since the computational complexity of these problems is reduced from $O(N^2)$ for a brute force method to $O(N \\log N)$ for the tree method, where $N$ is the number of particles. In this paper, we present a parallel implementation of the tree method running on a graphics processing unit (GPU). We present a detailed description of how we have implemented the tree method on a Cypress GPU. An optimization that we found important is localized particle ordering to effectively utilize cache memory. We present a number of test results and performance measurements. Our results show that the execution of the tree traversal in a force calculation on a GPU is practical and efficient. ", "introduction": "A technique for gravitational many-body simulations is a fundamental tool in astrophysical simulations because the gravitational force drives structure formation in the universe. The length scales that arise in structure formation range from less than 1 cm for the aggregation of dust to more than $10^{24}$ cm for the formation of cosmological structures. At all scales, gravity is a key physical process for the understanding of structure formation. The reason behind this is the long-range nature of gravity. Suppose we simulate structure formation with $N$ particles. The flow of the many-body simulation is as follows. First, we calculate the mutual gravitational forces between the $N$ particles, then integrate the orbits for the $N$ particles, and repeat this process as necessary. Although it is simple in principle, the force calculation is a challenging task from the point of view of computer science. A simple, exact method for the force calculation requires $O(N^2)$ computational complexity, which is prohibitively compute-intensive for large $N$. An exact force calculation is necessary in some types of simulations, such as few-body problems, the numerical integration of planets orbiting around a star (e.g., the Solar System), and the evolution of dense star clusters. For simulations that do not require exact forces, however, several approximation techniques have been proposed \\citep{Hockney_1981, Barnes_1986, Greengard_1987}. The particle--mesh/particle--particle--mesh method \\citep{Hockney_1981} and the tree method \\citep{Barnes_1986} reduce the computational complexity of the force calculation to $O(N \\log N)$. The fast multipole method (FMM) reduces it further to $O(N)$ \\citep{Greengard_1987}. Of these methods, the tree method has been used extensively in astrophysical simulations, since its adaptive nature is essential for dealing with clumpy structure in the universe \\citep[e.g.,][]{Bouchet_1988}. Despite the $O(N \\log N)$ complexity, computational optimization of the tree method by techniques such as vectorization and parallelization is necessary to accommodate demands for simulations with larger and larger $N$. \\cite{Hernquist_1990}, \\cite{Makino_1990}, and \\cite{Barnes_1990} have reported various techniques to vectorize the force calculation with the tree method. \\cite{Warren_1992}, \\cite{Dubinski_1996}, and \\cite{Yahagi_1999} have reported a parallel tree method for massively parallel processors (MPPs). In a recent publication \\citep{Springel_2005}, a simulation of large-scale structure formation in the universe with more than ten billion particles, using a parallel tree code running on an MPP, has been reported. Another computational technique to speed up the tree method utilizes the GRAPE special-purpose computer \\citep{Sugimoto_1990,Makino_1998}. Using a combination of vectorization techniques for the tree method, the tree method can be executed efficiently on a GRAPE system \\citep{Makino_1991}. Cosmological simulations are a ``grand challenge\" problem. The Gordon Bell prizes have been awarded many times for cosmological simulations \\citep{Warren_1992a, Fukushige_1996, Warren_1997, Warren_1998, Kawai_1999, Hamada_2009}. In those simulations, both parallel tree codes \\citep{Warren_1992a, Warren_1997, Warren_1998} and a tree code running on a GRAPE system \\citep{Fukushige_1996, Kawai_1999} and a graphics processing unit (GPU) \\citep{Hamada_2009} were used to perform cosmological simulations. In the present paper, we describe our implementation of the tree method on a GPU. The rise of the GPU forces us to rethink our way of doing parallel computing, since the performance of recent GPUs has reached the impressive level of $> 1$ Tflops. Acceleration techniques for many-body simulations with a GPU have already been reported \\citep[e.g.,][]{Nyland_2007, Portegies_2007, Hamada_2007,Belleman_2008}; however, these techniques have implemented an exact, brute force method with $O(N^2)$ complexity. It is apparent, however, that for applications that do not require exact forces, it is possible to do much more efficient computation by the tree method. We have directly implemented the tree method on a GPU so that we can enjoy the speed of an $O(N \\log N)$ algorithm. For small $N < 30,000$, the brute force method on a GPU is faster than the tree method owing to extra work concerning the tree data structure. However, our results show that the tree method significantly outperforms the brute force method on a GPU for $N \\gg 10,000$, which is the standard size in current astrophysical simulations is. Our code is simple and easy to extend to other numerical algorithms that require a neighbor list or a short-range force, such as algorithms for the smoothed particle hydrodynamics (SPH) method \\citep{Gingold_1977, Lucy_1977}. ", "conclusions": "In this paper, we have described our implementation of the tree method on a Cypress GPU. By transforming a recursive tree traversal into an iterative procedure, we have shown that the execution of a tree traversal together with a force calculation on a GPU can be practical and efficient. In addition, our implementation shows performance comparable to that of a recently reported FMM code on a GPU. We can expect to get further performance gains by fully utilizing the four-vector SIMD operations of the SCs of the GPU. Moreover, since 10--20\\% of $T_{\\rm total}$ is spent on the tree construction, parallelization of this part on a multicore CPU will be an effective way to boost the total performance. Provided that we can easily extend our code to implement a force calculation for short-range interactions by a method such as the SPH method, we believe that a future extended version of our code will enable us to do a realistic astrophysical simulation that involves baryon physics with $N > 1,000,000$ very rapidly. It is fairly easy to incorporate higher-order multipole expansion terms into our method, and it would be a natural extension to the present work. Another good application of the tree method on a GPU would be to simulations that adopt a symmetrized Plummer potential \\citep{Saitoh_2010}. We believe that our method is the best for implementing that proposal, and hence that we shall certainly obtain better accuracy and good performance in simulating galaxy evolution and formation with different mass resolutions." }, "1112/1112.3319_arXiv.txt": { "abstract": "We derive bounds on the mixing between the Standard Model (``active'') neutrinos and their right-chiral (``sterile'') counterparts in the see-saw models, by combining neutrino oscillation data and results of direct experimental searches. We demonstrate that the mixing of sterile neutrinos with any active flavour can be significantly suppressed for the values of the angle $\\theta_{13}$ measured recently by Daya Bay and RENO experiments. We reinterpret the results of searches for sterile neutrinos by the PS191 and CHARM experiments, considering not only charged current but also neutral current-mediated decays, as applicable in the case of see-saw models. The resulting \\emph{lower bounds} on sterile neutrino lifetime are up to an order of magnitude \\emph{stronger} than previously discussed in the literature. % Combination of these results with the upper bound on the lifetime coming from primordial nucleosynthesis rule out the possibility that two sterile neutrinos with the masses between 10 MeV and the pion mass are solely responsible for neutrino flavour oscillations. We discuss the implications of our results for the Neutrino Minimal Standard Model (the $\\nu$MSM). ", "introduction": "\\label{sec:introduction} Transitions between neutrinos of different flavours (see e.g.~\\cite{Strumia:06} for a review and Refs.~\\cite{Schwetz:08a,Schwetz:2011zk,Fogli:2011qn} for the recent update of experimental values) are among the few firmly established phenomena \\emph{beyond the Standard Model of elementary particles}. The simplest explanation is provided by the ``neutrino flavour oscillations'' -- non-diagonal matrix of neutrino propagation eigenstates in the weak charge basis. While the absolute scale of neutrino masses is not established, particle physics measurements put the sum of their masses below $2\\ev$~\\cite{PDG:11} while from the cosmological data one can infer an upper bound of $0.58\\ev$ at 95\\% CL ~\\cite{WMAP7}. A traditional explanation of the smallness of neutrino masses is provided by the \\emph{see-saw mechanism}~\\cite{Minkowski:77,Ramond:79,Mohapatra:79,Yanagida:80}. It assumes the existence of several \\emph{right-handed neutrinos}, coupled to their Standard Model (SM) counterparts via the Yukawa interaction, providing the Dirac masses, $M_\\dir$, for neutrinos. The Yukawa interaction terms dictate the SM charges of the right-handed particles: they turn out to carry no electric, weak and strong charges; therefore they are often termed ``singlet,'' or ``sterile'' fermions. Sterile neutrinos can thus have Majorana masses, $M_\\m$, consistent with the gauge symmetries of the Standard Model. If the Majorana masses are much larger than the Dirac ones, the \\emph{type I seesaw formula}~\\cite{Minkowski:77,Ramond:79,Mohapatra:79,Yanagida:80} holds, expressing the mass matrix of observed neutrinos $(\\mathcal{M}_\\nu)$ via \\begin{equation} \\mathcal{M}_\\nu = - M_\\dir M_\\m^{-1} M_\\dir^T, \\label{see-saw} \\end{equation} where $\\mathcal{M}_\\nu$ is a $3\\times 3$ matrix of active neutrino masses, mixings, and (possible) CP-violating phases. The masses of sterile neutrinos are given by the eigenvalues of their Majorana mass matrix (with the corrections of the order $M_\\dir^2/M^2_\\m$). They are much heavier than the active neutrino masses as a consequence of~(\\ref{see-saw}). Numerous searches for sterile neutrinos in the mass range up to $\\sim 100\\gev$ had been performed (see the corresponding section in Particle Data Group~\\cite{PDG:11},\\footnote{\\url{http://pdglive.lbl.gov/Rsummary.brl?nodein=S077&inscript=Y}} see also~\\cite{Atre:09} and refs. therein). These searches provided upper bounds on the strength of interaction of these neutral leptons with the SM neutrinos of different flavours -- active-sterile neutrino \\emph{mixing angles} $\\vartheta_\\alpha^2 \\propto \\left|\\frac{M_{\\dir,\\alpha}}{M_\\m}\\right|^2$ for sterile neutrino with the mass $M_\\m$.\\footnote{Here and below we use the letter $\\vartheta$ for \\emph{active-sterile mixing angles} (defined by Eq.~(\\ref{eq:19}) below) while reserving $\\theta_{12},\\theta_{13}$ and $\\theta_{23}$ for the measured parameters of the active neutrinos matrix $\\mathcal{M}_\\nu$. These quantities $\\vartheta_\\alpha$ are often denoted $|V_{4\\alpha}|^2$ or $|U_{x\\alpha}|^2$ in the experimental papers, to which we refer. Here and below the Greek letters $\\alpha,\\beta$ are flavour index $e,\\mu,\\tau$ and $i,j=1,2,3$ denote active neutrino mass eigenstates.} These bounds then can be interpreted as \\emph{lower bounds} on the lifetime of sterile neutrinos $\\tau_s$ via \\begin{equation} \\label{eq:lifetime_expression} \\tau_s^{-1} = \\frac{G_F^2 M_\\m^5}{96\\pi^3}\\sum_X \\vartheta_\\alpha^2 B_{X}^{(\\alpha)}, \\end{equation} where the sum runs over various particles to which sterile neutrino can decay, depending on their mass ($\\nu$, $e^\\pm$,$\\mu^\\pm$,$\\tau^\\pm$, $\\pi$, $K$, heavier mesons and baryons) and dimensionless quantities $B_X^\\alpha$ depend { on the branching ratios (see Appendix \\ref{app:lifetime} for details). The lower bound on the lifetime $\\tau_s$ is usually dominated by the least constrained mixing angle, $\\vartheta_\\tau^2$ (as will be shown later). This bound can be made \\emph{stronger} if one assumes that the same particles are also responsible for the neutrino oscillations. The see-saw formula~(\\ref{see-saw}) limits (at least partially) possible values of ratios of the mixing angles $\\vartheta_\\alpha^2/\\vartheta_\\beta^2$. In the simplest case when only two sterile neutrinos are present (the minimal number, required to explain two observed neutrino mass differences) the ratios of mixing angles varies within a limited range, see e.g.~\\cite{Gorbunov:07a,Shaposhnikov:08a}. While this range can be several orders of magnitude large (owing to our ignorance of certain oscillation parameters, such as e.g. a CP-violating phase~\\cite{Gorbunov:07a,Shaposhnikov:08a}), the implied (lower) bounds on the lifetime become much stronger, essentially being determined by the \\emph{strongest}, rather than the weakest direct bound on $\\vartheta_\\alpha$.} When confronted with the upper bound from Big Bang Nucleosynthesis~\\cite{Dolgov:00a,Dolgov:00b}, they seem to close the window of parameters for sterile neutrinos with the mass lighter than about $150\\mev$~\\cite{Gorbunov:07a,Boyarsky:09a}. It was argued in~\\cite{Asaka:2011pb}, however, that in the case of normal hierarchy there can be a small allowed window of parameters of sterile neutrinos with the mass below the pion mass. In this paper we reanalyze restrictions on sterile neutrino lifetime in view of the recent results of the Daya Bay~\\cite{An:2012eh} and RENO~\\cite{Ahn:2012nd} collaborations, that measured a non-zero mixing angle $\\theta_{13}$ (see also~\\cite{Adamson:2011qu,Abe:2011sj}). We demonstrate that in the case when there are only two sterile neutrinos, responsible for the observed neutrino oscillations, the oscillation data allow for such a choice of the active-sterile Yukawa couplings that the mixing of sterile neutrinos with any given flavour can be strongly suppressed. This happens \\emph{only} for a non-zero values of $\\theta_{13}$, in the range consistent with the current measurements~\\cite{Schwetz:08a,Fogli:2011qn,An:2012eh,Ahn:2012nd}. We also confront our results with the recently reanalyzed bounds from primordial nucleosynthesis~\\cite{Ruchayskiy:2012si} and show that the window in the parameter space of sterile neutrinos with masses $10\\MeV \\lesssim M_s \\lesssim 140\\MeV$ discussed in previous works~\\cite{Asaka:2011pb} (see also~\\cite{Gorbunov:07a}) is closed. For larger masses the window remains open. The results of this paper partially overlap with~\\cite{Asaka:2011pb} (also~\\cite{Gorkavenko:2009vd}), and we compare in the corresponding places to the previous works. The paper is organized as follows: in Section~\\ref{sec:numsm} we briefly describe the type I see-saw model we use. We then investigate the relations between different mixing angles imposed by the see-saw mechanism and demonstrate that the mixing with any flavour $\\vartheta_\\alpha^2$ can become suppressed (Section~\\ref{sec:solution-see-saw}). Section~\\ref{sec:experiments} is devoted to the overview of the experiments, searching for sterile neutrinos with the masses below $2\\GeV$, and the way one should interpret their results to apply to the see-saw models that we study. Section~\\ref{sec:discussion} summarizes our revised bounds on mixing angles and translates them into the resulting constraints on sterile neutrino lifetime (Figs.~\\ref{fig:lifetime_result}). We conclude in Section~\\ref{sec:conclusion}, discussing implications of our results and confronting them with the bounds from primordial nucleosynthesis. ", "conclusions": "In this Section we summarize our results: the upper bound on the (combination of) mixing angles of sterile and active neutrinos in the see-saw models~(\\ref{eq:lifetime_expression}) in the range 10~MeV -- 2~GeV and the lower bound on sterile neutrino lifetime, obtained in combination of these bounds with constraints, coming from neutrino oscillation data. \\begin{figure}[htp] \\centering % \\subfloat [Normal hierarchy, mass range $10\\mev - 2\\gev$] {\\includegraphics[width=.5\\textwidth]{figures/lifetime_NH+MI+NC-largemass-revDaya}}% ~\\subfloat [Normal hierarchy, zoom at the mass range $10\\mev - 140\\mev$] % {\\includegraphics[width=.5\\textwidth]{figures/lifetime_NH+MI+NC-smallmass-revDaya}}\\\\ \\subfloat% [Inverted hierarchy, mass range $10\\mev- 2\\gev$] { \\includegraphics[width=.5\\textwidth]{figures/lifetime_IH+MI+NC-largemass-revDaya} }% \\subfloat [Inverted hierarchy, zoom at the mass range $10\\mev - 140\\mev$] { \\includegraphics[width=.5\\textwidth]{figures/lifetime_IH+MI+NC-smallmass-revDaya}% } \\caption{The resulting lower bounds on sterile neutrino lifetime $\\tau_s$ as a function of their mass, obtained by requiring that two Majorana sterile neutrinos are responsible for neutrino oscillations and their parameters do not contradict the negative results of direct experimental searches. In all figures the upper curve comes from using of the best fit neutrino oscillation parameters, the middle one -- from their variation within the $3\\sigma$ limits, and the lower one does not take into account neutrino oscillation data and puts all three mixing angles equal to their direct experimental bounds. The dashed line for NH corresponds to the best-fit values of PMNS parameters with $\\theta_{13} = 0$ and shows how much the bounds on the lifetime relax for non-zero value of $\\theta_{13}$ (see text, Section~\\protect\\ref{sec:discussion} for discussion). } \\label{fig:lifetime_result} \\end{figure} \\subsection{Bounds on the mixing angles of sterile neutrinos} \\label{sec:bounds-mixing} For the models~(\\ref{eq:nuMSM_Lagrangian}) (two Majorana sterile neutrinos, interacting through both charged and neutral interactions), the compilation of constraints on various combinations of active-sterile mixing angles ($\\vartheta_e^2$, $\\vartheta_\\mu^2$, $\\vartheta_e \\sqrt{\\sum c_\\alpha \\vartheta_\\alpha^2}$, $\\vartheta_\\mu \\sqrt{\\sum c_\\alpha \\vartheta_\\alpha^2}$) that we used in this work are plotted in Figs.~\\ref{fig:PS191-bounds} and~\\ref{fig:mixing_constraints}.\\footnote{Notice that in the published results of the PS191 experiment~\\cite{Bernardi:1987ek} bounds are given up to $M_\\m=400\\MeV$. We extend these bounds up to $450\\MeV$, using the PhD Thesis of J.-M. Levy \\cite{Levy}.} \\subsection{The lower bound on the lifetime of sterile neutrinos} \\label{sec:bounds-lifetime} The result of the Sections~\\ref{sec:normal-hierarchy}--\\ref{sec:inverted-hierarchy}, combined with these experimental bounds can be translated into the \\emph{lower} limits on the lifetime of sterile neutrinos. These results are presented in Figs.~\\vref{fig:lifetime_result}. Additionally, we plot the lifetime bounds for the best-fit values of the PMNS parameters yet with $\\theta_{13}=0$ (as used e.g. in~\\cite{Gorbunov:07a,Canetti:10a}). For normal hierarchy we see that our bounds with $\\theta_{13} \\neq 0$ are relaxed by as much as the order of magnitude at some masses, compared to $\\theta_{13}=0$ case. The difference for IH is not so pronounced. \\emph{Notice}, that the bounds of~\\cite{Gorbunov:07a,Canetti:10a,Asaka:2011pb} were different from what we show as dashed line in Fig.~\\ref{fig:lifetime_result} because of ignoring the neutral current contributions to the results of PS191 experiment (for details see discussion in Section~\\ref{sec:experiments} and Figs.~\\ref{fig:lifetime-interpretations-peaks},~\\ref{fig:lifetime-CC+CCNC-prev-studies})." }, "1112/1112.3543_arXiv.txt": { "abstract": "{The earliest stages of high-mass star formation are still poorly characterized. Densities, temperatures and kinematics are crucial parameters for simulations of high-mass star formation. It is also unknown whether the initial conditions vary with environment.} {We want to investigate the youngest massive gas clumps in the environment of extremely active star formation. } {We selected the IRDC\\,18454 complex, directly associated with the W43 Galactic mini-starburst, and observed it in the continuum emission between 70\\,$\\mu$m and 1.2\\,mm with Herschel, APEX and the 30\\,m telescope, and in spectral line emission of N$_2$H$^+$ and $^{13}$CO with the Nobeyama 45\\,m, the IRAM 30\\,m and the Plateau de Bute Interferometer.} {The multi-wavelength continuum study allows us to identify clumps that are infrared dark even at 70\\,$\\mu$m and hence the best candidates to be genuine high-mass starless gas clumps. The spectral energy distributions reveal elevated temperatures and luminosities compared to more quiescent environments. Furthermore, we identify a temperature gradient from the W43 mini-starburst toward the starless clumps. We discuss whether the radiation impact of the nearby mini-starburst changes the fragmentation properties of the gas clumps and by that maybe favors more high-mass star formation in such an environment. The spectral line data reveal two different velocity components of the gas at 100 and 50\\,km\\,s$^{-1}$. While chance projection is a possibility to explain these components, the projected associations of the emission sources as well as the prominent location at the Galactic bar -- spiral arm interface also allow the possibility that these two components may be spatially associated and even interacting.} {High-mass starless gas clumps can exist in the close environment of very active star formation without being destroyed. The impact of the active star formation sites may even allow for more high-mass stars to form in these 2nd generation gas clumps. This particular region near the Galactic bar -- spiral arm interface has a broad distribution of gas velocities, and cloud interactions may be possible.} ", "introduction": "\\label{intro} The initial conditions required to form massive stars is one of the major topics in high-mass star formation research today. The advent of new observational capabilities often opened the window to earlier and earlier evolutionary stages in that field. While the IRAS satellite combined with cm continuum surveys of the galactic plane revealed the populations of ultracompact H{\\sc ii} regions and high-mass protostellar objects (HMPOs, e.g., \\citealt{wc89,wc1989b,kurtz1994,molinari1996,sridha}), the advent of the mid-infrared satellites ISO, MSX and Spitzer allowed to access even earlier evolutionary stages, namely the infrared dark clouds (IRDCs, e.g., \\citealt{perault1996,egan1998,carey2000,sridharan2005,simon2006,peretto2009}). However, it turned out that most of the so far studied IRDCs host weak 24\\,$\\mu$m emission sources and drive molecular outflows, both strong indicators for star formation activity (e.g., \\citealt{beuther2005d,beuther2007a,beuther2007g,rathborne2006,motte2007,cyganowski2008}). Therefore, we are still lacking a thorough understanding of the initial conditions for high-mass star formation prior to any star formation activity. The advent of the Herschel far-infrared satellite \\citep{A&ASpecialIssue-HERSCHEL} now offers the unique opportunity to identify targets for these initial conditions and to study their physical and chemical properties in detail. Major characteristics of these earliest evolutionary stages are, that the massive gas cores with masses in excess of several 100\\,M$_{\\odot}$ observed at (sub)mm wavelengths are still dark at 70\\,$\\mu$m, hence they have not formed any protostellar objects to Herschel sensitivity limits yet. The Herschel guaranteed time key project EPOS (Early Stages of Star Formation, P.I.~O.~Krause) is a dedicated observation campaign targeting 45 IRDCs that are promising candidates to search for and to study these initial conditions. Early results from the first data revealed already exciting results, e.g., we identified candidate high-mass starless cores (HMSCs) as well as very weak 70\\,$\\mu$m sources within IRDCs that may be the very first manifestation of high-mass star formation \\citep{beuther2010b,henning2010,linz2010}. \\begin{figure*}[htb] \\includegraphics[width=0.99\\textwidth]{17850f1.jpg} \\caption{Compilation of the continuum data from 70\\,$\\mu$m to 1.2\\,mm wavelength as labeled in each panel. The top-row presents the data on a smaller spatial scale as marked in the bottom-left panel. The color-scale is chosen in each image individually to highlight the most important features. Contour levels of the 870\\,$\\mu$m data start at the 3$\\sigma$ levels of 0.27\\,mJy\\,beam$^{-1}$ and continue in in 3$\\sigma$ steps to 2.7\\,Jy\\,beam$^{-1}$, from where they continue in 2.7\\,Jy\\,beam$^{-1}$ steps. The 1.2\\,mm data are contoured in 3$\\sigma$ levels of 30\\,mJy\\,beam$^{-1}$. The numbers in the top-left panel label the sub-cores, the white small box there marks the region shown in Fig.~\\ref{18454_spitzer}, and a scale-bar for a distance of 5.5\\,kpc is shown as well. The crosses in the bottom-left panel mark the positions of a double-component radio recombination line H{\\sc ii} region from \\citet{anderson2011}, the W43-MM1 position from \\citet{motte2003} as well as the approximate center of the Wolf-Rayet/OB cluster \\citep{blum1999,motte2003}.} \\label{spire_pacs} \\end{figure*} \\paragraph{The IRDC\\,18454 close to W43:} One particularly interesting region within the sample are the IRDCs associated with the IRAS source IRAS\\,18454-0158. Henceforth, we will call the region IRDC\\,18454. While the IRAS source was first studied in the framework of the HMPO survey by \\citet{sridha} and \\citet{beuther2002a}, it was soon recognized that several of the sub-sources are indeed infrared dark \\citep{sridharan2005}. Another curiosity of that region is that it harbors two distinctly different velocity components, one around 50 and one around 100\\,km\\,s$^{-1}$ \\citep{sridharan2005,beuther2007g}. It is also interesting to note that only the 100\\,km\\,s$^{-1}$ component shows SiO emission, indicative of a molecular outflow, whereas we do not detect any SiO from the 50\\,km\\,s$^{-1}$ component \\citep{beuther2007g}. Furthermore, the measured H$^{13}$CO$^+$(1--0) line-width from the 100\\,km\\,s$^{-1}$ component is also broader than that at 50\\,km\\,s$^{-1}$ (2.7 versus 1.7\\,km\\,s$^{-1}$, \\citealt{beuther2007g}), also indicating that the 50\\,km\\,s$^{-1}$ component is in a less turbulent state. The kinematic distances derived by \\citet{beuther2007g} for the 50 and 100\\,km\\,s$^{-1}$ components, using the \\citet{brand1993} rotation curve were 3.5 and 6.4\\,kpc, respectively. Applying the new rotation curve by \\citet{reid2009}, we now get kinematic distances of 3.3 and 5.5\\,kpc for both components, respectively. The 5.5\\,kpc distance corresponds well to the distance derived for the large H{\\sc ii} region W43 by \\citet{wilson1970}. A similar distance of $\\sim 6$\\,kpc was also recently derived for the nearby red supergiant cluster RSGC3 \\citep{negueruela2011}). We will discuss later whether these are really two different components just chance-projected together on the plane of the sky, or whether it may be interacting cloud components at similar distances. In the direct vicinity of that region we find the large Galactic mini-starburst region W43 (Fig.~\\ref{spire_pacs}), which was already studied in detail in the past with a broad wavelength coverage (e.g., \\citealt{smith1978,lester1985,blum1999,motte2003,nguyen2011}) including also early Herschel results \\citep{bally2010,elia2010}. The IRDC\\,18454 is located at the northeastern edge of the Z-shaped filament first discussed in \\citet{motte2003} (see also Fig.~\\ref{spire_pacs}). This region is very luminous with $L\\sim 3\\times 10^6$\\,L$_{\\odot}$, and it contains sources at different evolutionary stages from a Wolf-Rayet cluster to active young star-forming cores (e.g., \\citealt{blum1999,motte2003}). \\citet{nguyen2011} recently discussed several broad velocity components in that regions and suggested that this whole complex might have formed via converging gas flows due to its special location at the interface of the Galactic bar with the inner Scutum spiral arm \\citep{benjamin2005,lopez2007,rodriguez2008}. Different velocity components were also identified in atomic HI emission and absorption \\citep{liszt1993}. However, our target region IRDC\\,18454 which hosts supposedly the earliest evolutionary stages, has not been subject to a detailed investigation yet. Among others, we like to address the following questions: Do we identify bona-fide high-mass starless cores at all? If yes, what are their physical properties, e.g., their temperatures and turbulent line-widths? Are the two velocity components spatially associated or are they rather chance alignments along the line of sight? Does the mini-starburst W43 impose any impact on the earliest evolutionary stages or are the young cores evolving independently of that? Do we find any signature of triggering? ", "conclusions": "\\label{discussion} \\subsection{Luminous starless clumps} \\label{luminous} The measured high luminosities from the starless clump candidates (Fig.\\ref{seds}), in particular clumps 1, 3 and 8a are intriguing since -- down to our detection limits -- they do not have an internal radiation sources. Also other sources like clump 2 are interesting because this region is still dark at 24\\,$\\mu$m and only becomes an emission source from 70\\,$\\mu$m onwards. Hence, it has to be at a very young evolutionary stage as well. This implies that the bolometric luminosities of these regions must be largely caused by external radiation sources of which the prime candidates are the diverse sources comprising the neighboring mini-starburst W43 (see Introduction or, e.g., \\citealt{motte2003,nguyen2011}). Do we find differences between IRDCs in the vicinity of very active massive star-forming regions and more isolated IRDCs? For comparison, we select the IRDC\\,18223 where luminosities and temperatures were measured in a similar fashion via SED fits of Herschel data \\citep{beuther2010b}. The derived temperatures for the clumps in IRDC\\,18454 are slightly elevated compared to IRDC\\,18223 (between 19 and 20\\,K for IRDC\\,18454, and between 16 and 18\\,K in IRDC\\,18223), and the measured bolometric luminosities in IRDC\\,18454 are factors between 4 and 10 larger than found in IRDC\\,18223. While the observed aperture size and spatial scale directly translates into different luminosities (Fig.~\\ref{seds}), the smaller aperture of $19.2''$ (corresponding to $\\sim$0.5\\,pc) used for IRDC\\,18454 is even smaller than that used for IRDC\\,18223 ($\\sim$0.64\\,pc). Therefore, spatial scale arguments cannot account for that difference. Another important parameter for the total bolometric luminosity is the available gas mass. While clumps 1 and 3 are factors of a few more massive than IRDC\\,18223-s1 and IRDC\\,18223-s2, clump 8 presented here and IRDC\\,18223-s2 have approximately the same mass (140 and 170\\,M$_{\\odot}$, respectively). Nevertheless, the luminosity of clump 8 is about a factor 4 larger than that of IRDC\\,18223-s2. It is no surprise that the temperature differences between IRDC\\,18223 and IRDC\\,18454 are less strong than the luminosity differences because the luminosity has a strong exponential dependence on the temperature (Stefan-Boltzmann law $L \\propto T^4$). While all these parameters have to be taken a bit cautiously since the associated errors are large (e.g., for the SEDs the uncertain background subtraction may contribute significantly), in principle the errors should be comparable for the different regions because we applied the same techniques. Therefore, it is tempting to speculate what may cause the measured luminosity and temperature differences between infrared dark clumps in the vicinity of the active star-forming region W43 compared to the more quiescent region IRDC\\,18223. It could be possible that the neighboring mini-starburst injects considerable energy even into the starless clumps that the total bolometric luminosities are increased above their expected more quiescent values. To first order, with such a luminosity increase one also expects a temperature increase within the gas clumps. Although the measured temperature difference between the regions is less strong than the luminosity increase, nevertheless, we see elevated temperatures in IRDC\\,18454 compared to IRDC\\,18223. Considering the large apertures used for the SEDs, the derived average temperatures may still be dominated by the average temperature of the general ISM that is in the same range (e.g., \\citealt{reach1995}). We can also quantitatively estimate the influence of the total bolometric luminosity of the Wolf-Rayet/OB cluster ($\\sim 3.5\\times 10^6$\\,L$_{\\odot}$) on the gas and dust environment at the typical projected separations from the target IRDCs (Fig.~\\ref{spire_pacs}). Assuming that all radiation of the central cluster is reprocessed as far-infrared radiation, and using an approximate projected separation of the infrared dark clumps of 11\\,pc, as well as Gaussian source FWHM of $36.6''$ or $19.2''$, respectively (Sec.~\\ref{sedsection} and Fig.~\\ref{seds}), the received luminosity at the location of the IRDCs are on the order of 2500 and 700\\,L$_{\\odot}$, respectively. Comparing these values with the measured luminosities in the same given apertures (Fig.~\\ref{seds}), elevated luminosities caused by the nearby luminous cluster appear clearly feasible. Furthermore, one can estimate the approximate required interstallar radiation field to account for the observed luminosity. A clump with a surface corresponding to a radius $19.2''/2$ has to intercept an interstellar radiation field of $\\sim$306 (in units of the Habing field $G_0=1.6\\times 10^{-3}$\\,erg\\,s$^{-1}$cm$^{-2}$, \\citealt{habing1968}) to exhibit a luminoisty of 1000\\,L$_{\\odot}$. In comparison, if one just accounts for geometric dillution, the $\\sim 3.5\\times 10^6$\\,L$_{\\odot}$ of the Wolf-Rayet/OB cluster correspond at a distance of 11\\,pc to $\\sim$588 times the interstellar radiation field. Subtracting additional scattering and blockade by interfering gas clumps, these two values agree well with each other. Based on kinematic arguments, \\citet{motte2003} also discussed the influence of the luminous cluster on the environment, and they similarly infer that there should be considerable influence from that first generation of high-mass star formation on its environment. In addition to the individual SED fits presented in section \\ref{sedsection} and Figure \\ref{seds}, one can also smooth all maps between 70 and 870\\,$\\mu$m to the same spatial resolution of the 500\\,$\\mu$m data and fit the SEDs pixel by pixel (see also \\citealt{stutz2010} for details of the fitting). This large-scale fitting approach allows us to infer the dust temperature structure and gradients across the whole field. Figure \\ref{t_herschel} presents the results. As expected, we identify a clear temperature gradient from more than 35\\,K toward the W43 main complex down to low values below 20\\,K in the east of the region. In particular, we see that many of the submm continuum structures correlate well with the lowest temperatures. This is true for the dark clumps of this analysis as well as for the filament going approximately in northsouth direction at the eastern edge of the map. The large-scale east-west temperature gradient throughout the region is consistent with significant influence of the W43 complex on the here discussed infrared dark high-mass starless clump candidates. \\begin{figure}[htb] \\includegraphics[width=0.49\\textwidth]{17850f11.jpg} \\caption{The color-scale shows a temperature map in K derived for the whole region from the combined Herschel and ATLASGAL data between 70 and 870\\,$\\mu$m, all smoothed to the spatial resolution of the 500\\,$\\mu$m data. Contour levels of the overlaid 870\\,$\\mu$m data start at the 3$\\sigma$ levels of 0.27\\,mJy\\,beam$^{-1}$ and continue in in 3$\\sigma$ steps to 2.7\\,Jy\\,beam$^{-1}$, from where they continue in 2.7\\,Jy\\,beam$^{-1}$ steps. Our target clumps are marked.} \\label{t_herschel} \\end{figure} Since the Jeans length and Jeans mass as indicators of the fragmentation properties of star-forming gas clumps are proportional to the temperature ($\\lambda_J \\propto T^{1/2}$ and $M_J \\propto M^{3/2}$), increases in luminosity and temperature of young gas clumps could change their fragmentation properties to preferentially form larger and more massive fragments. Although still speculative, such a scenario could imply that nearby active star formation does not only have a destructive character destroying pristine gas clumps, but that the energy input of close-by active regions could even elevate the ability to form massive stars in a triggered fashion in the immediate environment of the earlier generations of stars. While the energy input from a first generation of low-mass stars appears to be too little to halt fragmentation \\citep{longmore2011}, the influence of the nearby luminous high-mass star-forming region may be more important. \\subsection{Colliding clouds or chance alignment?} \\label{converge} As indicated in sections \\ref{kinematic1} \\& \\ref{kinematic2}, the two different velocity components at 100 and 50\\,km\\,s$^{-1}$ may either be chance projections along the line of sight, or they could be associated and potentially interacting cloud complexes. Based mainly on large-scale $^{13}$CO(1--0) data from the Galactic Ring survey (GRS, \\citealt{jackson2006}), \\citet{nguyen2011} suggest that the velocities associated directly with W43 range between 60 and 120\\,km\\,s$^{-1}$, whereas lower velocity components are unlikely part of the W43 complex. \\begin{figure}[htb] \\includegraphics[width=0.49\\textwidth]{17850f12.jpg} \\caption{IRAM\\,30m $^{13}$CO(2--1) observations toward a larger part of W43 and IRDC\\,18454. The color-scale shows the $^{13}$CO emission around 53\\,km\\,s$^{-1}$ whereas the contours present gas at 93\\,km\\,s$^{-1}$. The stars mark the positions of our clumps 1 to 8, a scale-bar is shown at the bottom-left.} \\label{13co} \\end{figure} In contrast to this, the spatial structure of the two velocity components as presented in Figs.~\\ref{n2hplus} and \\ref{18454-1_100_50} indicates that the two components may also be spatially associated and interacting. Also the fact that the mm fluxes of clump 1 and 2 are so similar could be taken as circumstantial support for this idea. To further investigate whether we find additional signatures of associated clouds in that region, we investigated the $^{13}$CO(2--1) data of a recent IRAM 30\\,m large program to study the entire W43 complex (PI F.~Motte). The whole project, the data and more detailed analysis will be presented in forthcoming papers by Nguyen-Luong et al.~and Carlhoff et al., here we only show a small sub-set where two selected velocity components are compared. Figure \\ref{13co} presents an overlay of the $^{13}$CO(2--1) emission around 50 and 100\\,km\\,s$^{-1}$ over a larger area but also encompassing our region of interest. The two velocity components appear in projection as a layered structure where one velocity component mainly emits at locations that are devoid of emission from the other velocity component. Unfortunately, this is also no unambiguous proof for associated or interacting clouds. However, the spatial alignment of different velocity components for molecules with varying critical densities over a broad range of spatial scales is suggestive for the idea of connected and interacting structures. Additional interesting information about different velocity components in this region is found for other tracers and wavelengths as well. For example, CH$_3$OH Class {\\sc ii} maser were found in this region at both velocities (e.g., \\citealt{menten1991}). Furthermore, the recent H{\\sc ii} region radio recombination line survey of \\citet{anderson2011} detected toward 21 out of 23 targeted position within 0.5 degrees from W43 two velocity components as well (one of the positions is marked in Fig.~\\ref{spire_pacs}). While projection effects of molecular and atomic gas along the line of sight through the Galaxy are well known, having such a large number of multiple-velocity recombination line spectra is more surprising since one does not expect that many projected H{\\sc ii} regions along each line of sight. Hence interacting clouds and H{\\sc ii} regions are possible. Furthermore, it is interesting to note that the largest number of H{\\sc ii} regions in our Galaxy are also found toward Galactic longitudes around $l\\approx \\pm30$\\,degrees \\citep{wilson1970,lockman1979,anderson2011}, hence again including the W43/IRDC\\,18454 complex. All this evidence suggests special environmental physical properties in this region that may affect the star formation processes. How could such interacting velocity components be explained? The whole W43-IRDC\\,18454 complex is located at approximately a Galactic longitude of 30.85 degrees where the Galactic bar ends and the Scutum spiral arm starts (for a recent description of the different Galactic components, see e.g., \\citealt{churchwell2009,nguyen2011}). Hence, one could fathom interaction between Galactic bar and spiral arm structures, e.g., streaming motions between both Galactic components. Following the Galaxy modeling by \\citet{bissantz2003}, the inner spiral arms connecting to the bar and the bar should co-rotate. In this framework, the 100\\,km\\,s$^{-1}$ component should be associated with the beginning of the Scutum arm (see also \\citealt{vallee2008}). Spiral arm velocities for the Sagittarius and Perseus arm are, according to \\citet{vallee2008}, around 70 and 25\\,km\\,s$^{-1}$, respectively. However, there is no clear association with a 50\\,km\\,s$^{-1}$ gas component. Inspecting the Galactic plane CO survey by \\citet{dame2001}, at a Galactic longitude of $\\sim 30.85$ degrees one also find strong emission at 100\\,km\\,s$^{-1}$ and only less prominent features at 50\\,km\\,s$^{-1}$. Therefore, from the spiral arm structure of our Galaxy, it is not obvious to what distance one can associate the 50\\,km\\,s$^{-1}$ component. Furthermore, \\citet{nguyen2011} associate even 60\\,km\\,s$^{-1}$ gas with W43 and not so much with the Sagittarius arm as one would expect from the Galaxy models (e.g., \\citealt{vallee2008}). Although we cannot firmly establish whether the 100 and 50\\,km\\,s$^{-1}$ gas components are spatially associated or just chance alignments, the close projected association of them as well as the extremely busy and active nature of that part of the Galaxy at the interface of the Galactic bar with the inner spiral arm, where streaming motions between both components are likely to occur, makes spatial association and interaction of different cloud components here a fair possibility. See also \\citet{anderson2011} for a similar discussion of the origin of the ionized gas. One way to distinguish the different explanation is to derive accurate distances to the different components. And the best way to do that today is parallax measurements of maser sources (e.g., \\citealt{reid2009}). The VLBI now has a very large project to do that for many maser sources throughout the Galactic plane (PI M.~Reid), and the masers toward the W43 complex are part of that program (K.~Menten, priv.~comm.). Therefore, we expect to get accurate distances for the different velocity components in the coming years. That should then allow us to answer the question whether these different velocity components are indeed spatially associated or whether they are just chance projections along the lines of sight." }, "1112/1112.4363_arXiv.txt": { "abstract": "Using the gamma-ray bursts simultaneously detected by \\textit{Swift}/BAT and \\textit{Fermi}/GBM we performed a joint spectral and temporal analysis of the prompt emission data and confirm the rough correlation between the BAT-band photon index $\\Gamma^{\\rm BAT}$ and the peak spectral energy $E_{\\rm peak}$. With the redshift known sub-sample, we derived the isotropic gamma-ray energy $E_{\\rm \\gamma,iso}$ and also confirm the $E_{\\rm \\gamma,iso}-E_{\\rm peak,rest}$ relation, with a larger scatter than the Amati sample but consistent with GBM team analyses. We also compare the $T_{90}$ values derived in the GBM band with those derived in the BAT band and find that for long GRBs the BAT $T_{90}$ is usually longer than the GBM $T_{90}$, while for short GRBs the trend reverses. This is consistent with the soft/hard nature of long/short GRBs and suggests the importance of an energy-dependent temporal analysis of GRBs. ", "introduction": "Since the launch of the \\textit{Fermi} satellite in 2008, a significant number of gamma-ray bursts (GRBs) have been observed by both the \\fermi Gamma-ray Burst Monitor (GBM, 8 keV-40 MeV) and \\swift \\cite{gehrels04}. Since the energy band of \\textit{Swift} BAT is narrow (15-150 keV) and the peak of the spectrum is often outside the observed energy window \\cite{sakamoto09}, the joint BAT-GBM sample is valuable for \\textit{Swift} science by providing important information about the prompt emission, especially the characteristic peak energy of the $\\nu F_{\\nu}$ spectrum, $E_{\\rm peak}$. Using the \\textit{Fermi}/GBM-\\textit{Swift}/BAT joint sample we performed a two-part analysis on the prompt emission data. First, a joint spectral analysis where we fit the data from both detectors and investigated several empirical correlations. Second, a temporal analysis where we rigorously derived $T_{90}$ in the 8-1000 keV band and compared it to the published BAT values. A variety of empirical relations have been discussed in the literature \\citep{amati02,amati03,sakamoto04,sakamoto06,ghirlanda04,liangzhang05,yonetoku04}, many of which utilize $E_{\\rm peak}$. Since $E_{\\rm peak}$ cannot be well measured for most \\swift GRBs, efforts have previously been made to look for indicators of $E_{\\rm peak}$ based on the available observed quantities. In particular, the effective photon power law index in the BAT band, $\\Gamma^{\\rm BAT}$, has been found to be broadly correlated with $E_{\\rm peak}$. Two versions of this relation are found in the literature: one presented in Sakamoto et al. (2009) using simulations of BAT spectra and $E_{\\rm peak}$ measurements from broad-band detectors (e.g. Konus-Wind), and another presented in Zhang et al. (2007b) with $E_{\\rm peak}$ estimates based on the hardness ratio in the BAT band itself (Zhang et al. 2007a; see also Cui et al. 2005). In \\S3.2 we re-calibrate this correlation with the GBM-BAT joint sample. The $E_{\\rm \\gamma,iso} - E_{\\rm peak,rest}$ relation \\citep[or Amati relation;][]{amati02,amati03} is one of the most widely studied, and hotly debated, empirical correlations that may connect to GRB physics \\citep[e.g.][]{band05,nakar05,kocevski11}. In \\S3.3 we test the Amati relation with the redshift-known sub-sample of the GBM-BAT joint sample. Finally, traditionally long and short GRBs were defined in the BATSE band with a separation of about 2 seconds in the observed frame (Kouveliotou et al. 1993). When applied to \\swift GRBs and combined with multi-wavelength afterglow data, this definition creates confusion when used for GRB classification (e.g. Zhang et al. 2007b, 2009). Since the energy band of GBM is similar to that of BATSE, a comparison between the measured $T_{90}$ durations in the two detectors is of great interest and is performed in \\S4. ", "conclusions": "Using the data available from \\fermi and \\swift we performed an analysis of the jointly detected GRB sample. The spectral analysis indicates that fitting the joint spectra shows consistency with values reported in the literature based solely on GBM data \\cite{nava11,bissaldi11}, and is able to reproduce a variety of empirical relations presented in the literature, including the $\\Gamma^{\\rm BAT}-E_{\\rm peak}$ and Amati relations. The joint fits do not, in general, give substantially different values for the spectral parameters from the GBM-only analyses. The updated sample of peak energies shows agreement with the relation between the PL slope of the BAT spectrum and $E_{\\rm peak}$ derived previously by Sakamoto et al. (2009) and Zhang et al. (2007b) and indicates the relation's ability to estimate the value of the observed $E_{\\rm peak}$ from a fit to the rather narrow \\swift band. The relation shows a fair amount of scatter, however, and we caution against using it for robust measurements of $E_{\\rm peak}$, especially when compared to spectra from \\fermi or other missions that often contain the spectral peak within the energy range of the detectors. It may be suited, for example, to give a rough estimate of the values of $E_{\\rm peak}$ for \\swift bursts that have no additional observations. The redshift known subsample of bursts reproduces the general relationship in the $E_{\\rm iso}-E_{\\rm peak,rest}$ plane and further confirms the conclusions of Gruber et al. (2011) that show generally harder spectra and a larger scatter around the relation than previously believed. Li (2007) proposed the argument that the $E_{\\rm iso}-E_{\\rm peak,rest}$ relation might evolve with redshift. We believe that the sample of bursts with known redshift, especially the subsample of joint bursts, is still too small to test such a claim. The necessary binning to test the hypothesis would thin out the sample sufficiently to make firm conclusions tenuous at best. Finally, the interesting discrepancy between the $T_{90}$ derived in the GBM and BAT detectors confirms the soft/hard nature of long/short GRBs, respectively, and suggests the need for a detailed energy-dependent temporal analysis of GRBs." }, "1112/1112.1400.txt": { "abstract": "It was recently shown that our peculiar velocity $\\boldsymbol{\\beta}$ with respect to the CMB induces mixing among multipoles and off-diagonal correlations at all scales which can be used as a measurement of $\\boldsymbol{\\beta}$, which is independent of the standard measurement using the CMB temperature dipole. The proposed techniques rely however on a perturbative expansion which breaks down for $\\,\\ell \\gtrsim 1/\\beta \\approx 800$. Here we propose a technique which consists of deboosting the CMB temperature in the time-ordered data and show that it extends the validity of the perturbation analysis multipoles up to $\\,\\ell \\sim 10000$. We also obtain accurate fitting functions for the mixing between multipoles valid in a full non-linear treatment. Finally we forecast the achievable precision with which these correlations can be measured in a number of current and future CMB missions. We show that Planck could measure the velocity with a precision of around $60$ km/s, ACTPol in 4 years around $40$ km/s, while proposed future experiments could further shrink this error bar by over a factor of around 2. ", "introduction": "\\label{sec:intro} The dipole of the CMB is measured to be much larger than the other multipoles and this is usually attributed to a Doppler effect due to our peculiar velocity $\\boldsymbol{\\beta}$ with respect to the CMB rest frame. Under this assumption we can infer, by combining the measured temperatures of WMAP dipole~\\cite{Hinshaw:2008kr} with the COBE monopole~\\cite{Lineweaver:1996xa,Mather:1998gm}, its direction to be $\\, l = 263.99^\\circ \u00b1 0.14^\\circ$, $b= 48.26^\\circ \u00b1 0.03^\\circ\\,$ in galactic coordinates, and its modulus to be $\\beta \\equiv |\\boldsymbol{\\beta}| = (1.231 \\pm 0.003) \\times 10^{-3}$. These very precise numbers rely however entirely on the above assumption, but generically the CMB dipole is not necessarily due only to a relative velocity. One way to test this assumption was proposed in~\\cite{Burles:2006xf} where it was shown that our peculiar velocity could also be measured using the asymmetry in the location of the peaks of the power spectrum between forward and backward hemispheres, achieving a possible detection of $\\beta$ at $2 - 3 \\sigma$ for resolution $\\ell = 1000 - 1500$. Recently it has also been pointed out~\\cite{Challinor:2002zh,Kosowsky:2010jm,Amendola:2010ty} that {\\it all} the CMB multipoles $a_{\\ell m}$ have a correction due to our local peculiar velocity because the primordial anisotropies are distorted by the Doppler and aberration effects. This shows up as a correlation between different multipoles $\\ell$ and could be used as an alternative way of measuring our velocity, as an independent consistency check. This fact might offer also a way to test the isotropy of the Universe on very large scales: it is known in fact that the CMB sky and other observations seem to exhibit some anomalies on the very large scales and this effect offers an observational handle which could either confirm and make more robust the standard assumptions or perhaps point to global anisotropies of the Universe. The analysis of these CMB correlations has been performed~\\cite{Challinor:2002zh,Kosowsky:2010jm,Pereira:2010dn,Amendola:2010ty} relying upon a Taylor expansion in orders of $\\beta$ of boost effects (Doppler and aberration) on the multipole coefficients $a^{X}_{\\ell m}$, where $X$ stands for $T$ (temperature) or $E, B$ (the $E$ and $B$ modes of polarization). It turns out, however, that each of the series coefficients brings together ever higher powers of $\\,\\ell$, effectively transforming the expansion into one of powers of the product $\\,\\beta\\,\\ell$. Since $\\beta = 1.231 \\times 10^{-3}$ this means that for $\\ell \\gtrsim 800$ the series can no longer be relied upon. We shall therefore distinguish between two regimes of angular scales where the the above correlations show up. In the first regime on large scales, $\\ell \\ll 1/\\beta \\simeq 800$, we can use a perturbative approach and therefore precisely predict that there are correlations of ${\\cal O} (\\beta \\ell)$ between neighbor $\\ell$'s and look for a signal in the CMB in the form of correlators $\\,a_{\\ell m}^* a_{\\ell+1 m'}$. Moreover it was shown in~\\cite{Amendola:2010ty} that one can non-trivially and very efficiently define three different estimators (for $m'-m=0,\\pm 1$) in order to measure directly the three cartesian components of $\\boldsymbol{\\beta}$, without having to scan all the possible directions in the sky (i.e., without the need to compute and minimize a numerical $\\chi^2$ for the correlations for each direction in a grid of $\\{\\theta,\\phi\\}$ coordinates). The multipoles in the complementary regime ($\\ell \\gtrsim 1/\\beta$) cannot be so easily treated but nevertheless also carry information about our peculiar velocity in the form of similar correlations, and it turns out that such information is needed in order to reach a signal-to-noise ratio larger than 1. In this regime the deviation angle $\\theta-\\theta'$ due to aberration is larger than the angular scale of interest $1/\\ell$, so things are much more complicated. There are in fact correlations which are large (${\\cal O} (1)$!) and nonzero also for distant $\\ell$'s given by a transformation $ a^{[A]}_{\\ell m} \\;=\\; \\sum_{\\ell'=0}^\\infty K_{\\ell' \\, \\ell\\, m}\\, a^{[P]}_{\\ell' m}\\,,$ between the aberrated $[A]$ frame and the primordial $[P]$ correlations. One in principle would have to compute the matrix elements $\\,K_{\\ell' \\ell m}\\,$, sometimes referred to as the \\emph{aberration kernel}, which are integrals of spherical harmonics with different arguments, plug the matrix elements into all the possible two-point correlation functions and compare with the data. In fact, as we will show in Section~\\ref{sec:nonlinear} the correlations in non-neighboring multipoles ($a_{\\ell m}^* a_{\\ell\\pm n\\, m'}$, $n>1$) also carry a measurable signal, which should be taken into account to measure the velocity with reasonable precision. While in principle straightforward this procedure has some disadvantages: \\emph{(i)} this can be a heavy and delicate numerical task, because of the highly oscillating integrands and also because of the huge number of correlators that one would have to consider (future experiments propose to measure all multipoles as far as $\\ell \\sim 3000+$ for both temperature and polarization); \\emph{(ii)} it is not obvious to understand in this case whether the three simple estimators can be written explicitly for the three cartesian components of the velocity for any $n$ in $a_{\\ell m}^* a_{\\ell\\pm n\\, m'}$, so perhaps the procedure would have to be carried out scanning the sky in all possible directions which would probably make this approach even more expensive in terms of computational time. Even though we explore this approach further and propose a solution (in Section~\\ref{sec:nonlinear}) to the first of the two challenges listed above, the main point of this paper is to suggest a very simple trick which may be used to overcome all of these problems and measure directly the three components of $\\boldsymbol{\\beta}$ from a map in a much faster and straightforward way. Using this result we are able to predict with which precision we can measure $\\boldsymbol{\\beta}$ for several future experiments. This trick relies on the fact that we may already {\\it assume} to know the direction and modulus of the velocity $\\boldsymbol{\\beta}$ from the CMB dipole and that we can the use the other multipoles with their correlations as an independent consistency check, in order to confirm (or not) the assumed value of $\\boldsymbol{\\beta}$ up to some precision. The trick (described in more detail in Section~\\ref{sec:pre-deboost}) is as follows. Given the central assumed value of the velocity $\\,\\boldsymbol{\\beta}_{\\rm dip} = \\boldsymbol{\\beta}_{\\rm dip}^{\\rm fit} \\pm \\delta\\boldsymbol{\\beta}_{\\rm dip}$ indicated by the dipole, we may take the CMB map $T(\\theta,\\phi)$ and ``deboost'' it by a Lorentz transformation into the frame with opposite velocity $\\,-\\boldsymbol{\\beta}_{\\rm dip}^{\\rm fit}$. As discussed in~\\cite{Menzies:2004vr} such transformation should be carried in the CMB time-ordered data (TOD), i.e., before data treatment to extract the harmonic multipole coefficients \\alm\\ and thus before one constructs the CMB temperature maps. Thus we call this technique ``pre-deboosting'' the CMB data. In the new (pre-deboosted) frame the residual velocity $\\,\\boldsymbol{\\beta}_{\\rm res}$ compared to the CMB frame is expected to be given just by the residual error $\\,\\delta\\boldsymbol{\\beta}_{\\rm dip}\\,$ on the experimental determination of $\\,\\boldsymbol{\\beta}_{\\rm dip}$. As quoted above, such error is currently approximately $\\,|\\delta\\boldsymbol{\\beta}_{\\rm dip}| \\,\\simeq\\, 3\\,\\times\\,10^{-6}$. As a consequence the correlations due to Doppler and aberration in this new frame are expected to be of the order $\\,{\\cal O}(\\beta_{\\rm res} \\; \\ell)\\,$ for $\\,\\ell \\ll 1/(\\beta_{res})\\simeq 3\\times10^5\\,$ and so we could safely use the first order perturbative equations to compute the correlation functions $\\,a^*_{\\ell m} a_{\\ell+1 m'}\\,$ up to, say, $\\,\\ell \\simeq 10000\\,$ which is more than enough for all future experiments. %Now, in this new frame we should be able to check that the correlations functions are small and measured the residual velocity $\\beta_{\\rm res}$ with some error $\\delta\\beta_{\\rm res}$, %consistent with zero (if the subtraction has been made to very high precision) %just evaluating the linear estimators for the velocity as described in~\\cite{Amendola:2010ty}. Note that this allows us to use directly again the three efficient define estimators (for $m'-m=0,\\pm 1$) for measuring direction and modulus of the residual velocity. An unexpected but very interesting prospect is the case in which the measured velocity turns out to be different from the expected $\\beta_{\\rm res}$. This would be a clear indication that the CMB dipole is not completely (nor around $99\\%$, as sometimes stated) due to our peculiar velocity alone. Therefore, this would imply that there are other contributions to the CMB dipole, and it would be interesting to understand whether such correlations may distinguish even the nature of such contributions: adiabatic, isocurvature perturbations, dipolar lensing or other more exotic contributions. %contributions due to vector fields. The ability to measure exotic contributions to the dipole is of great interest to test Cosmology on very large scales, which could hide non-trivial phenomena, as suggested by some reported anomalies on the low-$\\ell$ CMB multipoles itself~\\cite{Copi:2010na, Bennett:2010jb}. For instance, it could provide valuable information about some proposed tilted cosmological models in which the dipole arises partly due to primordial superhorizon-scale isocurvature fluctuations~\\cite{Turner:1991dn,Grishchuk:1992wv,Langlois:1995ca,Erickcek:2008sm,Zibin:2008fe}, which could provide a possible explanation some recent controversial claims of a high galaxy cluster~\\cite{Kashlinsky:2008ut,Kashlinsky:2009dw,AtrioBarandela:2010wy} and galaxy~\\cite{Watkins:2008hf} bulk flow on large scales and it could be used as a test of non-standard cosmological models. In any case, if $\\beta_{\\rm res}\\neq 0$ we would also be able to measure the direction of our velocity with an error $\\delta\\theta=\\delta\\beta/\\beta_{\\rm res}$ (see~\\cite{Amendola:2010ty}). This paper is organized as follows. We start by discussing in Section~\\ref{sec:nonlinear} the full non-linear approach to estimate our velocity, and derive a very accurate fitting function for the oscillating integrals. We then discuss in Section~\\ref{sec:pre-deboost} the pre-deboost technique originally proposed by~\\cite{Menzies:2004vr}. In Section~\\ref{sec:applications} we forecast the sensitivity expected from present and future experiments. Finally, we draw our conclusions and summarize our results. ", "conclusions": "In this work we proposed a method to overcome the challenges posed to the measurement of our peculiar velocity $\\boldsymbol{\\beta}$ through its aberration and Doppler effect in the CMB at very small scales, due to the breakdown of the perturbative calculations when $\\ell \\gg 1/\\beta$. This technique consists of pre-deboosting the CMB (at the level of the TOD) and subsequently evaluate the estimators originally constructed in~\\cite{Amendola:2010ty}, which are based on a linear order Taylor expansion in $\\beta$. The expected small residual velocity after such deboost justifies the use of only the linear order terms, and greatly simplifies the analysis. In particular, the pre-deboost method validates analysis of the aberration effect constrained to the correlations between $\\{\\ell,\\,\\ell+1\\}$ only. Making use of the estimator of the velocity derived in~\\cite{Amendola:2010ty} for such correlations, we investigated the precision with which many of the current and future CMB experiments might be able to measure our velocity $v$. We find that Planck will put an error bar $\\,\\delta v \\simeq 60$ km/s, similar to what ACTPol could achieve in only 2 years with a survey covering $40\\%$ of the sky (4 years of observation would result in $\\,\\delta v \\simeq 40$ km/s). SPTPol, could also be competitive if it carried out a similar wide survey. Proposed future space experiments such as EPIC and Core could put a limit $\\,\\delta v < 30$ km/s. Even more precise experiments, able to measure the CMB signals (and get systematics under control) all the way to $\\ell \\simeq 5000$ could in principle achieve $\\,\\delta v < 10$ km/s. Since one expects from the CMB dipole that $v \\simeq 370$~km/s, with current (near future) experiments one would be able to detect a residual term of primordial origin which contributes to a fraction larger that $20\\%$ ($10\\%$) of the dipole. This is of course true unless the primordial effect itself also induces a dipolar kernel leading to correlations, with exactly the same coefficients as the peculiar velocity we analyze here, on the $\\ell>1$ multipoles. Apart from velocity, other possible components to the dipole and/or a dipolar kernel include intrinsic adiabatic, isocurvature and lensing effects, as well as more exotic effects, some of which we now briefly discuss. An example would be given by some fundamental vector field or perhaps magnetic fields coherent over the horizon scale, which could single out a preferred direction and therefore induce a dipole. Some inhomogeneous and anisotropic cosmological models are also expected to contribute to the dipole. For instance inhomogeneous gigaparsec-scale void models, embodied by the Lema\\^itre-Tolman-Bondi metric, are sometimes constructed to serve as a candidate for dark energy, and have recently drawn considerable attention from the literature (see, for instance,~\\cite{Marra:2011ct}). In these spherical symmetric models, the observer is sometimes assumed to be in the center for simplicity, but any off-center displacement induces a dipole contribution. This in fact happens to be the tightest constrain in this off-center distance, if one assumes the CMB dipole to be entirely due to this displacement and that any additional peculiar velocity is actually zero~\\cite{Quartin:2009xr}. Metrics which exhibit vorticity (e.g., Bianchi V and VII) also can contain a preferred direction and sense, and so also introduce a dipole~\\cite{Barrow1985}. On the other hand, homogenous anisotropic metrics without vorticity, such as Kantowsky-Sachs, Bianchi I and III, exhibit a preferred direction but no preferred sense (i.e., a preferred axis, but not an arrow), so they contribute to the CMB angular power spectrum at most at the quadrupole level~\\cite{Graham:2010hh,Koivisto:2010dr}. Finally, it is also possible that more generically large-scale vector perturbations in FLRW with unusually large amplitude may induce such dipoles, as well as a primordial power spectrum which explicitly depends on the direction of the Fourier mode $\\mathbf{k}$ instead of just $k=|\\mathbf{k}|$~\\cite{Hanson:2009gu}. %,BlancoPillado:2010uw %, such as non-trivial topologies, \\emph{et cetera}. It is an interesting question, which deserves future investigation, to check what kind of primordial perturbations or other effects such as the ones discussed above could induce similar correlations between different $\\ell$'s and if the produced correlations are proportional to the produced Doppler effect. Since it is natural to expect that in some cases there is no such proportionality, aberration can provide in principle an observational handle to distinguish between these different contributions. While the pre-deboost is probably the simplest way to proceeed, we have also proposed very precise fitting functions for the complicated integrals~\\eqref{eq:non-linear-coef}, valid also when $\\ell \\gg 1/\\beta $. %which allows not only to easily compute the correlation coefficients but to quickly their dependence on $\\ell$, $m$ and $\\beta$. Although another method was recently proposed in~\\cite{Chluba:2011zh} to compute this kernel elements, it relies on recursion relations between some of its elements and is not as simple to implement as just using the simple Bessel functions~\\eqref{eq:non-linear-fit-general} and~\\eqref{eq:non-linear-fit-n0}. %Still concerning the integrals~\\eqref{eq:non-linear-fit-general}, the very high precision and breadth of applicability of an arguably simple fitting function~\\eqref{eq:non-linear-fit-general} (which holds for any $n$ and either temperature or polarization) hints to the existence of an analytical solution involving these Bessel functions. Such solution could perhaps only be obtained in some special limit, such as for instance in the so-called flat-sky approximation, which should be a reasonable approximation for small-scales (although see~\\cite{hu:2000}). An analytical derivation of~\\eqref{eq:non-linear-fit-general} or similar equation, if it exists, could shed some light into the mathematics of this problem and into the limits of validity of the approximation needed to make such a derivation. Nevertheless, such analysis is beyond the scope of this work. Apart from introducing a correlation between different \\alm 's, it was showed in~\\cite{Challinor:2002zh} that our peculiar velocity would also cause a bias in the measured CMB \\emph{intensity} power spectrum proportional to $4\\beta^2$. Here we showed instead that for the \\emph{temperature} power spectrum (the one which is physically of more interest) the bias is exactly \\emph{zero} at ${\\cal O}\\big(\\beta^2\\big)$. We also showed that the small cross-correlation between $E$ and $B$ modes computed in~\\cite{Challinor:2002zh} is instead also \\emph{zero} -- at least up to ${\\cal O}\\big(\\beta^6\\big)$ -- when one corrects the aberration kernel to transform as temperature instead of as intensity. With some recent claims of the existence of possible unexpected bulk flows~\\cite{Watkins:2008hf,Kashlinsky:2008ut,Kashlinsky:2009dw,AtrioBarandela:2010wy} it is a very interesting prospect that we will soon have a complementary measurement of our peculiar velocity with respect to the CMB. A dipolar effect has already been measured outside the CMB: in supernovae~\\cite{Weyant:2011hs}, in the cosmic infrared background~\\cite{Fixsen:2011qk} and also marginally in X-rays~\\cite{Boughn:2002bs}; it could also be measured in the future using the cosmic parallax technique~\\cite{Quercellini:2008ty,Quartin:2009xr,Quercellini:2010zr}, and in the optical, radio and gamma bands. It would therefore be interesting to explore the interconnections and compare the future precision that these different approaches can achieve." }, "1112/1112.5686_arXiv.txt": { "abstract": "We examine the claim made by Hara et al.\\cite{Hara} in 1969 of the observation of a $10^{19}$eV cosmic ray extensive air shower using the air fluorescence technique. We find that it is likely that fluorescence light was observed, confirming this as the first such observation. The work of Hara et al. and their friendly competitors at Cornell University paved the way for modern experiments like the Pierre Auger Observatory and the Telescope Array. ", "introduction": "Investigations into the feasibility of detecting air fluorescence light from extensive air showers were conducted in the 1960's by groups led by Suga in Japan and Greisen in the United States. Results from the Japanese experiment, reported by Hara et al.\\cite{Hara} in 1969, are reviewed here. In that report, the authors say ``One event is very likely due to the atmospheric scintillation [fluorescence] light from an air shower whose primary energy and distance are about $10^{19}$eV and 3\\,km, respectively''. This was the first reported observation of fluorescence light from an air shower. The purpose of the present short note is to the review this observation in the light of our modern understanding of fluorescence detection. The Japanese experiment ran at the Dodaira Observatory (altitude 876\\,m) for a period of 5 months from December 1968. The fluorescence telescope consisted of a 1.6\\,m diameter Fresnel lens focussing light onto a camera of 24 PMTs, each of which imaged a 4.5$^\\circ$ degree portion of the sky. For the observation described here, the telescope with its field of view of $23^\\circ \\times 20^\\circ$ was centered at an elevation of 30$^\\circ$. The design was similar to Greisen's Cornell telescope \\cite{Cornell}. However the Japanese design had the advantage of a larger Fresnel lens, and faster electronics. The rise and fall-times of pulses on the cathode-ray tube displays were 0.12$\\mu$s and 0.2$\\mu$s respectively. The potential fluorescence observation (event \\#12 in Fig 3 of \\cite{Hara}) triggered 8 PMTs with an angular track length of $18.4^\\circ$ and a duration of 1.9$\\mu$s. In the next section we review the event geometry before considering the shape of the light profile received at the telescope. ", "conclusions": "It appears very likely that the signal detected by Hara et al. was one dominated by fluorescence light, and it is reasonable that they lay claim to the first such detection. While the timing information of the event is not sufficient to reconstruct a unique shower axis geometry, the event's light profile is consistent with a fluorescence profile, and quite unlike the peaked profile expected for a Cherenkov dominated signal. Of course some doubt does remain, and the first {\\em unambiguous} detection of fluorescence light from an air shower was with the Utah Fly's Eye prototype operated in coincidence with the Volcano Ranch surface array in 1977 \\cite{FE}. The $5\\times10^{18}$eV shower simulated in this note produced a light profile with a peak flux of about $5\\times10^{10}$\\,photons/m$^2$/s, the flux observed in \\cite{Hara}. This energy is a factor of two lower than that claimed by Hara et al., but this can be explained by their assumption of a fluorescence yield of 1.9 photons per metre of electron track, which is approximately a factor of two lower than the modern accepted value. One remaining uncertainty in this energy is the absolute light calibration of the experiment. Hara et al. discuss an inconsistency between the observed light intensity spectrum and that expected from simulations. The inconsistency led to them increasing the observed light fluxes by a factor of 5. The origin of the inconsistency is unclear, but it would probably be safe to say that $5\\times10^{18}$eV is an upper limit to the energy of the observed air shower." }, "1112/1112.5479_arXiv.txt": { "abstract": "Dark matter-dominated cluster-scale halos act as an important cosmological probe and provide a key testing ground for structure formation theory. Focusing on their mass profiles, we have carried out (gravity-only) simulations of the concordance $\\Lambda$CDM cosmology, covering a mass range of $2 \\times 10^{12}-2\\times 10^{15} h^{-1}$M$_\\odot$ and a redshift range of $z=0-2$, while satisfying the associated requirements of resolution and statistical control. When fitting to the Navarro-Frenk-White profile, our concentration-mass ($c-M$) relation differs in normalization and shape in comparison to previous studies that have limited statistics in the upper end of the mass range. We show that the flattening of the $c-M$ relation with redshift is naturally expressed if $c$ is viewed as a function of the peak height parameter, $\\nu$. Unlike the $c-M$ relation, the slope of the $c-\\nu$ relation is effectively constant over the redshift range $z=0-2$, while the amplitude varies by $\\sim 30\\%$ for massive clusters. This relation is, however, not universal: Using a simulation suite covering the allowed $w$CDM parameter space, we show that the $c-\\nu$ relation varies by about $\\pm$ 20\\% as cosmological parameters are varied. At fixed mass, the $c(M)$ distribution is well-fit by a Gaussian with $\\sigma_c/\\langle c\\rangle \\simeq 0.33$, independent of the radius at which the concentration is defined, the halo dynamical state, and the underlying cosmology. We compare the $\\Lambda$CDM predictions with observations of halo concentrations from strong lensing, weak lensing, galaxy kinematics, and X-ray data, finding good agreement for massive clusters ($M_{vir} > 4\\times 10^{14} \\mau$), but with some disagreements at lower masses. Because of uncertainty in observational systematics and modeling of baryonic physics, the significance of these discrepancies remains unclear. ", "introduction": "\\label{section:introduction} According to the current cosmological model, structure forms in the Universe primarily by the amplification of primordial fluctuations driven by the gravitational Jeans instability. The process of nonlinear structure formation is hierarchical and complex, the initial perturbations evolving eventually into a `cosmic web' network consisting of voids, filaments, and clumps. The clumps, termed halos in cosmological parlance, are dark matter dominated localized mass overdensities with their own complex substructure. Observed baryonic systems such as galaxies and hot gas reside in these halos. Although the dark matter within halos cannot be observed directly, its presence can be inferred by dynamical arguments, and much more directly, through gravitational lensing of background sources. The notion of the dark matter dominated halo is one of the fundamental building blocks in studies of the formation of individual galaxies, galaxy groups, and galaxy clusters (for an overview, see \\citealt{mo_book}). The structure of halos has been extensively studied using N-body simulations over a wide range of halo masses. Even though individual halos can be, and are, dynamically and morphologically complex, it was shown by \\cite{nfw1, nfw2} (NFW) that the spherically averaged density profiles of `relaxed' halos formed in cold dark matter (CDM) simulations can be described by a roughly universal functional form -- the NFW profile -- independent of their mass, the spectrum of initial fluctuations, and cosmological parameters. The NFW profile has a fixed shape, albeit with two scale parameters; as applied to individual halos it has been remarkably successful and is often applied to all halos, regardless of their dynamical state. (When applied to stacked or average halos, this profile is somewhat less succesful, as discussed later below.) The two parameters of the NFW profile are a mass and a scale radius. The scale radius, $r_s$, specifies the point where the logarithmic slope of the profile equals -2 (at small radii, the profile $\\sim 1/r$, while at large radii, it asymptotes to $\\sim 1/r^3$). Instead of $r_s$, one often uses the concentration, which is the radial scale set by the halo mass divided by $r_s$. In cluster cosmology, the usual key observable is the halo mass, rather than the profile per se. The cluster mass function (cluster abundance, more generally), is a sensitive probe of dark energy, since clusters form very late, during the epoch of dark energy dominance. However, measuring the concentration parameter, the simplest first measurement of a profile, can also be very useful. First, as shown originally by NFW, the concentration of a halo, $c$, has a strong correlation to its mass, $M$, therefore measuring the $c-M$ relation observationally is a direct test of the CDM paradigm. In fact, combining cluster $c-M$ predictions and measurements, and the measured gas mass fraction, one can aim to constrain $\\Omega_m$ and $\\sigma_8$ \\citep{ettori11}. As another example, lensing shear peak counts, a proposed weak lensing survey cosmological probe, is very sensitive to the form of the $c-M$ relation \\citep{king11}. Finally, future measurements of the weak lensing power spectrum will probe small enough spatial scales that results will be sensitive to baryonic effects on the halo profile, i.e., modifications to the gravity-only $c-M$ relation \\citep{mw04, zhan04}. We will return to these points in more detail below. The correlation of halo concentration with mass is based on the idea -- as first explicated by NFW -- that the concentration is determined by the mean density of the universe when the halo is assembled, with higher concentrations corresponding to higher densities. Thus cluster mass halos, which are assembling today, should have a lower concentration than halos of lower mass that were built up at an earlier epoch, where the mean density was higher. Furthermore, one may expect this trend to flatten out (sufficiently) beyond the nonlinear mass scale $M_*$, and therefore, since $M_*$ falls rapidly with redshift, flatten out over an extended range in mass as redshift increases. Although the general arguments are plausible and are broadly consistent with simulation results, a predictive theory for the mean of the $c-M$ relation, and its scatter, does not exist. Several simple heuristic models tuned to simulations have been suggested (NFW; \\citealt{bullock99}; \\citealt{eke01}; \\citealt{zhao09}) but their predictive status cannot be considered satisfactory, especially at the higher end of halo masses (see, e.g., \\citealt{gao07}; \\citealt{hayashi07}; \\citealt{maccio08}; \\citealt{zhao09}). Indeed there is sufficient uncertainty even when comparing simulation results from different groups, that the general problem is still open. However, as the mass resolution in large-volume simulations continues to improve, we may expect this situation to be merely temporary. On the observational front, cluster (and group) halo profiles can be studied using both strong and weak gravitational lensing, individually, and in combination (see, e.g., \\citealt{comerford07}; \\citealt{broadhurst08}; \\citealt{mandelbaum08}; \\citealt{okabe09}; \\citealt{oguri11}; \\citealt{zitrin11}; \\citealt{coe12}), projected gas density and temperature profiles from X-ray observations (see, e.g., \\citealt{vikhlinin05}; \\citealt{buote06}; \\citealt{schmidt06}; \\citealt{gastaldello07}; \\citealt{vikhlinin09}; \\citealt{sun09}; \\citealt{ettori11}), and galaxy kinematics (\\citealt{diaferio05}; \\citealt{rines06}; \\citealt{wojtak10} and references therein). Results from these observations have generally shown qualitative agreement with the $c-M$ relation obtained from simulations, although there have been difficulties with matching the shape and normalization. Additionally, there are discrepancies between different sets of observations, presumably because the underlying systematics are not fully understood and modeled. The purpose of this paper is to present a set of predictions for the NFW mass profile targeted primarily towards massive clusters. To do so, however, a fairly large mass range must be considered in order to obtain a sufficiently well-defined $c-M$ relation. Our simulations cover three orders of magnitude in mass ($\\sim 10^{12}-\\sim 10^{15} h^{-1}$M$_\\odot$) with very good control of statistics over the entire range. The high dynamic range and excellent statistics enable us to derive a new set of results for the mass profile, including profile evolution and probability distribution functions (PDFs) for the concentration as a function of mass. We compare our results with previous simulations and with a set of recent observations of the cluster mass profile. The paper is organized as follows. In Section~\\ref{section:halo} we discuss general features of the $c-M$ relation in the simulation context focusing on the role of differing definitions and analyses. In Section~\\ref{section:sims}, we describe the main features of the simulation runs. We present our results for the $c-M$ relation and its redshift evolution in Section~\\ref{section:results}. This is followed (Section~\\ref{sec:wcdm}) by a presentation of results from a suite of $w$CDM cosmologies in order to further study how the concentration depends on cosmology. Next, in Section~\\ref{sec:comp}, we provide a detailed comparison with recent observations, noting areas of agreement and disagreement. Finally, Section~\\ref{section:disc} is devoted to a summary of the results and further discussion. An Appendix discusses various systematic issues that need to be considered when deriving concentrations from simulation results. A number of tests are used to illustrate these points and to verify the robustness of the numerical procedures carried out in this paper. ", "conclusions": "\\label{section:disc} We presented results for the concentrations of dark matter halos using a set of large-volume simulations. With a total volume roughly 1-2 orders of magnitude larger compared to previous simulations, we focused on studying the $c-M$ relation for massive clusters. As shown in the past, at the high mass end, the $c-M$ relation becomes flatter at $z=0$ and the flattening becomes more significant at higher $z$. The mean concentration of the sample when expressed in terms of the peak height parameter, $\\nu(M,z)=\\delta_c/\\sigma(M,z)$, shows a roughly uniform slope at all redshifts. Indeed, the slope of the $c-\\nu$ relation does not change with redshift. The amplitude of the $c-\\nu$ relation evolves by about $30\\%$ at the high mass end from $z=0-2$. The $z$-evolution is consistent with the results of \\cite{gao07}, although the overall amplitude of the concentration differs because of the different choices of $\\sigma_8$. We do not observe a rise in concentration at higher masses as reported by \\cite{klypin10} and \\cite{prada11} (the Appendix includes further discussion). Because of our large halo sample, we can study the distribution of the concentration in individual mass bins; we find that the distribution of concentrations is well-described by a Gaussian PDF \\citep{lukic09,reed11}. Thus the halo profile shape can be described by two parameters -- the mean concentration and its standard deviation. By comparing results across a number of $w$CDM cosmologies, we find that the standard deviation is roughly universal, $\\sigma_c=0.33 c$, and does not change with redshift, halo dynamical state, or cosmological parameters. We investigated how the concentration changes as the cosmological parameters are varied using a set of 18 runs spanning the $w$CDM cosmology parameter space. The parameter range covers the 2$\\sigma$ variation around the best fit WMAP5 cosmology. We find that over this parameter range, the $c-M$ relation varies by $\\sim \\pm$ 20\\%, although the standard deviation $\\sigma_c$ follows the relation $\\sigma_c=0.33 c$. As suggested by our work on the $w$CDM models, and also previous studies of redshift evolution, the halo formation epoch, and hence the concentration, depends on the matter fluctuations, slope of the power spectrum and the growth factor. Thus calibrating the $c-M$ relation as a function of cosmology is important for a wide variety of problems, ranging from galaxy formation, the weak lensing shear power spectrum, to the case of assembly bias in clusters. We will address the cosmology dependence of the $c-M$ relation in detail using more simulations and analytical models elsewhere. The simulation predictions are in good agreement with observations from strong lensing, weak lensing, galaxy kinematics, and X-ray data for massive clusters with masses $M_{vir} > 4\\times 10^{14} \\mau$. At lower masses, different observations suffer from different sources of systematic error. For example, the lensing data need to account for bias due to the triaxiality of halos while the X-ray data typically ignore the non-thermal pressure component in galaxy clusters which can lead to a systematic underestimate of the cluster mass \\citep{lau09}. The simulations, on the other hand also need to account for baryonic effects which play a bigger role as the halo mass decreases. As a result, due to cooling, gravity-only simulations may predict $20-30\\%$ lower concentration for clusters with masses $M_{vir} < 4\\times 10^{14} \\mau$. The fact that most recent observations are in agreement with the simulation results (and amongst themselves) to better than $20\\%$ for massive clusters ($M_{vir} > 4\\times 10^{14} \\mau$) indicates that baryonic effects influencing the cluster mass profile are indeed small and that the individual observational systematics are under some level of control. \\appendix" }, "1112/1112.0540_arXiv.txt": { "abstract": "We present results from an analysis of the broad-band, 0.3--195\\,keV, X-ray spectra of 48 Seyfert 1--1.5 sources detected in the very hard X-rays with the Swift Burst Alert Telescope (BAT). This sample is selected in an all-sky survey conducted in the 14--195\\,keV band. Therefore, our sources are largely unbiased towards both obscuration and host galaxy properties. Our detailed and uniform model fits to Suzaku/BAT and XMM-Newton/BAT spectra include the neutral absorption, direct power-law, reflected emission, soft excess, warm absorption, and narrow \\ion{Fe}{1} K$\\alpha$ emission properties for the entire sample. We significantly detect O VII and O VIII edges in 52\\% of our sample. The strength of these detections are strongly correlated with the neutral column density measured in the spectrum. Among the strongest detections, X-ray grating and UV observations, where available, indicate outflowing material. The ionized column densities of sources with O VII and O VIII detections are clustered in a narrow range with N$_{\\rm warm} \\sim 10^{21}$\\,cm$^{-2}$, while sources without strong detections have column densities of ionized gas an order of magnitude lower. Therefore, we note that sources without strong detections likely have warm ionized outflows present but at low column densities that are not easily probed with current X-ray observations. Sources with strong complex absorption have a strong soft excess, which may or may not be due to difficulties in modeling the complex spectra of these sources. Still, the detection of a flat $\\Gamma \\sim 1$ and a strong soft excess may allow us to infer the presence of strong absorption in low signal-to-noise AGN spectra. Additionally, we include a useful correction from the Swift BAT luminosity to bolometric luminosity, based on a comparison of our spectral fitting results with published spectral energy distribution fits from 33 of our sources. ", "introduction": "The Swift Burst Alert Telescope (BAT) is conducting the first all-sky, very hard X-ray survey in thirty years. With hundreds of detections in a harder X-ray band than the previous survey by HEAO-1 \\citep{1982ApJ...253..485P}, the Swift survey presents an unprecedented sample of active galactic nuclei (AGN) sources. Due to their detection in the 14--195\\,keV, very hard X-ray band, the Swift sources are unbiased to all but the highest levels of obscuration. Therefore, the Swift BAT-detected AGN are an important sample for determining the global properties of AGN. In \\citet{2009ApJ...690.1322W}, the 0.3--10\\,keV X-ray properties for the 153 sources detected in the 9-month Swift BAT catalog \\citep{2008ApJ...681..113T} are presented. The 9-month catalog includes all sources with very hard X-ray detections of F$_{14 - 195 {\\rm keV}} > 10^{-11}$\\flux. These sources are local, with an average redshift of 0.03, and are bright IR/optical/UV/X-ray sources. The X-ray analysis of the 9-month sources uncovered many results of the properties of local AGNs, relying on simple models. However, since both broader X-ray band and higher signal-to-noise spectra are available for a majority of these sources, more detailed analyses are now possible. Our main goal in this paper is the detection and characterization of ``warm absorbers'', signatures of ionized gas potentially from outflows, in the Type 1 AGN. Previous studies searching for warm absorbers were biased -- selecting all the bright sources with archived X-ray or UV data \\citep{1997MNRAS.286..513R,1998ApJS..114...73G,1999ApJ...516..750C,2007AJ....134.1061D}. Therefore, the results of these studies may also be biased. Mass outflows or more specifically AGN feedback mechanisms are important in galaxy formation \\citep{1998AA...331L...1S,2003ARA&A..41..117C,2005ARA&A..43..769V} and a potential cause of the well-known relationship between galaxies and their black holes (i.e., the M-$\\sigma$ relation first described in \\citealt{2000ApJ...539L...9F,2000ApJ...539L..13G}). Thus, it is important to understand the AGN warm absorber properties in an unbiased sample. Since X-ray warm absorbers are most evident in the soft emission ($< 2$\\,keV), where more absorbed sources exhibit low flux levels, we rely on a study of the Type 1 sources detected by Swift's BAT. In this paper, we present the detailed broad-band spectral properties of 48/51 Seyfert 1--1.5 sources, with $| b | > 15$\\degr and high signal-to-noise X-ray CCD spectra available, which were detected in the Swift BAT 9-month survey. Our analysis relies preferentially on X-ray spectra from Suzaku, which provides simultaneous coverage from $\\sim 0.2$--$50$\\,keV, and time-averaged spectra from Swift's BAT in the 14--195\\,keV band. With a broad bandpass, we obtain tight constraints on the continuum emission, which is vital in determining the warm absorber properties. Where Suzaku spectra were not available, we use data from XMM-Newton, which covers the 0.3--10\\,keV band, with the time-averaged Swift BAT spectra. We describe our data reduction in \\S~\\ref{sect-data}. Our spectral analysis is detailed in \\S~\\ref{sect-spectra}. Discussion of our results is found in \\S~\\ref{sect-discussion}, followed by conclusions in \\S~\\ref{sect-conclusion}. ", "conclusions": "\\label{sect-discussion} We performed detailed, uniform spectral fitting on a sample of 48/51 Seyfert 1--1.5 sources selected in the 14-195\\,keV band with the Swift BAT. In this section, we discuss our findings, based on general spectral properties, including the continuum shape, luminosity, high energy emission lines (e.g., \\ion{Fe}{1} K), and soft excess. Additionally, we specifically highlight results of an analysis of the warm absorption properties and their implications for AGN feedback. \\subsection{Basic Spectral Properties} \\subsubsection{\\bf Intrinsic Neutral Column Density and Redshift}\\label{sect-nh} In the X-rays, type 1 AGN typically have low columns of obscuring material in the line of sight to the central source. Since the emission is largely unobscured, we view directly the central regions surrounding the black hole. In the optical, type 1 AGN exhibit broad permitted emission lines, particularly in H$\\alpha$ and H$\\beta$. The distinction between sub-types of Sy 1s relies on the strength of the narrow H$\\beta$ emission component, where a Sy 1.5 is intermediate between a Sy 1 and Sy 2 with a clearly distinguishable narrow emission-line component \\citep{1977ApJ...215..733O}. Since optical Sy 1.5 galaxies are in an intermediate state, in a simple unified model of AGN (e.g., \\citealt{1993ARA&A..31..473A}) we expect the Sy 1.5s to have intermediate column densities between the low column Sy 1s and the higher column Sy 2s. Our sample includes 22 Sy 1s, 9 Sy 1.2s, and 17 Sy 1.5s. Among the type 1s, there are several broad line radio galaxies (BLRGs), including 1H 0419-577, 3C 120, 3C 382, 3C 390.3, 4C +74.26, 2MASX J21140128+8204483, and MR 2251-178. Average values for the logarithm of neutral column density (using an upper limit of $10^{19}$\\,cm$^{-2}$ for spectra with no clearly detected neutral absorption component) correspond to $19.65 \\pm 0.53$ for Sy 1s, $20.12 \\pm 0.63$ for Sy 1.2s, and $20.61 \\pm 1.14$ for Sy 1.5s. Broad line radio galaxies, as a subset of Sy 1s, have an average neutral column density measured at $20.26 \\pm 0.74$. Therefore, Sy 1s do have the lowest column densities, while Sy 1.5s are more obscured in the X-rays, on average. However, if we exclude the four Sy 1.5s with columns higher than $10^{22}$\\,cm$^{-2}$, a K-S test shows a smaller probability of the Sy 1s and Sy 1.5s being drawn from different distributions ($D = 0.359$ and $P = 0.264$). % Further, the optical Sy 1--1.5 sources are, as expected, less obscured in the X-rays than the X-ray obscured sources (with typical values of \\nh\\,$> 10^{22}$\\,cm$^{-2}$, see \\citealt{2009ApJ...690.1322W}). In Figure~\\ref{fig-nhz}, we plot the measured column densities for our sample versus redshift. The highest redshift sources in our sample are the broad line radio galaxies (black squares), with $\\langle z \\rangle = 0.068$. The Sy 1.5s are the lowest redshift sources ($\\langle z \\rangle = 0.012$), while the Sy 1s have an average redshift of 0.026 and the Sy 1.2s are intermediary with a value of 0.017. In Figure~\\ref{fig-nhz}, we find that there are no Sy 1s at the lowest redshifts. From the plot, we find that there are 5 Sy 1.5s and 2 Sy 1.2s at $z < 0.008$. Since Sy 1s are relatively unobscured, we expect that Swift's BAT has no bias against selecting Sy 1s in this redshift range. The lack of Sy 1.5 detections at higher redshift may indicate that the Sy 1.5s are less luminous, since obscured sources in the BAT sample have lower luminosities than unobscured sources \\citep{2009ApJ...690.1322W}. However, at the low column densities measured from our broad-band X-ray fits (from 0.3--195\\,keV), we expect the obscuring material in Sy 1.5s to have little affect at the high energies to which BAT is sensitive. In the following section, we discuss the derived X-ray luminosities, with respect to optical classifications. While we find no Sy 1s at low redshift, it is difficult to make statistical claims on the paucity of low redshift Sy 1s due to the small number of objects in this redshift range. However, this result is intriguing since the X-ray analyses of the Swift sample reveal that the type 2 sources are less luminous than the type 1 objects \\citep{2009ApJ...690.1322W} and the fraction of obscured/unobscured AGN also changes with redshift (e.g., \\citealt{koss-thesis}). We will test this further with optical and X-ray follow-ups of the 600 AGNs detected in the 58-month Swift catalog (Baumgartner et al., submitted). If the same result of few Sy 1s at low-redshift is found, one possible explanation is that there is a physical difference in the broad line region/accretion state between Sy 1.5s and Sy 1s. \\citet{2009ApJ...701L..91E} posit that in the disk wind model, where the broad line region and torus are part of a wind from the accretion disk, that at low accretion rates the broad line region disappears. If the Sy 1.5s are accreting at a lower rate (see \\S~\\ref{lum}), the difference in the appearance of the optical broad lines could be correlated with a different accretion state. In the current study, there is the possibility, however, of our sources being mis-classified (i.e., between Sy 1 sub-types). This is a concern since the NED classifications we use come from heterogeneous sources. We will test this further in a follow-up paper examining our uniform optical spectra obtained with the Apache Point Observatory 3.5-m and CTIO SOAR 4-m telescopes. Using the optical follow-ups, we will carefully classify the optical spectra of our sample to determine whether the NED classifications are correct. We confirm that the optical follow-ups show broad Balmer lines in all of these sources, however, we will include more detailed analyses in the optical paper (Winter et al., in prep). \\subsubsection{\\bf Luminosity, Mass, and Accretion Rate}\\label{lum} The luminosity, black hole mass, and accretion rate are among the most basic properties of an AGN. Since the Swift BAT bandpass is at high enough energies (14--195\\,keV) to be unaffected by all but the highest levels of obscuration, it is a good proxy for the bolometric luminosity of the AGN. Our measurements of the BAT band luminosity are computed using the 14--195\\,keV flux listed in Table~\\ref{tbl-flux}. We find that the BLRGs are the most luminous sources in our sample, while the Sy 1.5s are least luminous. Average values and standard deviations for $\\log$\\,L$_{14-195\\,{\\rm keV}}$ are: $44.8 \\pm 0.2$ (BLRGs), $43.8 \\pm 0.4$ (Sy 1s), $43.4 \\pm 0.6$ (Sy 1.2s), and $43.3 \\pm 0.7$ (Sy 1.5s). The standard deviations demonstrate that there is no statistical difference between the luminosity of Sy 1s, 1.2s, or 1.5s. However, the BLRGs are significantly more luminous -- approximately 4 times more luminous than the Seyfert sources. The mass determinations recorded in Table~\\ref{tbl-flux} are derived from 2MASS bulge photometry of the host galaxies (see \\citealt{2008ApJ...684L..65M} and \\citealt{2009ApJ...690.1322W}) and are in good agreement with the independent analysis of 33 of our sources in \\citet{2007MNRAS.381.1235V} and \\citet{2009MNRAS.392.1124V}. Comparison of the 2MASS-derived masses with both reverberation mapping and H-$\\beta$ FWHM derived masses were discussed in \\citet{2010ApJ...710..503W}. We found masses derived from each of these methods to be well-correlated, with $\\log M_{2MASS} = (0.91 \\pm 0.14) \\times \\log M_{H\\beta} + (1.07 \\pm 1.13)$ and the H$\\beta$-derived masses equivalent to the reverberation results. The masses recorded in Table~\\ref{tbl-flux} are the 2MASS values corrected to match the more accurate H-$\\beta$ derived masses, with the exceptions of NGC 4051, NGC 4593, Mrk 279, and NGC 7469, where we use the reverberation mapping derived masses (in each of these cases there was a significant difference between the masses from the alternate method). Additionally, we note that the mass of the Sy 1.5 2MASX J11454045-1827149, derived from the 2MASS K-band photometry, is low ($\\log M/M_{\\sun} = 6.2$), requiring a very high accretion rate ($L_{\\rm bol}/L_{\\rm Edd} = 3.5$). We have obtained optical spectroscopy of the H-$\\beta$ region of this source from the CTIO SOAR telescope and are in the process of reducing the data. An alternative mass estimate will be included in Winter et al., in prep. Comparing the average masses for our sample, we find $\\log M/M_{\\sun}$ of $8.5 \\pm 0.2$ (BLRGs), $7.9 \\pm 0.6$ (Sy 1s), $7.9 \\pm 0.7$ (Sy 1.2s), and $7.8 \\pm 0.7$ (Sy 1.5s). We find that within the standard deviation the black hole masses are similar between all of our sources. This is also true for the obscured type 2 sources, shown in \\citet{2009ApJ...690.1322W}. To determine the accretion rate of our sources, we must first determine the bolometric luminosity. A reliable method for determining the bolometric luminosity is through broad-band fitting of the spectral energy distribution (e.g., \\citealt{2002ApJ...579..530W}). For 33 sources in our sample, bolometric luminosities and accretion rates were obtained through fitting simultaneous optical/UV/X-ray data from XMM-Newton \\citep{2007MNRAS.381.1235V} or Swift \\citep{2009MNRAS.392.1124V}. In Figure~\\ref{fig-comparebol}, we compare the Swift BAT band luminosities obtained from our spectral fits with the bolometric luminosities calculated from SED fitting in \\citet{2007MNRAS.381.1235V} and \\citet{2009MNRAS.392.1124V}. We expect that the 14--195\\,keV luminosity is the direct unobscured signature from the AGN for our Seyfert 1 sources and our figure confirms this. Fitting an ordinary least squares line to the data, we find that $\\log L_{\\rm bol} = (1.1157 \\pm 0.1172) \\log L_{14-195\\,{\\rm keV}} + (-4.2280 \\pm 5.1376)$. This fit is very significant, with a correlation coefficient of $R^2 = 0.82$. Using this relationship, we find that the average bolometric luminosities of our sources are $5.7 \\times 10^{45}$\\,\\lum (BLRGs), $4.4 \\times 10^{44}$\\,\\lum (Sy 1s), $1.6 \\times 10^{44}$\\,\\lum (Sy 1.2s), and $1.2 \\times 10^{44}$\\,\\lum (Sy 1.5s). The bolometric luminosities are included in Table~\\ref{tbl-flux}. The mass accretion rate, $\\dot{M} = \\frac{L}{\\eta c^2}$, ranges roughly from 1\\,M$_{\\sun}\\,{\\rm yr}^{-1}$ for the BLRGs and 0.02\\,M$_{\\sun}\\,{\\rm yr}^{-1}$ in the Sy 1.5s (assuming $\\eta = 0.1$). It is useful to parameterize the accretion rate relative to the Eddington limit. The Eddington ratio is defined as the ratio of the bolometric luminosity to the Eddington luminosity, the luminosity where the gravitational and radiation pressure balance ($L_{\\rm Edd} = 1.3 \\times 10^{38} \\times M/M_{\\sun}$\\,ergs\\,s$^{-1}$), or equivalently the ratio of the mass accretion rate to the Eddington accretion rate ($\\dot{M}_{\\rm Edd} = \\frac{L_{\\rm Edd}}{\\eta c^2}$). We include the Eddington ratio for our sample in Table~\\ref{tbl-flux}. We find values of $L_{\\rm bol}/L_{Edd}$ corresponding to $0.14 \\pm 0.08$ (BLRGs), $0.05 \\pm 0.14$ (Sy 1s), $0.03 \\pm 0.05$ (Sy 1.2s), and $0.02 \\pm 0.09$ (Sy 1.5s). The main uncertainties in these measurements are in the estimates of the mass. For one source, 2MASX J11454045-1827149, which does not have a measurement from alternative methods, the Eddington ratio is suspiciously high ($> 1$). As discussed above, we will determine a mass estimate based on the H-$\\beta$ width (Winter et al., in prep). The general picture, however, emerging from our estimates shows that local AGN selected in the very hard X-rays tend to have accretion rates from $10^{-3} - 0.5$ times the Eddington rate. The neutral column densities measured in the X-ray spectroscopy are a combination of the effects of absorption features of metals, which impose a soft X-ray cut-off in the spectrum. This material may be associated with intergalactic material, gas local to the host galaxy, and/or intrinsic to the AGN. In \\citet{2009ApJ...690.1322W}, we tested whether the obscuration was associated with the accretion rate of the AGN (Figure 10a) and the inclination of the host galaxy (Figure 16). We found that the bulk of the obscuring material is not associated with the host galaxy (highly inclined hosts have slightly higher neutral column densities, but not enough to account for all of the obscuration measured) and that there is no correlation between the column and accretion rate. In Figure~\\ref{fig-lumlledd}, we test the dependence of the measured column on both the luminosity and Eddington ratio measured from our broad-band spectral fits. There is no correlation found, which is consistent with the unified model, since our line of sight viewing angle with respect to the torus torus, a potential source of the neutral obscuration, is expected to only be an effect of our line of sight to the AGN (while this is true for low luminosity sources like the BAT-detected AGN, we note that this is not the case for higher luminosity sources). \\subsubsection{\\bf Direct and Reflected Continuum} The direct emission from our sources was modeled as a cutoff power-law. The power-law index, $\\Gamma$, measured values correspond to $1.90 \\pm 0.16$ (BLRGs), $1.78 \\pm 0.34$ (Sy 1s), $1.84 \\pm 0.18$ (Sy 1.2s), and $1.52 \\pm 0.61$ (Sy 1.5s). The low average photon index for the Sy 1.5s is a result of several outliers with $\\Gamma \\sim 1.0$, including NGC 526A, Mrk 6, NGC 3227, and NGC 3516, with NGC 4151 having $\\Gamma << 1.0$. Among these, NGC 3227, NGC 3516, and NGC 4151 are poorly fit with the base model (reduced $\\Delta\\chi^2 > 2.0$). Two Sy 1s also have low values: EXO 055620--3820.2 and NGC 3783. For these sources, we attribute the unusual $\\Gamma$ measurements to difficulty modeling the absorbing components. Warm absorber signatures, measured through the \\ion{O}{7}/\\ion{O}{8} edges (see Table~\\ref{tbl-warmabs}), are the strongest in our sample for all but EXO 055620--3820.2. While we do not measure strong absorbers in EXO 055620--3820.2, analysis of the archived X-ray data by \\citet{2009MNRAS.394L...1L} reveals obscuring clouds and complex structure (which we also detect in this observation). Removing these values, we find that the mean $\\Gamma$ value for Sy 1s and Sy 1.5s agree -- $1.90 \\pm 0.17$ and $1.90 \\pm 0.23$, respectively -- and a K-S test reveals a high probability ($P = 0.947$) that both are drawn from the same distribution. Further, these average values are consistent with the distribution of $\\Gamma$ for the BLRGs, demonstrating that there is no statistical indication of spectral hardening with luminosity/accretion rate in the BAT sample (see Figure~\\ref{fig-gammalum}). % In Figure~\\ref{fig-gammalum}, we also plot the average power-law index binned by luminosity/accretion rate (right plots). We divided the range of luminosities/accretion rates for our sources into equally-spaced bins and computed the average power-law index in each of these bins. The error-bars indicate the difference between the average value and the maximum/minimum in each bin, with average values of $\\Gamma = 1.07$, and are plotted as the logarithm of this value in order to illustrate the average values more clearly. In \\citet{2009ApJ...690.1322W}, we found that there was no correlation between power-law index and luminosity/accretion rate, consistent with results from previous low-redshift samples \\citep{2000ApJ...531...52G}. Since our study consisted of all of the Swift-BAT detected AGN (e.g., including both absorbed and lower luminosity sources), we concluded that we found no correlation because the properties of the Swift sources were more diverse than the higher-redshift, more luminous sources which showed these correlations \\citep{2004ApJ...605...45D,2008AJ....135.1505S}. We did, however, see correlations for individual sources where we had multiple X-ray observations \\citep{2008ApJ...674..686W}. As shown in Figure~\\ref{fig-gammalum}, in the current study we find that the photon-index is well-correlated with luminosity and accretion rate, when the values are binned. We find that: \\begin{equation} \\Gamma = (0.16 \\pm 0.03) \\log L_{\\rm bol} + (-5.31 \\pm 1.17), \\end{equation} with a correlation coefficient of $R^2 = 0.86$, and \\begin{equation} \\Gamma = (0.23 \\pm 0.03) \\log L_{\\rm bol}/L_{\\rm Edd} + (2.08 \\pm 0.04), \\end{equation} with a coefficient of determination (the square of the Pearson linear correlation coefficient) of $R^2 = 0.83$ (the probability for both corresponds to $P < 0.001$ for 48 degrees of freedom). Our linear correlation between binned $\\Gamma$-L$_{\\rm bol}$ is consistent with the rest-frame $\\Gamma$-(2-10\\,keV) luminosity relation found in \\citep{2004ApJ...605...45D} ($\\Gamma = (0.13 \\pm 0.04) L_{2-10\\,{\\rm keV}} + (-4.1 \\pm 1.7)$ for $0.3 \\la z \\la 0.96$ AGN). This suggests that the X-ray power-law index and luminosity/accretion rate are related for the Sy 1s and that a similar process is at work as with the higher redshift radio quiet quasars. However, the fact that a similar relation is not found for the entire Swift BAT-detected sample \\citep{2009ApJ...690.1322W} enforces the result that the lowest luminosity sources (absorbed sources whose optical emission line properties classify them as \\ion{H}{2} galaxies/composites/LINERs) are in a different accretion state \\citep{2010ApJ...710..503W}. Cutoff energies of the power-law component are poorly constrained for the majority of our sample (40/48). For many of the sources, the parameter was best-fit with the upper limit of $E_{cutoff} = 500$\\,keV. Of the sources with constrained cutoffs, EXO 055620--3820.2 and Mrk 79 have cutoffs between 20--60\\,keV, and the remaining five sources have cutoffs from the range $\\sim 60$--$400$\\,keV. By contrast, \\citet{2009MNRAS.399.1293M} find cutoff energies between 50--150\\,keV in their sample of Sy 1s, derived from the INTEGRAL survey. As with our analysis, \\citet{2009MNRAS.399.1293M} use broad-band X-ray spectral fits and use a very similar model, including reflection with the {\\tt pexrav} model. Of the overlapping sources between both the INTEGRAL and our Swift study, we find general agreement in the cutoff values, however, our error bars on the cutoff energies are much larger than those in the \\citet{2009MNRAS.399.1293M} analysis. It is unclear why this is the case, particularly since our study includes, on average, a higher number of degrees of freedom from using the joint Suzaku/BAT or XMM-Newton/BAT data points. A reflected spectrum is found to be significant in 37/48 (77\\%) of the sample, as determined by $\\Delta\\chi^2$ on adding this parameter (see Table~\\ref{tbl-fits}). We find good agreement with measured reflection parameters between our study and that of \\citet{2009MNRAS.399.1293M}, of which 10 sources overlap. The average and standard deviation for the measured reflection parameter ($R \\sim \\Omega/2\\pi$) correspond to $1.56 \\pm 1.45$ (BLRGs), $2.23 \\pm 1.47$ (Sy 1s), $1.96 \\pm 1.22$ (Sy 1.2s), and $1.75 \\pm 1.61$ (Sy 1.5s). There is no statistical difference in the reflection parameter between sources of different Sy type. In Figure~\\ref{fig-reflection}, we look for correlations between the reflection parameter and both AGN luminosity and power-law index, $\\Gamma$. Clearly, we find no correlation between $R$ and L$_{\\rm bol}$. A correlation does appear, however, between $R$ and $\\Gamma$, where higher $\\Gamma$ values coincide with higher reflection parameters (note that $R \\sim 0.0$ corresponds to no reflection, while more negative values correspond to larger reflection). Excluding sources with $\\Gamma < 1.5$, which include sources with complex absorbers, we find that $R = (-8.58 \\pm 1.84) + (5.74 \\pm 0.97) \\times \\Gamma$. This correlation is not strong, with a correlation coefficient of $R^2 = 0.49$. Correlations in $R - \\Gamma$ were previously seen in a number of studies, including those of \\citet{1999MNRAS.303L..11Z}, \\citet{2007ApJ...664..101M}, and \\citet{2009MNRAS.399.1293M}. Since both parameters are linked in the fitting process, this trend may not be physical. If the correlation is physical, \\citet{1999MNRAS.303L..11Z} interpret the $R - \\Gamma$ relation as follows. The reflection parameter is proportional to the angle subtended by the reflector. This ``cold'' reflector emits soft photons, which irradiate the X-ray source (i.e., the corona) and act as the seed for Compton upscattering. If $R$ is large, there is more reflecting material and therefore a stronger flux from the soft photons. This leads to stronger cooling of the plasma, a smaller Comptonization parameter ($y$), and a softer spectrum (or larger $\\Gamma$). \\citet{2001ApJ...556..716P} show that with the {\\tt pexrav} model larger reflection parameters also correspond to higher corona temperatures ($kT_e \\propto E_{cutoff}$) and lower optical depths towards inverse Compton scattering, as well as softer spectra (higher $\\Gamma$). For sources in our sample with well-constrained $R$ and $E_{cutoff}$, there is no correlation between cutoff energy, thereby corona temperature, and reflection parameter/$\\Gamma$. We conclude that the relationship between the direct and reflected emission is not easy to interpret in our sample. Particularly, degeneracies between the fitted values from the {\\tt pexrav} model make it impossible to determine whether the noted correlation between $R - \\Gamma$ is physical. Sources with low $\\Gamma$ tend to have higher N$_H$, low $R$, and a range of luminosities. Sources with high $\\Gamma$ tend to have low N$_H$ and high $R$. Cutoff energies, which are related to the Comptonization parameter ($y$), corona temperature, and optical depth towards Compton scattering, are not well-constrained for the majority of the sample. Where they are constrained, there is no correlation with $\\Gamma$. In the future, we will compare the {\\tt pexrav} model results with those from alternative reflection models, such as the dusty torus model MYTORUS \\citep{2009MNRAS.397.1549M}, but such analysis is beyond the scope of the present paper. \\subsubsection{\\bf Soft Excess}~\\label{softexcess} An X-ray soft excess, which we modeled with a simple blackbody, is statistically significant in 45/48 (94\\%) of our sources. Of the three sources without a significant blackbody component, the spectrum of Mrk 926 is poorly fit and the spectrum of 2MASX J11454045$-$1827149 has a relatively short exposure time with XMM-Newton (9\\,ks). Further, a soft excess is detected in a long XMM-Newton spectrum of the final source, NGC 5548 \\citep{2003MNRAS.341..953P}. Therefore, a soft excess is likely present in all local AGN. This contrasts with our previous analyses of the Swift BAT sources, which found a soft excess in $40-50$\\% of AGN \\citep{2008ApJ...674..686W,2009ApJ...690.1322W}. The higher detection rate in the current study is likely due to the higher signal-to-noise observations used in this study. In \\citet{2009ApJ...690.1322W}, we used data from a variety of sources, with the majority of spectra from ASCA and lower signal-to-noise Swift XRT spectra. This made it difficult to detect fainter blackbody components. The average and standard deviation for the best-fit blackbody temperature ($kT$) corresponds to $0.11 \\pm 0.06$\\,keV (BLRGs), $0.11 \\pm 0.04$\\,keV (Sy 1s), $0.32 \\pm 0.60$ (Sy 1.2s), and $0.15 \\pm 0.22$ (Sy 1.5s). The average value for the Sy 1.2s is skewed by one point (MCG -01-13-025 has a best-fit $kT = 2.0$\\,keV, with large error bars) and if this point is disregarded the average and standard deviation ($0.11 \\pm 0.04$\\,keV) are in line with the other sources in our sample. Both the measured blackbody temperature and the small amount of scatter in our sample are consistent with measurements of sources in the Lockman Hole from the XMM-Newton survey \\citep{2005AA...444...79M}, PG-selected QSOs \\citep{2004AA...422...85P, 2005AA...432...15P,2010ApJ...725.1848T}, and our previous analyses of the Swift-selected sources \\citep{2008ApJ...674..686W,2009ApJ...690.1322W}. The origin of the soft excess in AGN is as yet unknown. It may arise from thermal emission from star formation (particularly as seen in ULIRGs or AGN hosted in galaxies with strong nuclear starbursts), a population of near nuclear X-ray binaries/ULXs \\citep{2010MNRAS.407.2399M}, blurred reflected emission \\citep{2003AA...412..317C, 2005MNRAS.358..211R}, or blurred absorption \\citep{2004MNRAS.349L...7G}. In Figure~\\ref{fig-bbodytemp}, we test whether a thermal model is plausible. If the emission is thermal and associated with the black hole, we expect the blackbody temperature to be proportional to $M^{-1/4} L/L_{Edd}^{1/4}$, where $M$ is the black hole mass and $L/L_{Edd}$ is the accretion rate. We find no correlation between the blackbody temperature and either of these parameters, in agreement with the results of our previous analysis of the X-ray spectra of the Swift BAT-selected AGN \\citep{2009ApJ...690.1322W}. However, we do find a slight correlation between the flux in the power-law and the flux in the blackbody component. In Figure~\\ref{fig-ktnorm}, we plot the blackbody normalization versus the power-law normalization. The correlation ($\\log kT\\,{\\rm norm} = (-1.64 \\pm 0.41) + (1.16 \\pm 0.19) \\log \\Gamma\\,{\\rm norm}$; excluding the Sy 1.5s, which have more uncertainty in determining the soft excess parameters due to complex absorption) is weak, with $R^2 = 0.28$, but it indicates the same direct correlation shown from our more careful comparison of the luminosity in the blackbody and power-law components in \\citet{2009ApJ...690.1322W} (i.e., $L_{power-law} \\propto L_{soft excess}$). Similarly, this correlation is also seen in the PG QSOs \\citep{2010ApJ...725.1848T}. The relationship between the power-law and soft excess shows that for the Sy 1s, the soft excess is either created by or affected by the direct AGN emission and not from thermal emission from star formation. Further clues to the origin of the soft excess are seen in a comparison with the absorbing gas. Namely, we find that sources with the largest soft excess also have the highest neutral column densities and the strongest absorber signatures (measured through the optical depth of the O\\,VII absorption edge). These results are clearly shown in Figure~\\ref{fig-nhoviiktnorm}. Binning these relationships, shows that the correlations are very strong. We find that \\begin{equation} \\log A_{\\rm kT} = (1.74 \\pm 0.26) \\log N_{\\rm H} + (-38.77 \\pm 5.50), \\end{equation} with a correlation coefficient of $R^2 = 0.88$. We also find that \\begin{equation} \\log A_{\\rm kT} = (1.62 \\pm 0.19) \\log \\tau_{\\rm O\\,VII} + (-0.58 \\pm 0.35), \\end{equation} with $R^2 = 0.92$. Therefore, the soft excess, neutral column density, and warm ionized gas are connected. Since the soft excess is also correlated with the direct emission, these processes must all be related (i.e., the direct power-law emission, soft excess, absorption). % \\subsubsection{\\bf \\ion{Fe}{1} K$\\alpha$ Emission} The Fe K$\\alpha$ emission feature is typically the most prominent feature in AGN spectra. Recent work shows that the hard X-ray region surrounding the Fe K$\\alpha$ band can be quite complex, with both narrow and broad features from \\ion{Fe}{1} K$\\alpha$, additional narrow features from, e.g., \\ion{Fe}{25} K$\\alpha$, \\ion{Fe}{1} K$\\beta$, and \\ion{Ni}{1} K$\\alpha$, iron K-edges (7.11\\,keV), and absorption features from outflowing highly ionized gas. We focus only on the narrow \\ion{Fe}{1} K$\\alpha$ line in this work, and defer further analysis of the hard X-ray emission/absorption to a future study. We note, however, that many of our sources are included in more in-depth studies, such as the Chandra HETG study in \\citet{2004ApJ...604...63Y} and the XMM-Newton study in \\citet{2007MNRAS.382..194N}. Details of the best-fit parameters for the \\ion{Fe}{1} K$\\alpha$ emission are included in Table~\\ref{tbl-fek}. For many of our sources (40\\% or 19/48), we could not constrain the energy or the width of the emission line. Instead, we fixed these values to default values of 6.41\\,keV and 0.01\\,keV, respectively, to determine limits on the equivalent width. We find that for the sources with well-constrained energies and widths, the average and standard deviation correspond to $E = 6.41 \\pm 0.04$\\,keV and a range in width from $\\sigma = 0.06 - 0.14$\\,keV (corresponding to ranges from $FWHM \\approx 2780 - 9180$\\,km\\,s$^{-1}$). Therefore, while the energy of this line is very similar for our sources, we find that the width varies between sources. However, the average value is still below the resolution of the XIS at 6.4\\,keV, which is $\\sim 130$\\,eV. Still, our best-fit values from the Suzaku and XMM-Newton spectroscopy are in good agreement with Chandra grating results for 14 Seyferts, which found $E = 6.404 \\pm 0.005$\\,keV and a $FWHM = 2380 \\pm 760$\\,km\\,s$^{-1}$ \\citep{2004ApJ...604...63Y}. Estimates of the $EW$ were made for all of the sources in the sample. We find average and standard deviations of the equivalent width of the \\ion{Fe}{1} K$\\alpha$ emission of $108.96 \\pm 190.34$\\,eV. Therefore, there is also a large spread in the $EW$, as well as the width, of the emission feature. The source with the strongest $EW$ line in our sample, with $EW \\approx 1.3$\\,keV, is EXO 055620-3820.2, whose 2006 observation corresponds to a low flux, possibly Compton-thick phase (see \\citealt{2009MNRAS.394L...1L} for a comparison of this observation with previous data showing a high flux state). Disregarding this outlier, we find $EW = 83.10 \\pm 70.08$\\,eV. While past studies found an anti-correlation between the equivalent width and luminosity of AGN (the X-ray Baldwin effect \\citealt{1993ApJ...413L..15I}), we find no correlation -- in agreement with our earlier results on the Swift sources \\citep{2009ApJ...690.1322W}. In Figure~\\ref{fig-fekew}, we show the relationship between the \\ion{Fe}{1} K$\\alpha$ equivalent width and both the bolometric luminosity and Eddington ratio. Linear correlations are not seen with either parameter, with correlation coefficients of $R^2 = 0.25$ and $0.19$, respectively. We previously found this in our \\citet{2009ApJ...690.1322W} study, determining that a correlation exists for multiple observations of individual sources or when the luminosity/accretion rate proxy are binned. To test this further, we looked for evidence of the X-ray Baldwin effect, by binning the data by luminosity and accretion rate. As shown in Figure~\\ref{fig-fekew}, there is no strong correlation between the \\ion{Fe}{1} K$\\alpha$ equivalent width and the bolometric luminosity. We find that $EW \\propto L_{\\rm bol}^{-0.27 \\pm 0.08}$, but with a correlation coefficient of $R^2 = 0.36$ ($P \\sim 0.1$). This correlation is consistent, however, with the relationship we found ($EW \\propto L^{corr}_{2-10\\,{\\rm keV}}$) for the entire 9-month sample \\citep{2009ApJ...690.1322W}. We find a strong correlation between the \\ion{Fe}{1} K$\\alpha$ equivalent width and accretion rate. The linear relationship is parameterized as \\begin{equation} \\log EW = (-0.38 \\pm 0.07) \\log L_{\\rm bol}/L_{\\rm Edd} + (1.35 \\pm 0.08), \\end{equation} with $R^2 = 0.80$. This relationship is consistent with the\\\\ $EW \\propto (L^{corr}_{2-10\\,{\\rm keV}}/L_{\\rm Edd})^{(-0.26 \\pm 0.03)}$ relationship found for the entire Swift sample \\citep{2009ApJ...690.1322W}. This confirms that the primary driver for the observed X-ray Baldwin effect is the correlation between the $EW$ and accretion rate, as suggested by \\citet{2006ApJ...644..725J}. \\subsection{Warm Absorbers} Using the base continuum model described in the preceding section, we constrained the properties of potential warm absorber signatures in our sample. The simplest model for detecting potential outflows in X-ray CCD data is with the addition of absorption edge models fit to the 0.73\\,keV \\ion{O}{7} and 0.87\\,keV \\ion{O}{8} features. We describe details of these fits in \\S~\\ref{o7}. We then fit the spectra with more sophisticated models using the {\\tt warmabs} model, which determines the ionization state and column density of the warm absorbing gas, described in \\S~\\ref{warmabs}. We discuss our conclusions on the warm absorption fits in \\S~\\ref{subsect-conclusions}. \\subsubsection{\\ion{O}{7} and \\ion{O}{8} Absorbers}\\label{o7} The results of adding absorption edges to detect \\ion{O}{7} and \\ion{O}{8} edges are included in Table~\\ref{tbl-warmabs}. Among our sources, 25/48 (52\\%) have clear detections of the edge features. This fraction is higher than our original report of 18/44 (41\\%) in \\citet{2010ApJ...725L.126W}, since we added analysis of 4 sources with recently available spectra (NGC 985, 2MASX J11454045$-$1827149, NGC 6814, and 2MASX J21140128+8204483) and re-analyzed longer exposures available for 3 additional sources (IC 4329A, NGC 5548, and Mrk 926). As in \\citet{2010ApJ...725L.126W}, we classify a detection based on $\\Delta\\chi^2 \\ga 13.39$, which corresponds to a probability of $P = 0.01$ for the four additional degrees of freedom added. The detection rate for sources with Suzaku spectra is 18/34 (53\\%), while the detection rate for XMM-Newton spectra is 7/14 (50\\%). We find no relationship between $\\Delta\\chi^2$, on adding the edges, and the total number of counts in the spectrum (i.e., there are spectra with few counts and strong detections of absorption edges, as well as spectra with many counts and low $\\Delta\\chi^2$ on adding the absorption edges). Likewise, there is no relationship between the measured optical depth of the absorption features and the number of counts in the spectrum (e.g., the four sources with the highest $\\tau$ in the \\ion{O}{7} edge have between $\\sim 1.5 - 4.2 \\times 10^{5}$\\,counts, while spectra in our sample range from total counts of $\\sim 5 \\times 10^{5} - 2 \\times 10^{6}$). Therefore, we find no evidence for biases in our detection rates or measured edge strengths with the total counts in the spectra. In the companion paper, \\citet{2010ApJ...725L.126W}, we showed that the detection of \\ion{O}{7} and \\ion{O}{8} is dependent on luminosity, accretion rate, and column density. In particular, we found that detection rates are higher in sources with larger neutral column densities and lower in the more luminous sources (see Table 2 and Figure 2 in \\citealt{2010ApJ...725L.126W}). Additionally, we find that the strength of the warm absorber features are also dependent upon column density. This is illustrated in Figure~\\ref{fig-eddingtonlimit}, where the optical depth in the \\ion{O}{7} edge is conveyed by the size of the symbols (i.e., the largest symbols correspond to the highest optical depths). In the plot, the neutral column densities are from Table~\\ref{tbl-fits1}, with the exception of NGC 3783. Determinations of the neutral column density of NGC 3783 are available from several observations from 2001--2007 utilizing XMM-Newton, Chandra, and Suzaku. The published measurements range from $\\sim 5 \\times 10^{21} - 10^{22}$\\,cm$^{-2}$ \\citep{2004ApJ...602..648R,2007MNRAS.379.1359M,2009PASJ...61.1331M}. The source is variable in the X-rays and it is unclear whether our low measurement in the 2009 Suzaku observation represents a change in column density from the earlier epochs or is the result of difficulties measuring the column in such a complex spectrum. In the plot, we adopt N$_{\\rm H} = 5 \\times 10^{21}$\\,cm$^{-2}$ for NGC 3783. The N$_{\\rm H}$-Eddington ratio plot is a useful illustration of the effects of radiation pressure on dusty gas surrounding the AGN (see for instance \\citealt{2006MNRAS.373L..16F}, \\citealt{2008MNRAS.385L..43F}, and \\citealt{2010MNRAS.402.1081V}). In Figure~\\ref{fig-eddingtonlimit}, the blue line represents the {\\it effective} Eddington limit, where gravitational pressure balances with radiation pressure on dusty gas with solar abundances for the dust grains, as computed in \\citet{2006MNRAS.373L..16F}. The radiation pressure on dusty gas is enhanced relative to that of ionized gas; such that the the Eddington limit for dusty gas is lower. Outflows driven by radiation pressure in a dusty medium are expected when both the Eddington ratio and the neutral column density are high (i.e., where the effective Eddington limit is exceeded). Sources in this region are believed to be short-lived, as the radiation pressure will eventually drive away the obscuring dusty material, leaving unobscured, low Eddington ratio sources. The sources with the strongest absorption features, i.e., highest optical depth for \\ion{O}{7} absorption, clearly fit with the Fabian et al., model. They have high column densities and Eddington ratios near or above the effective Eddington limit for dusty gas. The estimates of the Eddington ratios for 4/6 of these sources are secure, since the mass estimates are based on reverberation mapping \\citep{2004ApJ...613..682P}. Therefore, a wind driven by radiation pressure on the dusty gas is a likely mechanism for creating the absorption features seen in these sources. Assuming that the absorption edges trace outflowing ionized gas, the sources with the strongest outflow detections are NGC 3516, NGC 4151, Mrk 6, NGC 3227, NGC 526A, and NGC 3783. For NGC 3516, NGC 4151, NGC 3227, and NGC 3783, UV and X-ray grating results confirm that the warm ionized gas is indeed outflowing, with measured absorption lines blue-shifted by $\\sim 100 - 600$\\km~(see, for instance, \\citealt{1999ApJ...516..750C,2001ApJ...555..633C} for analysis of UV Hubble spectroscopy; also note that the low-ionization lines of NGC 3227 are red-shifted, whereas the higher ionization lines are blue-shifted). Neither Mrk 6 or NGC 526A have available UV spectroscopy, as these are highly reddened sources, and X-ray grating observations of these sources do not have sufficient signal-to-noise to confirm that the gas is outflowing. Since the strong outflow sources also have complicated spectra, we searched the literature for confirmation that the high neutral column densities we measured were also detected in alternative spectral fits. We already described above the column density measurements for NGC 3783, which indicate N$_H \\ga 5 \\times 10^{21}$\\,cm$^{-2}$. \\citet{2011ApJ...731...21M} find that there is a variable absorber in Mrk 6, with N$_{\\rm H} \\sim 10^{21} - 10^{23}$\\,cm$^{-2}$ over the past 6 years of available X-ray spectroscopy. The ASCA and Chandra analyses of NGC 4151 also reveal variable high column density gas (e.g., \\citealt{2010ApJ...714.1497W}). For this source, we note that the soft X-rays are dominated by emission features, with a weak and heavily absorbed continuum, and that the grating observations reveal that the edges from \\ion{O}{7}, \\ion{O}{8}, \\ion{Ne}{9}, and \\ion{Ne}{10} are blended together \\citep{2005ApJ...633..693K}. Simple fits to the Suzaku data of NGC 4151 indicate that there is a strong component of photoionized emission, as in the earlier observations, but a more comprehensive analysis of the complex spectrum is beyond the scope of the current paper. Despite the complexity of NGC 4151's X-ray spectra, the grating observations clearly indicated a high column density ($\\sim 10^{22}$\\,cm$^{-2}$) of absorbed gas, consistent with this study, particularly when NGC 4151 is in a lower flux state (as is the case in the Suzaku spectrum presented here). Since the soft continuum is very weak and dominated by strong emission, we can not constrain the warm absorption properties of NGC 4151. We can say that it is heavily absorbed and that the complexity of this source is unique in our sample, as no other source shows as weak a continuum or as strong of emission features. An independent analysis of the XMM-Newton spectrum of NGC 526A agrees with our measured column of $\\sim 10^{22}$\\,cm$^{-2}$ \\citep{2011MNRAS.413.1206B}. NGC 3516 also is known to exhibit high column density ($\\ga 10^{22}$\\,cm$^{-2}$) gas in its X-ray spectrum \\citep{2008PASJ...60S.277M}. The XMM-Newton analysis by \\citet{2009ApJ...691..922M} of NGC 3227 uncovers a lower neutral component, $9 \\times 10^{20}$\\,cm$^{-2}$, than our measured value of $\\sim 1.6 \\times 10^{22}$\\,cm$^{-2}$, but since the source's column density has been observed to vary in the past and there is no published analysis of the Suzaku spectrum we analyzed, it is unclear how our results compare. As a whole, our literature search supports our conclusion that the sources with the strongest outflow detections are associated with higher column densities of gas. Our spectral fits allow insight into the spectral properties of sources with strong outflows. In particular, we identify signs of strong absorbers that can be used in lower quality data, for instance, for low flux or higher redshift sources where X-ray grating spectroscopy is infeasible. Two clear signs of sources with strong outflows are a flat power-law index and a strong soft excess. We find that among the sources with the strongest warm absorber detections, all of these sources have measured values for $\\Gamma \\approx 1.0$. Two other sources with $\\Gamma < 1.5$ also have complex absorption in their spectra: EXO 055620-3820.2 and NGC 5548. These low measured values are likely due to difficulties in modeling the spectra, which are absorbed by multiple components of dusty and ionized gas. Likely, the ``true'' direct emission of these sources is a $\\Gamma \\sim 1.9$ power-law, as found for the majority of our sources, but the absorption distorts the soft X-ray spectrum, flattening the fit. Even with the extended hard X-ray spectra from Suzaku HXD and Swift BAT, the intrinsic power-law continuum is difficult to uncover from the fitting process. While a flat spectrum is not physically representative of the direct emission, it is an easy diagnostic for identifying sources with complex absorbed spectra, both for Seyfert 1s and heavily absorbed Compton-thick spectra, in lower signal-to-noise data. Additionally, we find that sources with strongly detected outflows (confirmed through the archived grating and UV data) also have the strongest soft excesses (see also \\S~\\ref{softexcess}). As with the flat power-law index, a strong soft excess can be an effect of strong absorption features distorting the spectrum. Alternatively, the soft excess could be related to the warm ionized absorbing gas. If the soft excess is a feature of the outflow/ionized absorption, this is in agreement with the soft excess as complex absorption model \\citep{2004MNRAS.349L...7G,2007MNRAS.374..150S}. From our spectral fitting of \\ion{O}{7} and \\ion{O}{8} absorption edges, we find that the 0.73\\,keV \\ion{O}{7} edge tends to have the higher optical depth. For the sources classified as not having strong warm absorber detections, based on $\\Delta\\chi^2$, only 10 sources have an optical depth of zero in the \\ion{O}{7} feature. This suggests that warm ionized absorbing gas is present in at least 80\\% of the sample. However, it is unclear whether this material is outflowing, as high signal-to-noise UV or X-ray grating observations are needed to measure the velocity shifts in individual absorption lines. Among the sources with zero optical depth, three sources are BLRGs, four are Sy 1s, two are Sy 1.2s, and one is a Sy 1.5. All of these sources have low neutral column densities (N$_{\\rm H} \\la 10^{20}$\\,cm$^{-2}$). In \\citet{2010ApJ...725L.126W}, we discussed how the published observations of each of these three BLRGs reveal ionized absorption, with ionization parameters higher or lower than are expected to create the \\ion{O}{7} and \\ion{O}{8} edges. It is possible that this is the case for the remaining seven sources, as well, in which case the covering fraction of an outflow is $\\Omega \\sim 1$. In the following section, we explore this through a discussion of our analytic warm absorber fits. With these models, we determined both the column density and ionization parameter of ionized gas in our entire sample, including both the sources with and without clear detections of \\ion{O}{7} and \\ion{O}{8} absorption. \\subsubsection{Analytic Warm Absorption Models}\\label{warmabs} In \\citet{2010ApJ...725L.126W}, we claimed that while the detection rate of \\ion{O}{7} and \\ion{O}{8} edges was low in the highest luminosity sources (33\\%) and high in the low luminosity sources (60\\%), more detailed spectral fits reveal the presence of ionized gas that in many cases is either more or less ionized than required to produce strong \\ion{O}{7} and \\ion{O}{8} features. As a follow-up to this work, we fit analytic models to the X-ray CCD spectra. Results of these fits are included in Tables~\\ref{tbl-warmabs-outflow} and \\ref{tbl-warmabs-nonoutflow}. The X-ray CCD data does not have the energy resolution to accurately determine the velocity of outflow components. However, we were able to determine both the column density of warm ionized gas (N$_{\\rm warm}$) and the ionization parameter ($\\xi = {\\rm L}_{\\rm ion}/(n_e R^2)$; where L$_{\\rm ion}$ is the ionizing luminosity, $n_e$ is the electron density, and $R$ is the distance of the ionized gas from the central ionizing source) for our entire sample. For the sources with clear detections of absorbers through \\ion{O}{7} and \\ion{O}{8} edges, we fit the spectra with two ionized components. We add only one warm absorber component for the sources without strong detections of the edge features, to determine limits on ionized absorbers present. For the majority of sources in Table~\\ref{tbl-warmabs-outflow}, the two component analytic model improves the fits significantly, as indicated by the $\\Delta\\chi^2$ values (e.g., out of the 25 sources fit with a two-component model, only 8 show a second component to have low significance with $\\Delta\\chi^2 < 20$ upon adding the second warm absorber). \\subsubsection*{Caveats on the Analysis and Detection of Ionized Absorbers in CCD Data} While ideally we would compare our results to those from an X-ray grating analysis, the complication of X-ray variability makes such comparisons difficult. For instance, while 8 sources overlap between our analysis and the uniform Chandra grating analysis in \\citet{2007MNRAS.379.1359M}, a comparison of the observed soft fluxes from our observations and the Chandra observations reveals that 6 of the sources vary considerably. We find, for example, that NGC 4051 is ten times brighter in the Suzaku observation, while NGC 4593 is four times fainter. Therefore, there is no consistent way to compare our results unless we know that a source does not vary or we analyze simultaneous grating and CCD data. In \\citet{2010ApJ...719..737W}, such an analysis is presented using the 2009 XMM-Newton CCD and grating spectrum of NGC 6860, along with a Suzaku observation, taken a year earlier, which showed no variability from the XMM observation. In this analysis, the best-fit neutral and ionized column densities were consistent within the error-bars between the Suzaku CCD, XMM-Newton CCD, and XMM-Newton grating observations. We find, however, that the ionization parameter, consistent in both CCD analyses ($\\log \\xi \\sim 2.0$), is higher in the grating observation ($\\log \\xi = 2.4$). The second warm absorber fit had best-fit values that varied considerably between the CCD and grating measurements. Since multiple components are present in AGN spectra and the quality of the grating spectrum from NGC 6860 was relatively low signal-to-noise, it is unclear how these results reflect on our entire sample. However, it is likely that the measured column densities are accurate, particularly for significant warm absorption components. % It is also important to note that since we do not have velocity information on the detected ionized absorbers, due to the low resolution of the CCD data, we can not confirm that the absorbers originate in an AGN driven outflow (i.e., due to the low velocity resolution, we fixed the velocity of the ionized absorbers to the systemic velocity of the AGN). We have shown that for the sources with the strongest detections, outflows are likely present, since UV/X-ray grating observations confirm blue-shifted absorption features. For the remaining sources, the absorbing gas could be intrinsic to the AGN or a feature from ionized gas in our own Galaxy, the host galaxy, or intervening systems (signatures of the warm hot ionized medium). A Chandra grating study by \\citet{2004ApJ...617..232M}, for instance, detected oxygen absorption in half of a sample of 15 type 1 AGN sources that was identified as hot gas from local interstellar structures and potential intergalactic medium features. In order to test whether the ionized gas is outflowing from the AGN, high signal-to-noise X-ray grating observations and/or ultraviolet spectra are needed to measure the velocity of the absorption features. In future work, we will present an analysis of the X-ray grating and ultraviolet spectra of our sample. \\subsubsection*{Results of the Analytic Warm Absorption Spectral Fits} For the six sources with the strongest \\ion{O}{7} and \\ion{O}{8} absorption features, we find average column densities of warm ionized gas of a few $10^{21}$\\,cm$^{-2}$ in each of the two absorption components fit to the spectra. All but NGC 3227 have a warm ionized component with an ionization parameter near $\\xi \\approx 100$\\,ergs\\,s$^{-1}$ (NGC 3227's highest measured ionization component is $\\xi \\approx 28$\\,ergs\\,s$^{-1}$). Out of the entire sample of sources with detected \\ion{O}{7} and \\ion{O}{8} edges, we find that the average column and ionization parameters, for the component with the highest warm ionized column density, are $N_{\\rm warm} = (3.8 \\pm 9.1) \\times 10^{21}$\\,cm$^{-2}$ and $\\log \\xi = 2.04 \\pm 1.27$. We find $N_{\\rm warm} = (3.9 \\pm 4.1) \\times 10^{20}$\\,cm$^{-2}$ and $\\log \\xi = 1.59 \\pm 1.69$ for the second component. These values are in line with the range in column density (N$_{\\rm warm} \\sim 10^{20} - 10^{23}$\\,cm$^{-2}$) and ionization parameter ($\\xi \\sim 10^{0-4}$\\,ergs\\,s$^{-1}$) found from studies of X-ray grating observations (see \\citealt{2005AA...431..111B} and \\citealt{2007MNRAS.379.1359M}). For the sources without strong detections of \\ion{O}{7} and \\ion{O}{8}, we find $N_{\\rm warm} = (1.8 \\pm 2.2) \\times 10^{20}$\\,cm$^{-2}$ and $\\log \\xi = 1.09 \\pm 2.04$. Of these, three have significant warm absorption detections with the analytic model, including, 1H 0419-577 (low column and low ionization parameter), EXO 055620-3820.2 (low column and low ionization parameter), and 3C 382 (with low column and a higher ionization parameter). To test whether the analytic model result of a very low ionization parameter in the spectrum of 1H 0419-577 (the source with the highest significance with the analytic model in Table~\\ref{tbl-warmabs-nonoutflow} and lowest measured ionization parameter) was plausible, we alternatively fit the spectrum with an edge model for lowly ionized oxygen (\\ion{O}{1}/\\ion{O}{2}) at 0.545\\,keV. We found that the lowly ionized edge was very significant, with $\\Delta\\chi^2 = 73.3$, and find an optical depth of $\\tau = 0.15^{+0.03}_{-0.03}$. Our results, then, show that sources without strong detections of \\ion{O}{7} and \\ion{O}{8} absorption edges tend to have lower column densities and lower ionization parameters, on average, than sources with strong detections. In particular, the column densities of warm ionized gas are an order of magnitude lower in the non-detection sources. The ionization parameters, however, do show a large range of values, including both lowly ionized and highly ionized gas. In Figure~\\ref{fig-nwarm1}, we compare the results on the properties of the warm ionized gas in sources with strong \\ion{O}{7} and \\ion{O}{8} absorption edges. Component 1 is chosen as the model component with the highest column density from Table~\\ref{tbl-warmabs-outflow}. In the top plots, we compare the ionization parameters to the bolometric luminosity. We find that the ionization parameter, $\\xi$, has no dependence on the luminosity of the AGN. Instead, we find the majority of values clustered near $\\xi = 100$\\,ergs\\,s$^{-1}$ for component 1 and a broader range in $\\xi$ for component 2. This method is not particularly sensitive to very high $\\xi$ ($>3.5$) or very low $\\xi$ ($<1.0$) absorbers, since those absorber components are driven by fits to a small number of highly ionized lines (e.g. \\ion{Si}{14} Lya) or Fe UTAs, which require grating spectral resolution to fit properly. The Fe UTAs in particular may show up in the \\ion{O}{7} and \\ion{O}{8} edge fits. The most likely scenario to account for the different components is that the absorbing region consists of a warm ionized medium with higher density blobs or filaments embedded within it \\citep{2001ApJ...561..684K}. In this case, the sources without the strong detections of \\ion{O}{7} and \\ion{O}{8} have similar ionization parameters with the lower density, ionized medium also detected in the sources with 'strong' absorber detections. \\subsubsection{Conclusions on Warm Absorber Properties}\\label{subsect-conclusions} The most comparable previous studies to the work we present on the warm absorber properties in the Swift-detected Seyfert 1s are the ASCA studies of \\citet{1997MNRAS.286..513R} and \\citet{1998ApJS..114...73G}. In \\citet{2010ApJ...725L.126W}, we discussed the overlap in our sample with the \\citet{1997MNRAS.286..513R} study. Out of the 24 sources in the ASCA study, 18 are also in our sample. Between publication of \\citet{2010ApJ...725L.126W} and our present results, we revised the classification of two sources (IC 4329A and NGC 5548) to have warm absorption detections. Thus, we find good agreement in our classification scheme with the \\citet{1997MNRAS.286..513R} study, with 16/18 sources classified accordingly. Both 3C 382 and 3C 390.3 were classified by \\citet{1997MNRAS.286..513R} as exhibiting \\ion{O}{7} and \\ion{O}{8} features in the ASCA data, while our analysis shows that these sources do not have significant detections. Grating spectroscopy of these sources reveal that ionized outflows are present, but at higher ionization parameters \\citep{2010MNRAS.401L..10T,2009ApJ...700.1473S}. Our detection rate for warm absorption features, revealed by the significance of \\ion{O}{7} and \\ion{O}{8} absorption edge features, is 52\\%. This is in agreement both with the more biased samples presented in the ASCA study of \\citet{1997MNRAS.286..513R} and the UV outflow studies of \\citet{1999ApJ...516..750C}, but much lower than the $\\sim 70$\\% detection rate reported in \\citet{1998ApJS..114...73G}. While our detection rate, assuming that the warm absorbers are produced in outflowing gas, initially suggests a covering fraction of AGN outflows of $\\Omega \\sim 0.5$, there is clearly more that needs to be considered in this simple picture. In \\citet{2010ApJ...725L.126W}, we pointed out that the detection rate of \\ion{O}{7} and \\ion{O}{8} edges was related to the luminosity of the AGN. There are fewer detections in the most luminous ($\\sim 30$\\%) and higher detection rates for the least luminous ($\\sim 60$\\%) sources. Grating observations, however, reveal that the detection rate is higher -- consistent with outflows in at least 80\\% of the most luminous sources. In the current study, we find that the main difference between sources with and without \\ion{O}{7} and \\ion{O}{8} detections is the measured column density of potentially outflowing ionized gas. We find that sources with strong detections have an order of magnitude or higher ionized column densities than those without detections. Therefore, we can not rule out that sources without \\ion{O}{7} and \\ion{O}{8} detections do not have warm absorbers present. Our results support the hypothesis that they do have ionized gas, just at lower column densities than are easily detected in the X-ray data, which is most sensitive to absorption with N$_{\\rm H} > 10^{20}$\\,cm$^{-2}$. This is still orders of magnitude higher than the ionized gas detected in the UV, which has column densities from N$_{\\rm ion} \\sim 10^{12} - 10^{14}$\\,cm$^{-2}$. Through this work, we suggest a change in the paradigm of assuming a 50\\% covering fraction for AGN warm absorbers (and by extension outflows, since grating observations reveal that this gas is typically outflowing) to $\\sim$100\\%. Our analysis of an unbiased sample of type 1 AGNs selected in the very hard X-rays with Swift shows that while the covering fraction is likely unity, there are clear distinctions between the higher column density warm absorbers, which have neutral and ionized columns $> 10^{20}$\\,cm$^{-2}$, and lower column density warm absorbers. The higher column density sources make up half of the sample. Among these sources with strong detections of warm absorbers/outflows, we suggest that the strongest outflows are driven by radiation pressure on dusty gas, as claimed by \\citet{2006MNRAS.373L..16F}. A main goal of AGN outflow study work is to determine how much mass and energy is entrained in the outflow. While work such as the grating study of \\citet{2005AA...431..111B} places estimates on the energy in the outflows (e.g., \\citealt{2005AA...431..111B} estimate that outflows account for less than 1\\% of the bolometric AGN luminosity), many assumptions, particularly relating to the geometry of the outflow, are necessary to make these estimates (see, for instance, \\citealt{2007MNRAS.379.1359M} for a discussion of these assumptions). Therefore, we do not include estimates of the mass outflow rate or the total energy in the current paper. In future work, we will follow-up the current study with a uniform analysis of the archived grating observations, available for the majority of our sample but without as uniformly high a signal-to-noise ratio as in the CCD data presented in this work. Additionally, estimates of the total outflow rate/energy must rely on multi-wavelength data, characterizing the AGN outflow contribution from neutral through the most highly ionized gas (e.g., \\citealt{2011arXiv1109.2882T} detect Fe K-shell absorption lines in $> 35$\\% of their sample of 42 low redshift, radio-quiet AGN), and include studies of the obscured AGN (X-ray type 2) sources, whose UV and soft X-ray ionized absorber properties are difficult to determine. We will explore both of these lines of research in our future work." }, "1112/1112.2918_arXiv.txt": { "abstract": "Helium abundances and atmospheric parameters have been determined from high resolution spectra for a new sample of 46 bright hot subdwarf B (sdB) stars. The helium abundances have been measured with high accuracy. We confirm the correlation of helium abundance with temperature and the existence of two distinct sequences in helium abundance found previously. We focused on isotopic shifts of helium lines and found $^{3}$He to be strongly enriched in 8 of our programme stars. Most of these stars cluster in a small temperature range between $27\\,000\\,{\\rm K}$ and $31\\,000\\,{\\rm K}$ very similar to the known $^{3}$He-rich main sequence B stars, which cluster at somewhat lower temperatures. This phenomenon is most probably related to diffusion processes in the atmosphere, but poses a challenge to diffusion models. ", "introduction": "The atmospheric helium abundances of sdBs are poorly understood. From the typical abundance patterns of these stars, which show a depletion of light elements as well as an enrichment of heavy metals, it has been concluded that diffusion play an important role in their atmospheres. However, diffusion models predict an almost total depletion of helium in contrast to what is observed. The helium abundances of sdBs range from slightly above solar down to $\\log{y}<-4$. Mass loss caused by stellar winds as well as extra mixing in the atmosphere have been invoked to counteract gravitational settling and explain the observed helium abundances \\citep[see ][ and references therein]{hu11}. \\citet{edelmann03} found a correlation of helium abundance with temperature. The hotter the sdB, the more helium is present in its atmosphere. Similar correlations have been found by other groups (see e.g. Vennes et al. these proceedings). However, \\citet{edelmann03} also reported the discovery of two distinct sequences showing a similar correlation with temperature, the ''lower sequence'' being offset by about $2\\,{\\rm dex}$ from the upper sequence. The majority of stars lie on the ''upper sequence''. Those sequences could not be clearly identified in other datasets so far \\citep[e.g. ][]{lisker05, geier11}. \\citet{otoole08} combined the then published data sets and found the stars of the upper sequence to lie near the Extreme Horizontal Branch (EHB) band in the $T_{\\rm eff}-\\log{g}$-plane, as expected, whereas the lower-sequence stars lie in a much more dispersed area \\citep[see Figs.~2,3 in][]{otoole08}. Gravitational settling can also lead to isotopic anomalies in stellar atmospheres. In the case of helium the light isotope $^{3}$He can be enriched with respect to the usually much more abundant $^{4}$He. Such an enrichment has initially been found in main sequence B stars with subsolar helium abundance \\citep{hartoog79}. However, \\citet{heber91} detected strong line shifts in the sdB star SB\\,290 and the blue horizontal branch star PHL\\,25 indicating that basically the whole helium content of the atmosphere consists of $^{3}$He. Later on \\citet{edelmann01} and \\citet{heber04} found another three sdBs, where $^{3}$He is enriched in the atmosphere. Here we present the results of a quantitative spectral analysis of a sample of 46 sdB stars from high resolution spectra. \\begin{figure} \\begin{center} \\includegraphics[width=10cm]{nhevsteff.eps} \\caption{Helium abundance $\\log{y}$ plotted against effective temperature. The filled symbols mark the results from our study. Filled red diamonds mark objects where isotopic shifts due to an enrichment of $^{3}$He were detected, filled circles objects with atmospheres dominated by $^{4}$He. Upper limits are marked with triangles. The solid horizontal line is drawn at solar helium abundance. The two dotted lines are regression lines for the two distinct helium sequences taken from \\citet{edelmann03}. Results taken from the literature are plotted as grey symbols \\citep{saffer94, maxted01, edelmann03, morales03, lisker05, geier11, vennes11, oestensen10}.} \\label{fig:abun} \\end{center} \\end{figure} ", "conclusions": "The distribution of these stars in the $T_{\\rm eff}$-$\\log{g}$-diagram is shown in Fig.~\\ref{fig:tefflogg} including the three sdBs with isotopic shifts taken from literature. It can be clearly seen that they cluster in a narrow temperature range between $27\\,000\\,{\\rm K}$ and $31\\,000\\,{\\rm K}$ with BD+48\\,2721 ($T_{\\rm eff}=24\\,800\\,{\\rm K}$) being the only exception. Given the uncertainties, this $^{3}$He-strip may be pure. Most stars show clear shifts of the He\\,{\\sc i} line at $6678\\,{\\rm \\AA}$ indicating that almost all helium in the atmosphere is $^{3}$He. BD+48\\,2721, EC\\,12234$-$2607 and PG\\,1519+640 show strong lines of $^{3}$He blended with weak components of $^{4}$He. These three stars cover the whole $^{3}$He temperature strip. The isotope ratio is therefore not correlated to the effective temperature. An $^{3}$He isotope anomaly has first been found for chemically peculiar main sequence stars of spectral type B. The $^{3}$He-stars were found at effective temperatures between $18\\,000\\,{\\rm K}$ and $21\\,000\\,{\\rm K}$ separating helium-poor stars at lower $T_{\\rm eff}$ from helium-rich stars at higher $T_{\\rm eff}$ \\citep{hartoog79}. In Fig.~\\ref{fig:tefflogg} a similar pattern can be seen for the sdBs. The stars enriched in $^{3}$He occupy a small strip in $T_{\\rm eff}$, while the helium abundance decreases towards lower temperatures and rises towards higher temperatures. \\citet{michaud11} carried out diffusion calculations and predict a mild enrichment of $^{3}$He, but due to gravitational settling of the heavier isotope this should be the case in all sdBs. Hence, the $^{3}$He strip stars lacks an explanation. \\citet{hartoog79} argued that diffusion is responsible for this effect. At low temperatures the radiation pressure is not strong enough to support helium in the atmosphere. As soon as the temperature reaches a certain threshold value, the less massive $^{3}$He can be supported, but not the more abundant $^{4}$He. This leads to an enrichment of $^{3}$He in the atmosphere. At even higher temperatures both isotopes are enriched and the isotopic anomaly vanishes as the helium abundance rises." }, "1112/1112.2403_arXiv.txt": { "abstract": "{ The stellar metallicity is a direct measure of the amount of metals present in a galaxy, as a large part of the metals lie in its stars. In this paper we investigate new stellar metallicity indicators suitable for high-z galaxies studying the stellar photospheric absorption lines in the rest frame ultraviolet, hence sampling predominantly young hot stars. We defined these new indicators based on the equivalent widths (EW) of selected features using theoretical spectra created with the evolutionary population synthesis code {\\it Starburts99}. We used them to compute the stellar metallicity for a sample of UV-selected galaxies at $z>3$ from the AMAZE survey using very deep (37h per object) VLT/FORS spectra. Moreover, we applied the new metallicity indicators to eight additional high redshift galaxies found in literature. We then compared stellar and gas-phase metallicities measured from the emission lines for all these galaxies, finding that within the errors the two estimates are in good agreement, with possible tendency to have stellar metallicities lower than the gas phase ones. For the first time, we are able to study the stellar mass-metallicity relation at $z>3$. We find that the metallicity of young, hot stars in galaxies at $z\\sim3$ have similar values of the aged stars in local SDSS galaxies, contrary to what observed for the gas phase metallicity. } ", "introduction": "Metallicity is one of the important properties of galaxies, and its study is able to shed light on the details of galaxy evolution. It is, in fact, an integrated property, related to the whole past history of the galaxies. In particular, metallicity is sensitive to whole star formation history, and so to evolutionary stage of the galaxy. Moreover, it is affected by presence of infalls and outflows, i.e. by feedback processes and by the interplay between the forming galaxy and the intergalactic medium (see e.g. Erb et al. 2008, Mannucci et al. 2009, Cresci et al. 2010). As consequence, it has become an important test of galaxy evolution (e.g. Nagamine et al. 2001, Spitoni et al. 2010, Dav\\`e et al. 2011).\\\\ Local galaxies show a clear correlation between mass and metallicity (MZR), for which the galaxies with larger stellar mass have higher metallicities, and this correlation appears to hold both in term of gas-phase metallicity (e.g. Tremonti et al. 2004) and stellar metallicity (Gallazzi et al. 2005, Panter et al. 2008).\\\\ At high-redshift the gas-phase metallicity of the ISM of star-forming galaxies has been measured using primarily oxygen abundances. The most common techniques to determine the gas phase metallicity are based either on theoretical calibrations (see Kewley \\& Ellison 2002, and Kewley \\& Ellison 2008) or on empirical metallicity calibrations, the so-called ``strong line diagnostics'', which are based on the ratios of collisionally excited forbidden lines to hydrogen recombination lines. Previous studies have shown that the mass-gas phase metallicity relation presents evidence of strong redshift evolution. Among others, Savaglio et al. (2005) and Zahid et al. (2011) studied star forming galaxies at redshift $z\\sim0.7$ and demonstrated that, at given mass, these galaxies shown lower metallicity than the SDSS sample at $z\\sim0.1$. Erb et al. (2006) reported a more significant decrease of metallicity in galaxies at $z\\sim2.2$. Two projects were specifically designed to extend the investigation of MZR at $z>3$: LSD (Lyman-break Stellar population and Dynamic) and AMAZE (Assessing the Mass-Abundance redshift Evolution). With these projects, Maiolino et al. (2008) and Mannucci et al. (2009) showed for the first time the evolution of the mass-metallicity relation at $z>3$. However, the redshift evolution of the gas phase metallicity in galaxies have been questioned recently by Mannucci et al. (2010). They discovered that metallicity depends not only on the mass, but also from the Star Formation Rate (SFR): for a given stellar mass, galaxies with higher SFRs systematically show lower metallicities. This is the so-called ``Fundamental Metallicity Relation (FMR)'', i.e., a tight relation between stellar mass, gas-phase metallicity, and star formation rate (SFR). Local SDSS galaxies show very small residuals around this relation, of the order of 0.05dex. Yates et al. (2011) found a similar relation, with some differences due to the metallicity calibration adopted. According to Mannucci et al. (2010), the FMR does not appear to evolve with redshift up to z$\\sim$2.5, with the high redshift galaxies following the same FMR defined by the local SDSS galaxies. This suggests that the observed evolution of the mass-metallicity relation is due to selection effects and to the increase of the average SFR with redshift. In fact, the measured metallicity in several additional samples of high-z galaxies results to be in agreement with the predictions of the FMR given the mass and SFR of the galaxies: galaxies having lower SFRs than the general population at their redshifts also have higher metallcities (e.g., Richard et al. 2011, Nakajina et al. 2011), and galaxies with higher SFRs also have lower metallicities (e.g., Erb et al. 2010, Contini et al. 2011, Sanders et al. 2011), so that all these galaxies follow the FMR. Also, the FMR allowed Mannucci et al. 2011 and Campisi et al. 2011 to show that the hosts of the long- GRBs have the same metallicity properties of the other star-forming galaxies. However, they found some metallicity evolution of the FMR at $z\\sim3.3$, where galaxies tend to have lower metallicities.\\\\ All the observational studies mentioned in the previous paragraph refer to the gas-phase metallicity, as measured by emission lines. Gallazzi et al. (2005) presented the local mass-stellar metallicity relation based on $\\sim170000$ SDSS galaxies (Sloan Digital Sky Survey Data Release Two). The stellar metallicities were derived using the Lick system of spectral indices in the optical region which are sensitive to the overall metallicity of the stellar population, primarily dominated by intermediate/old stars. They adopted a Bayesian statistical approach and derived the stellar metallicities by comparing the observed spectrum of each galaxy with a comprehensive library of model spectra corresponding to different star formation histories. They found that at low masses, the stellar metallicity increase with mass, while above $\\sim3\\times10M_{\\odot}$ the relation flattens out. In addiction they noted that gas-phase metallicity is best determined for star-forming galaxies, whereas stellar metallicity is best determined for early-type galaxies, and found that the stellar metallicity is generally lower than the gas-phase metallicity (by \u223c0.5 dex). More recently Panter et al. (2008) inferred the stellar metallicity history of SDSS galaxies and determined their stellar mass-metallicity relation. They used a different approach respect to Gallazzi et al. (2005), but they found similar results. Moreover, considering only the younger population of galaxies ($\\leq$1 Gyr) they found good agreement also with the gas phase metallicities. \\\\ A very limited work has instead been done on stellar metallicity at high redshift, see Shapley (2011) for a recent review. As in local {\\it starbursts}, the strongest features in the rest-frame UV spectrum of distant galaxies are interstellar and photospheric absorption lines of C, N, O, Si, and Fe, produced by hot, young O-B stars (see Shapley et al. 2003). One advantage in using these spectra to measure the metallicity at $z\\sim3$ is that the UV rest frame is shifted into the optical spectral region, which is easier to observe from the ground-based telescope. However, very high signal to noise on the stellar continuum is required to study the relevant absorption features for metallicity measurements, and therefore such studies were obtained, until now, mainly for gravitationally lensed galaxies(Rix et al. 2004, Quider et al. 2009, Dessauges-Zavadsky et al. 2010) or for co-added star forming galaxy spectra (Halliday et al. 2008).\\\\ Several authors have presented calibrations for stellar metallicity based on UV absorption lines. Leitherer et al. (2001) investigated the influence of metallicity on the spectra of star forming galaxies. In particular, they investigated the existence of some blended photospheric lines whose strengths depend on metallicities only. They found that the two blends of lines near $\\lambda$1370 and $\\lambda$1425 (which they attributed to OV $\\lambda$1371 and FeV $\\lambda$1360-$\\lambda$1380 and to $\\lambda$SiIII 1417, CIII $\\lambda$1427 and FeV $\\lambda$1430 respectively) have equivalent widths that increase steadily with metallicity and do not depend on other stellar parameters, as age and IMF. Rix et al. (2004) using the {\\it Starburst99} plus their non-LTE model atmosphere code WM-basic, supported the conclusions that these lines are useful metallicity indicators, and suggested a new indicator at $\\lambda$ 1978. Rix et al. (2004) applied the new indicators to measure the stellar metallicity of two lensed galaxies, MS 1512-cB58 at z=2.73, and Q1307-BM1163 at z=1.411, finding good agreement with the gas phase metallicity.\\\\ On the basis of these works, Mehelert et al. (2006) calculated the EWs of the $\\lambda$1370 $\\lambda$1425 for 12 galaxies with $2.373$. In the next section we review the different indices used for stellar metallicity estimates at high-z, and define new rest frame UV feature suitable for such measurements to enlarge the number of available metallicity indicators. In Section 3 we present our sample of high redshift galaxies and we compute their stellar metallicity applying the calibrations found. In Section 4 we calculate the stellar and gas phase metallicity for some lensed galaxies found in literature, and we compare the two. Finally we present the first stellar mass-metallicity relation obtained at $z>3$, followed by the conclusions. \\begin{figure*} \\centering \\includegraphics[width=12cm,angle=270]{Fig1.ps} \\caption{ Variation of the line indices with stellar population age and metallicity. In each plot, the EW is shown in left vertical axis, while the right axis shows the average fractional depth below the continuum. For each feature we draw the models constructed with {\\it Starburts99} at five value of metallicity as shown in different colors. For each metallicity, the color regions represent the error due to the dependence on the IMF assumed. In this way we can highlight the dependence of the models and robustness of the indices from the IMF and stellar population age assumed. } \\label{linvar} \\end{figure*} ", "conclusions": "In this paper, we have investigated for the first time the stellar mass-metallicity relation at high redshift, $z\\sim3$.\\\\ Using the theoretical spectra created with the population synthesis code {\\it Starburts99}, we looked for photospheric absorption lines to be used as indicators of stellar metallicity.\\\\ First we tested the line indices proposed by Leitherer et al. (2001) and Rix et al. (2004), the F1370, F1425 and F1978 using the last version of {\\it Starburts99}, although the F1978 shows a strong dependence from the resolution and the IMF.\\\\ Then we defined two new photospheric lines, F1460 and F1501, and we found that these lines are sensitive to the metallicity and almost independent of the age and the IMF, and therefore useful stellar metallicity indicators. The F1501 index seems to be the most promising because it is defined on the narrowest wavelength range and less affected by the uncertainties on the continuum definition.\\\\ We provided the metallicity calibrations, see Fig.~\\ref{cal1}, with two different definitions of the continuum: the first relations are referred to the real continuum, that we suggest to use in case of spectra with low signal-to-noise ratio, the others were obtained using the definition of the ``pseudo-continuum'' provided by Rix et al. (2004), that we recommend in case of high signal-to-noise spectra, see Sec.~\\ref{mc}.\\\\ We applied the relations on one galaxy and a composite spectra comprised of three additional galaxies of the AMAZE sample at $z\\sim3.3$, for which the gas phase metallicity and the galaxy masses were already know.\\\\ We took from the literature the spectra of eight additional galaxies, and we recompute their stellar metallicity using the new calibrations.\\\\ At the end we compared the results found with the gas phase metallicity for each object, see Fig.~\\ref{dif}. The main conclusion of this work is that within the errors, the stellar and the gas phase metallicity are consistent, although there seems to be a tendency to find stellar metallicity lower than the gas phase one by $\\sim0.1dex$, as already found by Halliday et al. (2008). This result supports the low metal content derived for the gas phase of high-z galaxies from optical strong line ratios, as well as an evolution of the Fundamental Metallicity Relation at $z>3$ as found by Mannucci et al. (2010). \\\\ For the first time, we obtained the stellar mass-metallicity at redshift $z>2.5$, see Fig.~\\ref{MZ}. We notice that the stellar metallicities found at high redshift is comparable with those found by Panter et al. (2008) for local galaxies, although the two are not straightforwardly comparable as in high redshift galaxies the stellar metallicity are computed for hot, young stars, while in the local galaxies for cold, older stellar population.\\\\ In summary, the rest-frame UV is rich in metallicity dependent features, which are able to provide a measure of stellar metallicity in high redshift galaxies. This represent an independent measure of the chemical abundances in galaxies with respect to the more widespread gas phase metallicities, which can provide important constraints to the star formation histories of galaxies in the early Universe. Although this technique is currently limited to very bright or lensed galaxies by the high S/N required, the advent of next generation of telescopes will give us much higher quality spectra for high redshift galaxies, and the stellar metallicity indicators will play a more important role in chemical abundances studies at high redshift." }, "1112/1112.2479.txt": { "abstract": "{We use cosmological hydrodynamical simulations to show that a significant fraction of the gas in high redshift rare massive halos falls nearly radially to their very centre on extremely short timescales. This process results in the formation of very compact bulges with specific angular momentum a factor $5-30$ smaller than the average angular momentum of the baryons in the whole halo. Such low angular momentum originates both from segregation and effective cancellation when the gas flows to the centre of the halo along well defined cold filamentary streams. These filaments penetrate deep inside the halo and connect to the bulge from multiple rapidly changing directions. Structures falling in along the filaments (satellite galaxies) or formed by gravitational instabilities triggered by the inflow (star clusters) further reduce the angular momentum of the gas in the bulge. Finally, the fraction of gas radially falling to the centre appears to increase with the mass of the halo; we argue that this is most likely due to an enhanced cancellation of angular momentum in rarer halos which are fed by more isotropically distributed cold streams. Such an increasingly efficient funneling of low-angular momentum gas to the centre of very massive halos at high redshift may account for the rapid pace at which the most massive super massive black holes grow to reach observed masses around $10^9$M$_\\odot$ at an epoch when the Universe is barely 1 Gyr old.} ", "introduction": "Supermassive black holes (BH) have been established to be ubiquitous at the centre of local galactic bulges and their mass has been shown to correlate well with the stellar mass and perhaps even more strongly with the stellar velocity dispersion of these bulges \\citep{magorrianetal98, tremaineetal02, haring&rix04}. As the accretion of gas onto BHs can drive strong feedback from Active Galactic Nuclei (AGN) through spherical winds or collimated jets, and can potentially self-regulate the growth of the BH along with the cold baryon content of galaxies \\citep{silk&rees98, haehneltetal98, king03} AGN feedback is a natural candidate to explain the observed correlations. Indeed, this picture has been further substantiated by several numerical implementations of such feedback using either semi-analytical models \\citep{boweretal06, crotonetal06, somervilleetal08} or hydrodynamical simulations \\citep{dimatteoetal05, booth&schaye09, duboisetal12}. The issue is further complicated by the fact that the most supermassive BHs (several $10^9\\, \\rm M_\\odot$) seem to be already in place at $z\\approx 6-7$, i.e. less than a Gyr after the Big-Bang \\citep{willottetal03, fanetal06,jiangetal09,mortlocketal11}. Growth to such large masses in such a short time scale constitutes a significant challenge for any model of BH evolution, especially those with strong AGN feedback, as it requires sustained feeding at close to the Eddington accretion rate \\citep{haiman04, sijackietal09}. However, considering that a substantial fraction of the sky had to be surveyed to discover the luminous QSOs powered by these very massive high-redshift BHs ($10^{-9} \\, \\rm Mpc^{-3}$ comoving number density of high-redshift quasars according to~\\citealp{fanetal06}), it appears likely that the halos hosting these BHs are the most massive formed at these early times and thus are very rare objects. Obviously, the main caveat of this 'rareness' argument is that since these young BHs are thought to be growing fast because of the presence of large amounts of gas in their surroundings, a larger number of their quasar counterparts could possibly be obscured in optical wavebands \\citep{alexanderetal03, treisteretal11, willott11} and only be visible in X-rays~\\citep[e.g.][]{daddietal07} and the Infrared. This selection effect could increase the {\\it true} number of very bright quasars/super massive BHs at high redshift considerably. Nevertheless, even when adopting the view that these objects are rare, one still needs to propose an effective mechanism to funnel low angular momentum gas all the way down to the sphere of influence of the BHs at the very centre of their very massive halos hosts. Hydrodynamical simulations suggest that high-redshift quasars are accreting gas at rates close to their Eddington (or super-Eddington) limit because large amounts of cold gas are indeed trapped in the centre of galaxies, and that the fraction of Eddington-limited growth of BHs diminishes with time as this gas reservoir is depleted~\\citep{sijackietal07, dimatteoetal08, duboisetal12}. \\cite{dimatteoetal12} have emphasized that filamentary infall of cold gas is the major mode of gas supply for the continuous feeding of BHs similar to what is discussed for the build-up of galaxies at intermediate redshift \\citep{keresetal05, agertzetal09}. Note that this is a very different mode of gas supply from that due to the often invoked more episodic feeding of BHs due to large amounts of gas driven to galactic nuclei by (major) mergers \\citep{mihos&hernquist96, kauffmann&haehnelt00, hopkinsetal06, mayeretal10} \\cite{pichonetal11} and \\cite{kimmetal11} have recently used hydro cosmological Adaptive Mesh Refinement (AMR) simulations to investigate the angular momentum properties and the infall of gas in Milky Way class halos at low and intermediate redshifts with a view to study the formation and evolution of their central galaxy discs. With a similar technique~\\cite{bournaudetal11} have performed simulations of isolated galaxy discs with parsec resolution to show that Toomre instabilities drive strong inflows of gas to the centre of these discs. We build here on this work, by investigating the more massive and rarer halos expected to host the most massive supermassive BHs at high redshift. We therefore concentrate our exploratory study on the gas with the lowest angular momentum falling to the centre of these rare, massive halos. %--------------------------------------------------------------------------------------- \\begin{figure*} \\includegraphics[width=\\columnwidth]{fig/nice_SH_z6.eps} \\includegraphics[width=\\columnwidth]{fig/nice_LH_z6_zoom64.eps} \\includegraphics[width=\\columnwidth]{fig/nice_SH_z6_20pc_zoom1024.eps} \\includegraphics[width=\\columnwidth]{fig/nice_stars_ugr_zoom4096.eps} \\caption{{\\sl Top panels:} Polychromatic view of the two halos in the SH and the LH simulation at $z=6$. Gas density is color coded in green, gas temperature in red, and tracer particles in blue. The circles in the two top panels corresponds to the virial radius $r_{\\rm vir}$. The {\\sl bottom left panel} displays the inner region of the top left panel simulated with higher resolution (SHhr), showing the transient disc and its connected filaments as discussed in the text. The circle corresponds to 0.1 $r_{\\rm vir}$ (see table~\\ref{tab:feature} for numerical values). The {\\sl bottom right panel } displays the stellar emission in the inner region of the bottom left panel as it would be observed in ugr filter bands. Note the compact bulge (the saturated spherical region at the centre) and the significant number of satellites flowing in along the cold streams on their way to merge with the central galaxy. Finally, we point out that even though in projection the hot gas phase seems to dominate, this halo is still mainly accreting cold gas. } \\label{fig:visual} \\end{figure*} %--------------------------------------------------------------------------------------- Whilst \\cite{dimatteoetal12} have recently investigated this problem using a simulation with a large enough volume ($533\\, h^{-1}\\cdot\\rm Mpc$) to capture halos sufficiently massive to plausibly host these $\\sim 10^9\\, \\rm M_\\odot$ BHs, we take a different approach here and focus our study on two individual very massive halos at high redshift which we follow using a high-resolution cosmological hydrodynamics re-simulation technique. The main aim here is to better resolve the dynamics and angular momentum history of the inflowing cold gas as it falls to the centre of halos, forms a galactic bulge and potentially feeds the growth of the central supermassive BH. %-------------------------------------------------------------------------------------- \\begin{figure*} % \\includegraphics[width=\\columnwidth]{fig/vrot_vcirc_vdisp.ps} \\includegraphics[width=\\columnwidth]{fig/vrot_vcirc3.ps} % \\includegraphics[width=\\columnwidth]{fig/vrot_vcirc4.ps} \\includegraphics[width=\\columnwidth]{fig/vrot_vcirc_vdisp_LH.ps} \\caption{Circular velocities (black lines and inset), radial velocity dispersion (blue lines) and rotational velocities (red lines) of the DM (dotted lines), the stars (dashed lines), and the gas (dot-dashed lines) for the SHhr simulation (left panel), and the LH simulation (right panel) at $z=6$. The blue line corresponds to the SH run. The bulge radius for the SHhr simulation is shown as a vertical arrow. At this redshift, these galaxies have the rotation curve of very compact ellipticals. } \\label{fig:vrot_vcirc} \\end{figure*} %-------------------------------------------------------------------------------------- The paper is organized as follows. Section~\\ref{section:numerics} describes the numerical setup of our simulations. Section~\\ref{sec:mass} investigates the angular momentum evolution and the typical trajectories of the baryonic material forming the bulge. Section~\\ref{sec:discussion} discusses the implications for BH build up at high redshift while Section~\\ref{sec:conclusion} gives our conclusions. %=============================================== % Simulation setup %=============================================== ", "conclusions": "\\label{sec:conclusion} We have investigated here the formation of a compact central bulge in two rare very massive halos at redshift $z=6$. Our main results are as follows. \\begin{itemize} \\item{The typical path for a parcel of fluid entering a halo and eventually ending up in the bulge, is to stream out of a void into the surrounding wall, from there into the filaments defining the intersection of the cosmic walls/sheets, and then along the filaments which penetrate deep into the halo.} \\item{ Early on, most of the baryons ending up in the bulge stream nearly radially along the cold filamentary infall directly into a very compact bulge. As time progresses an increasing fraction of the baryons settles first into a compact gravitationally unstable transient central disc from where they progress rapidly into the bulge. By redshift $z=6$ about half the baryons forming the bulge move first through this surrounding disc before reaching the bulge.} \\item{In the outer part of the halo the angular distribution of the baryonic inflowing matter is stable and reflects the angular distribution of the cosmic web, while well inside the virial radius the angular distribution becomes much more random and changes rapidly.} \\item{A small fraction ($\\sim$ 10 \\%) of the baryons in the halos falls directly from the structures of the cosmic web into the centre of the halos in the form of proto-galactic clumps; these clumps are carried along the filamentary streams, that represent more than half of the total accretion of gas, and join the compact central bulge as minor mergers. } \\item{In our simulations most of the baryons in the bulge turn into stars with a remaining gas fraction of 5-30\\%. The gas fraction is, however, strongly dependent on the resolution and the star formation criterion of the simulation. The gas fraction should also be strongly affected by feedback both from the stars but even more by feedback from the accreting central supermassive BH once the BH has become sufficiently massive. Note again that the simulations used here do not include any feedback.} \\item{The baryon dominated bulge has angular momentum a factor 10-30 lower than the average specific angular momentum of the baryons in the two simulated halos. The baryonic material which is mainly in the form of stars is extremely strongly bound and compact with a peak of the circular velocity approaching 1000 km/s at a radius $< 100$pc. } \\item{The low value of the angular momentum is due to three effects. First the bulge is preferentially made from material which already has a somewhat lower specific angular momentum than average as it enters the halo along the filaments of the cosmic web. Second the baryonic matter entering the halo from the cosmic environment does so with a broad angular momentum distribution. As the gas streams in cold the lower angular momentum material falls quickly to the centre of the halo due to the lack of pressure support where the randomly oriented angular momentum effectively cancels. Third there is additional angular momentum loss due to dynamical friction on inspiralling proto-galactic clumps, exchange between spin and orbital angular momentum and gravitational instabilities in the disc.} \\item{The fraction of gas streaming nearly radially to the centre of the halo making up the low-angular momentum bulge appears to strongly increase for rare halos. We argue that this is due to the fact that isolated rare halos are connected to the cosmic web more isotropically and with a larger number of filaments. This leads both to lower specific angular momentum when the baryonic matter enters the halo and to more complete angular momentum cancellation when the low angular momentum material streams cold into the compact bulge. } \\end{itemize} Numerical simulations of galaxy formation are still subject to a range of uncertainties both numerical and in terms of the implemented physics in particularly with respect to feedback due to stars and AGN. The filamentary cold inflow of gas appears nevertheless to emerge as the key ingredient in the formation of the observed compact discs and bulges at high redshift. As we have discussed here it may also be key to the efficient build up of the several billion solar mass black holes inferred to exist as early as $z\\sim 6-7$. %=============================================== % Acknowledgments %===============================================" }, "1112/1112.0360_arXiv.txt": { "abstract": "The k=0 Friedmann Lemaitre Robertson Walker model with a positive cosmological constant and a massless scalar field is analyzed in detail. If one uses the scalar field as relational time, new features arise already in the Hamiltonian framework of classical general relativity: In a \\emph{finite} interval of relational time, the universe expands out to infinite proper time and zero matter density. In the deparameterized quantum theory, the true Hamiltonian now fails to be essentially self-adjoint both in the Wheeler DeWitt (\\WDW) approach and in LQC. Irrespective of the choice of the self-adjoint extension, the big bang singularity persists in the \\WDW theory while it is resolved and replaced by a big bounce in loop quantum cosmology (LQC). Furthermore, the quantum evolution is surprisingly insensitive to the choice of the self-adjoint extension. This may be a special case of an yet to be discovered general property of a certain class of symmetric operators that fail to be essentially self-adjoint. ", "introduction": "\\label{s1} Loop quantum cosmology (LQC) of the k=0, $\\Lambda=0$ Friedmann Lemaitre Robertson Walker (FLRW) model with a massless scalar field was discussed in detail in \\cite{aps3}. The scalar field serves as a viable internal time variable both in the classical and the quantum theory, with respect to which relational observables such as the matter density and curvature evolve \\cite{aps1,aps2}. This makes it possible to explicitly construct the physical Hilbert space and introduce relational Dirac observables to unravel physics of the Planck regime in a large number of cosmological models \\cite{as}, and a scheme has been sketched even for full general relativity \\cite{warsaw-full}. Using this setup it was rigorously established that, while the big bang singularity persists in the \\WDW theory of the k=0, $\\Lambda=0$ model, it is resolved due to the quantum geometry effects of loop quantum gravity (LQG) \\cite{aps3}.% \\footnote{Recently, this result has been conceptually sharpened using the consistent histories framework in which one can calculate probabilities for the occurrence of certain histories without recourse to external measurement devices or interaction with environment. Using appropriate coarse grained histories which completely decohere, it was shown that the probability of encountering a singularity in the distant past or future is 1 in the \\WDW theory and 0 in LQC for any state (which is in the domain of operators used to construct coarse-grained histories) \\cite{consistent2,consistent3,as}.} An appendix in \\cite{aps3} also outlined how the cosmological constant $\\Lambda$ with either sign can be incorporated. A subsequent, detailed discussion of the $\\Lambda<0$ case appeared in \\cite{bp}. It firmly established that, as in the $\\Lambda=0$ case, in LQC the big bang singularity is replaced by a quantum bounce which occurs when the total energy density $\\rho_{\\rm tot}$ reaches its maximum value $\\rcr$. Furthermore the numerical value of $\\rcr$ is the same as in the $\\Lambda=0$ case, $\\rcr \\approx 0.41 \\rho_{\\rm Pl}$, although now $\\rho_{\\rm tot}$ includes a contribution from the cosmological constant in addition to the matter density $\\rho$;\\, $\\rho_{\\rm tot} = \\rho + \\Lambda/8\\pi G$. It turns out that, by a suitable choice of time variable (or lapse function), the $\\Lambda=0$ model can be solved exactly \\cite{acs}. This is not the case for $\\Lambda\\not=0$. Therefore, results of \\cite{bp} for the $\\Lambda <0$ case are conceptually important also because they demonstrate that the LQC bounce and the qualitative features of the resulting Planck scale physics are not tied to exact solvability. Finally, although the situation with the bounce is the same, the presence of the cosmological constant does alter the underlying mathematical structure in non-trivial ways. In particular, in the deparameterized picture, while the spectrum of the true Hamiltonian is continuous in the $\\Lambda=0$ case, it is purely discrete in the $\\Lambda <0$ case. The goal of this paper is to present an analogous, detailed account of the $\\Lambda >0$ case. Even though we will again consider a massless scalar field, rather surprisingly, the flip of the sign of the cosmological constant changes the underlying mathematical and conceptual structure significantly. Let us begin with the classical theory. If one again uses the scalar field $\\phi$ for internal time, in contrast to the $\\Lambda =0$ and $\\Lambda <0$ cases \\cite{aps3,bp}, the Hamiltonian vector field on the phase space is now incomplete. As a result, volume of any compact co-moving region becomes infinite and the matter density vanishes at a finite instant $\\phi_o$ of internal time $\\phi$. This situation is qualitatively similar to that in the case of a non-relativistic particle in a steep negative potential whose dynamical trajectories reach infinity in a finite time. In such situations, typically, the Hamiltonian operator in Schr\\\"odinger quantum mechanics is symmetric but not essentially self-adjoint. Each self-adjoint extension then yields a unitary evolution but evolutions obtained from distinct operators are both mathematically and physically inequivalent. In the present case, one again finds that the true Hamiltonian operator generating evolution in the scalar field time is symmetric but not essentially self-adjoint. However, rather surprisingly, this ambiguity has negligible effect on states of physical interest: those that start out being peaked at a classical solution in a low curvature region. In particular, all these states undergo a quantum bounce and the total density $\\rho_{\\rm tot}$ at the bounce is again universal. Furthermore, while the evolution of expectation values of physical observables does depend on the choice of self-adjoint extension, the dependence is extremely weak. This robustness may be related to the fact that, on the classical phase space, one can extend both the evolution equations and the solutions simply by analytical continuation, without having to introduce specific boundary conditions at infinity. Our analysis raises the possibility that there may well be a general pattern and new results could be found on properties of certain sub-classes of operators that fail to be essentially self-adjoint. The paper is organized as follows. In Sec. \\ref{s2} we discuss the Hamiltonian framework for the k=0, $\\Lambda>0$ model. Sec. \\ref{s3} is devoted to the \\WDW quantum theory and Sec. \\ref{s4} to LQC. We conclude in \\ref{s5} with a brief summary and discussion. Because the numerical simulations in this paper were completed soon after the initial analysis in \\cite{aps3}, they use an older value of LQC area gap which turned out to be half the value that is relevant for states used in LQC \\cite{awe1}. In the main text we use this more recent value. Therefore, unfortunately, in Sec. \\ref{s4} there is an occasional mismatch of factors of two between the text and the figures. ", "conclusions": "\\label{s5} In this paper, we analyzed in detail the \\WDW theory and LQC of the k=0, $\\Lambda>0$ FLRW model along the lines of the treatment of the $\\Lambda <0$ case of \\cite{bp}, thereby completing the program outlined in an Appendix of \\cite{aps3}. As in the $\\Lambda \\le 0$ cases, the scalar field can be used as a global clock, providing us with a natural notion of relational time both in the classical and quantum theories. However, interestingly, there is a key difference in the physically most interesting case, that of $\\Lambda >0$: In classical general relativity, solutions that start with infinite matter density $\\rho$ at the big bang at time $\\phi = -\\infty$ expand out and now achieve $\\rho=0$ at some \\emph{finite} value $\\phi_o$ of internal time (when the volume $v$ of the fiducial cell $\\mathcal{C}$ becomes infinite). Thus, in the $\\rho$-$\\phi$ plane each of these dynamical trajectories starts out at $\\phi = -\\infty$ but ends at $\\phi=\\phi_o$. But it can be \\emph{analytically} extended beyond $\\phi=\\phi_o$ and the extension represents a universe which starts out with zero matter density at $\\phi=\\phi_o$ but contracts, ending in a big crunch singularity at $\\phi=\\infty$. >From the relational time perspective, then, one is led to regard the two branches as providing a single dynamical trajectory because it is artificial to simply end dynamics at a finite value of time. Now, in non-relativistic mechanics if the potential is negative and steep the particle may roll off to infinity in a finite amount of time. In that case, one has to choose from a one (or more) parameter family of boundary conditions at infinity to continue dynamics beyond that time. In the present case, by contrast, we did not have to resort to making a choice because the most interesting Dirac observable, $\\rho|_\\phi$, is analytic in $\\phi$. Nonetheless, the fact that the universe expands out to $\\rho=0$ at a finite value of the relational time introduces ambiguities in the \\emph{quantum} evolution: The operator $\\Theta_\\Lambda$ which generates dynamics with respect to $\\phi$ now fails to be essentially self-adjoint. Before discussing this point in detail, let us first note two aspects of this phenomenon. First, it is not a peculiarity of LQC; it occurs also in the \\WDW theory. Second, in both cases, the lack of essential self-adjointness is related directly with the behavior of the system at large $v$ and low matter density $\\rho$; its origin does \\emph{not} lie in the Planck scale physics. In both quantum theories, the dynamical operator admits a one parameter family of self-adjoint extensions. For a general system, different choices of extensions can give rise to very different dynamics. However, in this model the results are surprisingly robust with respect to this choice. Not only is the qualitative behavior of dynamics the same, but the differences in the dynamics of the expectation values of the most interesting Dirac observables in theories resulting from two different extensions are smaller than their dispersions in any one theory, even for general states. Furthermore, numerical simulations show that, irrespective of the choice of extension, quantum states which are semi-classical in the low curvature (or low total density $\\rho_{\\rm tot}$) regime remain sharply peaked at the \\emph{extended} classical trajectory in the low curvature regime both in the \\WDW theory and LQC. It is tempting to conjecture that this robustness of quantum dynamics is related to the fact that we did not have to choose a boundary condition at $\\phi=\\phi_o$ to extend the classical $\\rho-\\phi$ trajectory. In the remainder of this section, most of our discussion on the behavior of wave functions will refer to these states. As in the $\\Lambda \\le 0$ cases, there is a \\emph{pronounced} difference between the quantum dynamics of the two theories in the Planck regime. In the \\WDW theory, the wave function simply follows the extended classical trajectory into the big-bang and the big-crunch singularities. In LQC, by contrast, while these states remain peaked at the classical trajectory so long as the curvature (or $\\rho_{\\rm tot}$) is low compared to the Planck scale, there is a dramatic departure in the Planck regime. There is again a new repulsive force with origin in the quantum geometry that overwhelms classical gravity and cases a quantum bounce. Again, the numerical simulations show that, although the force is so strong in the Planck regime, it dies very quickly and becomes negligible once $\\rho_{\\rm tot}$ falls below $10^{-2}- 10^{-3}$ Planck density. In LQC then, even though we are in the k=0 case, we are led to a scenario that is approximately cyclic. As in the k=1 LQC models, the quantum evolution spans an infinite number of epochs. In each epoch the universe begins with a quantum bounce where $\\rho_{\\rm tot} \\approx 0.41 \\rho_{\\rm Pl}$, expands out till $\\rho_{\\rm tot} = \\Lambda/8\\pi G$ and then undergoes a collapse till it reaches another quantum bounce. For states under consideration, dynamics is nearly periodic. How does this dynamics appear in the space-time picture? Let us begin with the classical theory and consider a solution in which the universe starts out with a big-bang at $\\phi=-\\infty$. It expands out to $\\scri^+$ ---which is space-like for $\\Lambda>0$--- where the matter density vanishes and $\\rho_{\\rm tot} = \\Lambda/8\\pi G$. In terms of the physical metric, this space-time is future complete. However, the extended phase space trajectory analytically continues the space-time geometry across $\\scri^+$, effectively gluing it with $\\scri^-$ of a contracting solution.% \\footnote{The detailed gluing procedure will involve a conformal completion along the lines of \\cite{ar} where the normal component to $\\scri$ of the metric is rescaled by a different power of the conformal factor than the tangential one.} Quantum states under consideration remain peaked at these extended space-time geometries across $\\scri$. An extension is but to be expected both in the \\WDW theory and in LQC: quantum evolution in the internal time $\\phi$ is \\emph{unitary} and $\\phi$ achieves a finite value $\\phi_o$ at $\\scri$ of the given classical solution, unitary evolution could not just stop there. What is interesting is that, \\emph{irrespective of the choice of the self-adjoint extension}, the state remains sharply peaked on the analytically extended geometry. This extension, and the ensuing nearly cyclic scenario has some similarities with Penrose's recent proposal of a cyclic conformal cosmology \\cite{rp}. However, there are also key differences. In our case, $\\scri^+$ of the expanding branch is glued to the $\\scri^-$ of the contracting branch; not to the big-bang singularity of the next `aeon'. More importantly, quantum geometry effects are crucial in LQC. In particular $\\hbar$ appears in the denominator of the expression of the maximum density $\\rcr$ whence, as one would expect, $\\rcr$ would diverge in the classical limit $\\hbar \\to 0$. Therefore, quantum effects and a non-zero $\\hbar$ play an essential role in the approximately cyclic scenario of LQC. By contrast, a central feature of the cyclic conformal cosmology paradigm is that, although one does have unboundedly large curvatures, $\\hbar$ plays no role at all in this regime. Use of the scalar field as a relational time variable played a key role throughout our analysis, both in the classical and quantum theory. What would have happened if we had made some other choice? In a recent analysis \\cite{hp-qg,hp-qc} non-rotating dust has been used in place of the scalar field. In this case, the expression (\\ref{eq:constr-quant}) of the gravitational part $\\Theta_\\Lambda$ of the Hamiltonian constraint is modified because the lapse is now tailored to proper time. In particular, the coefficient of $\\Lambda$ is now \\emph{linear} rather than \\emph{quadratic} in $v$. Consequently, the analog of $\\Theta_\\Lambda$ is now essentially self-adjoint and the LQC evolution resembles that in the $\\Lambda=0$ case \\cite{aps3}: The universe starts out with infinite volume in the distant past, collapses, undergoes a quantum bounce and then expands out to infinite volume. However, in the Planck regime, quantum matter should be described using quantum field theory and for all standard quantum fields the kinetic term in the Hamiltonian is quadratic in momenta. Therefore the specific feature that simplifies the mathematics in the case of dust is no longer available and the overall situation is then the same as that in the case of the scalar field. But what if we return to using scalar field as matter source but let a geometric variable be the relational time? An obvious choice is volume. But in LQC, volume fails to be single valued making it difficult to introduce the `time-dependent' relational Dirac observables ---such as the matter density operator $\\hat\\rho|_v$--- especially near the bounce. On the other hand, the conjugate variable $b$ is single valued both on classical and effective trajectories and, as the form of the Hamiltonian constraint suggests, the function $y$ it determines is a possible candidate for relational time. However, it appears that the evolution would then be unambiguously unitary only for $|y| >y_o$. Furthermore, defining the physical state would require specification of the initial data at $y=y_o$ but the theory does not provide any selection principle for this task. These difficulties with $v$ and $b$ could well be surmountable with new ideas and more careful analysis. But as of now they seem to be more serious handicaps than the complications associated with the use of the scalar field as relational time we encountered in this paper. Could one perhaps retain scalar field as the matter source but set lapse $N=1$ in the quantum constraint and use a `timeless framework' in LQC? Results of \\cite{klp1,klp2} imply that we would have been led to a theory that is mathematically free of ambiguities associated with self-adjoint extensions. However, as we now explain, this theory is difficult to interpret and it is unclear whether it is physically viable. We will conclude our discussion with a detailed elaboration of this point. With lapse $N=1$, the quantum Hamiltonian constraint has the form \\, $\\hat{C}= B(v) \\otimes \\partial_\\phi^2 + \\hat{C}_{\\gr}\\otimes \\id$ where $C_{\\gr}$ is the gravitational part of the Hamiltonian constraint and $B(v) \\sim 1/v$ for large values of $v$ (see, e.g., \\cite{aps3}). This total constraint operator $\\hat{C}$ has been shown to be essentially self-adjoint \\cite{kp-posL} and for simplicity we will denote its self-adjoint extension also by $\\hat{C}$. One can therefore use it directly for group averaging and construct the physical Hilbert space $\\Hil^{\\phy}$ in the `timeless framework' without recourse to deparametrization. Then, although states can be represented as wave functions $\\Psi(v,\\phi)$, the physical scalar product is no longer given by an integral over $v$ at a fixed value of $\\phi$. What is the relation between this $\\Hil^{\\phy}$ and the Hilbert spaces $\\Hil^{\\phy}_{\\beta}$ associated with self-adjoint extensions $\\Theta_{\\Lambda,\\beta}$ we constructed in this paper? It turns out \\cite{klp2} that $\\Hil^{\\phy}$ is huge; it is given by the \\emph{direct integral} of all $\\Hil^{\\phy}_{\\beta}$;\\,\\, $\\Hil^{\\phy}= \\int_{\\oplus\\, I} \\Hil^{\\phy}_{\\beta} \\, \\rd\\beta$, where $I$ is the interval $(0, \\pi)$. We will briefly discuss a simple example ---due to Wojciech Kami\\'nski in \\cite{k-pres}--- to illustrate the relation between these Hilbert spaces. Consider a 2-dimensional strip, $M= \\mathbb{R} \\times [0,1]$, with coordinates $\\phi \\in \\mathbb{R}$ and $x\\in [0,1]$ and a constraint thereon in the form of the Schr\\\"odinger equation $(-i\\partial_\\phi + i\\partial_x)\\Psi(x,\\phi) =0$ (with a first-order Hamiltonian). As in LQC, $\\phi$ plays the role of time while $x$ is to be thought of as the analog of the compactified volume coordinate $\\theta$. The operator $\\Theta_x := -i\\partial_x$ fails to be essentially self-adjoint on the closed interval $[0,1]$; it admits a 1-parameter family of self-adjoint extensions $\\Theta_{x,\\beta}$, labeled by $\\beta \\in [0, 2\\pi)$ (with domain $\\Dom_\\beta$ given by wave functions $\\psi(x)$ satisfying $\\psi(1) = e^{i\\beta}\\psi(0)$). For each extension, we can construct a physical Hilbert space $\\Hil^{\\phy}_\\beta$ in the standard manner from solutions to the quantum constraint: $\\Psi(x,\\phi) \\in \\Hil^{\\phy}_\\beta$ if and only if $-i\\partial_\\phi \\Psi(x,\\phi) = \\Theta_{x,\\beta}\\, \\Psi(x,\\phi)$. Next, let us define $\\Hil := \\int_{\\oplus\\, I} \\Hil^{\\phy}_{\\beta} \\, \\rd\\beta$, with $I= (0,2\\pi)$. ($\\Hil$ is analogous to $\\Hil^{\\phy}$ obtained by group averaging in \\cite{klp2}). Every $\\Psi_\\beta(x,\\phi) \\in \\Hil^{\\phy}_\\beta$ can be expanded as \\be \\Psi_\\beta(x,\\phi) = \\sum_{n=-\\infty}^{\\infty} \\tilde{\\psi}_{\\beta, n}\\,\\, e^{i(2\\pi n+\\beta) x}\\,\\, e^{ik\\phi} \\ee where $k:=2\\pi n+\\beta$. On the other hand $\\Psi(x,\\phi) \\in \\Hil$ has the form \\begin{equation}\\begin{split} \\Psi(x,\\phi) &= \\int_0^{2\\pi}\\!\\! \\rd\\beta\\, \\Psi_\\beta(x,\\phi) = \\sum_{n=-\\infty}^{\\infty} \\, \\int_0^{2\\pi}\\!\\rd \\beta\\, \\tilde{\\psi}_{\\beta,n}\\, e^{i(2\\pi n +\\beta) x}\\, e^{ik\\phi} \\\\ &= \\int_{-\\infty}^{\\infty}\\!\\! \\rd k\\, \\tilde\\Psi(k)\\, e^{ikx + ik\\phi}\\, , \\end{split}\\end{equation} where we have set $k = 2\\pi n +\\beta$ as above. The norms in $\\Hil^{\\phy}$ are given by: \\be \\|\\Psi(x,\\phi)\\|^2 = \\int_{-\\infty}^{\\infty}\\!\\! \\rd k\\, |\\tilde{\\Psi}(k)|^2\\, . \\ee Note that $\\tilde{\\Psi}(k)$ has support on the entire $k$-axis and furthermore, the norm of $\\Psi(x,\\phi)$ also involves the integral of $|\\tilde{\\Psi}(k)|^2$ over the entire $k$ axis. Therefore, while elements $\\Psi_\\beta(x,\\phi)$ of any one $\\Hil^{\\phy}_\\beta$ are restricted to have support only on the physical configuration space $M = \\mathbb{R}\\times [0,1]$, $\\Psi(x,\\phi)$ in $\\Hil^{\\phy}$ are allowed to be non-zero all along the $x$-axis, even though points outside the $x$-interval $[0,1]$ have no physical interpretation in the model under consideration. In other words, although we restricted ourselves only to a `strip' $M= \\mathbb{R}\\times [0,1]$ of the Minkowski space in defining the classical system, in effect $\\Hil^{\\phy}$ describes a system on the entire 2-dimensional Minkowski space $\\mathbb{R}^2$. \\footnote{Note that, since the basis functions $e^{i(2\\pi n+\\beta)x}$ can be extended analytically to the entire $x$-axis, the quantum constraint can be uniquely extended to the full Minkowski space $\\mathbb{R}^2$.} To summarize, for the particle on the strip $M$, one would physically expect that the quantum theory should be formulated just on $M$ since values of $x$ outside $[0,1]$ have no physical meaning. This expectation is borne out if one works with any one self-adjoint extension $\\Theta_{x,\\beta}$ of $\\Theta_{x}$ but not if one works with the direct integral $\\Hil^{\\phy}$ of all the resulting Hilbert spaces. In LQC, the situation is analogous. Working with a specific self-adjoint extension allows us to remain in the interval $[0, \\infty]$ of the $|v|$-axis, introduce Dirac observables and track their evolution in the internal time $\\phi$ and compare it with classical trajectories. On the other hand, working in the timeless framework in effect requires us to extend the $v$-axis beyond $v=\\infty$ and this extension is difficult to interpret physically.% \\footnote{Furthermore, unlike in the simple example discussed above, here one cannot extend the constraint operator uniquely even if one expresses the basis vectors as functions of the `compactified volume' coordinate $\\theta$ of \\eqref{eq:p-theta}, as these functions are not analytic in $\\theta=\\pi/2$.} For the same reason, while one can introduce Dirac observables also in the timeless framework, we cannot ask for their `evolution' and it is difficult to compare predictions of the quantum theory with those of the classical. In particular, while the choice $N=1$ of the lapse yields evolution in proper time in the classical theory, unfortunately this interpretation does not extend to the quantum theory in a simple way. Thus, while at first it seems mathematically natural to work with the lapse $N=1$ because the full constraint is then essentially self-adjoint, the resulting Hilbert space $\\Hil^{\\phy}$ appears to be simply too large to be physically viable in the above context. Finally, our analysis brought out an unexpected robustness of the quantum evolution with respect to the choice of self-adjoint extensions $\\Theta_{\\Lambda,\\beta}$. As we noted above, this may be related to the fact that, in the physical sector associated with \\emph{any} self-adjoint extension, under unitary evolution quantum states of interest follow the natural and unambiguous analytic extension of classical trajectories. Is this perhaps a special case of as yet unknown general result? Is the lack of sensitivity of the quantum evolution on the choice of analytic extensions have its origin in some special features of the classical evolution?" }, "1112/1112.2365_arXiv.txt": { "abstract": " ", "introduction": "Red supergiants are long--period variables with semiregular light variations on a timescale $\\gtrsim 10^2$ day. The period--luminosity relation (Glass 1979; Feast et al. 1980) and the linear theory of adiabatic oscillations (Stothers 1969, 1972) allow us to suppose that such a type of variability is due to radial stellar pulsations. At the same time together with semiregular variability some red supergiants exhibit superimposed irregular light variations (Kiss et al. 2006). The secondary stochastic variability is thought to be due to the large--scale convection in the outer subphotospheric layers (Stothers and Leung 1971; Schwarzschild 1975; Stothers 2010). Red supergiants are also remarkable due to intensive mass loss revealed through a large infrared excess indicating dust production in the stellar wind (Verhoelst et al. 2009). The period--luminosity relation of radially pulsating red supergiants is used for determination of extra--galactic distances and in comparison with Cepheids the red supergiants allow us to substantially extend the distance scale due to their higher luminosities (Pierce et al. 2000; Jurcevic et al. 2000). Application of the theory of stellar pulsation to the analysis of observed variability of red supergiants allows us to verify some conclusions of the stellar evolution theory in a way similar to that employed earlier for Cepheids. It should be noted also that the growing bulk of recent observations indicate that the strong stellar wind of massive late--type supergiants is due to nonlinear stellar oscillations (van Loon et al. 2008) The nature of radial oscillations in red supergiants is still not completely clear yet. The linear analysis of pulsational instability of red supergiants with masses $15M_\\odot\\le M\\le 30M_\\odot$ was performed by Li and Gong (1994) and Guo and Li (2002). According to their calculations radial oscillations of red supergiants are due to instability of the fundamental mode and, perhaps, the first overtone. However the theoretical period--luminosity relation agrees only with fundamental mode oscillations. Nonlinear radial oscillations of red supergiants were considered only in two studies. In the first one (Heger et al. 1997) the authors investigated radial oscillations of the red supergiant with mass $M = 15M_\\odot$ at the final stage of the core helium burning. In the second work (Yoon and Cantiello 2010) the authors investigated pulsational instability of red supergiants with masses $15M_\\odot\\le M \\le 40M_\\odot$. It should be noted that in both these studies the self--exciting stellar oscillations were treated with modified methods of stellar evolution calculation and effects of interaction between pulsation motions and turbulent convection were not taken into account. Below we present results of investigation of nonlinear pulsations of red supergiants obtained from the self--consistent solution of the equations of radiation hydrodynamics and turbulent convection. The need for such an approach is due to the significant length and mass of the outer convection zone involved in pulsation motions. The treatment of convective heat transport uses the solution of the diffusion--type equations for the enthalpy and the mean turbulent energy obtained by Kuhfu\\ss (1986) for spherically--symmetric gas flows from the Navier--Stokes equation. Thus, the results presented below deal with modelling the semiregular variability and the secondary stochastic variability is not considered because this problem is beyond the approximation of spherical geometry. We consider the stars with masses at the zero--age main sequence $8M_\\odot\\le\\mzams\\le 20M_\\odot$ and initial fractional mass abundunces of hydrogen and elements heavier than helium $X=0.7$ and $Z=0.02$. ", "conclusions": "Given in the previous section estimates of masses of seven galactic red supergiants allow us to conclude that the theory of stellar evolution is in an agreement with observational estimates of stellar radii. To compare more stars with the theoretical period--luminosity relation one should consider pulsational instability of red supergiants in the wider interval of initial masses $\\mzams$. For more detailed theoretical period--luminosity relation one should consider the role of some parameters used in evolutionary computations. One of them is the overshooting parameter. In the present study the evolutionary computations were done for the ratio of the overshooting distance to the pressure scale height $l_\\mathrm{ov}/\\hp = 0.15$. The need to know the role of this parameter is due to the dependence of the mass--luminosity relation of helium burning stars on convective overshooting. In stars with masses $M\\ge 20M_\\odot$ effects of mass loss during the red supergiant evolutionary stage become significant. In the present study the evolutionary calculations were done with mass loss rates by Nieuwenhuijzen and de Jager (1990) however determination of the mass loss rate $\\dot M$ as a function of fundamental stellar parameters remains disputable (Mauron and Josselin, 2011) Therefore, one should employ parametrization of the expression for $\\dot M$ and consider the mass--luminosity and period--luminosity relations as a function of this parameter. Another parameter which significantly affects the period--luminosity relation of red supergiants is the mass fraction abundance of heavy elements $Z$. Of special interest is the period--luminosity relation for $Z=0.008$ which is typical for the Large and Small Magellanic Clouds. \\newpage" }, "1112/1112.0010_arXiv.txt": { "abstract": "In this work I investigate the statistical properties of a huge catalog of closely interacting pairs of dark matter haloes, extracted from the Millennium Simulation database. Only haloes that reach a minimum mass $\\geq 8.6 \\times 10^{10} M_{\\odot}\\, h^{-1}$ (corresponding to 100 particles) are considered. Close pairs are selected if they come within a critical distance $d_{\\rm crit}$. I explore the effects of replacing $d_{\\rm crit}=1\\, {\\rm Mpc}\\, h^{-1} \\rightarrow 200\\, {\\rm kpc}\\, h^{-1}$ on the evolution of separations, lifetimes, total masses and mass ratios of these pairs. ", "introduction": "Mergers of galaxies play a fundamental role in essentially all modern theories of galaxy formation. They are believed to determine the morphology of a galaxy, drive its star formation and even activate its nuclear supermassive black hole. This has motivated the development of very detailed numerical simulations of merging galaxies (often involving only two galaxies in isolation). Unfortunately, with very few exceptions \\cite[e.g.,][]{tonnesen11}, these simulations typically concentrate primarily on the post-merger aftermath, often neglecting the early stages of interaction. Observations, on the other hand, tell us a different story. Surveys like the SDSS and zCOSMOS have confirmed, in large numbers, that galaxies in pairs tend to be bluer, have their star formation enhanced, and are more likely to be active \\citep[e.g.,][]{ellison11,silverman11}. Moreover, the discovery rate of binary quasars has accelerated to unprecedented levels in the last few years \\citep[e.g.,][]{liu11}. For this reason, it is vital to re-focus our attention to the early stages of galactic interactions. ", "conclusions": "Using the Millennium Simulation, I analyze the behavior of a very large set of halo pairs selected by a proximity criterion. As this critical distance is reduced, (1) more mergers are pre-selected, (2) pairs end up with more elongated orbits, (3) interactions tend to be shorter, and (4) non-mergers split more violently. Using a smaller $d_{\\rm crit}$ probes smaller scales, enhancing the importance of pairs involving a massive member and the prevalence of massive neighbors capable of splitting pairs. With these results, I wish to highlight the following important lessons: \\begin{itemize} \\item {\\bf Extra care must be taken when using close galaxy pairs as proxies for mergers. This subtlety is important even in physical three-dimensional space!} \\item {\\bf Systems with two merging galaxies in isolation are just an approximation. In reality, the universe can nurture these duos, or split them altogether!} \\end{itemize} Next I will use this catalog to explore the symbiosis between interacting galaxies and binary quasars. Ultimately, this will set up realistic initial conditions for black hole mergers. This work is just a puzzle piece of an ambitious long-term research program centered on the evolution of galaxies, supermassive black holes, and their environment." }, "1112/1112.3338_arXiv.txt": { "abstract": "{\\normalsize The search for classically stable Type IIA de-Sitter vacua typically starts with an ansatz that gives Anti-de-Sitter supersymmetric vacua and then raises the cosmological constant by modifying the compactification. As one raises the cosmological constant, the couplings typically destabilize the classically stable vacuum, so the probability that this approach will lead to a classically stable de-Sitter vacuum is Gaussianly suppressed. This suggests that classically stable de-Sitter vacua in string theory (at least in the Type IIA region), especially those with relatively high cosmological constants, are very rare. The probability that a typical de-Sitter extremum is classically stable (i.e., tachyon-free) is argued to be Gaussianly suppressed as a function of the number of moduli. } \\vspace{3cm} \\begin{flushleft} \\today \\end{flushleft} \\end{center} \\end{titlepage} \\setcounter{page}{1} \\setcounter{footnote}{0} \\tableofcontents \\parskip=5pt ", "introduction": "Recent cosmological data strongly suggests that our universe is sitting at a vacuum state with a very small positive vacuum energy density, or cosmological constant. Further cosmological data also suggests that our universe went through an inflationary epoch in its very early stage. This epoch follows from the presence of a vacuum energy density much bigger than today's value. So it is very likely that our universe started with a relatively large vacuum energy density during the inflationary epoch; it subsequently moved down a ``potential landscape\" and landed at the present small value before nucleosynthesis time. A question of great interest is whether this picture is compatible with our present understanding of string theory. In particular, we would like to examine whether the stringy cosmic landscape has features that may suggest the above cosmological scenario. Although recent studies of flux compactification in string theory suggest that there are numerous solutions to the string/supergravity equations of motion with different vacuum energies \\cite{Bousso:2000xa,Kachru:2003aw,Susskind:2003kw}, we believe that most of them are only extrema of the resulting effective potential. In fact, explicit model building shows that a meta-stable (i.e., classically stable but may have a finite decay time due to quantum tunneling) de-Sitter ($dS$) vacuum is hard to come by. In particular, the search of a single $dS$ minimum (i.e., with only semi-positive scalar field mass-squares) in Type IIA models has so far come up empty. This is somewhat discouraging as the search includes a collection of exponentially many extrema (by varying the fluxes) in Type IIA vacua. On the other hand, this result is really not that surprising from the properties of multidimensional potentials \\cite{Aazami:2005jf,Easther:2005zr}. Consider a flux compactification with $N$ moduli. An extremum will be (meta-)stable if all the scalar mass-squares are semi-positive. The axionic component of each modulus presumably has an oscillating behavior, so it will hit a minimum half the time. If we hit a maximum, we expect a nearby minimum to which the wavefunction of the universe can easily move. To simplify the discussion, we may assume that it is easy to reach an axionic minimum and we focus mostly only on the real moduli. Now (with canonical kinetic terms), a typical stringy effective potential $V(\\phi_j)$ ($j=1,2, \\cdots ,N$) of $N$ (real) moduli have non-trivial behavior. Since none (or almost none) of the moduli takes a constant value, they are expected to have some extrema too. For a complicated potential $V(\\phi_j)$ sitting at a minimum, the Hessian (i.e., the $N \\times N$ mass-squared (symmetric) matrix, or simply mass matrix) must have only semi-positive eigenvalues. However, this likelihood is very small, as first pointed out by \\cite{Aazami:2005jf}. Let ${\\cal P}$ be the probability that a given de-Sitter solution (an extremum of a positive $V(\\phi_j)$) turns out to be a meta-stable $dS$ minimum (that is, the $dS$ vacuum is tachyon-free). To avoid de-compactification, we consider only the meta-stable vacua within the finite ranges of the moduli. Suppose all the real entries in the Hessian is random, then the probability ${\\cal P}$ that it has only positive eigenvalues is roughly given by \\begin{equation} \\label{intro1} {\\cal P} \\sim e^{- \\frac{\\ln 3}{4} (N+ 0.7)^2 } \\end{equation} where $\\ln (3)/4= 0.275$ is obtained in \\cite{Dean:2006wk, Dean2008}. For a relatively large $N$, ${\\cal P}$ is Gaussian-suppressed. Even if the moduli do not couple to each other so the Hessian is diagonal, the probability ${\\cal P}$ for large $N$ is still exponentially suppressed, \\begin{equation} \\label{intro2} {\\cal P} \\sim \\left( \\frac{1}{2} \\right)^N = e^{- N {\\ln 2}} \\end{equation} where $\\ln 2= 0.693$. These probabilities apply to searches via trial and errors only. One may argue that a generic potential $V(\\phi_j)$ must hit some minima somewhere. This is certainly true. In the Type IIA cases, we see that all known (meta-stable) minima happen to have zero (Minkowski) or negative (Anti-de Sitter) vacuum energy densities, where supersymmetry as well as other symmetries help to guide the search for minima. In fact, some searches start from an $AdS$ minimum with only positive mass squares and then lift it to de-Sitter space. However, upon lifting to de-Sitter space, tachyon generically appears. The couplings among the moduli introduces off-diagonal terms in the Hessian. As the cosmological constant increases, the magnitudes of the off-diagonal terms increase as well, and that tend to cause instability. To see the impact on the stability of the vacuum due to the increase of the vacuum energy, let us start with the diagonal positive mass-square matrix $A$ for an $AdS$ vacuum. Given the masses, we can determine the variance $\\sigma_A$ of $A$. As we lift the cosmological constant, the Hessian at the extremum becomes $A+B$ where $B$ may be treated as a random matrix for a complicated generic $V(\\phi_j)$. The matrix $B$ has variance $\\sigma_B$. This allows us to define the size of the average magnitude of the off-diagonal terms relative to the diagonal mass-squared terms in terms of $y= \\sigma_B/\\sigma_A$. The parameter $y$ essentially describes how the uplifting potential impact on the stability of the vacuum. Now let \\begin{equation} {\\cal P} = a\\, e^{-b N^2 - c N} \\end{equation} where the Gaussian suppression dominates the exponential suppression when $bN/c >1$. Numerically we find that, for small $y$, \\begin{equation} \\begin{split} b&= 0.000395 y + 1.05 y^2 - 2.39 y^3,\\\\ {b\\over c}&= 0.0120+ 2.99 y - 12.2 y^2 + 1650 y^3. \\end{split} \\label{p suppressed off-diagonal comp} \\end{equation} Here we see that Gaussian suppression becomes dominant when $y \\ge 0.0241$ for $N=10$ (where $b/c \\ge 0.1$ and $Nb/c \\ge 1$) and $y \\ge 0.00269$ for $N=50$. For fixed $N$, Gaussian suppression becomes more dominant as $y$ increases. As an example, we look at a concrete search for a $dS$ vacuum starting from an $AdS$ vacuum undertaken in \\cite{Caviezel:2008tf}. In this $SU(2) \\times SU(2)$ model, there are 14 moduli (7 complex moduli), i.e., $N=14$. At the extremum with positive vacuum energy, $y \\simeq 0.274$, so $bN/c \\gg 1$. This indicates that such a search has a Gaussianly small probability of success. Sure enough, tachyon appears at this extremum. On the other hand, some generic argument suggests that $dS$ vacua exist in Type IIB models, especially when non-perturbative effects are turned on. For example, a KKLT vacuum may be obtained by uplifting an $AdS$ minimum with non-perturbative effects. Attempts to construct $dS$ vacua in Type IIA models studied so far do not include non-perturbative effects. This is because the no-scale structure is present at tree level in Type IIB, while not in Type IIA. The no-scale structure may help to have a hierarchical structure in Type IIB with sub-leading corrections. We shall discuss Type IIB models in a separate paper. Even without going into the details here, we shall use the simple belief that there are metastable $dS$ vacua present in string theory, even though they may be very rare. Suppose all potentials of $dS$ vacua may be treated as ``uplifts\" of $AdS$ vacua with semi-positive mass-squares. A small uplifting will introduce relatively small off-diagonal terms into the Hessian so the chance of being a $dS$ minimum is relatively good. As we increase the uplifting to higher vacuum energy densities, the off-diagonal terms in the Hessian increase accordingly and the Hessian becomes complicated. The off-diagonal terms tend to push the lowest eigenvalues to negative values. So the chance of this being a $dS$ minimum becomes Gaussianly small (i.e., (\\ref{intro1})) as we go to higher cosmological constants. This leads to the conjecture that there are essentially no $dS$ minima in the relatively higher CC regions in the Type IIA cosmic landscape. The message leads to the following proposal : \\begin{quote} {\\it Raising the cosmological constant destabilizes the classically stable vacua}. \\end{quote} This suggests that, as the universe evolves down the potential, it encounters no $dS$ minima along its way for relatively large cosmological constant (CC) values. Presumably, this happened during the inflationary epoch. Towards the end of inflation, when the universe reaches regions with sufficiently small CC, there may be some $dS$ minima around for the universe to be trapped in one of them. That is, the percolation probability is of order unity for high CC, but decreases substantially by the time the universe reaches the small CC region. It is very likely that there are many more $AdS$ vacua than $dS$ minima around, but since the universe starts from a relatively high CC region (to generate enough inflation), it has to go through the small positive CC region before reaching the negative CC region. It is not unreasonable that it becomes trapped in a low $dS$ minimum on its way towards negative CC region if the probability of finding a $dS$ minimum with a small CC is not too suppressed. As we shall see, it is reasonable to expect that, for the small positive CC region, some of the moduli with large masses essentially decouple, thus reducing $N$ to a smaller effective value, so the probability of finding a low $dS$ minimum may not be exponentially suppressed compared to the number of $AdS$ vacua nearby. Since no classically stable Type IIA $dS$ vacuum has been found so far, the above proposal cannot be (non-trivially) checked at the moment. However, we do know some Type IIB solutions; so presumably such a check can be performed for Type IIB regions of the landscape. This study (which involves some subtleties) is under way. This paper is organized as follows. In section \\ref{sec:prob-estim-rand}, we review and discuss the properties of a real symmetric random matrix as a typical example of a Hessian. We then discuss a Hessian that is the sum of a random diagonal positive matrix plus a random symmetric matrix. This mimics the Hessian that generically appears in the search for a $dS$ minimum in Type IIA models. In Sec. 3, we consider a few examples to illustrate the main point of this paper. Sec. 4 contains some discussions. In particular, we shall comment on an earlier estimate of the probability of obtaining $dS$ vacua in \\cite{Denef:2004cf}. Some details are relegated to the appendix. ", "conclusions": "In this work, we use random matrix theory to estimate the probability of metastable $dS$ vacua in IIA string theory. Our purpose here is to quantify how unlikely we can stabilize all of the K\\\"ahler moduli and the complex structure moduli simultaneously at a $dS$ minimum of a tree level potential. Including quantum effects will simply yield a more complicated potential, which in turn tends to improve the randomness approximation. Assuming the possibility of introducing some hierarchies in the mass matrix in this setup, we expect different probabilities for different situations. Relatively large off-diagonal terms tend to make the probability ${\\cal P}$ of locating a meta-stable $dS$ vacuum more Gaussianly suppressed than exponentially suppressed. In the IIA models, we explained why tachyons are ubiquitous in meta-stable de-Sitter extrema in terms of the appearance of the off-diagonal terms in the Hessian. An increase in the cosmological constant typically destabilizes the vacuum as a result of the increase of off-diagonal components. Note that, in the search for meta-stable $dS$ vacua, the potential used (see (\\ref{potentials in two moduli}) for example) in the search typically introduces multiple free parameters. In attempts to find meta-stable $dS$ vacua, the search allows any appropriate values for these parameters. In reality, these parameters and other hidden ones must be dynamically determined. In the present searches, it is assumed that the modes associated with these parameters (there can be a number of modes associated with each parameter) are heavy and have been dynamically stabilized already. Furthermore, they have enough multiple stable solutions so that their values needed for $dS$ vacuum stability will be among the possible stabilized solutions. This may be a reasonable assumption for low energy scales. However, as the vacuum energy scale is increased to scales comparable to some of lighter masses of these modes, we can no longer ignore the dynamics of these modes. For example, consider the supergravity potential (\\ref{SUGRAV}) (for simplicity without the D term). Suppose one is allowed to increase the energy scale simply by adding a positive constant to $K$, without changing the spectrum. By doing so, we may have to include the lighter part of heavy moduli into the potential that have not been included in $K$ and $W$ so far. The resulting relevant Hessian with increased number of moduli will not be simple in general. So the typical relevant Hessian will grow in size and complication as the vacuum energy is increased. Even though changing the overall scale does not alter the hierarchy between the diagonal and off-diagonal entries in the original mass matrix (which is a subset of the entire mass matrix), the increase of the number of relevant moduli increases the likelihood of instability as we argued in section \\ref{sec:random-matrix-with}. The implication of this viewpoint is as follows: As the vacuum energy scale increases, the number of moduli to be included in the analysis of the relevant mass-squared matrix increases. As a result, the unavoidably complicated form of the matrix will imply that the probability of having a meta-stable $dS$ vacuum becomes increasingly Gaussianly unlikely. This reinforces our view that: \\begin{quote} {\\it The probability of finding a meta-stable de-Sitter vacuum in the string landscape becomes asymptotically Gaussianly unlikely as the vacuum energy is increased.} \\end{quote} Since not a single classically stable Type IIA $dS$ vacuum has been found so far, this proposal/conjecture is compatible with this fact, but at the same time, we cannot check the proposal/conjecture in a non-trivial way. On the other hand, we do know some Type IIB approximate solutions; so the above proposal may be checked (to some extent) for the Type IIB regions of the landscape. However, all known meta-stable Type IIB vacua involve non-perturbative contributions to the 4-dimensional effective potential, thus the analysis is somewhat more subtle. The probability of obtaining a $dS$ vacuum in string theory was previously considered in \\cite{Denef:2004cf}. In this seminal work, the authors estimated such probability in the context of ${\\cal N}=1$ supergravity (and for simplicity, they considered cases where the D-terms are absent). To maintain stability, the authors required the off-diagonal terms to be suppressed relative to the positive diagonal terms. Since this potential is a function of the superpotential $W$ and its derivatives, such off-diagonal suppression is achieved if we satisfy the condition $|D_A D_B D_C W \\bar{\\psi}_1^A \\bar{\\psi}_1^B \\bar{\\psi}_1^C| < {\\cal O}(DW)$ where the vector $\\psi_1$ specifies a direction for extrema with the smallest eigenvalue, under the assumption of $|DW| \\sim |W|$ (in Planck unit). Since a term proportional to $D_A D_B D_C W$ shows up in the off-diagonal components of the Hermitian mass matrix, such a constraint provides a hierarchy between the diagonal and off-diagonal entries. Now, we have shown that the probability of stability is Gaussianly suppressed (as a function of the number of moduli) when the off-diagonal terms are within an order of magnitude of the diagonal terms; and we typically expect one of the lightest modes to turn tachyonic first. This implies that the above condition will be Gaussianly unlikely to satisfy as the number of moduli increases. Since the Hessian is built from $W$ and its derivatives, the elements in the Hessian are not totally random. However, for a complicated system with a large number of moduli, we believe that some form of the central limit theorem should hold and our approach is valid. While this paper was in preparation, we were informed that a more detailed model, based on \\cite{Denef:2004cf} using some random distributions of $W$ and its derivatives has recently been performed \\cite{Marsh:2011aa}. Furthermore, in our setup, we had in mind a much wider class of constructions, some of which cannot be written in the form of an ${\\cal N}=1$ supergravity potential without $D$ terms; in these cases, our approach should provide a good estimates of the probability ${\\cal P}$ when the hierarchy between the diagonal and the off-diagonal components of the Hessian can be approximately estimated. We will report on our findings for the probability of metastable $dS$ vacua in IIB string theory in a forthcoming paper." }, "1112/1112.1084_arXiv.txt": { "abstract": "We have obtained long-slit spectra of 174 star-forming galaxies with stellar masses greater than $10^{10} M_{\\odot}$ from the {\\it GALEX} Arecibo SDSS (GASS) survey. These galaxies have both HI and H$_2$ mass measurements. The average metallicity profile is strikingly flat out to $R_{90}$, the radius enclosing 90\\% of the $r$-band light. Metallicity profiles which decline steadily with radius are found primarily for galaxies in our sample with low stellar mass (Log$(M_*)<10.2$), concentration, and/or mean stellar mass density. Beyond $\\sim R_{90}$, however, around 10 percent of the galaxies in our sample exhibit a sharp downturn in metallicity. Remarkably, we find that the magnitude of the outer metallicity drop is well correlated with the {\\it total} HI content of the galaxy (measured as $f_{HI}=M_{HI}/M_*$). We examine the radial profiles of stellar population ages and star formation rate densities, and conclude that the galaxies with largest outer metallicity drops are actively growing their stellar disks, with mass doubling times across the whole disk only one third as long as a typical GASS galaxy. We also describe a correlation between {\\em local} stellar mass density and metallicity, which is valid across all galaxies in our sample. We argue that much of the recent stellar mass growth at the edges of these galaxies can be linked to the accretion or radial transport of relatively pristine gas from beyond the galaxies' stellar disks. ", "introduction": "A proper characterization of the ages and metallicities of the stars in spiral galaxies, as well as the radial dependence of the metallicity of stars and gas in these systems, has long been recognized as a key stepping-stone to unravelling disk galaxy formation processes, including the roles of gas accretion, supernovae-driven outflows and the radial migration of stars \\citep{quirk73, tinsley80, lacey85, wyse89, kauffmann96, chiappini97, schonrich09}. There have, however, been rather few systematic studies of how radial star formation and metal abundance gradients vary across {\\em populations} of disk galaxies. \\citet{vilacostas92} carried out an analysis of abundance gradients in a sample of 30 spiral galaxies with spectroscopy from the literature. Barred galaxies were found to have flatter abundance gradients than un-barred galaxies. When the analysis was restricted to un-barred galaxies, gradients were found to be stronger in late-type galaxies (see also \\citep[see also][]{oey93}. Vila-Costas \\& Edmunds also found the the central metallicities of spiral galaxies were correlated with both the total stellar mass of the galaxy, and the local central surface density of stars. Zaritsky, Kennicutt \\& Huchra (1994) obtained uniform HII-region based abundance gradients for a sample of 39 nearby disk galaxies. In this study, the slopes of radial abundance gradients, when expressed in units of dex/isophotal radius, did not exhibit any significant correlation with luminosity or Hubble type, but it was later noted by \\citet{garnett97} that gradients expressed in dex/kpc steepen with decreasing luminosity. Further observations by \\citet{vanzee98} focused on increasing the number of measurements in the outer regions of galaxies; they identified several galaxies where the fitted gradient changed significantly with the addition of higher-radius data. Recently, \\citet{moustakas10} presented resolved metallicity measurements for a sample of nearby galaxies from the {\\it Spitzer} Nearby Galaxies Survey \\citep[SINGS,][]{kennicutt03a}, and pointed out that derived metallicity gradients can be sensitive to the methodology used to calibrate the strong-line abundance estimates. A complete, systematic and carefully-executed survey of metallicity gradients in a large sample of nearby disk galaxies is urgently needed to help put into context some recent studies that have reported unexpected or peculiar metallicity profiles in a number of individual galaxies. A number of these \\citep[e.g.,][]{werk11, rupke10, kewley10, bresolin09} find surprisingly flat metallicity gradients, even out to extreme distances, both for interacting galaxies and peculiarly gas-rich objects, in contrast with earlier work documenting low metallicities or large drops in the extreme outskirts of galaxies \\citep[e.g.,][]{ferguson98, kennicutt03}. The more recent papers argue that these differences reflect the important role of gas inflows and outflows, as well as mixing, in determining how metallicity changes as a function of radius in galaxies. In order to learn more about the relations between cold gas and star formation in nearby galaxies, we are carrying out the {\\it GALEX} Arecibo SDSS Survey (GASS)\\footnote{http://www.mpa-garching.mpg.de/GASS} \\citep[][hereafter C10]{catinella10}. GASS is designed to measure the neutral hydrogen content of a representative sample of $\\sim 1000$ galaxies uniformly selected from the Sloan Digital Sky Survey \\citep[SDSS,][]{york00} and {\\it Galaxy Evolution Explorer} ({\\it GALEX}, Martin et al. 2005) imaging survey, with stellar masses in the range $10^{10}-10^{11.5} {\\rm M}_{\\odot}$ and redshifts in the range $0.0250.7R_{90}$, O3N2 scale) for each galaxy. We measure a correlation coefficient $|\\rho|=0.53$, significant at the $5\\sigma$ level. We include upper limits as if they were detections for this calculation, but excluding these limits---either in $f_{HI}$, metallicity, or both---does not significantly change the result. The C10 HI plane can be used to predict $f_{HI}$ with rms accuracy of $\\sim0.3$~dex, and for comparison the scatter in our outer metallicity-$f_{HI}$ relation is 0.4~dex in $f_{HI}$. We note, however, that outer metallicities that are near solar provide little predictive power for HI. In contrast, low metallicity points {\\it do} seem to be universally associated with gas-rich galaxies. Whether or not Figure~\\ref{FP} reflects a {\\it continuous} relation between outer metallicity and $f_{HI}$, or whether it is more properly described as a {\\it threshold} in $f_{HI}$, above which outer metallicities are suppressed in proportion to the gas content, is unclear from the current data. We hope to revisit this issue with the full GASS data-set once it is in hand. The quantity plotted in the left panel of Figure~\\ref{FP}, being defined as the lowest metallicity we measure in a galaxy's outskirts, could be subject to a number of biases due to the imprecise definition. To check whether the way we define the metallicity value is driving this relation in any way, we also calculated metallicity values by coadding all spectroscopic flux from $R>R_{90}$ into a single high-S/N spectrum for each galaxy. Metallicities measured from these spectra, being integrated over a much larger area of the galaxy, often do not reach such low values as our more finely sampled points. Even so, when plotting these metallicity values against $f_{HI}$, in Figure~\\ref{FP}, right panel, we find a nearly identical relation as before, with a marginally higher $|\\rho|=0.57$ and $6\\sigma$ significance. We note, also, that plotting the size of the outer metallicity {\\it drop}, rather than the outer metallicity directly, gives largely the same results. Considering that the inner metallicities of most GASS galaxies are nearly solar everywhere, this is not surprising. To understand what might be driving this peculiar relation between a very {\\it local} quantity, the outer metallicity, and the {\\it global} gas content, we must examine how outer metallicity relates to other global galaxy characteristics. In Figure~\\ref{global_plots}, we plot the minimum outer metallicity on the O3N2 scale (as in the left panel of Figure~\\ref{FP}), as a function of several such quantities. The first quantity of interest is the molecular gas fraction, $f_{H2}$. As can be seen in the lower-left panel, $f_{H2}$ by itself does not correlate with outer metallicity. Since H$_2$ is generally more centrally concentrated in galaxies than HI \\citep[e.g.,][]{leroy09}, this result may not be surprising. In fact, we do detect a weak correlation ($|\\rho|=0.39$, $3.9\\sigma$) between $f_{H2}$ and {\\it central} metallicity. The outer metallicity may simply be more closely associated with the size of a galaxy's gas {\\it reservoir}, as traced by HI, rather than by the amount of currently star-forming gas, traced by H$_2$. Interestingly, the {\\it ratio} of molecular to atomic gas, shown in the top left panel, {\\it does} correlate with outer metallicity, in the sense that galaxies more dominated by their HI exhibit lower outer metallicity. However, this correlation is weaker ($|\\rho|=0.44$) and less statistically significant ($3.7\\sigma$) than the HI relation in Figure~\\ref{FP}, and so we may only be seeing here a reflection of the underlying relation between HI and outer metallicity. \\begin{figure*}[t] \\includegraphics[width=\\columnwidth]{f10a.eps} \\includegraphics[width=\\columnwidth]{f10b.eps} \\caption{\\label{metal_gradients} Left: Metallicity versus $R/R_{90}$ for the 13 galaxies with the largest outer metal drops. Right: Metallicity versus $R/R_{90}$ for the matched comparison galaxies. In both panels, solid lines connect adjacent measurements for each individual galaxy, and both halves of each galaxy are plotted separately, `folded' onto the same $R/R_{90}$ scale. Points and lines are color coded for galaxy HI content, according to the color bar at top, which indicates $Log(f_{HI})$ associated with each shade. Galaxies that do not yet have an HI measurement are plotted in black.} \\end{figure*} In the upper middle panel of Figure~\\ref{global_plots}, we plot outer metallicity as a function of the same combination of global NUV--{\\it r} and $\\mu_*$ that goes into the C10 HI plane--i.e., the x-axis can be thought of as the $f_{HI}$ that is predicted given the galaxy's NUV--{\\it r} and $\\mu_*$. We do this to evaluate the possibility that the $f_{HI}$--metallicity relation is simply an induced correlation: if it is really the star formation rates and stellar mass densities that set metallicity, as our results above seem to suggest is true locally, it may be that the correlation with HI content is not fundamental, but instead driven by the dependence of HI on the same underlying parameters (sSFR and $\\mu_*$, but galaxy-averaged now) that drive metallicity. What the upper middle panel of Figure~\\ref{global_plots} shows, however, is that the correlation of outer metallicity with global $\\mu_*$ and NUV--{\\it r} is relatively weak compared to Figure~\\ref{FP}, with $|\\rho|=0.37$ and a significance of $4\\sigma$. Since C10 found the relation between $f_{HI}$ and NUV--{\\it r} color alone was nearly as tight as the one including $\\mu_*$, we also plot in the upper right panel outer metallicity versus NUV--{\\it r}. The correlation here is more significant, at 4.8$\\sigma$, but again the correlation coefficient is not quite as strong ($|\\rho|=0.44$) as we saw in Figure~\\ref{FP}. The {\\it residuals} from the C10 HI plane are useful for identifying galaxies that have either abnormally high or low atomic gas content, compared to that expected for galaxies with their specific combination of $\\mu_*$ and NUV--{\\it r}. In the lower middle panel of Figure~\\ref{global_plots}, we plot these residuals versus outer metallicity. The correlation is quite low and only marginally significant ($2.8\\sigma$), suggesting that it is the absolute level of HI content that matters, rather than the {\\it excess} or {\\it deficit} of HI compared to other similar galaxies. Finally, we checked for relations between metallicity and other global galaxy properties such as concentration, $\\mu_*$, and $M_*$. None show any significant correlation; the lower right panel of Figure~\\ref{global_plots} shows metallicity versus Log$(M_*)$, which is representative of the others. Outer metallicity thus appears to depend hardly at all on the structural properties of galaxies in GASS. \\begin{figure*} \\centering \\includegraphics[width=2\\columnwidth]{f11a.eps} \\medskip \\includegraphics[width=2\\columnwidth]{f11b.eps} \\caption{\\label{low_stamps}SDSS postage stamp images of galaxies with the steepest outer metallicity drops (top), and the matched control sample selected without regard for metallicity (bottom). Images are $80\\arcsec\\times80\\arcsec$, and are arranged so that the match to a galaxy in the top panel is in the same relative position in the lower panel. White bars in each image indicate a projected physical length of 10kpc at the redshift of that galaxy.} \\end{figure*} In short, we have found that the relation between $f_{HI}$, a global property of the galaxy, and outer metallicity $12+Log(O/H)$, a very local one, is tighter and therefore likely {\\it more} fundamental than any relation between global SFR/$\\mu_*$ and metallicity. The question, then, is why this should be true. One possibility is that metallicity at the outer edge of the star-forming disk is simply a sensitive `thermometer' of sorts for measuring the amount of new gas accreting onto the disk of the galaxy, or perhaps for the rate that existing gas is transported inward. Lower metallicities would simply be reflecting a higher proportion of pristine or relatively unenriched gas residing in or flowing through the outer stellar disk. If this is correct, then varying rates of gas accretion/flow may leave other signatures in, for example, the stellar populations and star formation rates of galaxies, or on their radial gradients. To evaluate this possibility, in the following section we will examine in detail the subset of galaxies exhibiting the strongest metallicity drops in our sample, and compare to a control sample selected without regard to outer metallicity. \\subsection{The Galaxies with Steepest Metallicity Drops} In order to determine whether a low outer metallicity is associated with any distinctive features in the star formation rates and histories of our galaxies, either at their outer edges or across their disks, we select for detailed study a small sample containing only those galaxies with the steepest drops in outer metallicity. We will refer to this as the `low metallicity' or `large drop' sample, and we will compare it to a `control' sample selected without regard to metallicity, but matched in global characteristics one-for-one with galaxies in the large-drop sample. We note that both samples are selected from the subset of 119 galaxies that exhibit measurable star formation in their outskirts at $R>0.7R_{90}$, which ensures that all the galaxies in the control sample have significant star formation and measurable metallicities at the same large radii as the large-drop galaxies. In total, we identify 13 galaxies with outer-disk metallicities $12+Log(O/H)<8.4$, a threshold we chose both because no measured point in the inner region of any galaxy reaches this low, and because it is significantly below solar at the $>3\\sigma$ level (0.25~dex). Such galaxies make up about 10\\% of the total sample, and so even though the number identified is small, they represent a significant proportion of all massive star-forming galaxies. For the control sample, we select each matched counterpart by requiring that it be within 0.2~dex in stellar mass, 0.4~dex in global $\\mu_*$, and 0.3 in NUV-{\\it r} color, similar to the procedure in \\citet{wang11a}. These limits were chosen to ensure that each large-drop galaxy has at least one match within our spectroscopic sample, and for cases with more than one match we select randomly among the choices. In Figure~\\ref{metal_gradients}, in the left-hand panel we plot the radial metallicity profiles of all 13 of our low-metallicity galaxies, as a function of $R/R_{90}$. Solid lines connect adjacent points to more easily follow the two folded halves of each profile (i.e., from both sides of each galaxy). Points are color coded for galaxy HI content, as illustrated by the color bar. It is clear from this plot that even those galaxies with the strongest metallicity drops show typically flat profiles in their inner regions (though some variation can, indeed, be seen). The median difference between the inner metallicity and the lowest measured point is $\\sim0.3$ dex for these objects. In the right-hand panel of Figure~\\ref{metal_gradients}, we now plot the metallicity profiles of the control sample. Again, we see quite flat inner profiles, though there may be a hint that even these galaxies show subtle drops in metallicity at the highest measured points. Consistent with the overall correlation shown in Figure~\\ref{FP}, galaxies with large metallicity drops have on average higher $f_{HI}$, which can be seen by noting the quite different range of colors between the two panels. \\subsubsection{Morphologies} Our low-metallicity and control samples of galaxies by design have certain features in common, in particular their stellar masses and colors, as well as widespread star-formation extending out to $R>R_{90}$. Yet the disparity in the magnitude of the metallicity drop, as well as the wide variation in HI content, suggests that it is worthwhile to examine the images of both samples directly, to check for any subtle structural differences between the two that are not apparent when considering just the broadest characteristics. In Figure~\\ref{low_stamps}, we show SDSS postage stamp images of these galaxies, with large-drop galaxies in the top panel, and the control galaxies at bottom. Images are arranged so that each matched pair of galaxies is in the same relative location in the top panel and bottom. We remind the reader that galaxy pairs were not matched in redshift (beyond the normal GASS constraint of $0.0250.5$ vs $1.4R_{90}$ for $b/a<0.5$), but it is unclear if this difference is large enough to cause the observed bias. The median radius reached for the control sample galaxies is only slightly lower than the large drop galaxies: $1.2R_{90}$ and $1.3R_{90}$, respectively. A few galaxies in Figure~\\ref{low_stamps} seem to harbor close-in tidally disturbed companions, which could suggest that new gas accreted or cannibalized from these companions is the source of the low-metallicity material. However, the majority of the large-drop galaxies do not have obvious companions, and the frequency of companions is not obviously larger than that seen in the control sample images. Indeed, the one large-drop galaxy that we have previously studied in detail \\citep{moran10}, shows remarkably little evidence for any sort of dynamical disturbance, such as one would expect in the case of an accreted companion. Though beyond the scope of this paper, rotation curves are available for all of these galaxies, and we expect to address the relation between gas content and galaxy dynamics and/or mergers in a future paper. We measured a number of quantitative morphological parameters for both sets of galaxies to search for any differences. These included asymmetry, bar fraction, and an asymmetry variant weighted to the outer disk \\citep{wang11b}. Though such small samples make it difficult to assess any statistical difference between the two sets of galaxies, we find no clear evidence that the populations are different in any of the key morphological measures. This remains true whether or not we exclude the most edge-on galaxies (which could have peculiar values of these morphological parameters). Besides the obvious difference in orientation, then, we find no dramatic differences in appearance between the large-drop and control sample galaxies in Figure~\\ref{low_stamps}. Below, we will examine the radial profiles of quantities detailing the star formation histories and current star formation rates as a function of radius across each galaxy type. \\subsubsection{Star Formation Histories} So far, then, our large-drop sample of galaxies appears to differ from the control sample only in two important properties: metallicity drop and HI content. But the key question remains: why does a high HI content appear to drive low-metallicity star formation in galaxies' outskirts? To help answer this question, we can look to the star formation rates and histories of our two groups of galaxies. Specifically, we can examine the radial profiles of a number of spectroscopic diagnostics of both past-averaged and present-day star formation. \\begin{figure*}[t] \\includegraphics[width=0.66\\columnwidth]{f12a.eps} \\includegraphics[width=0.66\\columnwidth]{f12b.eps} \\includegraphics[width=0.66\\columnwidth]{f12c.eps} \\\\ \\includegraphics[width=0.66\\columnwidth]{f12d.eps} \\includegraphics[width=0.66\\columnwidth]{f12e.eps} \\includegraphics[width=0.66\\columnwidth]{f12f.eps} \\caption{\\label{other_gradients} From left to right, $D4000_n$, SFR density, and sSFR gradients as a function of $R/R_{90}$, for galaxies in the large-metallicity-drop sample (top row), and those in the control sample (bottom row). Grey points show gradients with adjacent points connected by lines, as in Figure~10. Thick black points with error bars show the mean and 1-sigma values for bins of $\\sim25$ points each. Light grey shading shows the $\\pm1-\\sigma$ region occupied by the full GASS sample with extended star formation, divided into radial bins of $\\sim150$ points each, and including only star-forming regions for $D4000_n$. We plot $D4000_n$ only for points with continuum S/N greater than 5 per angstrom, and uncertainty in $D4000_n$ less than 0.1. Typical uncertainties are less than 0.05.} \\end{figure*} In Figure~\\ref{other_gradients}, we plot, from left to right, the $D4000_n$ index, SFR density (in $M_\\odot$~yr$^{-1}$~kpc$^{-2}$), and sSFR (yr$^{-1}$) as a function of $R/R_{90}$, for both the galaxies with large metallicity drops (top), and the control sample (bottom). As in Figure~\\ref{metal_gradients}, grey lines and dots connect the points for individual galaxies. Black dots with error bars show the average and scatter in bins of radius, for all the galaxies in each subsample. Light grey shading denotes the $\\pm1\\sigma$ region occupied by our {\\it entire} 119-galaxy GASS sample with extended star formation, as a function of radius. We note that $D4000_n$ averages for the full sample are calculated only from points with detected star-formation, to ensure the curves reflect the same set of galaxies (and individual points) as those in the other two panels. The top panels of Figure~\\ref{other_gradients} show that, for the large-drop galaxies, the radial profiles of all three quantities show rather similar trends from galaxy to galaxy. $D4000_n$ generally decreases monotonically from center to outskirts (signifying decreasing stellar population ages), specific star formation rate {\\it increases} monotonically towards the outside, and SFR density seems to exhibit largely flat radial profiles for most galaxies (though the precise level of SFR density varies considerably). Moreover, it is clear by comparing to the underlying shaded region marking the full GASS sample, that, while the trends for decreasing age and increasing sSFR with radius are typical of all star-forming galaxies in this mass range, large-drop galaxies appear overall younger and are building up stellar mass faster than the typical GASS galaxy. Large-drop galaxies have $D4000_n$ shifted an average of 0.16 lower than the full-sample mean, and sSFR is fully 0.5~dex higher, which translates to mass-doubling times three times shorter---as low as 1~Gyr at the galaxy edges in the most extreme cases. Both of these differences are statistically significant in all radial bins. SFR density is also higher by an average of 0.27~dex, though the difference in the innermost bin is not significant. We note that all three plots for the large-drop galaxies seem remarkably similar to the detailed properties of UGC8802 described in \\citet{moran10}, where we argued that the combination of flat SFR density profile and declining $D4000_n$ profile could easily be replicated by a toy model featuring a recent episode of constant star formation spread evenly across the galaxy, on top of an older stellar population that built up most of the pre-existing stellar mass at an earlier time ($>1-2$ Gyr ago). The fact that $D4000_n$ indicates stellar populations that are everywhere younger than the full-sample average, while sSFR is everywhere higher than the full-sample average, lends support to this scenario. One of the most striking features in Figure~\\ref{other_gradients} is the large difference in scatter between the large-drop and control samples, visible in all three measured quantities. In contrast to the fairly uniform profiles of the large-drop galaxies, those of the control sample are quite heterogeneous. For each of $D4000_n$, SFR density, and sSFR, the formal rms scatter at every radius bin in the large-drop sample is very nearly half what we measure in the corresponding control sample bin (with the precise ratio ranging from 50\\% to 65\\%). The level of heterogeneity displayed by the control sample is essentially indistinguishable from that of the full GASS sample, and the two samples have the same mean values as well, with one exception discussed below. Thus, the spread in star formation rates and stellar population ages of the control sample can be considered typical for star-forming galaxies in this mass range, while those for the large-drop sample are abnormally uniform and offset. Only in the outermost bin (or two bins, for sSFR) do the mean properties of the control sample of galaxies differ significantly from those of the full GASS sample. In each measured parameter, the mean value of this outermost bin (or two) is instead consistent with that of the large-drop sample. In other words, the high sSFRs and young ages at the very edges of our big-drop galaxies are indistinguishable from those of other galaxies selected to have the same global NUV-{\\it r}, $M_*$, and $\\mu_*$, {\\it even though the metallicities are dramatically different.} What this means is that we cannot definitively link the high sSFRs and young ages to the high HI content and/or low metallicities at these locations, because we cannot exclude the possibility that these features are generic for galaxies with this combination of global properties. In contrast, the uniformly younger and more vigorously star-forming profiles at lower radii can perhaps {\\it only} be attributed to the high HI and/or low outer metallicities, since all the other parameters are identical between the two samples. Can these two statements, seemingly at odds, be tied together into a uniform picture of what is happening in the high HI, large-metallicity drop galaxies? We believe the following scenario is plausible: \\bigskip In galaxies with a large HI content, it has been shown that much of the gas often resides beyond the optical disk of the galaxy \\citep[e.g.,][]{bigiel08}. During the process of galaxy growth, much of this gas must eventually be transported inward, where it will form molecular gas and then stars. At the same time, it has been known for a long time that radial flows of gas can naturally lead to metallicity gradients \\citep[e.g.,][]{lacey85}, and that an `inside--out' buildup of galaxy disks---like we see in the sSFR profiles of big drop galaxies---may be required to explain the observed strengths of gradients \\citep{boissier00}. We speculate that it is this transport of gas inward that sets the level of metallicity suppression at the outer edges of our galaxies: star-forming gas at the optical edge of the galaxy is diluted in simple proportion to the total amount of gas residing in the extended reservoir, and the continuous flow of such gas serves to {\\it keep} the metallicity low. Under this scenario, the homogeneity of the radial profiles in the `large-drop' sample might arise because the star formation across the entire galaxy becomes dominated by the dynamics of this inwardly-transported gas. A dense flow of gas, by providing ample fuel to all corners of the galaxy, could well act to suppress the normal spatial variations in star formation rate seen in more typical galaxies, and at the same time cause a period of intense disk-building that elevates sSFRs everywhere, including at the outskirts where the disk is building fastest. In the control sample, small amounts of gas may also be building up the disk outskirts, but with a more-nearly-complete metal enrichment (or, rather, a less effective dilution of metals). In these systems, however, the lower quantities of gas involved are not enough to homogenize the star formation rates or stellar population age gradients along the lines seen for the large-drop objects. Let us return briefly to the question of flat inner metallicity gradients, because it is puzzling why gradients should `saturate' and flatten out so effectively in our high-mass galaxies. Generically, our observational result implies that the new metals produced by star formation are everywhere (except at the outskirts) precisely counterbalanced by either the net inflow of gas, or outflowing winds, or both. Since inner metallicity profiles are flat for galaxies with both high and low $f_{HI}$ (Figure~\\ref{metal_gradients}), this balance presumably holds over a range of gas densities. Such an equilibrium in metallicity may seem implausible, but on the other hand could just be the latest of many such observational `conspiracies' in the properties of galaxies. In any case, it appears that models of the Milky Way with varying assumptions about cosmological gas infall and the details of `inside--out' formation can produce a range of flat to sloping inner gradients \\citep[][and references therein]{colavitti09}. A full comparison to models will not be undertaken here, but it is important to recognize that at least some models {\\it can} reproduce both the flat inner gradients and the steep outer drops we see. To recap, we have clearly shown that galaxies with large outer metal drops not only have high HI content, but show evidence that this gas is currently involved in a substantial and widespread disk-building phase. Though the fates of these galaxies as their gas depletes and metallicities rise is still unclear, it does seem to be the case that a low outer metallicity is a sensitive signpost or thermometer for the presence of large amounts of active gas in a galaxy. Given that 10\\% of our overall GASS sample appears to be in this phase actively building disks, even though GASS probes a stellar mass range where such activity is thought to be {\\it decreasing}, it will be interesting to see whether the abundance of these systems is compatible with theoretical models of galaxy formation and evolution." }, "1112/1112.4497_arXiv.txt": { "abstract": "Using the Inamori Magellan Areal Camera and Spectrograph (IMACS) integral-field unit (IFU) on the 6.5\\,m Magellan telescope, we have designed the first statistically significant investigation of the two-dimensional distribution and kinematics of ionised gas and stars in the central kiloparsec regions of a well-matched sample of Seyfert and inactive control galaxies selected from the Sloan Digital Sky Survey. The goals of the project are to use the fine spatial sampling (0.2 arcsec~pixel$^{-1}$) and large wavelength coverage (4000--7000\\AA) of the IMACS-IFU to search for dynamical triggers of nuclear activity in the central region where active galactic nucleus (AGN) activity and dynamical timescales become comparable, to identify and assess the impact of AGN-driven outflows on the host galaxy and to provide a definitive sample of local galaxy kinematics for comparison with future three-dimensional kinematic studies of high-redshift systems. In this paper, we provide the first detailed description of the procedure to reduce and calibrate data from the IMACS-IFU in `long-mode' to obtain two-dimensional maps of the distribution and kinematics of ionised gas and stars. The sample selection criteria are presented, observing strategy described and resulting maps of the sample galaxies presented along with a description of the observed properties of each galaxy and the overall observed properties of the sample. ", "introduction": "Supermassive black holes are widely accepted to lie at the centres of all bulge-dominated galaxies, with the observed correlation between black hole mass and bulge velocity dispersion evident in both active and inactive galaxies confirming a causal link between previous phases of accretion-driven nuclear activity in the galactic life cycle \\citep{geb00,merr01}. Although, such active galactic nuclei (AGN) are recognised to be integral to galaxy formation and evolution, the AGN triggering and fuelling mechanisms remain unknown. Further, the transportation of fuel to the AGN and any corresponding ejection or feedback are key components of cosmological models and are tuned to deliver observed properties of today's galaxies (e.g., \\citealt{spring05}). Merger-driven galaxy evolution and nuclear activity is central to these models but is unlikely to be a valid mechanism in the local Universe, where the merger rate has declined and the $\\sim$20\\% of galaxies currently exhibiting nuclear activity \\citep{gould10} are commonly found in early type spiral galaxies in non-cluster environments and do not show signatures of strong interactions \\citep{west07,gabor09}. Identifying the triggering and fuelling mechanism and the origin and transportation of fuel in AGN therefore remain an important goal in astronomy. Optical/IR {\\em imaging} studies have remained ambiguous in identifying a single fuelling mechanism \\citep{86bus,fuentes88,deRob98,hunt04,tang08,kuo08}. External perturbations, such as tidal interactions or minor mergers, and internal instabilities in non-axisymmetric potentials such as bars or $m=1$ spirals have long been suggested as viable mechanisms for removing angular momentum from host galaxy gas to allow it to move closer to the AGN \\citep{wada04,witold97}, and more recently hydrodynamic turbulence in the nuclear interstellar medium has been suggested to contribute to low-level accretion \\citep{alig11}, but direct evidence of such physical mechanisms has been hard to obtain (e.g., \\citealt{cgm99,martini01}). The availability of large numbers of uniformly derived galaxy properties from the Sloan Digital Sky Survey (SDSS; \\citealt{york00}) has allowed statistical comparisons of active and inactive galaxies to be performed and evolutionary trends to be investigated in an attempt to explain the location of active and inactive galaxies across the so-called blue and red sequences in galaxy color-magnitude space (\\citealt{bald04,west07,schaw09}). These studies build on earlier work on smaller samples, such as that by \\cite{hunt04}, who suggested an evolutionary sequence driven by an underlying, but unidentified, dynamical instability to explain structural differences observed in a sample of 250 active and inactive host galaxies imaged with NICMOS on the \\textit{Hubble Space Telescope} (\\textit{HST}). Direct determination of the dynamical properties of active and inactive host galaxies is therefore vital. Radio interferometric spectroscopy provides valuable information on the large-scale gaseous environment and gaseous structure and kinematics of the host galaxy disks to large radii (e.g., \\citealt{cgm07,haan08,haan09}) but surface brightness sensitivity constraints of current interferometers limits achievable angular resolutions to $\\sim5$\\as\\ at best (\\citealt{mund99,things08}), but more typically $\\sim$18\\arcsec. Millimetric observations can sample dense gas, such as CO, closer to the nucleus, but this gas may be clumpy and discontinuous, limiting the ease with which the velocity fields can be interpreted (e.g., \\citealt{dumas10}). Optical observations of galactic kinematics offer the capability of sampling the velocity field close to the centre of a galaxy; in Seyferts, such kinematic observations were traditionally performed with a combination of narrowband imaging to identify line-emitting regions and long-slit spectroscopy in a preferred position angle (e.g., \\citealt{ekers83,tad89a,wilson94}), but disentangling the kinematics of the host galaxy from non-gravitational AGN-driven outflows is difficult and model dependent. With recent developments in integral-field unit (IFU) technology on large telescopes, which provide 2D imaging spectroscopy, it is now feasible to study the distribution and kinematics of gas {\\em and} stars on scales ever closer to the galactic nucleus (e.g., \\citealt{fathi06,barb06,riffel08}). Spurious misinterpretation of one-dimensional velocity fields is less likely when full two-dimensional velocity fields are sampled; the comparison of stellar and gaseous velocity fields is vital to identify kinematics associated with the galactic gravitational potential rather than non-circular gaseous streaming motion that may be erroneously interpreted in long-slit data e.g., mistakenly implying the presence of black holes offset from their galactic centres \\citep{wilson85,ferruit04}. The two-dimensional approach was particularly effective in the comparison of stellar and gaseous kinematics in a small matched sample of active and inactive galaxies by \\cite{dumas07}, who used the SAURON IFU \\citep{bacon01} on the 4.2-m William Herschel Telescope to tentatively identify a kinematic difference between Seyfert and inactive galaxy hosts, after removal of non-gravitational AGN dynamics. The SAURON IFU provides a large field of view (FOV, 33\\as $\\times$ 41\\as) but at the cost of pixel sampling ($\\sim 1$\\as) and the sample was selected before SDSS data were available. We, therefore, carefully selected a larger, distance-limited and well-matched sample of active and inactive control galaxies from the SDSS (see \\citealt{west07}) and observed them with the IMACS-IFU on the 6.5-m Magellan telescope \\citep{bigelow98,schmoll04}. The IMACS-IFU is particularly well-suited to the study of galactic nuclei as it has small pixels (\\das02) to sample its (4\\as $\\times$ 5\\as) field of view and an unusually large wavelength coverage ($\\sim 4000 - 7000$~\\AA), which provides access to all major emission lines from \\hb\\ to \\ha\\ as well as the underlying stellar continuum and stellar absorption lines such as Mg\\,$b$ and \\feii. Table~\\ref{tab:IFUs} shows a comparison of selected integral-field spectrographs. IMACS is most commonly used as a multi-object spectrograph (MOS) and published documentation concentrates on this MOS mode. In this paper, we describe the full procedure for obtaining stellar and gaseous distribution and kinematic maps from Magellan IMACS-IFU data. In particular we present, for the first time, the detailed procedure for reducing and calibrating IMACS in IFU `long-mode'. We present the resultant gaseous and stellar maps in catalogue form for our SDSS-selected IMACS-IFU sample of active and matched-inactive galaxies. The detailed dynamical analysis, interpretation and comparison will be presented in Paper II. ", "conclusions": "We have presented the first two-dimensional maps of ionised gas and stars in an SDSS-selected sample of active and inactive galaxies obtained using the IMACS-IFU on the Magellan-I 6.5m telescope. This study required the development of a full reduction pipeline for IMACS-IFU, which was previously unavailable. In this paper we have provided a detailed description of the reduction process and analysis of IMACS-IFU data, from observing strategy to final extracted maps, with particular emphasis on aspects that affect IMACS-IFU such as telescope flexure. Following extensive testing of the IMACS-IFU pipeline, two-dimensional stellar and gas kinematics were derived for 26 of the 28 galaxies in the sample thus far. In contrast to other IFUs, the unusually large wavelength coverage ($\\sim 4000-7000$ \\AA) provided by the IMACS-IFU, coupled with the fine pixel sampling across the \\das415 $\\times$ \\das500 FOV has allowed the extraction of gaseous and stellar kinematics that probe both the AGN-related regions and the host galaxy. In particular, simultaneous observation of \\oiii\\ and \\ha\\ emission lines provide independent probes of galaxy rotation and AGN-driven outflow. Evidence of rotation was found in the stellar velocity fields of 11 out of the 26 galaxies analysed, while the S/N of the remaining galaxies was deemed insufficient to make any conclusions. Fifteen of twenty-six galaxies showed clear evidence of rotation in the \\ha\\ velocity fields. Seven Seyfert galaxies show possible evidence of an \\oiii\\ outflow component, the most extreme of which exceeds a blue-shifted line-of-sight velocity of 200\\,\\kms\\ and extends over 2\\as\\ ($\\sim$1.5~kpc). We have demonstrated the value of large simultaneous wavelength coverage, which allows for the derivation of the underlying host-galaxy dynamics from \\ha\\ kinematics, regardless of the perturbations revealed in \\oiii\\ kinematics. Full kinematics, modelling and its interpretation will be presented in Paper II (P. B. Westoby et al., in preparation)." }, "1112/1112.0561_arXiv.txt": { "abstract": "{Future cosmological surveys will provide 3D large scale structure maps with large sky coverage, for which a 3D Spherical Fourier-Bessel (SFB) analysis in spherical coordinates is natural. Wavelets are particularly well-suited to the analysis and denoising of cosmological data, but a spherical 3D isotropic wavelet transform does not currently exist to analyse spherical 3D data.}{The aim of this paper is to present a new formalism for a spherical 3D isotropic wavelet, i.e. one based on the SFB decomposition of a 3D field and accompany the formalism with a public code to perform wavelet transforms.}{We describe a new 3D isotropic spherical wavelet decomposition based on the undecimated wavelet transform (UWT) described in Starck et al. 2006. We also present a new fast Discrete Spherical Fourier-Bessel Transform (DSFBT) based on both a discrete Bessel Transform and the HEALPIX angular pixelisation scheme. We test the 3D wavelet transform and as a toy-application, apply a denoising algorithm in wavelet space to the Virgo large box cosmological simulations and find we can successfully remove noise without much loss to the large scale structure.}{We have described a new spherical 3D isotropic wavelet transform, ideally suited to analyse and denoise future 3D spherical cosmological surveys, which uses a novel Discrete Spherical Fourier-Bessel Transform. We illustrate its potential use for denoising using a toy model. All the algorithms presented in this paper are available for download as a public code called {\\tt MRS3D} at \\url{http://jstarck.free.fr/mrs3d.html}}{} ", "introduction": "\\subsection*{Challenges in Modern Cosmology} The wealth of cosmological data in the last few decades \\citep{wmap7,Schrabback:2010,Percival:2007} has led to the establishment of a standard model of cosmology, which describes the Universe as composed today of approximately 4\\% baryons, 22\\% dark matter and 74\\% dark energy. The main challenges in modern cosmology are to understand the nature of both dark energy and dark matter, as well as the initial conditions of the Universe \\citep{DETF,WGFC}. A thorough understanding of these three topics may lead to a revision of Einstein's theory of General Relativity and our view of the early Universe. New surveys are planned who aim to answer these important questions e.g. Planck for the CMB \\citep{Planck}, DES \\citep[Dark Energy Survey,][]{DES:2005}, BOSS \\citep[Baryon Oscillation Spectroscopic Survey,][]{Schlegel:2007}, LSST \\citep[Large Synoptic Survey Telescope,][]{LSST} and Euclid \\citep{Euclid:2011} for weak lensing and the study of large scale structure with galaxy surveys. The challenge with these upcoming large data-sets is to extract the cosmological information in the most suitable manner in order to test the cosmological paradigm. Depending on the signal one wishes to extract, and/or survey geometry, different bases may be more or less suitable (e.g., Fourier, Spherical Harmonic, Configuration or Wavelet Space). Moreover, future surveys may be in 2D (e.g. Planck) or in 3D (e.g. galaxy or weak lensing surveys). Where 3D data is available, a tomographic analysis is possible (also known as 2D 1/2), or a full 3D analysis can be done. For data in spherical coordinates, this corresponds to a Spherical Fourier-Bessel (SFB) decomposition \\citep{Heavens:1995,Fisher:1995,weak3d,Erdogdu:2005wi,Erdogdu:2006dv,Leistedt2011,Rassat:2011bao}. \\subsection*{Wavelet Transform on the Sphere} Wavelets are particularly well suited to the analysis of cosmological data \\citep{aw:martinez93,starck:sta05_2}, since cosmological data can often be sparsely represented in wavelet space. 2D Wavelets have been used in many astrophysical studies \\citep{starck:book06} for a broad range of applications such as denoising, deconvolution, detection, etc. CMB studies have motivated the development of 2D spherical wavelet decompositions. Continuous wavelet transforms on the sphere~\\citep{wave:antoine99,wave:tenerio99,wave:cayon01,wave:holschneider96} have been proposed, mainly for non Gaussianity studies. In \\citet{starck:sta05_2}, an invertible isotropic undecimated wavelet transform (UWT) on the sphere based on spherical harmonics was described, that can be also used for other applications such as deconvolution, component separation \\citep{starck:yassir05,bobin-gmca-cmb,delabrouille08}, inpainting \\citep{inpainting:abrial06,starck:abrial08}, or Poisson denoising \\citep{schmitt2010}. A similar wavelet construction has been published in \\citep{marinucci08,fay08a,fay08} using so-called ``needlet filters\", and in \\citet{wiaux08}, an algorithm was proposed which allows us to reconstruct an image from its steerable wavelet transform. Other multiscale transforms on the sphere such as ridgelets and curvelets have been developed \\citep{starck:sta05_2}, and are well adapted to detect anisotropic features. Other multiscale transforms on the sphere, such as ridgelets and curvelets, have been developed \\citep{starck:sta05_2} and are well adapted to detect anisotropic features. An extension of this UWT has also been developed for polarised CMB data in \\citet{starck:pola09}. In this paper, we describe a new 3D isotropic spherical wavelet decomposition, which is reversible, and could therefore be useful for many different applications. It is based on the UWT proposed by \\cite{starck:sta05_2} and extended into 3D. The 3D-UWT proposed here can be used to analyse 3D data in spherical coordinates, such as a 3D galaxy or weak lensing survey with large (but not necessarily full) sky coverage. ", "conclusions": "Modern cosmology requires the analysis of 3D fields on large areas of the sky, i.e. where the field is best viewed in spherical coordinates. In this configuration, a Spherical Fourier Bessel (SFB) transform is the most natural way to statistically analyse the field. Wavelet transforms have been shown to be ideally suited for cosmological fields, which tend to be sparse in wavelet space. The wavelet transform can be used e.g. for denoising, but there is yet no 3D wavelet transform in spherical coordinates. We present in this paper a new 3D spherical wavelet transform, based on the undecimated wavelet transform (UWT) described in \\citep{starck:sta05_2}. In order to perform operations on the wavelet transforms (such as denoising), we require a discrete version of the SFB transform for both the direct and inverse transforms. We show a novel way to perform such a fast Discrete Spherical Fourier-Bessel Transform (DSFBT) based on both a discrete Bessel transform and the HEALPIX angular pixelisation scheme. Using the 3D wavelet transform and the DSFBT, both introduced in this paper, we denoise a test large scale structure data set, taken from the Virgo large box simulations\\footnote{\\url{http://www.mpa-garching.mpg.de/Virgo/VLS.html}}. We find we can satisfactorily remove artificially added Gaussian noise without much loss to the large scale structure. All the algorithms presented in this paper are available for download as a public code called {\\tt MRS3D} at \\url{http://jstarck.free.fr/mrs3d.html}." }, "1112/1112.2087_arXiv.txt": { "abstract": "{Transit and radial velocity observations continuously discover an increasing number of exoplanets. However, when it comes to the composition of the observed planets the data are compatible with several interior structure models. Thus, a planetary parameter sensitive to the planet's density distribution could help constrain this large number of possible models even further.} {We aim to investigate to what extent an exoplanet's interior can be constrained in terms of core mass and envelope metallicity by taking the tidal Love number $k_2$ into account as an additional, possibly observable parameter.} {Because it is the only planet with an observationally determined $k_2$, we constructed interior models for the Hot Jupiter exoplanet HAT-P-13b by solving the equations of hydrostatic equilibrium and mass conservation for different boundary conditions. In particular, we varied the surface temperature and the outer temperature profile, as well as the envelope metallicity within the widest possible parameter range. We also considered atmospheric conditions that are consistent with nongray atmosphere models. For all these models we calculated the Love number $k_2$ and compared it to the allowed range of $k_2$ values that could be obtained from eccentricity measurements of HAT-P-13b.} {We use the example of HAT-P-13b to show the general relationships between the quantities temperature, envelope metallicity, core mass, and Love number of a planet. For any given $k_2$ value a maximum possible core mass can be determined. For HAT-P-13b we find $M_\\mathrm{core}<27$\\,\\ME, based on the latest eccentricity measurement. We favor models that are consistent with our model atmosphere, which gives us the temperature of the isothermal region as $\\sim2100$\\,K. With this external boundary condition and our new $k_2$-interval we are able to constrain both the envelope and bulk metallicity of HAT-P-13b to 1 -- 11 times stellar metallicity and the extension of the isothermal layer in the planet's atmosphere to 3 -- 44\\,bar. Assuming equilibrium tidal theory, we find lower limits on the tidal $Q$ consistent with $10^3 - 10^5$.} {Our analysis shows that the tidal Love number $k_2$ is a very useful parameter for studying the interior of exoplanets. It allows one to place limits on the core mass and estimate the metallicity of a planet's envelope.} ", "introduction": "Today more than 180 transiting exoplanets have been discovered. Within the exoplanet family planets that are found to transit their host star are especially important. The knowledge of mass \\emph{and} radius enables us to infer the density and bulk composition of a planet. Further, transit and secondary eclipse observations reveal information about an exoplanet's atmosphere. Despite all the information given, we are still far away from knowing the composition of the deep interior of exoplanets. Characteristics like the existence and mass of a potential core or the amount and distribution of heavy elements are quite ambiguous. Yet, this information is highly desired in order to understand planet formation. For the solar system planets, the ambiguity of interior models can be reduced by taking into account information from the gravity field, which is quantified by the gravitational moments $J_2$, $J_4$, and $J_6$. In our solar system these parameters are measureable and improved methods continuously increase their accuracy. However, for an extrasolar planet gravitational moments cannot be determined. Hence, we need a similar parameter that is accessible and will also provide us with information about the interior density distribution of the planet. This can be accomplished with the tidal Love number $k_2$ \\citep{Love11,Gavrilovetal75,GavrilovZharkov77,ZharkovTrubitsyn78}. As it is equivalent to $J_2$ for the solar system planets \\citep{Hubbard84}, it is promising that $k_2$ will help us to further constrain the interior models of exoplanets. Recent studies have shown that $k_2$ is itself in general a degenerate quantity if considered in a three-layer model \\citep{Krammetal11}. In this study, we focus on the transiting Hot Jupiter HAT-P-13b \\citep{Bakosetal09}. This planet is of great interest because it is the only planet so far which has an observationally determined value for the Love number $k_2$ \\citep{Batyginetal09}. Hence, it gives us the chance to investigate how much information about the interior structure can be inferred from an actually measured $k_2$ and serves as a case study in this work. We especially evaluate the impact of the new eccentricity measurement for HAT-P-13b by \\citet{Winnetal10} as it gives new information about allowed $k_2$ values. In \\S\\ref{sec:Methods} we describe our modeling procedure, the equation of state (EOS) used, and our calculation of the Love number. The known observational constraints of the examined planet HAT-P-13b are given in \\S\\ref{sec:inputmodelassumptions} where we also discuss the effect of different eccentricity measurements, and describe the setup of our interior models. The results are explained in \\S\\ref{sec:Results}, and in \\S\\ref{sec:Discussion} we compare with previous results. The applicability of the underlying theory is discussed in the Appendix. A summary is given in \\S\\ref{sec:Summary}. ", "conclusions": "\\label{sec:Summary} In this work, we presented new interior models of the transiting Hot Jupiter HAT-P-13b with the aim of showing to what extent interior models of extrasolar giant planets can be constrained by using the tidal Love number $k_2$ in addition to the known observables mass and radius. We also varied the envelope temperature and metallicity in order to demonstrate the uncertainties imposed on the inferred interior models. One main result of our work is that based on the Love number $k_2$ one cannot draw a conclusion on the precise core mass of the planet. Only a \\emph{maximum possible} core mass can be inferred which is given by adiabatic zero-envelope-metallicity models. Taking into account the allowed $k_2$ interval ($0.1166500$\\,K can be ruled out (compare Fig.~\\ref{fig:HAT-P-13b_k2Z}). On the other hand, the very cold models with $T_\\mathrm{1\\,kbar}<4000\\,$K are unlikely because they have a metallicity that is smaller than stellar. We favor models with $T_\\mathrm{1\\,bar}=2080$\\,K, an isothermal outer layer of 3 -- 44\\,bar, and an envelope metallicity of $1 < Z/Z_* < 11$ because they (i) are placed in our \\emph{new $k_2$ interval}, (ii) are consistent with the model atmospheres from \\citet{Fortneyetal07}, and (iii) have an envelope metallicity that is above the stellar metallicity. Assuming these conditions really apply for HAT-P-13b the planet would have a core mass of 27\\,\\ME~or less. Given the calculated tidal $Q$, if HAT-P-13b turns out to have a small core and an envelope enrichment similar to Jupiter, it could represent a Jupiter-like extrasolar planet. By comparing with the previous results from \\citet{Batyginetal09} we have seen that the core EOS affects the core mass and Love number, resulting into a difference of about 28\\% in the prediction of a maximum possible core mass of HAT-P-13b. As pointed out in \\S\\ref{subsec:e}, the eccentricity of the inner planet is a crucial parameter because it determines the Love number $k_2$. The value of the eccentricity is still uncertain. We analyzed the consequences of an $e_b = 0.021\\pm0.009$ \\citep{Bakosetal09} and $e_b = 0.0133\\pm0.0041$ \\citep{Winnetal10}. Finally, despite the uncertainty of the inner planet's eccentricity it is still unclear whether the tidal fixed point theory from \\citet{Mardling07} can really be applied to HAT-P-13b. Further observations are necessary to prove the co-planarity and apsidal alignment configuration of the system. Also, TTV measurements could provide clues about the existence of other long-period companions in the system." }, "1112/1112.5458_arXiv.txt": { "abstract": "We report the first detection of the intrinsic velocity dispersion of the Arches cluster - a young ($\\sim$~2 Myr), massive ($10^4 M_{\\odot}$) starburst cluster located only 26 pc in projection from the Galactic center. This was accomplished using proper motion measurements within the central $10'' \\times 10''$~of the cluster, obtained with the laser guide star adaptive optics system at Keck Observatory over a 3 year time baseline (2006-2009). This uniform dataset results in proper motion measurements that are improved by a factor $\\sim$~5 over previous measurements from heterogeneous instruments. By careful, simultaneous accounting of the cluster and field contaminant distributions as well as the possible sources of measurement uncertainties, we estimate the internal velocity dispersion to be $0.15 \\pm 0.01$~mas yr$^{-1}$, which corresponds to $5.4 \\pm 0.4$~km s$^{-1}$~at a distance of 8.4 kpc. Projecting a simple model for the cluster onto the sky to compare with our proper motion dataset, in conjunction with surface density data, we estimate the total {\\it present-day}~mass of the cluster to be \\besttot. The mass in stars observed within a cylinder of radius $R$~(for comparison to photometric estimates) is found to be \\bestmass~at formal $3\\sigma$~confidence. This mass measurement is free from assumptions about the mass function of the cluster, and thus may be used to check mass estimates from photometry and simulation. Photometric mass estimates assuming an initially Salpeter mass function ($\\IMF = 1.35$, or $\\PDMF\\sim 1.0$~at present, where $dN/d(\\log M) \\propto M^{\\PDMF}$)~suggest a total cluster mass $M_{cl} \\sim (4-6) \\times 10^4 M_{\\odot}$~and projected mass ($\\sim 2 \\le M(R<0.4~{\\rm pc}) \\le 3$)~$\\times 10^4 M_{\\odot}$. Photometric mass estimates assuming a globally top-heavy or strongly truncated present-day mass function (PDMF, with $\\PDMF \\sim 0.6$) yield mass estimates closer to $\\MProjObs \\sim 1-1.2 \\times 10^4 M_{\\odot}$. Consequently, our results support a PDMF that is either top-heavy or truncated at low-mass, or both. Collateral benefits of our data and analysis include: (i) cluster membership probabilities, which may be used to extract a clean cluster sample for future photometric work; (ii) a refined estimate of the bulk motion of the Arches cluster with respect to the field, which we find to be 172 $\\pm$~15 km s$^{-1}$, which is slightly slower than suggested by previous VLT-Keck measurements; and (iii) a velocity dispersion estimate for the field itself, which is likely dominated by the inner galactic bulge and the nuclear disk. ", "introduction": "The spectrum of masses produced during the star formation process (the Initial Mass Function, or IMF) is a key prediction of the star formation process as it indirectly links to the observable Present-Day Mass Function (PDMF) of the population \\citep[for example, see][for review]{millerscalo79, mckee07, bastian10}. Because star formation depends on collapse by self-gravity out of a turbulent medium threaded with a magnetic field, there is some expectation that the physical conditions in the parent cloud should affect the slope of the IMF, its minimum mass, or both \\citep[e.g.][]{morris93}. Numerical modeling provides some support for environment-dependent IMF variations, particularly in the unusual environment of the Galactic center \\citep[e.g.][]{bonnell04, klessen07, krumholz08, bonnell08}. However, the resulting IMF variations may be so small as to be observable only in extreme environments \\citep[e.g. ][]{elmegreen08}. There is some observational support for a varying IMF and turn-over mass in the extreme environments of the young ($\\sim$~few Myr), massive ($\\sim 10^4 M_{\\odot}$) starburst clusters NGC 3603 \\citep{harayama08} and for the stellar cluster at the Galactic center itself \\citep{bartko10}. The young, massive cluster (YMC) known as the Arches Cluster \\citep[e.g.][]{nagata95, cotera96}\\footnote{Throughout this communication, ``the Arches'' refers to the star cluster, not the arched radio filaments \\citep{yusef84, morris89}, against which the cluster is projected, and with which it is physically associated \\citep{lang04}.} is a particularly well-studied example of an extreme environment for star formation. It is massive \\citep[Total mass $M_{cl}\\sim (2-7) \\times 10^{4} M_{\\odot}$;][]{figer99,figer02} dense \\citep[$\\rho_c \\sim 10^5 M_{\\odot}$~pc$^{-3}$][]{espinoza09} and young \\citep[$\\sim 2-2.5$~My;][]{najarro04, martins08}. It contains a substantial number of massive stars \\citep{serabyn98} which both contribute to and heat the surrounding medium \\citep[e.g.][]{figer02, yusef02, lang04}. The Arches cluster is located only 26 pc in projection from the Galactic center~(hereafter the GC). It therefore likely formed in an environment characterised by high gas pressure and velocity dispersion in the parent cloud, and high ambient temperature, particularly when compared to the relatively more benign environment of NGC 3603. As these parameters are thought to impact the IMF \\citep{morris93, klessen07}, the Arches cluster is expected to be an excellent candidate for observing a non-canonical IMF, whether in its mass function exponent, low-mass turnover, a low-mass cut-off, or all three \\citep[e.g.][]{stolte02, klessen07}. It is also young enough that the most massive main-sequence stars are still present, making it one of the few clusters in which the upper mass-limit to the star formation process may be observationally tested \\citep{figer05, crowther10}. It has thus received substantial observational attention, with efforts focused particularly on estimates of its IMF. Indeed, the Arches was originally the prototypical object for a non-standard IMF \\citep{figer99}, with an observed present-day luminosity function indicating an overabundance of massive stars compared to the canonical Salpeter IMF (parameterized as $dN/d(\\log M) \\propto M^{-\\Gamma}$, with $\\Gamma = 1.35$; see \\citealt{bastian10} for a review). However, a number of effects conspire to obscure the true IMF from observation, complicating the interpretation of the PDMF, and indeed the present consensus seems to be that the Arches began with an IMF that is consistent with the canonical Salpeter IMF found in most environments. Photometric efforts to chart the present-day luminosity function of the cluster suffer from two important limitations. Firstly, the observations are difficult; strong crowding and high, spatially-variable extinction are observed across the field of view, so that the photometric completeness is challenging to estimate for masses lower than a few $M_{\\odot}$. There is evidence for mass segregation in the cluster \\citep{figer99, stolte05}, seen as a steepening of the present-day luminosity function towards the cluster center, implying strong spatial selection effects when attempting to constrain the IMF. Secondly, the relationship of the PDMF to the IMF is not trivial to evaluate. Stellar evolution must be taken into account when relating the PDMF to the IMF, requiring a prescription for mass-loss from high-mass objects \\citep[e.g.,][]{espinoza09}. In addition, mass segregation and tidal stripping are both likely to have been important for the evolution of the Arches. Present-day mass segregation need not be primordial, since the Arches is likely already in a post-collapse phase \\citep[e.g.][]{pz07, allison09}. The Arches cluster sits in a strong tidal field, such that as much as half of its stars may already have been stripped into the field over the $\\sim 2.5$~My of its history \\citep{kim00, pz02}. Mass segregation and tidal stripping together imply that the true IMF of the cluster may differ from the IMF drawn from the subset of stars that have remained within the Arches cluster to the present day. We review the literature mass determinations in Section \\ref{ss_masscomp}. A {\\it kinematic} mass measurement provides a direct test of the PDMF of the Arches cluster, because its selection effects are somewhat less stringent. Stars below the typical photometric completeness limit of $\\sim 1-2 M_{\\odot}$~are observable through their contribution to the total cluster mass. \\citet{figer02}~were the first to attempt this, by estimating the radial velocity dispersion of a sample of emission-line stars and assuming the cluster is spherically symmetric and in virial equilibrium. However, the estimate is complicated by the difficulty of resolving the blended lines, their high width, and intrinsic line-profile variation among the sample, so that the resulting mass estimate is strongly dependent on atmosphere models. Mass estimates using the velocity dispersion derived from proper motions are independent of the details of the atmospheres of the tracer stars, and in principle allow for the mass distribution to be derived in a more assumption-free manner \\citep{lm89}. The advent of adaptive optics on large telescopes in the near-infrared has enabled the measurement of precise proper motions of a large number of stars in the Galactic center clusters. In a pioneering proper motion study of the Arches, \\citet{stolte08} used one epoch each of VLT/NACO and Keck/NIRC2 separated by 4.3 years to measure the motion of the cluster. However, differential distortion between the cameras limited the proper motion precision to $\\sim 0.7$~mas yr$^{-1}$, somewhat too coarse to measure the internal velocity dispersion\\footnote{We use the term ``velocity dispersion'' to refer to both the dispersion in mas yr$^{-1}$~and km s$^{-1}$~throughout.}, for which the expected order of magnitude is about $\\sim 0.2$~mas yr$^{-1}$. We have observed the central $10'' \\times 10''$~of the Arches across five epochs in three years (2006-2009) with a uniform observational setup (PI Morris). Using NIRC2 on Keck-2, behind the LGS Adaptive Optics facility \\citep{ghez05, wizinowich06}, these cross-instrument systematics encountered by \\citet{stolte08} are not present in our observations, and we are able to attain proper motion measurements with error lower than the expected velocity dispersion. We report here on our results, which provide the first kinematic mass estimate of the Arches cluster from proper motions. This communication is organised as follows. Section \\ref{s_obs} describes the observations and positional measurement technique, while Section \\ref{s_analyse} describes the process of proper motion measurement and error assignment. Section \\ref{s_res} describes the techniques used to fit the cluster membership probabilities and kinematic parameters. Section \\ref{s_discuss} provides our mass measurement and new bulk motion measurement for the Arches, and briefly discusses the implications. ", "conclusions": "With uniform observational setup over a sufficient time baseline and careful accounting for a number of sources of proper motion error, we have measured the internal velocity dispersion of the Arches cluster for the first time, finding $\\sigma = 0.15 \\pm 0.01$~mas yr$^{-1}$, which corresponds to $5.4 \\pm 0.4$~km s$^{-1}$~at a distance of 8.4 kpc. We have used this dispersion to test the photometric estimates of the present-day mass function (PDMF) of the Arches cluster. The total mass is likely in the range \\besttot, but this is only weakly constrained by kinematic data and is consistent with nearly all suggestions of total cluster mass from the modeling literature. The projected mass (i.e., mass contained within a cylinder of radius $R$~on the sky) is rather better constrained; we find \\bestmass~at formal $3\\sigma$~confidence. The upper bound of this range is $3\\sigma$~below the photometric estimate for \\MProjObs~estimated by \\citet{espinoza09}~under the assumptions of a non-top-heavy mass function without low-mass truncation. If a substantial contribution from massive binaries were unknowingly included in our measurement, this would strengthen our conclusion because the upper cluster mass bound would accordingly be reduced. This is the first mass estimate for the Arches based on proper motion velocity-dispersion. We have also revised the bulk motion of the Arches slightly downward. Our updated motion of $172 \\pm 15$~km s$^{-1}$~is only slightly lower than the $212 \\pm 29$~km s$^{-1}$~determined previously (Stolte et al. 2008). Taken at face value, this supports the previous conclusion that the Arches Cluster is unlikely to pass within 10~pc of the GC. Finally, we have provided the first estimate (to our knowledge) of the velocity dispersion of the Bulge along such a close sight-line to the Galactic center; this estimate is (103, 64)~$\\pm$~(7.7,5.0)~km s$^{-1}$, with the major axis coincident with the Galactic plane, to within the uncertainties. \\begin{figure} \\begin{center} \\includegraphics[width=14cm]{f1.jpg} \\caption{NIRC2 $K'$~mosaic of the core field of the Arches in May 2009. This is our best map in terms of both resolution (51 mas) and sensitivity ($K'_{lim} = 20.59$~mag; Table 1). All stars on which we report in this paper fall within the field of view indicated here. The scale-bar is two arcseconds in length. Stars used as PSF reference-stars are indicated by circles. When stellar membership probabilities are reported, positions are reported as offsets from the reference star indicated by the square in this figure.} \\label{fig_findingchart} \\end{center} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[width=12cm]{f2.png} \\caption{Positional errors as measured for an example epoch (2008.50). Positions are those in the image-stack with centroiding errors assessed as the rms of measurements within an epoch (Section \\ref{ss_centroiding}). Top row: centroiding errors along detector-X and detector-Y (top-left and top-middle respectively), and the average of the two as a function of distance from the field center (top-right). Bottom row: alignment errors along X and Y (bottom-left and bottom-middle; Section \\ref{ss_align}). The magnitude histogram is given in the bottom-right panel.} \\label{fig_additive_det} \\end{center} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[height=120mm]{f3.jpg} \\caption{Five example positional time-series. Left columns show motion along X, right columns along Y. Object IDs and $K'$~magnitudes are indicated in the right top and bottom corners respectively. Note that each vertical axis is scaled to accommodate the motion of the star and is in units of pixels in the \\treference~reference frame. The best-fit straight line to the motions are indicated in each case, as are 1$\\sigma$~positional error curves. Object 154 is likely a field object, as indicated by its large proper motion relative to the reference frame.} \\label{f_example_curves} \\end{center} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[width=15cm]{f4.png} \\caption{The distribution of adopted proper motion precision (Section \\ref{ss_errors} and Table \\ref{tab_budget}), for all objects with five position-measurements. Outliers due to likely mismatches are indicated by squares and were removed from the analysis. An object qualifies as an outlier if the rms in either co-ordinate falls obviously outside the sequence defined by most of the points.} \\label{f_motion_errors} \\end{center} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[height=18cm]{f5.png} \\caption{Vector Point Diagrams for the overlapping magnitude-ranges of Section \\ref{ss_fit} and Table \\ref{tab_kinparams}. Shaded ellipses give the $1\\sigma$~contours for the two-dimensional gaussian components fit to the field and cluster components. Within each ellipse, the lines indicate the length and direction of the semimajor (thick red line) and semiminor (thin red line) axes.} \\label{f_VPD_compilation} \\end{center} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[height=10cm]{f6.jpg} \\caption{Observed velocity dispersion in each coordinate for stars with $14.0 \\le K' < 17.0$~compared to a gaussian of width equal to the mean measurement error over this range (Section \\ref{ss_errors} and Table \\ref{tab_budget}). Panels correspond to detector-X (Top) and detector-Y (bottom). This figure was constructed after removing likely field objects (Section \\ref{ss_membprob})} \\label{f_compare_erro} \\end{center} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[width=15cm]{f7.png} \\caption{Color-magnitude diagram (CMD) and Vector Point Diagram (VPD) for all objects with proper motion error $< 0.5$~mas yr$^{-1}$~and five epochs of measurement. The CMD presented here was constructed by matching $K'$~measurements to photometry taken in $H$-band in 2006 May with Keck-2/NIRC2-LGS \\citep{mccrady11}, which limits the depth in the CMD. Objects with $P_{\\rm{cluster}} > 0.995$~are shown in black, all other objects denoted with open circles. Red objects in the VPD correspond to the stars within the red dot-dashed box in the CMD, and represent well-measured objects with a possible $H-K'$~excess. Of these objects, those with $P_{\\rm cluster} > 0.995$~are shown with a red circle; their field counterparts are shown with red squares. See Section \\ref{ss_excess}~and \\citet{stolte10}~for more information on these objects.} \\label{fig_CMD} \\end{center} \\end{figure} \\begin{figure} \\includegraphics[width=8cm]{f8a.png} \\includegraphics[width=8cm]{f8b.png} \\caption{Views of the $\\Delta \\chi^2_{full} < 7.82$~region when both kinematic and surface density data (for stars of mass $10 \\le M/M_{\\odot} \\le 30$; Espinoza et al. 2009) are included in the assessment. Axes are: $R_c, R_t, M_{cl}$, with total cluster mass $M_{cl}$~vertical in each case. Limits shown are: $0.05 \\le R_c \\le 0.8$~pc; $1.0 \\le R_t \\le 30$~pc; $0.5 \\le M_{cl} \\le 6.0 \\times 10^4 M_{\\odot}$.} \\label{fig_chibubble_mtot_10-30} \\end{figure} \\begin{figure} \\includegraphics[width=8cm]{f9a.png} \\includegraphics[width=8cm]{f9b.png} \\caption{As Figure \\ref{fig_chibubble_mtot_10-30}, but with \\MProjObs~along the vertical axis. Limits shown are: $0.05 \\le R_c \\le 0.8$~pc; $1.0 \\le R_t \\le 30$~pc; $0.5 \\le \\MProjObs \\le 2.5 \\times 10^4 M_{\\odot}$.} \\label{fig_chibubble_mproj_10-30} \\end{figure} \\begin{figure} \\includegraphics[width=16cm]{f10.png} \\caption{Radial profiles corresponding to parameter-sets within the $\\Delta \\chi^2_{full} < 7.82$~surface; both our own kinematic data and surface density (by number) \\SurfdensN~data are used (corresponding to stars of mass $10 \\le M/M_{\\odot} \\le 30$; Espinoza et al. 2009). Top-left and top-middle panels show radial and tangential velocity dispersions from proper motions (points) over the projected profiles corresponding to model parameters (lines). Top-right panel shows the \\SurfdensN~dataset with model predictions. Bottom-left and bottom-middle panels show the total mass within cylindrical radius $R$~on the sky, with $R=0.4$~pc indicated by the vertical dashed line. Bottom-right panel shows the mass enclosed within a sphere of radius $r$~pc from the cluster center. See also Table \\ref{table_massest_10-30_full}.} \\label{fig_radprofs_finer_10-30} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[width=16cm]{f11.png} \\caption{Photometric mass estimates for the Arches cluster from the literature. Symbols give the directly-observed photometric mass (circles) and extrapolated mass (where reported; squares and pentagons) depending on the low-mass cut-off assumed. The citation for each estimate is shown inside the symbols (using the numbering of Table \\ref{table_all_mass_estimates}). The horizontal bands show our model-dependent mass estimate using our dispersion data (at 1$\\sigma$, 2$\\sigma$~and 3$\\sigma$), and the \\SurfdensN~dataset of \\citet{espinoza09}; see Section \\ref{ss_massmodel} for more detail on the mass modeling used. All masses are reported as \\MProjObs.} \\label{fig_literature_estimates} \\end{center} \\end{figure} \\begin{table} \\small \\begin{center} \\begin{tabular}{lcccccccc} % Epoch & (t$_{int} \\times N_{coadd}$) & $N_{images}$ & $N_{use}$ & FWHM & Strehl & $N_{\\ast}$ & $N_{\\ast, {\\rm uncrowd}}$ & $K'_{lim}$ \\\\ % & (s) & & & (mas) & & & & (mag) \\\\ % \\hline 2006 May 21 & 3.00 $\\times$ 10 & 15 & 15 & 61.05 & 0.261 & 660 & 649 & 19.43 \\\\ % 2006 Jul 18 & 3.00 $\\times$ 10 & 52 & 38 & 56.95 & 0.349 & 657 & 642 & 19.89 \\\\ % 2008 May 13 & 3.00 $\\times$ 10 & 146 & 72 & 66.66 & 0.219 & 556 & 536 & 19.74 \\\\ % 2008 Jun 01 & 3.00 $\\times$ 10 & 89 & 83 & 54.96 & 0.373 & 845 & 810 & 20.42 \\\\ % 2009 May 02 & 2.80 $\\times$ 10 & 119 & 108 & 51.47 & 0.442 & 968 & 917 & 20.59 \\\\ % \\end{tabular} \\end{center} \\normalsize \\caption{Summary of observations. Reading left-right, the columns are: Epoch of observation, the total integration time for each image, the number of images observed, the number of images used, the median FWHM and Strehl ratio over the set of accepted images $N_{use}$, the number of stars measured within the mean image stack in each epoch; the number surviving the cut on proximity to a known neighbour, and finally the magnitude $K'_{lim}$~at which the cumulative distribution function of the observed $K'$~magnitudes reaches 90$\\%$~of the total number of stars in the sample at each epoch.} \\label{tab_obsns} \\end{table} \\scriptsize \\begin{deluxetable}{lcccccccc} \\tablewidth{0pt} \\tablecaption{PSF stars. Reading left-right, columns are: sequential star number in the master table of membership probabilities, estimated brightness, and finally the position of the star expressed as an offset in arcseconds from the reference star along the (E-W) and (S-N) directions. See Figure 1 for the locations of these stars on the field of view.} \\tablehead{ Row & $K'$ & $\\Delta x$ & $\\Delta y$ \\\\ & (mag) & ($\"$) & ($\"$) \\\\ } \\startdata 1 & 10.24 & 2.736 & -3.943 \\\\ 2 & 10.48 & 2.063 & -1.193 \\\\ 3 & 10.49 & 0.791 & 0.755 \\\\ 4 & 10.66 & 3.150 & -2.899 \\\\ 5 & 11.08 & -0.633 & -4.252 \\\\ 7 & 11.22 & -1.650 & 1.730 \\\\ 8 & 11.25 & 1.385 & -2.334 \\\\ 16 & 12.18 & 1.012 & -5.199 \\\\ 24 & 12.49 & 0 & 0 \\\\ 25 & 12.50 & 5.407 & -0.218 \\\\ 32 & 12.88 & 2.499 & -5.402 \\\\ \\enddata \\label{tab_psfstars} \\end{deluxetable} \\normalsize \\begin{sidewaystable} \\centering \\scriptsize \\begin{tabular}{l|r||l|l|l|l|l||} \\multicolumn{2}{l||}{Parameter} & 2006.38 & 2006.54 & 2008.37 & 2008.50 & 2009.33 \\\\ \\hline \\hline \\multicolumn{2}{l||}{$N_{ref}$} & 238 & 239 & 235 & 241 & 233 \\\\ \\multicolumn{2}{l||}{$N_{<4}$} & 9 & 10 & 21 & 11 & 6 \\\\ \\hline \\hline $\\Delta$ & $x'$ & ~-83.94 $\\pm$~$6.03 \\times 10^{-3}$ & ~9.26 $\\pm$~$4.01 \\times 10^{-3}$ & ~4.16 $\\pm$~$3.71 \\times 10^{-3}$ & (-4.06 $\\pm$~17.716)$\\times 10^{-4}$ & ~6.53 $\\pm$~$3.52 \\times 10^{-3}$\\\\(pix)& $y'$ & ~-16.88 $\\pm$~$4.74 \\times 10^{-3}$ & ~5.69 $\\pm$~$4.39 \\times 10^{-3}$ & ~3.35 $\\pm$~$5.64 \\times 10^{-3}$ & (3.66 $\\pm$~25.421)$\\times 10^{-4}$ & ~3.51 $\\pm$~$3.66 \\times 10^{-3}$\\\\ \\hline $x$ & $x'$ & ~1.00005 $\\pm$~$2.11 \\times 10^{-5}$ & ~0.99994 $\\pm$~$9.12 \\times 10^{-6}$ & ~1.00025 $\\pm$~$1.34 \\times 10^{-5}$ & ~0.99999 $\\pm$~$6.18 \\times 10^{-6}$ & ~1.00002 $\\pm$~$8.24 \\times 10^{-6}$\\\\()& $y'$ & (6.64 $\\pm$~0.185)$\\times 10^{-4}$ & (7.73 $\\pm$~0.079)$\\times 10^{-4}$ & (1.33 $\\pm$~1.425)$\\times 10^{-5}$ & (2.32 $\\pm$~5.858)$\\times 10^{-6}$ & (2.75 $\\pm$~0.090)$\\times 10^{-4}$\\\\ \\hline $y$ & $x'$ & (-3.15 $\\pm$~0.177)$\\times 10^{-4}$ & (-3.53 $\\pm$~0.089)$\\times 10^{-4}$ & (1.23 $\\pm$~0.105)$\\times 10^{-4}$ & (-2.40 $\\pm$~4.766)$\\times 10^{-6}$ & (-2.54 $\\pm$~0.072)$\\times 10^{-4}$\\\\()& $y'$ & ~1.00046 $\\pm$~$1.97 \\times 10^{-5}$ & ~1.00028 $\\pm$~$9.62 \\times 10^{-6}$ & ~1.00014 $\\pm$~$1.34 \\times 10^{-5}$ & ~1.00000 $\\pm$~$6.66 \\times 10^{-6}$ & ~1.00009 $\\pm$~$7.99 \\times 10^{-6}$\\\\ \\hline $x^2$ & $x'$ & ~28.90 $\\pm$~5.946 & ~20.61 $\\pm$~3.463 & ~-18.24 $\\pm$~3.989 & ~-0.27 $\\pm$~1.986 & ~4.63 $\\pm$~2.818\\\\($\\times 10^{-8}$ pix$^{-1}$)& $y'$ & ~11.78 $\\pm$~4.751 & ~15.66 $\\pm$~3.266 & ~23.10 $\\pm$~4.404 & ~0.48 $\\pm$~1.937 & ~20.65 $\\pm$~2.938\\\\ \\hline $xy$ & $x'$ & ~1.80 $\\pm$~5.818 & ~20.39 $\\pm$~3.772 & ~58.96 $\\pm$~4.594 & ~-0.28 $\\pm$~2.040 & ~-0.58 $\\pm$~2.973\\\\($\\times 10^{-8}$ pix$^{-1}$)& $y'$ & ~-5.05 $\\pm$~6.935 & ~-10.19 $\\pm$~3.985 & ~-28.61 $\\pm$~5.203 & ~-1.52 $\\pm$~2.680 & ~-20.03 $\\pm$~3.335\\\\ \\hline $y^2$ & $x'$ & ~-3.04 $\\pm$~5.917 & ~5.55 $\\pm$~3.048 & ~1.09 $\\pm$~3.540 & ~1.71 $\\pm$~1.686 & ~-0.59 $\\pm$~2.959\\\\($\\times 10^{-8}$ pix$^{-1}$)& $y'$ & ~33.22 $\\pm$~6.228 & ~52.07 $\\pm$~3.664 & ~75.11 $\\pm$~5.451 & ~-1.13 $\\pm$~2.488 & ~-0.47 $\\pm$~2.633\\\\ \\hline \\hline \\multicolumn{2}{l||}{$1.0$-$M$} & ~-2.55 $\\pm$~0.150 & ~-1.14 $\\pm$~0.064 & ~-1.95 $\\pm$~0.095 & (1.97 $\\pm$~4.303)$\\times 10^{-2}$ & ~-0.58 $\\pm$~0.062\\\\ \\multicolumn{2}{l||}{($\\times 10^{-4}$ )} & & & & &\\\\ \\hline \\multicolumn{2}{l||}{$1.0$-$M_y/M_x$} & ~-4.08 $\\pm$~0.275 & ~-3.34 $\\pm$~0.135 & ~1.10 $\\pm$~0.189 & (-5.72 $\\pm$~9.564)$\\times 10^{-2}$ & ~-0.71 $\\pm$~0.104\\\\ \\multicolumn{2}{l||}{($\\times 10^{-4}$ )} & & & & &\\\\ \\hline \\multicolumn{2}{l||}{$\\theta_{rot}$} & ~-10.11 $\\pm$~0.250 & ~-11.62 $\\pm$~0.129 & ~1.13 $\\pm$~0.182 & (-4.88 $\\pm$~7.925)$\\times 10^{-2}$ & ~-5.46 $\\pm$~0.116\\\\ \\multicolumn{2}{l||}{($''$)} & & & & &\\\\ \\hline \\multicolumn{2}{l||}{$\\theta_{skew}$} & ~7.20 $\\pm$~0.553 & ~8.66 $\\pm$~0.234 & ~2.82 $\\pm$~0.364 & -0.002 $\\pm$~0.153 & ~0.43 $\\pm$~0.245\\\\ \\multicolumn{2}{l||}{($''$)} & & & & &\\\\ \\hline \\end{tabular} \\normalsize \\caption{Transformation parameters taking the starlist in each epoch into the reference frame $t_{ref}$. {\\it Top row}: Number of reference stars $N_{ref}$~used in the mapping, along with the number of reference stars $N_{<4}$~that appear in fewer than four epochs. {\\it Next six rows:} Coefficients of the polynomial fits $x' = f(x,y)$~and $y' = g(x,y)$~(top and bottom rows respectively in each pair). For reference, a quadratic term of size $10.0\\times10^{-8}$~pix$^{-1}$~would introduce displacement 0.25 mas at the edges of the detector, comparable to the centroiding error for bright ($K' < 16$)~objects (Table \\ref{tab_budget}). {\\it Bottom four rows:} The linear parts of the transformations re-expressed as a global scaling $M$, nonuniform magnification $M_y/M_x$, rotation $\\theta_{rot}$~and departure from perpendicular axes $\\theta_{skew}$. (Global shifts $\\Delta$~appear in the polynomial fits and are not repeated.)} \\label{tab_transform} \\end{sidewaystable} \\begin{table} \\scriptsize \\begin{center} \\begin{tabular}{l|l||l|l|l|l|l||c||l||l||} \\multicolumn{2}{l||}{$K'$} & \\multicolumn{5}{|l||}{Centroiding, Alignment (mas)} & Additive & Confusion & Motion \\\\ \\multicolumn{2}{l||}{} & 2006.39 & 2006.54 & 2008.37 & 2008.50 & 2009.33 & (mas) & bias (mas) & (mas/y) \\\\ \\hline \\hline \\hline \\multirow{2}{*}{$10-16$} & $x$ & 0.25, 0.09 & 0.10, 0.05 & 0.25, 0.06 & 0.06, 0.04 & 0.08, 0.03 & 0.16 $\\pm$~0.02& 0.0 - 0.03 & {\\bf 0.076 } \\\\ & $y$ & 0.23, 0.08 & 0.07, 0.05 & 0.19, 0.08 & 0.11, 0.04 & 0.09, 0.05 & 0.15 $\\pm$~0.02 & & {\\bf 0.074 } \\\\ \\hline \\multirow{2}{*}{$16-18$} & $x$ & 0.41, 0.11 & 0.23, 0.05 & 0.40, 0.08 & 0.14, 0.05 & 0.14, 0.04 & 0.24 $\\pm$~0.02& 0.03 - 0.10 & {\\bf 0.130 } \\\\ & $y$ & 0.42, 0.10 & 0.20, 0.05 & 0.43, 0.09 & 0.17, 0.05 & 0.18, 0.05 & 0.30 $\\pm$~0.03 & & {\\bf 0.153 } \\\\ \\hline \\multirow{2}{*}{$18-20$} & $x$ & 1.10, 0.11 & 0.92, 0.05 & 1.03, 0.08 & 0.60, 0.05 & 0.59, 0.04 & 0.59 $\\pm$~0.06& 0.1- 1.0 & {\\bf 0.378 } \\\\ & $y$ & 1.38, 0.10 & 1.05, 0.05 & 1.35, 0.09 & 0.81, 0.05 & 0.77, 0.06 & 0.71 $\\pm$~0.08 & & {\\bf 0.478 } \\\\ \\end{tabular} \\caption{Astrometric error budget. For each magnitude bin, the top (bottom) row gives errors in X (Y). For each star, centroiding, alignment and additive error describe random variation between epochs. The effect of confusion bias on motions depends on its variation between epochs; random variation is already included in the additive error, while linear trends masquerading as spurious motions are expected to be $\\lesssim 10\\%$~of the confusion bias across the epochs for all objects (Section \\ref{ss_additive}).} \\label{tab_budget} \\end{center} \\normalsize\\end{table} \\scriptsize \\begin{deluxetable}{lcccccccc} \\tablewidth{0pt} \\tablecaption{Membership table for objects in the Arches Central field. Reading left-right, columns are: Sequential star number, estimated brightness, offset from reference star (E-W and S-N), proper motion and error, and the formal probability that the object is associated with the cluster and field, respectively.$^\\dagger$} \\tablehead{ Row & $K'$ & $\\Delta x$ & $\\Delta y$ & $\\mu_x$ & $\\mu_y$ & $P({\\rm cluster})$ & $P({\\rm field})$ \\\\ & (mag) & ($\"$) & ($\"$) & (mas yr$^{-1}$) & (mas yr$^{-1}$) & & \\\\ } \\startdata 1$^{\\ast}$ & 10.24 & 2.736 & -3.943 & 0.14 $\\pm$ 0.07 & 0.13 $\\pm$ 0.08 & 0.999 & $8.54 \\times 10^{-4}$ \\\\ 2$^{\\ast}$ & 10.48 & 2.063 & -1.193 & 0.08 $\\pm$ 0.07 & 0.08 $\\pm$ 0.07 & 0.999 & $5.85 \\times 10^{-4}$ \\\\ 3$^{\\ast}$ & 10.49 & 0.791 & 0.755 & 0.15 $\\pm$ 0.07 & -0.09 $\\pm$ 0.07 & 0.999 & $1.06 \\times 10^{-3}$ \\\\ 4$^{\\ast}$ & 10.66 & 3.150 & -2.899 & -0.01 $\\pm$ 0.07 & 0.20 $\\pm$ 0.07 & 0.999 & $8.91 \\times 10^{-4}$ \\\\ 5$^{\\ast}$ & 11.08 & -0.633 & -4.252 & 0.03 $\\pm$ 0.07 & -0.11 $\\pm$ 0.08 & 0.999 & $6.86 \\times 10^{-4}$ \\\\ 6 & 11.16 & 4.603 & 1.092 & 0.00 $\\pm$ 0.08 & 0.07 $\\pm$ 0.08 & 1.000 & $4.88 \\times 10^{-4}$ \\\\ 7$^{\\ast}$ & 11.22 & -1.650 & 1.730 & 0.21 $\\pm$ 0.09 & 0.05 $\\pm$ 0.08 & 0.999 & $1.11 \\times 10^{-3}$ \\\\ 8$^{\\ast}$ & 11.25 & 1.385 & -2.334 & -0.28 $\\pm$ 0.07 & 0.06 $\\pm$ 0.07 & 0.999 & $1.34 \\times 10^{-3}$ \\\\ 9 & 11.63 & -1.758 & -1.287 & -0.00 $\\pm$ 0.07 & 0.05 $\\pm$ 0.08 & 1.000 & $4.72 \\times 10^{-4}$ \\\\ 10 & 11.67 & 2.038 & 0.445 & 0.05 $\\pm$ 0.07 & 0.03 $\\pm$ 0.07 & 0.999 & $5.04 \\times 10^{-4}$ \\\\ 11 & 11.81 & -2.337 & -0.540 & -0.28 $\\pm$ 0.08 & -0.20 $\\pm$ 0.08 & 0.997 & $2.59 \\times 10^{-3}$ \\\\ 12 & 11.88 & 5.528 & -3.874 & -0.04 $\\pm$ 0.08 & 0.04 $\\pm$ 0.11 & 1.000 & $4.66 \\times 10^{-4}$ \\\\ 13 & 11.89 & 0.285 & -1.191 & 0.04 $\\pm$ 0.07 & 0.01 $\\pm$ 0.07 & 1.000 & $4.98 \\times 10^{-4}$ \\\\ 14 & 12.00 & -0.158 & -3.382 & -0.06 $\\pm$ 0.07 & -0.08 $\\pm$ 0.07 & 0.999 & $5.54 \\times 10^{-4}$ \\\\ 15 & 12.18 & 5.362 & 1.667 & -0.02 $\\pm$ 0.08 & 0.09 $\\pm$ 0.08 & 0.999 & $5.02 \\times 10^{-4}$ \\\\ 16$^{\\ast}$ & 12.18 & 1.012 & -5.199 & 0.02 $\\pm$ 0.07 & -0.13 $\\pm$ 0.08 & 0.999 & $7.41 \\times 10^{-4}$ \\\\ 17 & 12.19 & -1.490 & 0.681 & 0.06 $\\pm$ 0.08 & -0.07 $\\pm$ 0.07 & 0.999 & $6.26 \\times 10^{-4}$ \\\\ \\enddata \\tablenotetext{\\ast}{PSF Star} \\tablenotetext{\\dagger}{This Table will be published in its entirety in the electronic edition of the Astrophysical Journal, A portion is shown here for guidance regarding its form and content. Until publication, an electronic copy of this table is available from the first author.} \\label{table_memprob} \\end{deluxetable} \\normalsize \\begin{table} \\scriptsize \\begin{center} \\begin{tabular}{c||c|c|c|c|c} $K'$ & 14.0-16.0 & 15.0-17.0 & 16.0-18.0 & 17.0-19.0 & 18.0-20.0 \\\\ & & & & \\\\ \\hline \\hline $N$ & 75 & 105 & 135 & 165 & 135 \\\\ & & & & \\\\ \\hline $\\pi_{cl}$ & 0.80 $\\pm$~0.036 & 0.72 $\\pm$~0.035 & 0.72 $\\pm$~0.030 & 0.59 $\\pm$~0.029 & 0.52 $\\pm$~0.032 \\\\ & & & & \\\\ \\hline $\\Delta \\mu$ & 4.28 $\\pm$~0.526 & 4.54 $\\pm$~0.430 & 5.15 $\\pm$~0.356 & 3.68 $\\pm$~0.293 & 3.51 $\\pm$~0.320 \\\\ (mas yr$^{-1}$) & & & & \\\\ \\hline $\\phi_{f}$ & 30.9 $\\pm$~4.8 & 30.5 $\\pm$~3.9 & 37.1 $\\pm$~3.1 & 36.6 $\\pm$~2.5 & 32.5 $\\pm$~2.7 \\\\ ($^o$) & & & & \\\\ \\hline $\\sigma_{a,f}$ & 2.21 $\\pm$~0.338 & 2.56 $\\pm$~0.280 & 2.87 $\\pm$~0.240 & 2.89 $\\pm$~0.198 & 2.80 $\\pm$~0.216 \\\\ (mas yr$^{-1}$) & & & & \\\\ \\hline $\\sigma_{b,f}$ & 1.50 $\\pm$~0.231 & 1.64 $\\pm$~0.193 & 1.85 $\\pm$~0.159 & 1.81 $\\pm$~0.130 & 1.64 $\\pm$~0.137 \\\\ (mas yr$^{-1}$) & & & & \\\\ \\hline $\\sigma_{a,cl}$ & 0.15 $\\pm$~0.013 & 0.17 $\\pm$~0.012 & 0.16 $\\pm$~0.014 & 0.24 $\\pm$~0.022 & 0.45 $\\pm$~0.034 \\\\ (mas yr$^{-1}$) & & & & \\\\ \\hline $\\sigma_{b,cl}$ & 0.12 $\\pm$~0.010 & 0.16 $\\pm$~0.010 & 0.16 $\\pm$~0.012 & 0.16 $\\pm$~0.019 & 0.17 $\\pm$~0.029 \\\\ (mas yr$^{-1}$) & & & & \\\\ \\hline $\\theta_{f}$ & 33.9 $\\pm$~17.3 & 27.8 $\\pm$~14.2 & 35.0 $\\pm$~11.8 & 28.5 $\\pm$~8.8 & 26.7 $\\pm$~10.3 \\\\ ($^o$) & & & & \\\\ \\hline $\\theta_{cl}$ & 70.2 $\\pm$~21.7 & 78.6 $\\pm$~30.3 & 67.5 $\\pm$~48.3 & 117.1 $\\pm$~63.5 & 114.0 $\\pm$~66.4 \\\\ ($^o$) & & & & \\\\ \\hline $\\sigma_{b,cl} / \\sigma_{a,cl}$ & 0.83 $\\pm$~0.086 & 0.91 $\\pm$~0.078 & 0.96 $\\pm$~0.069 & 0.69 $\\pm$~0.075 & 0.37 $\\pm$~0.085 \\\\ & & & & \\\\ \\hline $\\sigma_{b,f} / \\sigma_{a,f}$ & 0.68 $\\pm$~0.130 & 0.64 $\\pm$~0.113 & 0.65 $\\pm$~0.099 & 0.63 $\\pm$~0.083 & 0.58 $\\pm$~0.090 \\\\ & & & & \\\\ \\hline \\end{tabular} \\end{center} \\caption{Fitted kinematic parameters of cluster and field. For each magnitude range, rows give: the cluster fraction, the separation between cluster and field centers in the Vector Point Diagram (VPD), the orientation of the separation vector from the cluster center to the field center in the VPD, the semimajor and minor axes of the field component, the semimajor and minor axes of the cluster component, the orientation of the semimajor axis of the field component, the orientation of the semimajor axis of the cluster component, and finally the axis ratio (minor/major) of the cluster and field components. Errors are estimated from Monte Carlo simulations: populations in the VPD are simulated under the intrinsic kinematic parameters estimated from observation, perturbed by the measured errors for stars in each magnitude range, and re-fitted. Orientations are position-angles reported in degrees East of North.} \\label{tab_kinparams} \\normalsize \\end{table} \\begin{table} \\begin{center} \\begin{tabular}{l|c||c|c||c|c} $K'$ & $N$ & $\\sigma_x$ & $\\sigma_y$ & $\\sigma_x$ & $\\sigma_y$ \\\\ & & (mas yr$^{-1}$) & (mas yr$^{-1}$) & (km s$^{-1})$ & (km s$^{-1}$) \\\\ \\hline 10.0-14.0 & 67 & 0.130 $\\pm$~0.017 & 0.123 $\\pm$~0.016 & 4.912 $\\pm$~0.639 & 4.680 $\\pm$~0.593 \\\\ 14.0-16.0 & 72 & 0.161 $\\pm$~0.019 & 0.129 $\\pm$~0.016 & 6.088 $\\pm$~0.739 & 4.878 $\\pm$~0.606 \\\\ 16.0-18.0 & 107 & 0.177 $\\pm$~0.027 & 0.180 $\\pm$~0.030 & 6.721 $\\pm$~1.034 & 6.839 $\\pm$~1.142 \\\\ 18.0-20.0 & 97 & 0.224 $\\pm$~0.039 & 0.148 $\\pm$~0.046 & 8.508 $\\pm$~1.498 & 5.629 $\\pm$~1.753 \\\\ \\end{tabular} \\end{center} \\caption{Arches velocity dispersion in each co-ordinate. Reading left-right, columns are: Magnitude range of interest, number of cluster stars in this magnitude range, intrinsic velocity dispersion and error in each coordinate, first in mas yr$^{-1}$~and then km s$^{-1}$~assuming the Arches is at 8.4~kpc.} \\label{tab_dispersions} \\end{table} \\include{table8} \\begin{table} \\begin{tabular}{r|c|c|c} $\\Delta \\chi^2_{full}$ & 3.50 & 7.82 & 13.93 \\\\ Confidence & 68\\% & 95\\% & 99.7\\% \\\\ & \"$1\\sigma$\" & \"$2\\sigma$\" & \"$3\\sigma$\" \\\\ \\hline $M(R< 0.40~{\\rm pc})$ & 0.69 - 1.10 & 0.62 - 1.20 & 0.55 - 1.30\\\\ ($10^4~M_{\\odot}$) & & & \\\\ \\hline $M(r < 1.0~{\\rm pc})$ & 1.16 - 1.88 & 1.04 - 2.06 & 0.91 - 2.24\\\\ ($10^4~M_{\\odot}$) & & & \\\\ \\hline $\\rho_0$ & 0.45 - 1.66 & 0.30 - 2.34 & 0.20 - 3.19\\\\ ($10^5 M_{\\odot}~{\\rm pc^{-3}}$) & & & \\\\ \\hline $R_c$ & 0.18 - 0.31 & 0.15 - 0.44 & 0.13 - 0.80\\\\ (pc) & & & \\\\ \\hline $R_t$ & 2.00 - 30.00 & 2.00 - 30.00 & 1.00 - 30.00\\\\ (pc) & & & \\\\ \\hline $M_{cluster}$ & 1.64 - 4.29 & 1.45 - 4.86 & 1.07 - 5.62\\\\ ($10^4~M_{\\odot}$) & & & \\\\ \\hline $1000 \\times \\Sigma_{N,0} / \\rho_0$ & 0.00 - 0.07 & 0.00 - 0.18 & 0.00 - 15.62 \\\\ (stars pc$^{-2} / M_{\\odot}~{\\rm pc}^{-3}$) & & & \\\\ \\hline \\end{tabular} \\caption{Significance regions for isotropic King modeling of the Arches cluster. Ranges of each parameter corresponding to the stated significance level are given, when $R_c, R_t, M_{cluster}$~are all allowed to vary. The quantity $\\chi^2_{full}$~denotes the badness-of-fit when comparing model predictions to both the Arches kinematic dataset and the surface density dataset of Espinoza et al. (2009), over the mass range ($10 \\le M \\le 30$)~$M_{\\odot}$.} \\label{table_massest_10-30_full} \\end{table} \\begin{sidewaystable} \\scriptsize \\begin{tabular}{lcccccccccl} Ref& $R_{in}$& $R_{out}$& $R_{ext}$& $\\Gamma$& $M_{low}$& $M_{obs}$& $M_{calc}$& $\\delta (M_{obs})$& $\\delta(M_{calc})$& Notes \\\\ & (pc) & (pc) & (pc) & & $(M_{\\odot})$ & $(\\times 10^4 M_{\\odot})$ & $(\\times 10^4 M_{\\odot})$ & $(\\times 10^4 M_{\\odot})$ & & \\\\ \\tableline \\tableline 1 & - & 1.15 & - & - & - & 0.08 & 0.24 & - & - & Lower limit on total mass \\\\ 2 & - & 0.35 & - & 1.35 & 2.0 & 0.50 & 1.5 & 0.1 & - & PDMF Salpeter; \\MProjRad \\\\ 2 & - & 0.35 & - & 1.35 & 0.1 & 0.50 & 6.0 & 0.1 & - & PDMF Salpeter; \\MProjRad \\\\ 3 & 0.12 & 0.35 & - & 0.6 &\t 1.0 & 0.51 & 1.08 & - & - & PDMF Top-heavy; \\MProjRad \\\\ 3 & 0.12 & 0.35 & - & 0.6 &\t 0.1 & 0.51 & 1.20 & - & - & PDMF Top-heavy; \\MProjRad \\\\ 4 & - & 0.40 & - & - & 2.0 & 0.63 & - &\t - & - & Rough limit on \\MProjRad~reported \\\\ 5 & -\t & 0.23 & - & - & - & 7.0 & - & - & - & Upper limit on \\MProjRad~from radial velocities \\\\ 6 & - \t & 0.40 & - & 1.1 &\t 1.0 & 0.557 & 2.0 & - & 0.6 & low-mass truncation; \\MProjRad \\\\ 6 & -\t & 0.40 & - & 1.35 & 0.08 & 0.557 & 3.1 & - & 0.6 & Kroupa PDMF; no low-mass truncation; \\MProjRad \\\\ \\tableline 7 & - & 2.50 & - & 0.5 &\t 1.0 & - & 1.60 & - & \t- & N-body; IMF top-heavy; $M_0$~reported \\\\ 7 & - & 2.50 & - & 0.75 & 1.0 & - & 2.00 & - &\t- &\t `` `` \\\\ 7 & - \t& 2.50 & - & 1.0 &\t 1.0 & - & 2.80 & - &\t- &\t `` `` \\\\ 8 & -\t& 1.26 & 3.0 & 2.8 & - & - & 4.00 & - &\t- & N-body; Multi-component IMF \\\\ 9 & -\t& 0.35 & 2.5 & 0.9 &\t 1.3 & - & 4.00 & - &\t- & N-body; Representative $M_0$~reported in Figure caption \\\\ 10 & -\t& 0.35 & 2.5 & 0.9 &\t 0.1 & - & 14.50 & - &\t- & Turbulent-fragmentation calculation \\\\ 11 & -\t& 0.35 & 2.4 & 1.1 & 0.9 &\t- & 5.90 & - &\t- & N-body; Observations reported in Kim et al. (2006) \\\\ 12 & -\t& 0.40 & - & 1.35 & 0.5 & - & 1.80 & - &\t- & N-body, Salpeter IMF; Present-day \\MProjObs \\\\ 12 & -\t& 0.40 & 1.0 & 1.35 & 0.5 & - & 3.60 & - &\t- & Present-day simulated mass within projected radius $R = 1.0$~pc \\\\ 12 & -\t& 0.40 & 2.8 & 1.35 & 0.5 & - & 4.90 & - &\t0.8 & Total initial cluster mass $M_0$~with low-mass truncation \\\\ 12 & -\t& 0.40 & 2.8 & 1.35 & 1.0 & - & 3.60 & - &\t0.6 & Total initial cluster mass $M_0$~with moderate-mass truncation \\\\ 12 & -\t& 0.40 & 2.8 & 1.35 & 4.0 & - & 1.90 & - &\t0.3 & Total initial cluster mass $M_0$~with moderate-mass truncation \\\\ \\tableline \\tableline \\end{tabular} \\caption{All Arches literature mass estimates of which the authors are aware, current as of September 2011. Observational estimates are listed first, followed by cluster mass estimates from models. Reading left-right, columns give: 1: Reference cited. 2,3: Inner and outer radii within which stars were observed $R_{in}, R_{out}$. 4: Radius to which mass function has been extrapolated $R_{ext}$. 5: Mass function slope $\\Gamma$. 6, lower stellar mass used for IMF. 7:Total mass of stars directly observed, $M_{obs}$. 8: Extrapolated mass $M_{calc}$~where appropriate. 9,10: errors (where given) in the observed and extrapolated masses; 11. Brief description. References are: 1. \\citet{cotera96}; 2. \\citet{serabyn98}; 3. \\citet{figer99}; 4. \\citet{stolte02}; 5. \\citet{figer02}; 6. \\citet{espinoza09}. 7. \\citet{kim00}; 8. \\citet{pz02}; 9. \\citet{kim06}; 10. \\citet{dib07}; 11. \\citet{chatterjee09}; 12. \\citet{harfst10}} \\label{table_all_mass_estimates} \\end{sidewaystable}" }, "1112/1112.2752_arXiv.txt": { "abstract": "The spatial distribution of matter in clusters of galaxies is mainly determined by the dominant dark matter component, however, physical processes involving baryonic matter are able to modify it significantly. We analyse a set of 500 pc resolution cosmological simulations of a cluster of galaxies with mass comparable to Virgo, performed with the AMR code RAMSES. We compare the mass density profiles of the dark, stellar and gaseous matter components of the cluster that result from different assumptions for the subgrid baryonic physics and galaxy formation processes. First, the prediction of a gravity only N-body simulation is compared to that of a hydrodynamical simulation with standard galaxy formation recipes, then all results are compared to a hydrodynamical simulation which includes thermal AGN feedback from Super Massive Black Holes (SMBH). We find the usual effects of overcooling and adiabatic contraction in the run with standard galaxy formation physics, but very different results are found when implementing SMBHs and AGN feedback. Star formation is strongly quenched, producing lower stellar densities throughout the cluster, and much less cold gas is available for star formation at low redshifts. At redshift $z=0$ we find a flat density core of radius 10 kpc in both of the dark and stellar matter density profiles. We speculate on the possible formation mechanisms able to produce such cores and we conclude that they can be produced through the coupling of different processes: (I) dynamical friction from the decay of black hole orbits during galaxy mergers; (II) AGN driven gas outflows producing fluctuations of the gravitational potential causing the removal of collisionless matter from the central region of the cluster; (III) adiabatic expansion in response to the slow expulsion of gas from the central region of the cluster during the quiescent mode of AGN activity. ", "introduction": "Clusters of galaxies are the most massive virialised structures observed in the Universe and provide a wonderful laboratory to test astrophysical theories. In the $\\Lambda$CDM cosmological scenario, clusters are assembled via a hierarchy of mergers of less massive structures like galaxies and groups of galaxies. Many physical processes play a role during the formation of a cluster. When satellite galaxies are accreted into a cluster, their properties can be changed by tidal and ram pressure stripping, leading to the formation of a wide variety of galaxy morphologies. Furthermore, clusters are known to be dark matter dominated structures, with most of the baryonic matter residing in a hot diffuse X-ray emitting gaseous phase, the Intracluster Medium (ICM). The stellar mass is less significant and mainly contained in the massive central elliptical galaxy. Since they are dominated by dark matter, this mass component determines the global properties of the mass distribution in the cluster. However, from the theoretical side, it is well known that baryonic processes can produce significant differences in the distribution of matter in collapsed structures with respect to models including only collisionless cold dark matter. For example, baryons are known to condense the centre of dark matter halos due to dissipative processes, producing adiabatic contraction of the total mass distribution \\citep{Gnedin:2004p569}. Several models including baryonic physics have been invoked to solve the so--called cusp/core problem in the $\\Lambda$CDM cosmological framework, i.e. the discrepancy between the centrally cuspy dark matter profiles observed in dark matter halos in numerical N-body simulations and the centrally cored dark matter profiles inferred by observations in dwarf galaxies and low surface brightness galaxies (see the recent review by \\citealp{2010AdAst2010E...5D} and references therein). The study of baryon physics induced modifications in the mass distribution in collapsed structures, and clusters of galaxies in particular, is still a field with many open issues. The present paper is dedicated to the study of the effects of baryonic processes on the mass distribution in clusters of galaxies. In particular, we use a set of cosmological hydrodynamical simulations performed using the AMR code RAMSES \\citep{Teyssier:2002p451} to study the effect of different models for baryons and galaxy formation physics on the mass density profile of a cluster of galaxies comparable to the Virgo cluster. This work can be considered as an extension of the analysis performed by \\cite{2011MNRAS.414..195T} on the same simulations, and is complementary to the analysis presented by \\cite{2012MNRAS.420.2859M}. Here, we focus on the peculiar properties produced in the mass density profile when including Super Massive Black Holes (SMBHs) and the related AGN feedback in the recipes for galaxy formation physics. We stress that AGN feedback was initially introduced to solve the so--called \"overcooling problem\", namely the fact that too much stellar mass is produced in massive collapsed structure in hydrodynamical simulations with respect to what is observed in the real Universe \\citep{Borgani:2009p728}. The strong quenching of star formation produced by processes that couple AGN activity with the gas is expected to improve the match between simulations and observations \\citep{Tabor:1993p1080, Ciotti:1997p1087, Silk:1998p941}. Strong evidence for the existence of AGN feedback is provided by observations of X-ray cavities and radio blobs in galaxy clusters, typically interpreted as buoyantly rising bubbles of high entropy material injected in the central region of clusters by jets of relativistic particles. In this paper, we show that by including SMBH physics and AGN feedback it is possible to obtain interesting predictions on the modifications they can induce on the mass distribution in massive dark matter halos. The paper is organized as follows: the first section is dedicated to the numerical methods and the galaxy formation recipes adopted for our simulations; we show our main results and provide our interpretation in the second section; the last section is left for a short summary of our results and to a discussion. ", "conclusions": "\\label{sec:summary} We used the results of a set of cosmological simulations to study the effect of baryons in massive dark matter halos, focusing on phenomena involving SMBHs and AGN feedback. We use the zoom-in technique to simulate the formation of a galaxy cluster of mass comparable to Virgo at high spatial and mass resolution. In this work we analyse three different simulations: the first is a gravity only run with no baryons (DMO), the second is a hydrodynamical run with standard galaxy formation physics (AGN-OFF), the third is a hydrodynamical run with standard galaxy formation physics plus SMBHs and AGN feedeback (AGN-ON). We adopt a modified version of the \\cite{Booth:2009p501} model to implement thermal AGN feedback in the AGN-ON run. This model allows to reproduce the self-regulated growth of black holes throughout the cosmic ages and it is tuned to reproduce the observed $M_{\\rm BH}-\\sigma$ relation. Our analysis focuses on the evolution of the 3D mass density profiles of the cluster from redshift $z=5$ to redshift $z=0$. We find a number of interesting differences between our models that highlight the importance of accounting for all the physical processes taking place during the formation and evolution of the most massive bound structures in the Universe. Our results can be summarized in the following points: \\begin{itemize} \\item At redshift $z=0$ the dark matter density profiles of the three simulations are consistent with each other at distances $r>10$ kpc from the centre, but they differ significantly in the central region. The DMO profile is consistent with a standard Einasto model. The AGN-OFF profile shows a mass excess with respect to an Einasto fit due to adiabatic contraction of the baryons at the centre of the cluster. The AGN-ON profile presents a flat core of radius 10 kpc, i.e. a mass deficiency with respect to an Einasto fit. The dark matter core forms gradually from $z=5$ to $z=0$. We stress that dark matter cores in dark matter profiles have been claimed to be observed in several clusters \\citep{2004ApJ...604...88S, 2008ApJ...674..711S, 2009ApJ...706.1078N, 2011ApJ...728L..39N, 2011A&A...531A.119R}. \\item The stellar density profile at $z=0$ is also very different between the AGN-OFF and AGN-ON runs. The AGN-OFF profile is very peaked at the centre, while the AGN-ON profile has a core of the same size as that of the dark one. Unlike the dark core, the stellar one forms between $z=1$ and $z=0$. Note that, cores in the surface brightness profiles of massive ellipticals and cluster central galaxies have been observed by several authors \\citep{1999ASPC..182..124K, 2000ApJS..128...85Q, 2003AJ....125..478L, 2004AJ....127.1917T, 2005AJ....129.2138L, 2007ApJ...671.1456C, 2009ApJS..182..216K, 2011arXiv1108.0997G}, so our model is not in disagreement with observations and it {\\it predicts} the existence of dark matter cores associated to stellar cores in massive ellipticals. Furthermore, in the AGN-OFF run the stellar density is larger than in the AGN-ON run at all radii. This is a clear effect of overcooling of gas leading to enhanced (and too efficient) star formation. In the AGN-ON run, AGN feedback strongly quenches star formation and the overcooling problem is avoided \\citep{2011MNRAS.414..195T,2012MNRAS.420.2859M}. The result is also that less cool gas is available at the centre of the AGN-ON cluster with respect to the AGN-OFF run: gas is heated by AGN feedback and carried away via convective motions and slow adiabatic expansion. \\item The core in the dark matter end stellar mass density profiles is a peculiar feature of our simulation that includes the physics associated with active galactic nuclei. In this paper, we suggest that the coupling of several mechanisms is responsible for their formation. First, SMBHs transfer part of their orbital energy to collisionless matter via dynamical friction during dry galaxy mergers \\citep{2010ApJ...725.1707G}, especially at redshift $z<1$. Second, AGN driven gas outflows modify the gravitational potential in regions close to SMBHs with resulting ejection of collisionless matter from the central region of the cluster; subsequent gas outflows followed by the central 'revirialisation' of the central material are expected to produce cores \\citep{2011arXiv1106.0499P}; due to the stronger AGN feedback at $z>1$, this mechanism is more effective at high redshift, but preserves an important role at low redshift. Third, the central hot gas slowly cools radiatively, falling onto the SMBHs in convective flows and is subsequently ejected impulsively; the slow loss of mass from the central region will produce an expansion of the inner mass distribution. \\end{itemize} Despite assertions that AGN feedback can affect central cluster dynamics, in fact our results seem to be complementary to the those obtained by other authors for less massive dark matter halos \\citep{Governato:2010p1442, 2011arXiv1106.0499P, 2011arXiv1111.5620M}. Related processes involving baryons seem to be active at the low and high mass end of the galaxy mass function, especially mechanisms involving gas outflows produced by feedback. Given the fact that we perform the analysis on zoom simulations of one cluster, we need to stress that our result needs support from a suite of dedicated cosmological simulation aimed at exploring the effect of SMBHs, AGN feedback and baryon physics in general on the most massive clusters in the universe. Furthermore, a number of numerical experiments is also needed to explore in detail the core formation processes that have been considered in our discussion, with particular attention on their coupling. {The role of baryon mass outflows in galaxy clusters has been recently studied by \\cite{2012arXiv1202.1527R} using collisionless matter simulations where gas is modeled as a time varying external contribution to the gravitational potential; their results provide important support to ours.} To conclude, we are convinced that our results are robust enough to assert that the implementation of AGN feedback and SMBHs in cosmological hydrodynamical simulations is an important ingredient for modeling massive clusters of galaxies." }, "1112/1112.4338_arXiv.txt": { "abstract": "Stellar wind-emission features in the spectrum of eta Carinae have decreased by factors of 1.5--3 relative to the continuum within the last 10 years. We investigate a large data set from several instruments (STIS, GMOS, UVES) obtained between 1998 and 2011 and we analyze the progression of spectral changes in the direct view of the star, in the reflected polar-on spectra at FOS4, and at the Weigelt knots. We find that the spectral changes occurred gradually on a time scale of about 10 years and that they are dependent on the viewing angle. The line strengths declined most in our direct view of the star. About a decade ago, broad stellar wind-emission features were much stronger in our line-of-sight view of the star than at FOS4. After the 2009 event, the wind-emission line strengths are now very similar at both locations. High-excitation \\ion{He}{1} and \\ion{N}{2} absorption lines in direct view of the star strengthened gradually. The terminal velocity of Balmer P Cyg absorption lines now appears to be less latitude-dependent and the absorption strength may have weakened at FOS4. Latitude-dependent alterations in the mass-loss rate and the ionization structure of eta Carinae's wind are likely explanations for the observed spectral changes. ", "introduction": "\\label{sec:intro} Eta Carinae, one of the most massive and most luminous stars in our Galaxy, is famous for its Great Eruption about 170 years ago. Its recovery has been unsteady with unexplained photometric and spectral changes in the 1890s and 1940s (\\citealt{2008AJ....135.1249H}, and references therein). The spectral changes described in this paper may represent another rapid step in \\ec's recovery from its Great Eruption. Eta Car has a complex spectroscopic cycle, most likely regulated by a companion star in an eccentric orbit (\\citealt{1997NewA....2..107D}, and many references in \\citealt{2005ASPC..332.....H} and \\citealt{2012eta}). So-called spectroscopic events occur every 5.54 years since 1948 \\citep{2001MNRAS.322..741F,1996ApJ...460L..49D,2008MNRAS.384.1649D}. The events are characterized by drastic changes in \\ec's spectrum and photometry, e.g., high-excitation emission lines disappear for a few months (e.g., \\citealt{1953ApJ...118..234G,1984A&A...137...79Z}) and light curves at all wavelength regions show significant variations (e.g., \\citealt{1994MNRAS.270..364W,1997Natur.390..587C,2001MNRAS.322..741F,2006JAD....12....3V,2009A&A...493.1093F,2004AJ....127.2352M}). In a previous paper \\citep{2010ApJ...717L..22M} we compared spectra at corresponding phases of successive spectroscopic cycles and found dramatic changes in observations after the 2009 event.\\footnote{ We define ``phase'' by $P = 2023.0$ days and $t_0 =$ MJD 50814.0 = J1998.00, consistent with the Eta Carinae Treasury Program Archive (http://etacar.umn.edu/). Phases 0.00, 1.00, and 2.00 mark the 1998.0, 2003.5, and 2009.0 spectroscopic events. } Major stellar-wind emission features in the spectrum of \\ec\\ had decreased by factors of order 2 relative to the continuum within 10 years and helium P Cyg absorption had become stronger. Most of the broad emission lines in \\ec's spectrum originate in the primary star's wind, see many papers and refs.\\ in \\citet{2005ASPC..332.....H}, and the simplest explanation for the observed spectral changes is a decrease in \\ec's wind density, by a factor of 2 or more. The early exit from \\ec's 2009 X-ray minimum and the observed decrease of the 2--10 keV photons over the last two cycles are consistent with this interpretation \\citep{2009ApJ...701L..59K,2010ApJ...725.1528C,2011ApJ...740...80M}. In this paper we analyze spectra obtained between 1998 and 2011 with several instruments to investigate in detail spectral changes in \\ec's wind. We are not concerned here with the temporary spectral changes observed during the events -- the spectral changes discussed are of secular nature. In \\citet{2010ApJ...717L..22M} we noted only a few examples; here we explore a wider range of effects, and whether or not they have developed gradually as opposed to sporadically. Section {\\ref{sec:obs}} describes the observations. In Section {\\ref{sec:change}} we confirm the observations made by \\citet{2010ApJ...717L..22M} and show that the broad stellar wind features were still weak in {\\it HST\\/} STIS data obtained several months after our initial discovery in 2010 March data. The temporal progression of spectral changes and the dependence on the viewing direction is discussed. High-excitation emission and continuum from the nearby Weigelt knots, which are thought to be photoionized by a hot companion star, reveal additional information. In Section {\\ref{sec:massloss}} we discuss the implications of these observations and estimate the decrease in mass-loss rate over the last 10 years. In Section {\\ref{sec:discussion}} we give a short summary. ", "conclusions": "\\label{sec:discussion} In this paper we analyzed spectral data obtained with several instruments between 1998 and 2012. We confirmed the spectral changes in the wind emission lines first reported by \\citet{2010ApJ...717L..22M}; {\\it HST} STIS spectra obtained in 2010 August, $\\sim 170$ days after the first discovery, are comparable to the observations in 2010 March. Furthermore, we analyzed the long-term development of spectral changes in our direct line-of-sight view of the star, at FOS4, and the Weigelt knots. Eta Car's recent spectral changes involve both emission and absorption lines: \\begin{enumerate} \\item Broad stellar wind-emission features in our line-of-sight to the star have decreased by factors of 1.5--3 relative to the continuum within the last 10 years. These changes occurred gradually and are dependent on the viewing angle; spectra at higher stellar latitudes and from the outlying ejecta show smaller changes. The simplest explanation is a decrease in $\\eta$ Car's primary wind density. However, the decrease in wind density appears to be latitude dependent, with emission features showing much less change at higher latitudes. After the 2009 event, emission line strengths are now very similar in our direct line-of-sight view and in the reflected polar-on spectrum at FOS4 suggesting a more spherical wind and/or a more uniform distribution of circumstellar extinction. \\item High-excitation \\ion{He}{1} and \\ion{N}{2} absorption lines strengthened gradually over the last decade indicating a change in \\ec's wind ionization structure. Hydrogen P Cyg absorption at FOS4 might have weakened after the 2009 event. The terminal velocity of hydrogen P Cyg lines was found to be similar at all stellar latitudes. Those findings provide additional clues for a more spherical wind. \\end{enumerate} The observational results presented here are difficult to reconcile with a decrease in mass-loss rate primarily at lower stellar latitudes since it is generally assumed that \\ec's wind had higher densities at the poles \\citep{2003ApJ...586..432S}. Our observations may be more readily reconciled with alternative explanations for latitude-dependent spectral features, such as a complex ionization structure of \\ec's wind modulated by the secondary star's UV radiation \\citep{2010AJ....139.1534R} or the presence of a wind cavity in the primary wind caused by the secondary star \\citep{2010ApJ...716L.223G,2011arXiv1111.2280M}. Using H$\\alpha$ emission and the method by \\citet{1988ApJ...326..356L} we found that \\ec's mass-loss rate decreased by a factor of 2--3 between 1999 and 2010. A decrease in mass-loss rate on the order of 2--3 is consistent with changes in the X-ray light curve \\citep{2009ApJ...701L..59K,2010ApJ...725.1528C}. We did not attempt to derive the absolute value with any accuracy because there are too many unknown factors, such as latitudinal dependence and clumping of the wind. New theoretical models updating \\cite{2001ApJ...553..837H} are needed. Observations in 2012 and 2013 will be extremely valuable to further analyze the nature of the spectral changes in \\ec's wind. It is of great importance to monitor the star consistently since spectral changes may occur on time scales of only weeks to months. For the long-term recovery of \\ec\\ it is important to investigate if the wind will further decline or if it will stabilize or even recover to its former strength. But by mid-2013, the onset of the next event will dominate the spectrum, so observations in 2012 are needed. The last three events all differed from each other and considering the long-term spectral changes described in this paper we can expect many interesting new results from \\ec's 2014.5 event. {\\it Acknowledgement} We thank the staff and observers of the Gemini-South Observatory in La Serena for their help in preparing and conducting the observations, and Beth Perriello at STScI for assistance with {\\it HST\\/} observing plans. We also thank Otmar Stahl and Kerstin Weis for their effort in planning and obtaining the UVES spectra. AM was co-funded under the Marie Curie Actions of the European Commission (FP7-COFUND). MTR received partial support from Center for Astrophysics FONDAP and PB06 CATA (CONICYT)." }, "1112/1112.3342_arXiv.txt": { "abstract": "We present results from a suite of axisymmetric, core-collapse supernova simulations in which hydrodynamic recoil from an asymmetric explosion produces large proto-neutron star (PNS) velocities. We use the adaptive-mesh refinement code CASTRO to self-consistently follow core-collapse, the formation of the PNS and its subsequent acceleration. We obtain recoil velocities of up to 620 km\\,s$^{-1}$ at $\\sim$1 s after bounce. These velocities are consistent with the observed distribution of pulsar kicks and with PNS velocities obtained in other theoretical calculations. Our PNSs are still accelerating at several hundred km\\,s$^{-1}$ at the end of our calculations, suggesting that even the highest velocity pulsars may be explained by hydrodynamic recoil in generic, core-collapse supernovae. ", "introduction": "At birth, pulsars achieve velocities well above those of their progenitor population \\citep{Gunn:1970ys,Lyne:1994rr}. These pulsar ``kicks\" typically range from 200 km\\,s$^{-1}$ to 400 $\\rm km$\\,$\\rm s^{-1}$, with the fastest neutron stars achieving velocities near, or in excess of, 1000 km\\,s$^{-1}$ \\citep{Lyne:1994rr,Chatterjee:2005uq,Hobbs:2005fk}. The current distribution of observed pulsar velocities is Maxwellian, hinting at a common acceleration mechanism \\citep{Hansen:1997dq,Hobbs:2005fk,Zou:2005yq,Faucher-Giguere:2006lr}. Many scenarios have been proposed for the origin of pulsar recoil and neutron star kicks. Popular mechanisms often require strongly magnetized systems, exotic neutrino physics, and/or rapid rotation to produce substantial kicks. For example, in the presence of strong magnetic fields, neutrino-matter interactions can generate neutron star velocities on the order of a few hundred km\\,s$^{-1}$ by producing $\\sim$1\\% dipole asymmetries \\citep{Lai:1998tg,Nardi:2001hc,Lai:2001ij,Kusenko:1999kx,Lambiase:2005yq,Barkovich:2004vn,Fuller:2003yq,Kishimoto:2011fk}. Many of these scenarios require magnetic fields in the magnetar range (i.e.~10$^{14-16}$ G) and may not produce substantial kicks in typical core-collapse supernovae. Other scenarios involve jet/counter-jet misalignment launched near the proto-neutron star (PNS). In such situations, the jets accompany an associated gamma-ray burst (GRB) or form through magneto-rotational processes during core collapse \\citep{Cen:1998oq,Khokhlov:1999qy,Sawai:2008kx,Papish:2011qy}. These scenarios require rapid rotation and therefore, may only manifest in a small subset of core-collapse events. If neutron star kicks are a generic outcome of core collapse, then a natural explanation is recoil during an asymmetric supernova explosion. In the current, most sophisticated simulations, the bounce shock, launched when the equation of state stiffens at nuclear densities, stalls due to thermal energy losses from neutrino emission and dissociation of nuclei into nucleons. The stalled shock itself is subject to hydrodynamic and neutrino-driven instabilities, which manifest as prominent low-order $\\ell$-mode oscillations in axisymmetric simulations of non-rotating progenitors \\citep{Blondin:2003ul,Scheck:2004rt,Buras:2006qf,Scheck:2006vn,Burrows:2007eu,Blondin:2007mz,Ott:2008cr,Fernandez:2009uq,Nordhaus:2010fr,Nordhaus:2010kx,Fernandez:2010ly,Brandt:2011fj}. At the onset of shock revival, the PNS may recoil if large-scale asymmetries are present during the ensuing supernova explosion. While the mechanism by which core-collapse supernova progenitors explode is not fully understood, the most probable scenario involves absorption of neutrinos in the post-shock ``gain region\" \\citep{Bethe:1985fr} and likely requires the development of multi-dimensional instabilities in fully three-dimensional radiation-hydrodynamic simulations to succeed \\citep{Nordhaus:2010fr}. Nonetheless, recoil from an asymmetric neutrino-driven explosion presents a natural mechanism by which PNSs achieve high velocities \\citep{Burrows:1996th,Scheck:2004rt,Scheck:2006vn,Nordhaus:2010kx,Wongwathanarat:2010yq} and appears to be supported by recent X-ray observations of the Cassiopeia A supernova remnant \\citep{Hwang:2011lr}. Computational studies of recoil require multi-dimensional, radiation-hydrodynamics calculations which start at the onset of collapse and follow the dynamics self-consistently. This includes the formation of the PNS, the explosion, the propagation of the shock front through the stellar envelope and eventually, decoupling of the PNS from the surrounding material. Such an approach is computationally challenging and as such, various techniques have been adopted to make the computations tractable. One popular approach is to commence the calculations after bounce by mapping spherically symmetric solutions onto a multi-dimensional grid and excising the PNS from the computational domain \\citep{Scheck:2004rt,Scheck:2006vn,Wongwathanarat:2010yq}. This requires one to infer a PNS kick through a rigid, impenetrable inner boundary, but allows one to track the supernova explosion for several seconds and to distances greater than 10,000 km. While this approach is appealing, it must be checked by simulations which include the PNS in the computational domain. Recently, we have carried out the first axisymmetric, radiation-hydrodynamic simulation of recoil with the multi-group, arbitrary, Lagrangian-Eulerian code {\\sc VULCAN/2D} \\citep{Nordhaus:2010kx}. By transitioning from a spherical-polar mesh to a pseudo-Cartesian mesh at the center of the domain, we self-consistently tracked the PNS's formation and off-center motion. This calculation was computationally expensive, as it employed multi-group flux limited diffusion neutrino transport and followed the supernova explosion until the shock reached the 5,000 km radial outer boundary of the domain. At that time, the PNS had reached a velocity of $\\sim$150 km\\,s$^{-1}$ but had yet to fully decouple from the ejecta. The PNS recoil was due almost entirely to hydrodynamical processes and was consistent with previous excised-core calculations \\citep{Scheck:2004rt,Scheck:2006vn}. Asymmetric neutrino emission contributed $\\sim$2\\% to the overall kick magnitude. This suggests that neutrinos play no significant role in accelerating neutron stars to high velocities during typical core-collapse supernovae. At the end of the calculation, significant acceleration ($\\sim$350 km\\,s$^{-2}$), in addition to the degree of asymmetry in the ejecta, suggested that higher PNS velocities were possible. Verifying these estimates requires tracking the supernova shock to greater radial distances and later times. To expand upon the work of \\cite{Nordhaus:2010kx}, we carry out a suite of axisymmetric collapse calculations with the adaptive-mesh-refinement (AMR), radiation-hydrodynamics code, CASTRO \\citep{Almgren:2010fk,Zhang:2011lr}. By employing AMR and a simplified form of transport (see Sec. \\ref{2}), we expand the domain to a radial distance of 10,000 km and perform multiple calculations. Our PNSs achieve recoil velocities ranging from tens of km\\,s$^{-1}$ up to $\\sim$620 km\\,s$^{-1}$. In general, the magnitude of the recoil depends on the degree of asymmetry at the time of explosion and the energy of the explosion itself \\citep{Burrows:2007ly}. In Sec.~3, we discuss the physical processes that accelerate the PNS. In Sec. \\ref{4}, we compare our results with previous work before concluding and discussing future work in Sec. \\ref{5}. ", "conclusions": "} We have carried out a suite of axisymmetric simulations of the collapse of a massive star's core with the AMR, radiation-hydrodynamic code {\\sc CASTRO}. For each calculation, we follow the core collapse, PNS formation, ensuing neutrino-driven supernova explosion and PNS recoil. By incorporating the effects of neutrino heating and cooling in place of more detailed and computationally expensive neutrino transport, we are able to perform multiple calculations that simultaneously follow the evolution of the PNS and the global explosion for $\\sim$1 second and to distances of $\\sim$10,000 km. The PNSs in our simulations achieved recoil velocities comparable to the those of observed pulsars. After $\\sim$1 second of post-bounce evolution, the highest PNS velocity obtained was 620 km\\,s$^{-1}$ (model $\\rm L_{2.3}$). After $\\sim$0.6 seconds of post-bounce evolution, this acceleration was supplied primarily by the gravitational pull of slow-moving ejecta in front of the PNS. This gravitational effect dominates the late-time PNS acceleration in all of our calculations. While our PNSs have started to decouple from the surrounding fluid (see Fig.~\\ref{fig:accel_3panel}), the substantial and ongoing gravitational acceleration suggests that higher velocities may ultimately be achievable. Our results suggest that hydrodynamic recoil during an asymmetric supernova explosion provides a natural explanation for pulsar kicks. After the bounce shock stalls, hydrodynamic instabilities deform the shocked material and ensure that the ensuing explosion is asymmetric. Recoil during the supernova explosion and gravitational interaction with the expanding ejecta subsequently accelerate the PNS to high velocities. The results presented in this work are consistent with the findings of \\cite{Nordhaus:2010kx} and previous axisymmetric simulations that excised the PNS from the computational domain \\citep{Scheck:2004rt,Scheck:2006vn}. Taken together, these studies strongly suggest that generic core-collapse supernovae can accelerate neutron stars to the high velocities observed in the pulsar population. Additionally, these studies demonstrate that hydrodynamic processes, and not asymmetric neutrino emission, are responsible for this acceleration \\citep{Scheck:2006vn,Burrows:2007ly,Nordhaus:2010kx}. In fact, recent simulations of neutron star kicks in three dimensions suggest that velocities comparable to those from axisymmetric calculations are achievable \\citep{Wongwathanarat:2010yq}. In this work, we have provided substantial numerical support to the hydrodynamic mechanism of pulsar kicks. Recoil due to a neutrino-driven, core-collapse supernova explosion provides a natural explanation for pulsar kicks without appealing to more exotic scenarios. As computational methods and resources improve, self-consistent three-dimensional studies will enable a full comparison of theoretical models to observed distributions of pulsar kicks and spins." }, "1112/1112.2728_arXiv.txt": { "abstract": "A dedicated mission to investigate exoplanetary atmospheres represents a major milestone in our quest to understand our place in the universe by placing our Solar System in context and by addressing the suitability of planets for the presence of life. EChO -- the Exoplanet Characterisation Observatory -- is a mission concept specifically geared for this purpose. EChO will provide simultaneous, multi-wavelength spectroscopic observations on a stable platform that will allow very long exposures. The use of passive cooling, few moving parts and well established technology gives a low-risk and potentially long-lived mission. EChO will build on observations by Hubble, Spitzer and ground-based telescopes, which discovered the first molecules and atoms in exoplanetary atmospheres. However, EChO's configuration and specifications are designed to study a number of systems in a consistent manner that will eliminate the ambiguities affecting prior observations. EChO will simultaneously observe a broad enough spectral region -- from the visible to the mid-infrared -- to constrain from one single spectrum the temperature structure of the atmosphere, the abundances of the major carbon and oxygen bearing species, the expected photochemically-produced species and magnetospheric signatures. The spectral range and resolution are tailored to separate bands belonging to up to 30 molecules and retrieve the composition and temperature structure of planetary atmospheres. The target list for EChO includes planets ranging from Jupiter-sized with equilibrium temperatures $T_{eq}$ up to 2000 K, to those of a few Earth masses, with $T_{eq}$ $\\sim$300~K. The list will include planets with no Solar System analog, such as the recently discovered planets GJ1214b, whose density lies between that of terrestrial and gaseous planets, or the rocky-iron planet 55 Cnc~e, with day-side temperature close to 3000~K. As the number of detected exoplanets is growing rapidly each year, and the mass and radius of those detected steadily decreases, the target list will be constantly adjusted to include the most interesting systems. We have baselined a dispersive spectrograph design covering continuously the 0.4--16\\,$\\mu$m spectral range in 6 channels (1 in the visible, 5 in the InfraRed), which allows the spectral resolution to be adapted from several tens to several hundreds, depending on the target brightness. The instrument will be mounted behind a 1.5 m class telescope, passively cooled to 50~K, with the instrument structure and optics passively cooled to $\\sim$45~K. EChO will be placed in a grand halo orbit around L2. This orbit, in combination with an optimised thermal shield design, provides a highly stable thermal environment and a high degree of visibility of the sky to observe repeatedly several tens of targets over the year. Both the baseline and alternative designs have been evaluated and no critical items with Technology Readiness Level (TRL) less than 4 to 5 have been identified. We have also undertaken a first-order cost and development plan analysis and find that EChO is easily compatible with the ESA M-class mission framework. ", "introduction": " ", "conclusions": "" }, "1112/1112.0025_arXiv.txt": { "abstract": "{Observations of the $\\gamma$-ray sky with {\\textit Fermi} led to significant advances towards understanding blazars, the most extreme class of Active Galactic Nuclei. A large fraction of the population detected by {\\textit Fermi} is formed by BL~Lacertae (BL~Lac) objects, whose sample has always suffered from a severe redshift incompleteness due to the quasi-featureless optical spectra.} {Our goal is to provide a significant increase of the number of confirmed high-redshift BL~Lac objects contained in the 2~LAC {\\textit Fermi}/LAT cataloge.} {For 103 {\\textit Fermi}/LAT blazars, photometric redshifts using spectral energy distribution fitting have been obtained. The photometry includes 13 broad-band filters from the far ultraviolet to the near-IR observed with {\\it Swift}/UVOT and the multi-channel imager GROND at the MPG/ESO 2.2m telescope. Data have been taken quasi-simultaneously and the remaining source-intrinsic variability has been corrected for.} {We release the UV-to-near-IR 13-band photometry for all 103 sources and provide redshift constraints for 75 sources without previously known redshift. Out of those, eight have reliable photometric redshifts at $z\\gtrsim1.3$, while for the other 67 sources we provide upper limits. Six of the former eight are BL~Lac objects, which quadruples the sample of confirmed high-redshift BL~Lac. This includes three sources with redshifts higher than the previous record for BL~Lac, including CRATES~J0402-2615, with the best-fit solution at $z\\approx1.9$.} {} ", "introduction": "Since its launch in 2008, the {\\it Fermi} Space Laboratory has dramatically extended our view of the high-energy sky. The recently released 24-month catalog (2LAC) of Active Galactic Nuclei (AGN) detected by the Large Area Telescope \\citep[LAT; ][]{Atwood:2009aa} revealed 885 high-significance sources, the large majority of them being blazars \\citep{Ackermann:2011lr}. The latter form the most extreme class of AGN with their observational characteristics governed by the small angle between their relativistic jets and the observer's sight line \\citep{Blandford:1978lr}. The resulting Doppler boosting makes blazars exceptionally bright sources at nearly all wavelengths and therefore visible out to high redshift. The scientific relevance of blazars is very broad, ranging from laboratories for the physics and structure of relativistic jets (and thus the extraction of energy from the central massive black hole) \\citep[e.g.,][]{Abdo:2010ab} to probes of the extra-galactic background light (EBL) through attenuation of $\\gamma$-ray photons \\citep[e.g.,][]{Abdo:2010ac}. One of the crucial parameters for these applications is the distance to the objects, which unfortunately is not easy to obtain in most cases. Two classes of blazars dominate the 2LAC population. These are the flat-spectrum radio quasars (FSRQs, 310 sources) and BL~Lac objects (395), named after the prototype BL~Lacertae \\citep[][]{Hoffmeister:1929lr}. While for the former redshift measurements are routinely performed using their strong emission lines at UV-optical wavelengths, the featureless, power-law optical spectra of BL~Lac objects have proven to be a challenge \\citep[e.g.,][]{Shaw:2009fk}. Indeed, 220 of the 395 BL~Lacs in the 2LAC (55\\,\\%) lack redshift estimates. Until this incompleteness is resolved, conclusions about the EBL, the blazar sequence \\citep[e.g.,][]{Fossati:1998lr,Ghisellini:1998fk}, and the blazar population in general remain tentative at best. Several methods have been exploited to increase the BL~Lac redshift sample. At low distance, one can utilize the remarkably uniform absolute brightness of the giant elliptical BL~Lac host galaxies \\citep{Sbarufatti:2005qy,Meisner:2010uq} and very-high signal-to-noise optical spectroscopy to help to identify weak emission or absorption features in a few other cases \\citep[e.g.,][]{Shaw:2009fk}. An alternative method, applicable to the more distant sources, is the photometric redshift technique, which consists of fitting spectral energy distribution (SED) templates to multi-band photometry. Neutral hydrogen along the line of sight to the blazar will imprint a clear attenuation signature at the Lyman limit and thus allows an accurate estimate of the redshift of the absorber. Even though the absorber will be located somewhere along the line of sight and its redshift will thus not necessarily correpond to that of the blazar, photometric redshifts, $z_{\\rm phot}$, will provide a reliable lower limit on the blazar redshift. In this paper we explore the use of ultra-violet to near-infrared quasi-simultaneous photometry to measure the redshift of 103 2LAC blazars, 86 of them without previous redshift constraints. The sample selection is described in Sect.~\\ref{sec:sample} while the observations and data reduction are detailed in Sect.~\\ref{sec:obs} and Sect.~\\ref{sec:data}. The results and their discussion are presented in Sect.~\\ref{sec:results} and Sect.~\\ref{sec:discussion}, respectively. ", "conclusions": "\\label{sec:discussion} In this paper we presented redshift constraints for 103 blazars from the 2LAC catalog using UV-to-near-IR multi-band photometry obtained quasi-simultaneously with {\\it Swift}/UVOT and GROND. We provided the first reliable redshift measurements for eight sources and new upper limits for an additional 66 targets. Of the eight sources with reliable redshift, seven are located at $z_{\\rm phot}>1.3$. Six of those belong to the BL Lac population. For comparison, out of the total 395 BL Lac in the 2LAC sample only two sources have previously been known to lie at $z>1.3$ (Fig.~\\ref{fig:zDis}). Redshifts for these two, \\object{PKS~0332-403} and \\object{CRATES~J1312-2156}, have also been confirmed with our photometric observations. \\begin{figure} \\centering \\includegraphics[width=0.5\\textwidth]{zDis.eps} \\caption{Redshift distribution for the nine BL Lac objects with reliable photometric redshifts (blue filled histogram) together with the distribution for the BL~Lac in the 2LAC catalog \\citep[empty histogram;][]{Ackermann:2011lr}. Sources with $z<0.5$ have been omitted for clarity.} \\label{fig:zDis} \\end{figure} The six new BL Lac redshifts at $z>1.3$ represent a dramatic (from two to eight) increase of the confirmed high-z {\\it Fermi} sample and thus demonstrate the opportunity that the SED template fitting technique holds for obtaining photometric redshifts for BL Lac sources. For our sample, this was possible due to the densely-covered wide spectral range (160--2200\\,nm), necessary for a reliable constraint of the spectral slope, the excellent ultra-violet coverage to measure the Lyman-limit, and the quasi-simultaneity of the observations, important for minimizing the impact of source-intrinsic variability. The method applied in this work overcomes significant challenges inherent in other redshift techniques, namely the simplicity of the optical emission of BL Lacs manifested as the power-law-shaped synchrotron spectral component. This complicates spectroscopic redshift measurements, which are in most cases limited to the optical wavelength regime and thus insensitive to the Lyman-limit. Instead, spectroscopy relies on the detection of very faint emission features, mainly from the underlying host galaxy. It therefore requires very high signal-to-noise, often at the cost of long exposure times. On the other hand, the photometric redshift method for BL Lac does not require significant telescope time, in particular when data from efficient multi-channel instruments like GROND are available. Its weakness, however, is that the method relies on the detection of a particular spectral feature, which, albeit strong, is located in the rest-frame far-ultra-violet. For $z\\lesssim3$, this is only accessible with space-based ultra-violet telescopes, and even then usually limited to redshifts above $z\\approx1.2$, as shown above. The sensitivity of this method to high-redshift sources also allows the placement of upper limits for those SEDs that do not show the imprint of the Lyman-limit. Only for three targets is this boundary above $z=2$. In one case, \\object{CRATES~J0250+1708} ($z<3.1$), the optical counterpart is too faint ($r^\\prime\\approx20.7$\\,mag) and only upper limits in all six UVOT bands could be obtained. The counterpart of \\object{CRATES~J0705-4847} is only detected in the GROND $J$ ($\\approx19.7$\\,mag$_{\\rm AB}$) and $H$ bands and its SED and redshift are thus poorly constrained. Finally, the SED of \\object{ATG20~J0124-0625} ($z<2.46$) shows a significant break at $z\\sim1.9\\pm0.5$. However, as the redshift probability distribution is broad, and the $\\chi^2$ of the power law fit comparable to that of a $z\\approx0.4$ solution for a galaxy template, the photometric redshift is considered to be unreliable. Except for those three, no other source in our sample has a best-fit photometric redshift at $z>2$. Thus, \\object{CRATES~J0402-2615} can now be considered the most distant known BL Lac with a measured redshift of $z\\approx1.92$. Fig.~\\ref{fig:zDis} indicates that our high-redshift findings are the natural extension of the existing 2LAC redshift sample. However, due to the incompletenesses of both samples, we refrain from drawing any quantitative conclusions at this stage. We note, however, the our result is general agreement with the theoretical predictions from \\cite{Giommi:2011lr}. A more detailed physical interpretation of the establishment of an increased fraction of high-z BL Lacs in the 2LAC sample will be reported separately." }, "1112/1112.0355_arXiv.txt": { "abstract": "In accretion systems, outflows may have significant influence on the luminosity fluctuations. In this paper, following the Lyubarskii's general scheme, we revisit the power spectral density of luminosity fluctuations by taking into account the role of outflows. Our analysis is based on the assumption that the coupling between the local outflow and inflow is weak on the accretion rate fluctuations. We find that, for the inflow mass accretion rate $\\dot M \\propto r^{s}$, the power spectrum of flicker noise component will present a power-law distribution $p(f) \\propto f^{-(1+4s/3)}$ for advection-dominated flows. We also obtain descriptions of $p(f)$ for both standard thin discs and neutrino-cooled discs, which show that the power-law index of a neutrino-cooled disc is generally larger than that of a photon-cooled disc. Furthermore, the obtained relationship between $p(f)$ and $s$ indicates the possibility of evaluating the strength of outflows by the power spectrum in X-ray binaries and gamma-ray bursts. In addition, we discuss the possible influence of the outflow-inflow coupling on our results. ", "introduction": "The emission of Galactic Black Hole Binaries (BHBs) and active galactic nuclei (AGN) displays a significant aperiodic variability on a broad range of time-scales. The Power Spectral Density (PSD) of such variability is generally modeled with a power law, $p(f) \\propto f^{-\\beta}$, where $p(f)$ is the power at frequency $f$, and the power-law index ${\\beta}$ keeps a constant in a certain range of $f$, but changes among different ranges. At high frequencies, the PSDs of both BHBs and AGN present a steep slope with $\\beta \\sim 2$. On the contrary, below a break frequency, typically at a few Hz for BHBs, they flatten to a slope with $\\beta \\sim 1$, representing the flicker noise (see \\citealp{King04} and references therein). Several models have been proposed in order to understand this nearly featureless character of power spectra. The so-called ``shot noise models\" (\\citealp{Terrell72}) attempted to describe the light curves as a series of independent overlapping shots with specific time-scales, amplitudes, and occurence rates. Due to lack of physical picture in this scenario, various physically motivated ideas have been put forward subsequently, such as the fluctuations of hydrodynamic or magnetohydrodynamic turbulence (\\citealp{Nowak95}; \\citealp{Hawley01}), magnetic flares or density fluctuation in the corona (\\citealp{Galeev79}; \\citealp{Poutanen99}; \\citealp{Goosmann06}; \\citealp{Kawanaka08}), and Lyubarskii's general scheme (\\citealp{Lyubarskii97}; \\citealp{King04}). In the Lyubarskii's scheme, it was noted that any variation of accretion rate, which is caused by small amplitude variations in the viscosity, would induce a variation in the accretion rate at the inner radius of the disc, where most of the energy is released. Moreover, observations showed that the variability is non-linear and the rms variability is proportional to the average flux over a wide range of time-scales (e.g., \\citealp{Uttley01}; \\citealp{Uttley05}; \\citealp{Gleissner04}). It indicates that the short time-scale variations are modulated by the longer time-scales, which favors the Lyubarskii's scheme. However, the observed power spectra often deviate from the form $f^{-1}$. For example, the PSD of Cyg X-1 is well described with the form $f^{-1}$ in the soft state, however, exhibits the form $f^{-1.3}$ in the hard state (\\citealp{Gilfanov10}). In particular, it is shown that the power-law index is around $0.8-1.3$ both in the soft state of BHBs and in narrow-line Seyfert 1 galaxies (\\citealp{Janiuk07}). Such a dispersion of the power-law index reveals that there must exist some other mechanism. A radius-dependent amplitudes of $\\alpha$ fluctuations may help to alleviate the discrepancy between theories and observations. However, it remains unclear why the fluctuations of $\\alpha$ should have a strong radius-dependent form. In the present paper, we will take into account another mechanism, outflows, which is a popular phenomenon in accretion systems and has strong observational evidence. One of the best examples comes from Sgr A*, whose center harbors a supermassive black hole surrounded by an accretion flow that is likely to be in the form of the advection-dominated accretion flow (ADAF, \\citealp{Narayan94}). Radio polarization observations constrain the accretion rate in the innermost region is nearly two orders of magnitude lower than that measured at the Bondi radius (e.g., \\citealp{Marrone06}), which indicates that intense outflows may present in this system. Besides, the absorption lines from highly ionized elements, which have been detected in the X-ray spectrum of some microquasars such as GRO J1655-40 (\\citealp{Ueda98}; \\citealp{Yamaoka01}; \\citealp{Miller06}), GRS 1915+105 (\\citealp{Kotani00}; \\citealp{Lee02}) and Atoll sources (see e.g. the review by \\citealp{Diaz Trigo06} and references therein), also indicate the existence of outflows. On the other hand, \\citet{Jiao11} found that outflows generally exist in accretion discs no matter that the flow is advection-dominated such as the slim disc (\\citealp{Abramowicz88}) and the ADAF, or is radiation-dominated such as the standard thin disc (\\citealp{Shakura73}). In particular, for the three types of advection-dominated flows: ADAFs (gas internal energy dominant), slim discs (trapped photon energy dominant), and hyper-accretion discs (trapped neutrino energy dominant), outflows may be significantly strong due to positive Bernoulli parameters (e.g., \\citealp{Narayan97}; \\citealp{Liu11}) or the large radiation pressure (e.g., \\citealp{Gu07}). Furthermore, outflows have generally been found in many simulation works (e.g., \\citealp{Ohsuga11} and references therein). In the Lyubarskii's scheme, the power spectrum of luminosity fluctuations is sensitive to the varying mass accretion rate, thus we expect that outflows may have significant effects on the power-law index. The paper is organized as follows. In Section 2, we investigate the fluctuation power spectrum under a radius-dependent accretion rate following the method of Lyubarskii. The potential application of our results to observations is discussed in Section 3. ", "conclusions": "In this paper, we evaluate the effects of outflows on the luminosity fluctuations with the Lyubarskii's general scheme. With a radius-dependent accretion rate $\\dot M \\propto r^{s}$, the power spectrum of the luminosity fluctuations is $p(f) \\propto f^{-\\beta}$, where the value of $\\beta$ varies with the disc structure. By assuming that the coupling between the local outflow and inflow is weak on the accretion rate fluctuations, we obtain the following explicit expressions of $\\beta$ for different disc models: $\\beta = 1 + 4s/3$ for advection-dominated discs, $\\beta = 1 + 40s/(25-6s)$ for the outer region of standard thin discs, $\\beta = 1 + {10s}/(7-2s)$ for the middle region, $\\beta = 1 + {4}{s}/({7-4s})$ for the inner region, and $\\beta = 1 + 5s/2$ for NDAFs. The above expressions imply that $\\beta$ in a GRB is generally larger than that in a BHB for comparable $s$. The expressions of $\\beta$ indicate the possibility of evaluating the strength of outflows by the power spectrum in X-ray binaries and GRBs. In addition, if the coupling is not negligible, the value of $\\beta$ will probably be located between unity and the value presented in the above expressions. In both BHBs and AGN, ADAFs are usually adopted to describe the quiescent state, the low/hard state, and the corona which lies above a cold disc. ADAFs may produce significant outflows, and therefore the power spectrum can deviate from $f^{-1}$ based on the present analysis. The exact value of ${s}$ is, however, difficult to estimate from the theoretical point of view, except for the general constraint $0 < s < 1$ (\\citealp{Narayan08}). On the other hand, some observations indicate $s \\sim 0.3$ (\\citealp{Yuan03}; \\citealp{Zhang10}). Taking this value, we obtain $\\beta=1.4$ for ADAFs, which is close to $1.3$, the power-law index of PSDs presented in the low mass X-ray binary systems (\\citealp{Gilfanov05}). The quantitative difference may be relevant to the coupling between the outflow and inflow as discussed in \\S2.4. If there is only outflow that operates in the accreting system, $s$ should be positive. However, $s$ can also be negative due to the evaporation mechanism of a cold disc. For the accreting black hole in BHBs, the observed power-law components in the X-ray spectra are generally attributed to hot, tenuous plasmas, namely accretion disc coronae. Due to the high temperature in the corona, the interaction between the disc and corona would lead to mass evaporating from the disc to the corona (\\citealp{Meyer00}; \\citealp{Spruit02}). In this case, the value of $s$ for the corona should be negative if outflows are not strong, and therefore it is quite possible for $\\beta$ to be less than unity. Consequently, in this scenario $s$ for the underneath cold disc should be positive." }, "1112/1112.2350_arXiv.txt": { "abstract": "We extend the holographic Ricci dark energy model to include some direct, non-gravitational interaction between dark energy and dark matter. We consider three phenomenological forms for the interaction term $Q$ in the model, namely, $Q$ is taken proportional to the Hubble expansion rate and the energy densities of dark sectors (taken to be $\\rho_{\\rm de}$, $\\rho_{\\rm m}$, and $\\rho_{\\rm de}+\\rho_{\\rm m}$, respectively). We obtain a uniform analytical solution to the three interacting models. Furthermore, we constrain the models by using the latest observational data, including the 557 Union2 type Ia supernovae data, the cosmic microwave background anisotropy data from the 7-yr WMAP, and the baryon acoustic oscillation data from the SDSS. We show that in the interacting models of the holographic Ricci dark energy, a more reasonable value of $\\Omega_{\\rm m0}$ will be obtained, and the observations favor a rather strong coupling between dark energy and dark matter. ", "introduction": " ", "conclusions": "" }, "1112/1112.0952_arXiv.txt": { "abstract": "We present a brief overview of the theory of stellar winds with a strong emphasis on the radiation-driven outflows from massive stars. The resulting implications for the evolution and fate of massive stars are also discussed. Furthermore, we relate the effects of mass loss to the angular momentum evolution, which is particularly relevant for the production of long and soft gamma-ray bursts. Mass-loss rates are not only a function of the metallicity, but are also found to depend on temperature, particularly in the region of the bi-stability jump at 21 000 Kelvin. We highlight the role of the bi-stability jump for Luminous Blue Variable (LBV) stars, and discuss suggestions that LBVs might be direct progenitors of supernovae. We emphasize that radiation-driven wind studies rely heavily on the input opacity data and linelists, and that these are thus of fundamental importance to both the mass-loss predictions themselves, as well as to our overall understanding of the lives and deaths of massive stars. ", "introduction": "\\label{s:intro} Despite their relative rarity, massive stars are dominantly dangerous for their environments. This is a result of their mass and energy input via stellar winds and subsequent core-collapse supernovae (SN). A overarching parameter for the life expectancy of a massive star concerns its strong mass outflow which is driven by radiative forces on millions of ionic spectral line transitions, providing it with its name ``line-driven wind''. On the one hand, mass loss is thought to be a key agent in revealing chemically processed material at the stellar surface, making it responsible for evolutionary scenarios such as the O $\\rightarrow$ Luminous Blue Variable (LBV) $\\rightarrow$ Wolf-Rayet (WR) star $\\rightarrow$ SN sequence (e.g. Conti 1976, Chiosi \\& Maeder 1986, Langer et al. 1994). Furthermore, it determines the stellar mass before collapse and is thus relevant for the type of compact remnant that is left behind (i.e. neutron star or black hole). On the other hand, the role of mass loss may be equally relevant for the loss of angular momentum (e.g. Meynet \\& Maeder 2003). With respect to the latter, it has been suggested that low metallicity (actually low ``iron'' contents; Vink \\& de Koter 2005) leads to less mass and angular momentum loss in low metallicity environments, perhaps resulting in a preference of long gamma-ray bursts (GRB) in the early Universe, but however interesting the metallicity dependence and the long GRB puzzle may be, the temperature dependence of stellar winds and its role in the angular momentum evolution of massive stars has been highlighted more recently with respect to the possibility of {\\it bi-stability braking} (Vink et al. 2010). Given the crucial role that mass loss plays for massive star evolution, we discuss the theory of massive star mass loss and its implications, with a focus on the metallicity ($Z$) and effective temperature ($T_{\\rm eff}$) dependence. We will see that the {\\it iron line opacity} plays a dominant role in both cases. ", "conclusions": "" }, "1112/1112.1165_arXiv.txt": { "abstract": "Using the {\\hf Yunnan} evolutionary population synthesis (EPS) models with and without binary interactions, we present the luminosity of $\\rm H\\alpha$ recombination line ($L_{\\rm H\\alpha}$), the luminosity of [OII]$\\lambda$3727 forbidden-line doublet ($L_{\\rm [OII]}$), the ultraviolet (UV) fluxes at 1500 and 2800\\,$\\rm \\AA$ ($L_{i, {\\rm UV}}$) and far-infrared flux ($L_{\\rm FIR}$) for Burst, S0, Sa-Sd and Irr galaxies, and present the calibrations of star formation rate (SFR) in terms of these diagnostics. By comparison, we find that binary interactions lower the SFR.vs.$L_{\\rm H\\alpha}$ and SFR.vs.$L_{\\rm [OII]}$ conversion factors by $\\sim$0.2\\,dex. The main reason is that binary interactions raise the UV flux (shortward of the Lyman limit) of the stellar population (SP) in the age range 6.7$<$ log$t$/yr $<$8.4 and thus more ionizing photons are present in the nebula. Moreover, binary interactions do not significantly vary the calibrations of SFR in terms of $L_{i, {\\rm UV}}$. This is because binary interactions raise the flux at 1500\\,$\\rm \\AA$ of the SP in the range 8.75 $<$ log$t$/yr $<$ 9.2 and the maximal difference is about 1\\,dex. In addition, binary interactions have little effect on the flux at 2800\\,$\\rm \\AA$. At last, the calibration of SFR from $L_{\\rm FIR}$ is almost unaffected by binary interactions. This is caused by the fact that binary interactions almost do not affect the bolometric magnitudes of SPs. We also discuss the effects of initial mass function (IMF), gas-recycle assumption and EPS models (including {\\hf GISSEL98, BC03, STARBURST99, PopSTAR and P\\'{E}GASE} models) on these SFR calibrations. Comparing the results by using Salpeter (S55) IMF with those by using Miller \\& Scalo (MS79) IMF, we find that the SFR.vs.$L_{\\rm H\\alpha}$ and SFR.vs.$L_{\\rm [OII]}$ conversion factors by using S55 IMF are greater by 0.4 and 0.2\\,dex than those by using MS79 IMF for the {\\hf Yunnan} models with and without binary interactions, respectively. The SFR.vs.$L_{i,{\\rm UV}}$ and SFR.vs.$L_{\\rm FIR}$ conversion factors by using S55 IMF are larger by an amount of 0.2\\,dex than the corresponding ones by using MS79 IMF. The inclusion of gas-recycle assumption only lowers these SFR calibrations at faint SFR. Moreover, comparing the results when using different EPS models, we find that the differences in the SFR.vs.$L_{\\rm H\\alpha}$ and SFR.vs.$L_{\\rm [OII]}$ conversion factors reach $\\sim$ 0.7 and 0.9\\,dex, the difference in the SFR.vs.$L_{\\rm FIR}$ conversion factor reaches 0.4 and 0.8\\,dex, and the differences in the SFR.vs.$L_{i, {\\rm UV}}$ conversion factors reach 0.3 and 0.2\\,dex when using S55 and NON-S55 IMF (including Cha03, K01, K93' and MS79 IMFs, partly caused by the difference in the IMF), respectively. At last, we give the conversion coefficients between SFR and these diagnostics for all models. ", "introduction": "One of the most recognizable features of galaxies along the Hubble sequence (loose definition, including not only morphological type but also gas content, mass, bar structure and dynamical environment) is the wide range in young stellar content and star formation activity. Understanding its physical nature and origin of the variation in stellar content are fundamental to understand evolution of galaxies \\citep[][hereafter K98]{ken98}. Star formation rate (SFR) can be used to compare with those distant galaxies at cosmological lookback times, and extrapolate the future timescales for star formation in galaxies by combining with HI and CO measurements \\citep{ken94}. Moreover, star formation activity is usually correlated with cold gas and stars in galaxies: stars continuously produce mass, energy and metals during their evolution processes, and return them to galactic medium (gas), affecting the status of the next generation of stars. SFR carries the information on the evolution of galaxies. Therefore, it is important to determine SFR and its variation with Hubble type (loose definition) and environment, which can help us to understand the evolution of galaxies. The commonly used SFR tracers include the flux of H$\\alpha$ nebular recombination line, the ultraviolet (UV) continuum flux, the flux of [OII]$\\lambda$\\,3727 forbidden-line doublet and far infra-red (FIR) continuum flux (K98). These SFR tracers are more or less correlated with the UV passband. \\begin{itemize} \\item First, the UV flux is directly tied to the photospheric emission of the young stellar population (SP). \\item Second, the integrated luminosity of galaxy shortward of the Lyman limit (Far-UV) can ionize the hydrogen in the nebula and produce the recombination lines (such as H$\\alpha$, H$\\beta$, and so on). Thus the luminosities of these lines can be used to trace SFR. \\item Third, the luminosity of the strong [OII]$\\lambda$\\,3727 forbidden-line doublet is often empirically obtained through the H$\\alpha$ luminosity, although it is not coupled to the ionizing luminosity and the excitation of this line is sensitive to abundance and the ionization state of the gas. \\item Finally, the last SFR trace, the FIR luminosity, is also correlated with the UV passband. The interstellar dust can absorb the bolometric luminosity of galaxy and re-emit it in the thermal IR passband. The absorption cross section of the dust peaks in the UV passband, and, since the UV flux is considered as a tracer of young SP, the FIR luminosity can also diagnose SFR. \\end{itemize} Furthermore, the last three diagnostics and the corresponding calibrations of SFR are correlated with the UV flux. Since in our previous studies, we found that the inclusion of binary interactions in evolutionary population synthesis (EPS) models can raise the UV flux by $\\sim$ 2-3 magnitudes for SP at an age of $\\sim 1$\\,Gyr \\citep{zha04,zha05}, in this study we will discuss the effect of binary interactions on these calibrations. The outline of the paper is as follows. In Section 2 we describe the used EPS models and algorithm. In Section 3 we overview the previous results about SFR calibrations, and the advantages and disadvantages of these SFR tracers. In Section 4 we give the effect of binary interactions on these SFR calibrations. In Section 5 we discuss the effects of initial mass function (IMF), gas-recycle assumption and the EPS models on these SFR calibrations, and give the conversion coefficients between SFR and these tracers for all models. Finally we present a summary and conclusions in Section 6. ", "conclusions": "We use the {\\hf Yunnan} EPS models with and without binary interactions to present the luminosities of the $\\rm H\\alpha$ recombination line, the [OII]$\\lambda$3727 forbidden-line doublet, the UV (at $1500$ and $2800$\\,$\\rm \\AA$) and the FIR continuum for Burst, E, S0, Sa-Sd and Irr galaxies, and present the calibrations of SFR in terms of these diagnostics. By comparison, we find that binary interactions lower the SFR.vs.$L_{\\rm H\\alpha}$ and SFR.vs.$L_{\\rm [OII]}$ conversion factors by $\\sim $0.2\\,dex, and do not significantly vary the SFR.vs.$L_{i,{\\rm UV}}$ (at 1500 and 2800 $\\rm \\AA$) and SFR.vs.$L_{\\rm FIR}$ calibrations. We also consider the effects of IMF, the gas-recycle assumption and EPS models on these calibrations. By comparison, we find that the SFR.vs.$L_{\\rm H\\alpha}$ and SFR.vs.$L_{\\rm [OII]}$ conversion factors of Models A/B-S55 are larger by 0.4 and 0.2\\,dex than the corresponding ones of Models A/B-MS79, and that the SFR.vs.$L_{i,{\\rm UV}}$ and SFR.vs.$L_{\\rm FIR}$ conversion factors are larger by 0.2\\,dex. By comparing the results between Models C and Cr, we find that the inclusion of gas-recycle assumption only lowers the SFR calibrations at faint SFR. Also we use the other EPS models ({\\hf BC03}, {\\hf GISSEL98}, {\\hf PopSTAR}, {\\hf P\\'{E}GASE} and {\\hf STARBURST99}) to obtain these SFR calibrations. By comparison, we find that the differences in the SFR($L_{\\rm H\\alpha}$) and SFR($L_{\\rm [OII]}$) calibrations reach $\\sim$ 0.7 and 0.9\\,dex, the difference in the SFR($L_{\\rm FIR}$) calibration reaches 0.4 and 0.8\\,dex, and the differences in the SFR($L_{i,{\\rm UV}}$) calibration reach 0.3 and 0.2\\,dex when using S55 and NON-S55 (partly caused by the difference in the IMF) IMFs, respectively. At last, in this paper we give the conversion coefficients between $SFR$ and these diagnostics for all models. In this paper we have only considered the effects of binary interactions for solar metallicity galaxies - more detailed studies will be given." }, "1112/1112.1954_arXiv.txt": { "abstract": "The current dynamical structure of the Kuiper belt was shaped by the orbital evolution of the giant planets, especially Neptune, during the era following planet formation, when the giant planets may have undergone planet-planet scattering and/or planetesimal-driven migration. Numerical simulations of this process, while reproducing many properties of the belt, fail to generate the high inclinations and eccentricities observed for some objects while maintaining the observed dynamically ``cold\" population. We present the first of a three-part parameter study of how different dynamical histories of Neptune sculpt the planetesimal disk. Here we identify which dynamical histories allow an \\emph{in situ} planetesimal disk to remain dynamically cold, becoming today's cold Kuiper belt population. We find that if Neptune undergoes a period of elevated eccentricity and/or inclination, it secularly excites the eccentricities and inclinations of the planetesimal disk. We demonstrate that there are several well-defined regimes for this secular excitation, depending on the relative timescales of Neptune's migration, the damping of Neptune's orbital inclination and/or eccentricity, and the secular evolution of the planetesimals. We model this secular excitation analytically in each regime, allowing for a thorough exploration of parameter space. Neptune's eccentricity and inclination can remain high for a limited amount of time without disrupting the cold classical belt. In the regime of slow damping and slow migration, if Neptune is located (for example) at 20 AU, then its eccentricity must stay below 0.18 and its inclination below 6$^\\circ$. ", "introduction": "\\label{sec:intro} The solar system is often used as a case study for the formation of planetary systems from proto-planetary disks. The current configuration of Kuiper belt objects (KBOs) provides a map for how the dynamical evolution of the giant planets in our solar system sculpted the disk of planetesimals. Therefore it is possible to use the orbital properties of the Kuiper belt to constrain how the orbits of the giant planets evolved in the early solar system, particularly for Neptune, the primary sculptor of the Kuiper belt. Models employing N-body integrations to trace the effects of the giant planets' migration and orbital eccentricity evolution on the planetesimal disk have enjoyed substantial success in reproducing the dynamical populations of KBOs observed today \\citep[e.g.][]{1993M,1995M,1999H,2003G,2005H,2008L,2008M}. These populations include objects near orbital resonance with Neptune, objects scattering off Neptune, and ``classical\" objects decoupled from Neptune. \\citep[See][for definitions of the dynamical classes.]{2008G} Yet substantial discrepancies still exist between the simulations and observations, particularly for the classical population. The bias-corrected inclination distribution of observed classical KBOs is bimodal \\citep{2001B,2010G,2011V}. The low-inclination, dynamically ``cold\" component and the high-inclination, dynamically ``hot\" component also have distinct physical properties, including colors \\citep{2000T,2003T,2008P}, sizes \\citep{2001L,2010F}, albedos \\citep{2009B}, and binary fractions \\citep{2006S,2008N}. To date no simulations have been able to produce both the high and low inclination classical objects while qualitatively matching their observed eccentricity distribution. Several theories of the dynamical history of the giant planets in our solar system have been proposed, inspired by the dynamical populations within the Kuiper belt. One of the most widely accepted models, the Nice Model, stems from a postulated large scale instability in the early solar system \\citep[e.g.][]{1999T,2003G,2008M}. Inspired by the Nice Model, \\citet{2008L} proposes a scenario in which Neptune is scattered outward onto an eccentric orbit to near its current location from a formation location closer to the sun \\citep[e.g.][]{1999T,2002T}. In this scenario, Neptune's eccentricity subsequently damps due to dynamical friction with a remnant disk of planetesimals. This model reproduces the observed resonant population, scattered population, and the hot classical population, but, we argue, does not satisfactorily match the low eccentricities of the observed population of dynamically cold objects in the classical region. Furthermore, it does not produce a sufficient number of high-inclination objects. Other incarnations of the Nice model \\citep[e.g.][]{2003G,2008M} include an \\emph{in situ} population of cold objects, but, over the course of Neptune's evolution, these objects become excited to higher eccentricities. Whether or not the particular history described by the Nice Model is correct, the prevalence of large eccentricities among extrasolar giant planets suggests that large-scale orbital excitation is common during the formation of planetary systems. In contrast to the upheaval of the Nice Model, another model \\citep{1993M} proposes a period of extensive, smooth migration of the giant planets. Planetesimal-driven migration \\citep{1984F} is likely important in shaping the architecture of many planetary systems. \\citet{1993M,1995M,1999H} demonstrated that the migration of the giant planets in our solar system on flat, circular orbits would have perturbed the disk of planetesimals, scattering some to more eccentric orbits and capturing others into resonance. They also found that migration results in a large number of planetesimals in the location known today as the ``Scattered Disk,\" which encompasses those objects that have had gravitational interactions with one or more of the giant planets and have high orbital eccentricities and inclinations. The key inconsistency between this model and current observations is that it cannot produce, without additional processes such as stochasticity \\citep[e.g][]{2003L,2006M}, both the cold and hot classical populations from a single set of initial disk conditions or account for the differences in their physical properties. Because detailed N-body simulations are computationally expensive, previous works have investigated a limited number of solar system planetary histories. Currently, no comprehensive parameter study has been done exploring the evolution of Neptune's orbit given constraints from the sculpting of the Kuiper belt. In addition, although planet-planet scattering often produces large mutual inclinations \\citep{2011C}, as in observed in the Upsilon Andromedae system \\citep{2010M}, no serious, detailed treatment has included the possibility that Neptune underwent a period of high orbital inclination, a conceivable outcome of the orbital instability in the early solar system. Here we consider a general model that can encompass the two detailed models described above -- the Nice-Model-inspired scenario in which Neptune is scattered to a high eccentricity \\citep{2008L} and the scenario of extensive migration of Neptune on a low-eccentricity, low-inclination orbit \\citep[e.g.][]{1995M} -- as well as potential scenarios in which Neptune undergoes a period of high inclination. In this generalized model, Neptune undergoes some combination of migration and/or evolution of its eccentricity and/or inclination. Here we take a step toward understanding the qualitative differences in the dynamics of the Kuiper belt generated by a wide range of planetary histories. In a series of three papers, we perform a parameter study of the effects on a disk of planetesimals of the migration and the eccentricity and inclination evolution of Neptune in the early solar system. No component of this parameterization is new; rather, our goal is to comprehensively consider all possible parameters for Neptune's history. In this first paper, we introduce our method for computationally and analytically modeling Neptune's orbital evolution and its effects on the planetesimal population. Then we demonstrate several concepts that will allow us to thoroughly explore the parameter space of Neptune's dynamical history: \\begin{itemize} \\item If Neptune undergoes a period of elevated orbital eccentricity, it will secularly excite the eccentricities and inclinations of an \\emph{in situ} planetesimal population. We can analytically model this secular excitation using a simple expression. \\item As Neptune's eccentricity and inclination damp, the planetesimals evolve to final eccentricities and inclinations that depend on Neptune's initial eccentricity and inclination and damping timescales. \\item The effects of Neptune's eccentricity and inclination evolution on the planetesimals can be treated separately to first order. \\item The migration rate sets Neptune's effective location for the secular evolution of the planetesimals. \\end{itemize} Having demonstrated these points, we can place robust constraints on Neptune's semi-major axis, eccentricity, and inclination during its late evolution (after any period of planet-planet scattering), and on its migration, eccentricity damping, and inclination damping rates, identifying which parameters are consistent with maintaining the low eccentricities and inclinations of the cold classicals. In the second paper \\citep{2012D}, we place even stronger constraints by incorporating additional effects, including the effects of other planets on Neptune's orbital evolution and the effects of proximity to mean-motion resonance with Neptune on the secular excitation of the planetesimals, as well as additional constraints from the hot classical population. A third paper (Dawson and Murray-Clay 2012b, in prep) will focus on the inclinations of the classicals. In Section \\ref{sec:nep}, we present our parameterization of Neptune's orbital evolution. In Section \\ref{sec:small}, we discuss the observational constraints that today's cold classical KBOs place on past sculpting of the planetesimal disk and describe our computational and analytical models of the excitation of the planetesimal disk by an inclined and eccentric Neptune. We present the results of the parameter study in Section \\ref{sec:results}. In Section \\ref{sec:conclusions}, we present our conclusions and describe how the companion papers will expand upon this work. ", "conclusions": "\\label{sec:conclusions} As a first step in a comprehensive study of the impact on the Kuiper belt of a wide range of possible dynamical histories of the outer solar system, we have performed a suite of numerical integrations probing the impact of the orbital evolution of a single a planet on a disk of planetesimals. We have presented the observational evidence for a population of dynamically cold objects in the Kuiper belt in the region from 42.5 to 45 AU that are confined to $e < 0.1$ and $i < 6^\\circ$. We argued that recent models of Kuiper belt sculpting -- which explain many of the observed dynamical and physical properties of the Kuiper Belt -- do not generate or preserve sufficiently low eccentricities for the cold classicals \\citep[e.g.][]{2003G,2008L,2008M}. Our results have revealed several principles key to constraining which orbital histories of Neptune are consistent with preserving the \\emph{in situ} planetesimal population at the low eccentricities and inclinations required by the observations: \\begin{itemize} \\item If Neptune is scattered onto an eccentric and/or inclined orbit, it will secularly excite the eccentricities and inclinations of an \\emph{in situ} planetesimal population. \\item Planetesimals starting with $e = i = 0$ reach eccentricities and inclinations up to twice their forced eccentricity and inclination, on timescales given by Eqn. (\\ref{eqn:sece}) and (\\ref{eqn:seci}). \\item As Neptune's eccentricity and inclination damp, the planetesimals evolve to their final eccentricities and inclinations. If the damping timescales $\\tau_e$ and $\\tau_i$ of Neptune's orbit are slow compared to the secular evolution time, a planetesimal evolves to its initial free eccentricity and inclination, which are set by Neptune's initial eccentricity and inclination. If the damping is fast compared to the secular evolution time, the planetesimal's $e$ and $i$ effectively freeze at the values they reach after one damping time, reaching final values of $\\efree \\sin(\\gkbo \\tau_e)$ and $\\eforced \\sin(\\gkbo \\tau_i)$ respectively. See Table \\ref{tab:constrain} for constraints on Neptune's eccentricity and inclination in the two damping regimes. \\item The effects of Neptune's: 1) eccentricity evolution, and 2) inclination evolution, on the planetesimals can be treated separately to first order. \\item At a given location in the planetesimal disk, the secular excitation timescales and forced eccentricity (but not the forced inclination) depend on Neptune's location, which is altered by Neptune's migration. When Neptune's migration is slow relative to the damping time, the secular evolution effectively takes place at Neptune's initial location. When Neptune's migration is fast relative to the damping time, the secular evolution effectively takes place at Neptune's final location. \\end{itemize} From these principles, it is evident that the three models described in \\citet{2008L} would not be able to retain a cold classical population. Neptune begins with an eccentricity of 0.3 and a semi-major axis of 27.5 AU (run A and C) or 28.9 AU (run B) and damps on a timescale of 1 Myr (run A and B) or 3 Myr (run C), in the slow migration regime. When Neptune is at 27.5 AU, a planetesimal at 42.5 AU has a forced eccentricity of $e = 0.76 e_N = 0.23$ and a secular evolution timescale of 6 Myr. The final eccentricity of the planetesimal evolves to is reduced by a factor of $\\sin(\\gkbo \\tau_e) = 0.85$ for a damping timescale of 1 Myr (fast damping) and is not reduced for a damping timescale of 3 Myr (slow damping). Thus the final eccentricity of the planetesimal is 0.19 (run A) or 0.23 (run C), well above the observational limit. When Neptune is at 28.9 AU (run B), a planetesimal at 42.5 AU has a forced eccentricity of $e = 0.79 e_N = 0.24$ and a secular evolution timescale of 5 Myr. The final eccentricity the planetesimal evolves to is reduced by a factor of $\\sin(\\gkbo \\tau_e) = 0.97$ for a damping timescale of 1 Myr. Thus the final eccentricity of the planetesimal is 0.23, well above the observational limit. According to the constraints established in our paper, any of these three initial conditions for Neptune could retain the cold classicals if Neptune's eccentricity were to damp more quickly. Based on these principles, we can place robust constraints on Neptune's dynamical history. For example, in the regime of slow damping and slow migration, if Neptune's initial semi-major axis is 20 AU, then its eccentricity must stay below 0.18 and inclination below 6 degrees. If Neptune's initial semi-major axis is 30 AU -- or if it migrates quickly, relative to the damping time, from its initial location to 30 AU -- then its eccentricity must stay below 0.12. In the case of fast damping -- on a timescale shorter than a planetesimal's secular excitation time -- the initial eccentricity and inclination of Neptune can be even higher. Having established these principles, we complete our parameter study in two companion papers \\citet{2012D} and Dawson and Murray-Clay (2012b), in prep. In the first companion paper, we consider more generally the constraints on Neptune's dynamical history from the eccentricity distribution of the classical KBOs and incorporate several other important effects in the constraints from the cold classicals: \\begin{itemize} \\item A more accurate model for secular excitation that includes higher-order terms. \\item The greatly increased secular frequency near Neptune's mean-motion resonances, which places strong constraints on Neptune's semi-major axis when its eccentricity is high. \\item The effects of the other giant planets, including precession of Neptune, which can reduce a planetesimal's forced eccentricity \\citep{2011B}, and oscillations in Neptune's semi-major axis, which can create a chaotic sea in the Kuiper belt region. \\end{itemize} Then we combine these constraints for retaining the cold classicals with constraints for creating the hot classical population and identify which regions of parameter space of Neptune's dynamical history can produce the hot classical population without disrupting the cold population. In the second companion paper, we place constraints on Neptune's inclination and inclination damping time. Our parameterization of Neptune's orbital history allows us to constrain which orbital histories are consistent with maintaining the cold classical population. By combining the observational constraints (Sec. \\ref{sec:obs}) with the principles derived in this paper, we can immediately check whether a particular set of initial conditions for Neptune's semi-major axis, eccentricity, and inclination, and the rates of its migration and eccentricity and inclination damping, are consistent with maintaining the observed dynamically cold population, without performing computationally expensive integrations. A picture, both qualitative and quantitative, is emerging of what dynamical histories of Neptune allow a promising general scenario -- Neptune's delivery of the hot classicals from the inner disk to classical region, where the cold population has formed \\emph{in situ} -- to be consistent with observed low eccentricities and inclinations of the cold classical population. Barring fast precession of Neptune's orbit, which we do not consider in this work, the existence of the cold classical population implies that Neptune could have spent only a limited time at high eccentricity and/or inclination during its dynamical history." }, "1112/1112.4789_arXiv.txt": { "abstract": "We determine the observational signatures of protostellar cores by coupling two-dimensional radiative transfer calculations with numerical hydrodynamical simulations that predict accretion rates that both decline with time and feature short-term variability and episodic bursts caused by disk gravitational instability and fragmentation. We calculate the radiative transfer of the collapsing cores throughout the full duration of the collapse, using as inputs the core, disk, and protostellar masses, radii, and mass accretion rates predicted by the hydrodynamical simulations. From the resulting spectral energy distributions, we calculate standard observational signatures (\\lbol, \\tbol, \\lbolsmm) to directly compare to observations. We show that the accretion process predicted by these models reproduces the full spread of observed protostars in both \\lbol\\ $-$ \\tbol\\ and \\lbol\\ $-$ \\mcore\\ space, including very low luminosity objects, provides a reasonable match to the observed protostellar luminosity distribution, and resolves the long-standing luminosity problem. These models predict an embedded phase duration shorter than recent observationally determined estimates (0.12 Myr vs.~0.44 Myr), and a fraction of total time spent in Stage 0 of 23\\%, consistent with the range of values determined by observations. On average, the models spend 1.3\\% of their total time in accretion bursts, during which 5.3\\% of the final stellar mass accretes, with maximum values being 11.8\\% and 35.5\\% for the total time and accreted stellar mass, respectively. Time-averaged models that filter out the accretion variability and bursts do not provide as good of a match to the observed luminosity problem, suggesting that the bursts are required. ", "introduction": "Low-mass stars form from the gravitational collapse of dense cores of gas and dust (e.g., Beichman et al.~1986; Di Francesco et al.~2007; Ward-Thompson et al.~2007a). In the simplest model, the collapse of a singular isothermal sphere initially at rest as first considered by Shu (1977) and later extended by Terebey, Shu, \\& Cassen (1984; TSC84) to include rotation (often called the ``standard model'' of star formation), collapse proceeds in an ``inside-out'' fashion, beginning in the center of the core, moving outward at the sound speed, and giving rise to a constant mass accretion rate of $\\sim 2 \\times 10^{-6}$ \\msun\\ yr$^{-1}$. Many modifications to this model have been explored, including non-zero initial inward motions (Larson 1969; Penston 1969; Hunter 1977; Fatuzzao, Adams, \\& Myers 2004), magnetic fields (Galli \\& Shu 1993a, 1993b; Li \\& Shu 1997; Basu 1997), isothermal spheres that are \\emph{not} singular but feature flattened density profiles at small radii (Foster \\& Chevalier 1993; Henriksen, \\andre, \\& Bontemps 1997), and a finite outer boundary (Henriksen, \\andre, \\& Bontemps 1997; Vorobyov \\& Basu 2005a). All but the latter (finite outer boundary) generally increase the accretion rate over that predicted by the standard model. A significant shortcoming of the standard model is the classic ``luminosity problem,'' whereby accretion at the above rate produces accretion luminosities ($L_{acc} \\propto M_* \\dot{M}$) factors of $10-100$ higher than typically observed for embedded protostars. First noticed by Kenyon et al.~(1990) and further investigated by Kenyon et al.~(1994) and Kenyon \\& Hartmann (1995), this problem has recently been emphasized by studies presenting results from the \\emph{Spitzer Space Telescope} ``From Molecular Cores to Planet Forming Disks'' (cores to disks, hereafter c2d; Evans et al.~2003) Legacy Program. One of the first results to come from the c2d project was the discovery of Very Low Luminosity Objects (VeLLOs), objects embedded within dense cores with \\lint\\footnote{The internal luminosity, \\lint, is the luminosity of the central source and excludes luminosity arising from external heating.} $\\leq$ 0.1 \\lsun, most in cores previously classified as starless (Young et al.~2004; Bourke et al.~2006; Dunham et al.~2006, 2008, 2010a; Di Francesco et al.~2007; Terebey et al.~2009; Lee et al.~2009). The discovery of such low luminosity protostars only exacerbated the existing luminosity problem. Furthermore, Dunham et al.~(2008), Enoch et al.~(2009a), and Evans et al.~(2009) all examined protostellar luminosities from the c2d survey and showed that the protostellar luminosity distribution spans more than three orders of magnitude, is strongly skewed towards low luminosities (greater than 50\\% of protostars feature luminosities indicating $\\dot{M} \\la 10^{-6}$ \\msun\\ yr$^{-1}$), and is inconsistent with the standard model as well as with the modifications described above, which tend to increase the mass accretion rate and thus make the problem worse. One possible resolution to the luminosity problem is the idea that mass accretion is not constant. As noted by Kenyon et al.~(1990), either accretion rates that decline with time or accretion rates that are very low most of the time but occasionally very high could resolve the luminosity problem. The latter process, commonly referred to as episodic accretion, features prolonged periods of lower-than-average accretion punctuated by short bursts of higher-than-average accretion, a scenario already invoked to explain the luminosity flares seen in FU Orionis objects (Hartmann \\& Kenyon 1985). First proposed by Kenyon et al.~(1990), such a solution was also suggested by Dunham et al.~(2008), Enoch et al.~(2009a), and Evans et al.~(2009) as a plausible explanation for both the large spread in observed luminosities and the significant population of sources at low luminosities. Theoretical studies have provided several mechanisms for such a process in the embedded protostellar phase. Hydrodynamical and MHD simulations have demonstrated that material accreting from a core can pile up in a circumstellar disk until the disk becomes gravitationally unstable and fragments into spiral structure and dense clumps, which are then driven onto the protostar in short-lived accretion bursts generated through the gravitational torques associated with the spiral structure (Vorobyov \\& Basu 2005b, 2006, 2010; Machida et al.~2011). Numerical hydrodynamic simulations without self-consistent disk-core interaction but with gravitationally overstable massive disks have also shown quick migration of dense clumps into the disk inner region and probably onto the star (Boss 2002, Cha \\& Nayakshin 2011). Alternatively, several authors have explored a scenario where a similar process involving gravitational instabilities in the outer disk triggers rapid accretion into the inner disk; the subsequent heating of the inner disk by this accretion activates magnetorotational instabilities (MRI) that then drive material onto the protostar in short accretion bursts (Armitage et al.~2001; Zhu et al.~2009a, 2009b, 2010). Other possible mechanisms have been proposed as well, including quasi-periodic magnetically driven outflows in the envelope causing magnetically controlled accretion bursts (Tassis \\& Mouschovias 2005), decay and regrowth of MRI turbulence (Simon et al.~2011), close interaction in binary systems or in dense stellar clusters (Bonnel \\& Bastein 1992; Pfalzner et al.~2008), and disk-planet interaction (Lodato \\& Clarke 2004; Nayakshin \\& Lodato 2011). Direct observational evidence for episodic mass accretion bursts in the embedded phase is less clear. Several Class I sources, including V1647 Ori (e.g., \\'{A}brah\\'{a}m et al.~2004; Acosta-Pulido et al.~2007; Fedele et al.~2007; Aspin et al.~2009), OO Serpentis (K\\'{o}sp\\'{a}l et al.~2007), [CTF93]216-2 (Caratti o Garatti et al.~2011), and VSX J205126.1 (Covey et al.~2011; K{\\'o}sp{\\'a}l et al.~2011), have undergone recent flares attributed to accretion bursts. However, these sources only flared in \\lbol\\ by factors of $\\sim$ $2 - 10$, and all appear to be very late Class I sources near the end of the embedded phase with little remaining envelope material. On the other hand, the detection of silicate features and CO$_2$ ice bands in absorption in several known FU-Ori-like objects suggests the presence of massive envelopes (Quanz et al.~2007) and one of these objects, RHO~1B, has recently brightened by a factor of 1000 (Staude \\& Neckel 1991). Episodic mass ejection is seen in jets ejected from some protostellar systems, suggesting an underlying variability in the mass accretion, although the combination of jet velocities and spacing between knots often suggest shorter periods of episodicity than found by the above theoretical studies (e.g., Lee et al.~2007; Devine et al.~2009). Additionally, several Class 0 sources drive strong outflows implying higher average mass accretion rates than expected from their current luminosities (\\andre\\ et al.~1999; Dunham et al.~2006, 2010a; Lee et al.~2010) and Watson et al.~(2007) showed a mismatch between the accretion rates onto the disk and protostar of NGC 1333-IRAS 4B, although their results are currently under debate (Herczeg et al.~2011). Finally, White \\& Hillenbrand (2004) showed that a sample of optically visible Class I sources in Taurus have very low accretion rates comparable to values observed for typical Class II sources rather than those expected for accreting Class I objects, although it is unclear if this result is simply a consequence of the Class I objects in their study being optically visible and thus very near the end of the embedded stage. In this paper we test the hypothesis that the accretion process predicted by Vorobyov \\& Basu (2005b, 2006, 2010), which includes both accretion rates that decline with time and episodic accretion bursts, can resolve the luminosity problem and match the observed properties of embedded protostars. We couple the hydrodynamic simulations presented by Vorobyov \\& Basu with radiative transfer models to calculate the observational signatures of cores collapsing to form protostars in the manner predicted by these simulations, and use these results to directly compare to observations. This work is a direct follow-up to two previous studies. In the first, Young \\& Evans (2005; hereafter Paper I) used a one-dimensional dust radiative transfer code to calculate the observational signatures of cores undergoing inside-out collapse following Shu (1977). They followed three different cores with initial masses of 0.3, 1, and 3 \\msun\\ and showed that such models match only the upper end of the protostellar luminosity distribution, reconfirming the luminosity problem. In the second study, Dunham et al.~(2010b; hereafter Paper II) revisited the models from Paper I with a two-dimensional radiative transfer code and showed that including improved dust opacities, a circumstellar disk and rotationally flattened envelope structure, mass-loss, and outflow cavities improved the match to observations but did not fully resolve the luminosity problem, whereas a toy-model representation of episodic accretion could in fact resolve the luminosity problem. However, the latter conclusion is only tentative since the episodic accretion was included in a very simple manner where all mass accreted from the core builds up in the disk until the disk reaches 20\\% of the protostellar mass, at which point the accretion rate from the disk onto the star jumps from 0 \\msun\\ yr$^{-1}$ to $10^{-4}$ \\msun\\ yr$^{-1}$ until the disk is fully drained of its mass. The accretion rate from the disk onto the star then drops back down to 0 \\msun\\ yr$^{-1}$ and the cycle begins anew. This simplification likely exaggerates the fraction of total mass accreted in bursts, as pointed out by Offner \\& McKee (2011) (see also \\S \\ref{sec_discussion_bursts}). Here we will revisit the Paper II results using accurate predictions for the evolution of collapsing cores from hydrodynamical simulations. This paper is complementary to several other recent studies. Myers et al.~(1998) presented simple radiative transfer calculations predicting observational signatures of collapsing cores, including an exponentially declining accretion rate with time and the effects of mass-loss, and showed that such models could generally reproduce the median observed protostellar luminosities but not the full range of values. However, their model evolution is not based on a fully self-consistent analytic or numerical model. Lee (2007) modified the Paper I model to include episodic accretion in a very simple manner in order to study the effects such a process has on the chemical evolution of collapsing cores. Vorobyov (2009b) compared the distribution of mass accretion rates in their simulations featuring episodic accretion (Vorobyov \\& Basu 2005b, 2006) to those inferred from the luminosities of protostars in the Perseus, Serpens, and Ophiuchus molecular clouds compiled by Enoch et al.~(2009a) and showed that their simulations reproduced some of the basic features of the observed distribution of mass accretion rates. However, the observed distribution of accretion rates is quite uncertain since it is calculated from the observed protostellar luminosity distribution with assumed values for the protostellar masses and radii and with the assumption that all observed luminosity is accretion luminosity. Finally, Offer \\& McKee (2011) presented analytic derivations of the protostellar luminosity distribution for different models and concluded that models that tend toward a constant accretion time rather than constant accretion rate produce a greater spread in luminosities and are in better agreement with observations, similar to the result obtained by Myers (2010) that analytic models with accretion rates that increase with mass can at least partially resolve the luminosity problem. However, both Offner \\& McKee (2011) and Myers (2010) simply compared theoretical luminosities with the observed protostellar luminosity distribution, whereas this study uses radiative transfer calculations to ``observe'' the underlying theory (in this case, the Vorobyov \\& Basu simulations) in a manner consistent with, and with direct comparison to, the observations. The organization of this paper is as follows. A brief description of the models is provided in \\S \\ref{sec_model}, focusing on the hydrodynamical simulations in \\S \\ref{sec_model_sim} and the radiative transfer models in \\S \\ref{sec_model_radtrans}. \\S \\ref{sec_comparing} describes how the models are turned into observational signatures and the observational dataset to which the models are compared, while \\S \\ref{sec_results} describes the basic results. A discussion of these results is presented in \\S \\ref{sec_discussion}, focusing on the degree to which the models resolve the luminosity problem in \\S \\ref{sec_discussion_lum}, the match between observed and model bolometric temperatures in \\S \\ref{sec_discussion_tbol}, the duration of the embedded phase in \\S \\ref{sec_discussion_durations}, and the number, duration, and importance of the mass accretion bursts in \\S \\ref{sec_discussion_bursts}. Finally, a basic summary of our results and conclusions is given in \\S \\ref{sec_summary}. ", "conclusions": "\\label{sec_discussion} \\subsection{Resolving the Luminosity Problem}\\label{sec_discussion_lum} As shown above in Figures \\ref{fig_blt} and \\ref{fig_lboltbolhist}, the models considered here reproduce the full range of observations in \\lbol\\ $-$ \\tbol\\ space and provide a reasonable match to the observed protostellar luminosity distribution. Thus we conclude that the accretion process predicted by the Vorobyov \\& (2005b, 2006, 2010) simulations of collapsing cores resolves the luminosity problem, although we caution that these results must be revisited in the future once simulations that feature smaller sink cells closer to the stellar surface and fully capture both the physics in the inner disk and the actual protostellar accretion rates are possible (see \\S \\ref{sec_simulations_accretion} for further discussion). In Paper II we argued that episodic accretion is both necessary and sufficient to resolve the luminosity problem, but that conclusion is subject to uncertainty since the actual prescription for episodic accretion included in that paper was quite simple and did not fully capture the accretion process in the simulations. Here we have coupled the exact evolution of the collapsing cores with radiative transfer calculations and demonstrated that the Vorobyov \\& Basu simulations resolve the luminosity problem, although we defer the question of whether or not episodic accretion itself is necessary to Section \\ref{sec_discussion_bursts} below. We also note here that these models predict a smooth distribution in \\lbol\\ $-$ \\tbol\\ space, whereas the models in Paper II featured white ``excluded zones'' where the models spent no time but sources were observed to exist. The fact that only 3 initial mass cores were considered in Paper II (0.3, 1, and 3 \\msun) was argued to artifically create these zones; the increased sampling of core masses in this paper confirms this argument. Most of the remaining discrepancy between the model predictions and observations of protostellar luminosities actually appears in the form of a ``reverse luminosity problem'', evident in both Figures \\ref{fig_blt} and \\ref{fig_lboltbolhist} as an overabundance of time spent at \\lbol\\ $\\la$ 0.1 \\lsun\\ compared to observations. Indeed, a K-S test on the observed and model luminosity distributions that only compares the distributions above 0.1 \\lsun\\ returns a value of 0.87, slightly higher than the value of 0.85 obtained for the full distributions. What causes the disagreement at low luminosities? Only 7\\% of the observed sources have \\lbol\\ $\\leq 0.1$ \\lsun\\ whereas the models spent 21\\% of the total time at such luminosities. Furthermore, the majority of this time (19\\% out of 21\\%) is spent at \\tbol\\ $\\leq 100$ K, where there are no observed sources. At least some, and possibly all, of this discrepancy can be explained by observational completeness limits. As noted above, the c2d survey is generally complete to protostars with \\lbol\\ $\\ga$ 0.05 \\lsun, although the exact completeness limit varies depending on distance and the detailed shape of the SED of each source. Indeed, several extremely low luminosity sources undetected in the \\emph{Spitzer} c2d survey, with both internal and bolometric luminosities significantly below 0.1 \\lsun, have recently been discovered, most through detections of outflows driven by cores with no associated \\emph{Spitzer} c2d sources (Chen et al.~2010; Enoch et al.~2010; Dunham et al.~2011; Pineda et al.~2011). With such low luminosities at least some of these sources may be first hydrostatic cores. Sensitive interferometer outflow surveys and very deep \\emph{Herschel} and \\emph{James Webb Space Telescope} infrared surveys directed towards cores currently classified as starless are needed to fully identify and characterize the population of such extremely low luminosity protostars and/or first cores before an accurate comparison to the models can be made for luminosities below 0.1 \\lsun. \\subsection{Bolometric Temperatures}\\label{sec_discussion_tbol} As noted above in \\S \\ref{sec_results}, the models do not provide a good match to the observed \\tbol\\ distribution, with most of the discrepancy in a population of embedded objects at high \\tbol\\ ($\\ga 1000$ K) predicted by the models but lacking in the observations. In the models, most of this time spent at \\tbol\\ $\\ga 1000$ K arises when \\rdisk\\ is larger than a few hundred AU and the line-of-sight does not pass through the disk. As described in detail in \\S \\ref{sec_model_radtrans}, we adopt analytic profiles for the disk and core density profiles since the hydro simulations do not provide the full vertical density structure. In order to do this, we define the core inner radius to be equal to the disk outer radius so that there is no overlap between where the disk and core density profiles are defined. However, as a consequence of this method, large cavities devoid of material exist above the surface of the disk but within the core inner radius, and these cavities increase as the disk sizes increase. Lines of sight that pass through these cavities have reduced optical depths, allowing more short-wavelength emission to escape and thus increasing \\tbol. In reality, such large cavities are unlikely to exist; instead, the disk and core density profiles should smoothly join together. As we have mentioned in \\S \\ref{sec_model_radtrans}, an improved methodology that reconstructs the disk and core vertical structure and incorporates this exact structure into the radiative transfer calculations (rather than adopting simple analytic profiles) is currently under development and will be presented in a future paper. This method will result in a more accurate distribution of material above the disk surface and will likely remove much of the discrepancy between observed and model values of \\tbol. Since the distribution of luminosities is set mainly by the accretion rates and protostellar masses, we argue that including a more accurate physical structure should not significantly alter our results on the resolution of the luminosity problem, although this will be explicitly tested in a future paper. We also note here the possibility that these models contain too much rotation and angular momentum, since this would cause both the rotational flattening of the cores and sizes of the disks to be overestimated, resulting in the models overpredicting \\tbol\\ compared to observations. As discussed in some detail in \\S \\ref{sec_model_initial} above we do not consider this to be very likely, but we acknowledge that it is a possibility since $\\beta$, the ratio of rotational to gravitational energy, is not a directly observable quantity and observed ranges of $\\beta$ could be overestimated if infall and/or outflow motion is mistakenly attributed to rotation. Even after including a more accurate physical structure, some discrepancy between observed and model values of \\tbol\\ may remain, particularly once the effects of outflow cavities are included. Figure \\ref{fig_tbolsed}, which plots an example spectral energy distribution (SED) for a model with high \\tbol, shows that such models feature a double-peaked SED. This figure also demonstrates that, from optical wavelengths to either the 24 or 70 \\um\\ bands probed by \\emph{Spitzer} observations, such models resemble transition disk Class II sources (sometimes also referred to as cold disk sources; see \\merin\\ et al.~[2010] and references therein). Given that our models predict that 40\\% of embedded sources are classified as Class II by \\tbol, and that the observed fraction of Class 0+I to Class II sources is 0.19 when classifying via extinction-corrected values of \\tbol\\ (Evans et al.~2009), our models predict that 12\\% of Class II young stellar objects are actually embedded objects with SEDs like that shown in Figure \\ref{fig_tbolsed}. While this is consistent with the upper range of the observed fraction of Class II sources with transition disk SEDs (3\\% -- 12\\%; \\merin\\ et al.~2010; Furlan et al.~2011), the models clearly overpredict this fraction, as described above. Nevertheless, we caution that a small fraction of Class II sources with transition disk SEDs could in fact be embedded sources. \\begin{figure} \\epsscale{1.0} \\plotone{f11.eps} \\caption{\\label{fig_tbolsed}Example of a Spectral Energy Distribution (SED) for a model with high \\tbol. This particular SED is for the $i=$ 45$^{\\rm o}$ line-of-sight at 210,000 yr into the collapse of model 3, when the core mass is 0.9649 \\msun, the disk mass is 0.1656 \\msun, the protostellar mass is 0.5122 \\msun, and the disk radius is 744 AU. This SED has \\tbol\\ $= 1037$ K. The solid line plots the model SED over all wavelengths, while the points show the model SED at standard optical and near-infrared wavelengths and the $3.6-70$ \\um\\ wavelengths provided by the \\emph{Spitzer Space Telescope}. The points have been shifted down by a factor of 10 for clarity.} \\end{figure} If the above statement is true, why are they missing from the observed sample of embedded sources? As discussed in detail in Paper II, whether or not a population of embedded objects with high \\tbol\\ exists remains an open question. The Evans et al.~(2009) sample is based primarily on the association of \\emph{Spitzer} sources featuring rising SEDs and red colors with millimeter continuum sources (Enoch et al.~2009a; Dunham et al.~2008) and is likely biased against such objects since they would often not meet these criteria and would instead be assumed to be chance alignments between millimeter continuum sources and background sources and/or later-stage YSOs. Furthermore, the extinction corrections applied by Evans et al.~(2009) to the embedded sources are average extinctions over each individual cloud rather than true line-of-sight extinctions, and thus may underestimate the true extinction since current, active star formation (and thus the position of the youngest, embedded sources) is associated with the densest parts of molecular clouds (e.g., Heiderman et al.~2010; Lada et al.~2010). As shown in Paper II, such underestimates could, in the worst cases, artifically lower \\tbol\\ from several thousand K to several hundred K. Future work must revisit the observational samples and carefully evaluate whether or not a population of embedded sources with high enough \\tbol\\ to be classified as Class II or Class III exists. \\subsection{Duration of the Embedded Phase}\\label{sec_discussion_durations} \\begin{figure} \\epsscale{1.0} \\plotone{f12.eps} \\caption{\\label{fig_duration}\\emph{Left:} $t_{emb}$ vs.~\\mstar\\ for the 23 models listed in Table \\ref{tab_models} and considered in this paper. \\emph{Right:} Average \\mdotstar\\ over the embedded phase duration vs.~\\mstar\\ for the same models.} \\end{figure} For the 23 models listed in Table \\ref{tab_models} and considered in this paper, the duration of the embedded phase ($t_{emb}$) ranges from 0.052 (model 19) to 0.951 Myr (model 11). For each model, Figure \\ref{fig_duration} plots both $t_{emb}$ (left panel) and the average accretion rate over the embedded phase duration (right panel, calculated as final stellar mass divided by $t_{emb}$) vs.~the final stellar mass produced by the model. The duration of the embedded phase increases approximately linearly with the final stellar mass produced, and thus the average mass accretion rate (not to be confused with the \\emph{instantaneous} mass accretion rate, which is highly variable [see Figure \\ref{fig_mdot}]) is approximately constant and does not depend on final stellar mass. Offner \\& McKee (2011) argued that models that tend toward constant accretion time rather than constant accretion rate are necessary to match the observed protostellar luminosity distribution, consistent with the conclusion by Myers (2010) that accretion rates that increase with mass can at least partially resolve the luminosity problem. We disagree that such models are necessary. Our models tend toward a constant average accretion rate of $\\sim 1-3 \\times 10^{-6}$ \\msun\\ yr$^{-1}$ rather than a constant accretion time yet still provide an excellent match to the observed protostellar luminosity distribution. The average duration of the embedded phase, weighted by final stellar mass as described in \\S \\ref{sec_results}, is 0.12 Myr. In contrast, Evans et al.~(2009) derived an embedded phase lifetime of $t_{emb} = 0.44$ Myr based on the ratio of Class 0+I sources to Class II sources in the c2d sample and the assumption of a Class II lifetime of 2 Myr. Thus these models predict a significantly shorter $t_{emb}$ than suggested by recent observations. However, a number of caveats apply to this comparison: (1) We have taken the Class I/II boundary to be the point at which 10\\% of the initial core mass remains and terminate the models at this point. In reality the exact point at which to set this Class boundary is uncertain and could easily shift the duration of the embedded phase by factors of $\\sim$ 2 in either direction; (2) The observationally determined $t_{emb}$ is pinned to a Class II lifetime of 2 Myr, but the uncertainty in this lifetime is about 1 Myr (50\\%; see discussion in Evans et al.~2009); and (3) The number of Class 0+I sources in the Evans et al.~sample may be overestimated. The third point above is emphasized by three recent studies. First, van Kempen et al.~(2009) observed 22 Class I sources in Ophiuchus in \\hcop\\ $J = 4-3$, a tracer of warm and dense gas, and showed that 11 (50\\%) are not detected and thus show no evidence of being surrounded by a dense core. Second, McClure et al.~(2010) used their revised extinction law and classification method to show that greater than 50\\% of the Class I sources in Ophiuchus are highly extincted disk sources no longer embedded within cores. Third, Heiderman et al.~(2010) observed 53 Class I sources in a variety of nearby star-forming regions in \\hcop\\ $J = 3-2$, a tracer of dense gas, and showed that 31 (58\\%) are not detected and thus not associated with a dense core. All three studies likely represent upper limits to the true fraction of ``fake'' Class I sources. In the cases of van Kempen et al.~(2009) and McClure et al.~(2010) this is because there is substantial evidence that Ophiuchus is located behind an extinction screen that, if not properly accounted for, will redden source SEDs and artifically increase the number of Class I sources (see, e.g., Figure 12 of Evans et al.~2009). In the case of Heiderman et al.~(2010) this is because their study specifically targeted suspiscious Class I sources located in low extinction regions of clouds. We thus consider 50\\% as an upper limit to the fraction of Class I sources in the Evans et al.~(2009) sample that are actually misclassified Class II sources. Shifting 50\\% of the Class I sources to Class II would decrease the Evans et al.~(2009) observationally determined value of $t_{emb}$ from 0.44 Myr to 0.2 Myr. Combined with the other caveats mentioned above, the duration of the embedded phase predicted by our models is marginally consistent with observations but should be revisited in the future as uncertainties in the observations are improved. Finally, the Stage 0 ([\\mstar+\\mdisk] / [\\mstar+\\mdisk+\\mcore] $\\leq 0.5$; \\andre\\ et al.~1993) duration for the 23 models considered in this paper ranges from 0.009 (model 19) to 0.256 (model 15) Myr, with an average (again weighted by final stellar mass as described in \\S \\ref{sec_discussion}) of 0.027 Myr. Compared to the total embedded duration of 0.12 Myr, our models predict that the Stage 0 phase is only 23\\% of the total embedded duration. This is a natural consequence of the fact that these models feature average mass accretion rates that decrease with time (Vorobyov \\& Basu 2010; see Figure \\ref{fig_mdot}), thus the first 50\\% of the mass will accrete from the core faster than the second 50\\% of the mass. Our results are generally consistent with recent observationally determined estimates of the lifetime of Stage 0 relative to the total embedded phase. For example, Enoch et al.~(2009a) found 39 Class 0 and 89 Class I sources in Perseus, Ophiuchus, and Serpens, giving a relative Stage 0 lifetime of 30\\% of the total embedded phase duration. Additionally, Maury et al.~(2011) found that between $9-12$ of the 57 protostars they identified in the Serpens South cluster (Gutermuth et al.~2008) were Class 0 sources, giving a relative Stage 0 lifetime of 16\\%$-$21\\% of the total embedded phase duration. While the Enoch et al.~and Maury et al.~results are difficult to compare quantitatively since they use different classification methods that trace the underlying physical Stage to different degrees of accuracy (\\tbol\\ in the case of Enoch et al.~and position in \\lbol$-$\\mcore\\ space in the case of Maury et al.), the general agreement between our models and these observational results is encouraging. \\subsection{Number, Duration, and Importance of Bursts} \\label{sec_discussion_bursts} The 23 models considered in this paper feature between 0 (model 16) and 97 (model 15) accretion bursts, where the exact criteria for defining bursts is given in Appendix \\ref{app_resamp}. The percentage of total time spent in bursts ranges from 0\\% (model 16) to 11.8\\% (model 15), and the percentage of total mass accreted in bursts ranges from 0\\% (model 16) to 35.5\\% (model 4). On average (where the average is weighted by final stellar mass as described in \\S \\ref{sec_results}), 1.3\\% of the total time is spent in bursts and 5.3\\% of the total mass is accreted in these bursts. These values represent the statistical average values of the fraction of total time spent and mass accreted in bursts assuming a standard Kroupa IMF (see Section \\ref{sec_results} for details). We caution that the exact values depend on the criteria used for defining bursts and would increase if a lower accretion rate floor were used (see Appendix \\ref{app_resamp}). The simple, parameterized models presented in Paper II spend between $1.5\\%-2$\\% of their total time in bursts, consistent with the 1.3\\% of total time featured by these models. However, in the Paper II models between $50\\%-91\\%$ of the final stellar mass accretes in bursts, in stark contrast to the 5.3\\% of total mass that accretes in bursts in the models considered here. The explanation for this large difference lies in the detailed implementation of accretion bursts in Paper II. As described in \\S \\ref{sec_intro}, a burst was triggered each time the ratio of \\mdisk\\ to \\mstar\\ exceeded 0.2. At this point the accretion rate onto the star was increased from 0 \\msun\\ yr$^{-1}$ to $10^{-4}$ \\msun\\ yr$^{-1}$ until the disk was fully drained of mass, allowing the cycle to begin anew. However, in reality only the most extreme bursts reach accretion rates of $10^{-4}$ \\msun\\ yr$^{-1}$, as evident from Figure \\ref{fig_mdot}. Furthermore, only a very small amount of the mass in the disk accretes in a given burst (about $10^{-2}$ \\msun, on average), whereas in Paper II the assumption that the disk was fully drained in each burst led to situations where up to $0.1-0.2$ \\msun\\ were accreting in single bursts. The models considered here, based on actual hydrodynamic simulations rather than simple parameterizations, are significantly more realistic. Comparison to the results presented in Papers I and II demonstrate that the accretion process predicted by the Vorobyov \\& Basu (2005b, 2006, 2010) simulations essentially resolves the luminosity problem inherent in models with constant mass accretion. As first noted by Kenyon et al.~(1990), there are two types of non-steady mass accretion that could potentially resolve the luminosity problem: (1) accretion rates that start high and then decrease with time, and (2) generally low accretion rates punctuated by short, episodic bursts of high accretion. Figure \\ref{fig_mdot} clearly illustrates that the Vorobyov \\& Basu simulations feature both declining accretion rates with time \\emph{and} short, episodic accretion bursts. Which of these is responsible for resolving the luminosity problem? \\begin{figure} \\epsscale{1.0} \\plotone{f13.eps} \\caption{\\label{fig_lboltbolhist_avg}Histograms showing the fraction of total sources (observations; solid filled histogram) and fraction of total time spent by all models after time-averaging as described in \\S \\ref{sec_discussion_bursts} (dashed unfilled histogram; calculated from Equation \\ref{eq_bins}) at various \\lbol\\ (left) and \\tbol\\ (right). The binsize is 1/3 dex in both quantities. For the observations, only the 112 embedded sources (plotted as filled circles on the BLT diagrams) are included.} \\end{figure} Given that, on average, only 5.3\\% of the total mass accretes in bursts, one might suspect that it is the declining accretion rates with time rather than the bursts that lower model luminosities and improve the match to observations. However, this 5.3\\% excludes all of the mass that accretes in lower-amplitude accretion rate increases that do not meet the criteria for a burst as defined in \\S \\ref{app_resamp}. Thus, to properly evaluate whether the bursts are required in order to resolve the luminosity problem, we have time-averaged the accretion rates to filter out the effects of the bursts and variability and re-run all 23 models. We averaged all models over 20,000 yr durations unless the total model duration was less than 0.2 Myr, in which case we decreased the averaging window to either 10,000 or 5,000 yr in order to preserve at least 10 timesteps between the formation of the protostar and the end of the embedded phase. Figure \\ref{fig_lboltbolhist_avg} plots histograms comparing the fraction of total time the time-averaged models spend at various \\lbol\\ and \\tbol\\ to the observed distributions, similar to Figure \\ref{fig_lboltbolhist}. While the time-averaged models provide a much better match to observations than models with constant mass accretion (see Papers I and II for such models), comparing Figures \\ref{fig_lboltbolhist} and \\ref{fig_lboltbolhist_avg} shows that the time-averaged models feature a small shift to higher luminosities and do not do quite as good of a job resolving the luminosity problem. This is confirmed by a K-S test on the observed and time-averaged model luminosity distributions, which returns a value of 0.59, lower than the value of 0.85 returned for the original models. \\begin{figure} \\epsscale{1.0} \\plotone{f14.eps} \\caption{\\label{fig_arate_avg}\\mdotstar\\ vs.~time for the original (thin black line) and time-averaged (thick gray line) timesteps for Model 11. A window of 20,000 yr was used for the time-averaging. Note that the variability and bursts have not been fully filtered out.} \\end{figure} These results indicate that the declining accretion rates are not solely responsible for resolving the luminosity problem, a finding that is only reinforced by the fact that, even with the large windows over which we have time-averaged, the variability and bursts are not fully filtered out\\footnote{In theory we could adopt even larger averaging windows, but to do so would decrease the total number of timesteps below 10 for many models and risk not fully sampling the decline of the accretion rate with time.} (see Figure \\ref{fig_arate_avg} for an example). We thus conclude that it is a combination of both the accretion rates that decline with time and the variability and episodic bursts that resolve the luminosity problem. We consider this to be a plausible result given that the Vorobyov \\& Basu simulations self-consistently predict both, and we argue that the fact that the bursts are required is in general agreement with the other indirect evidence for accretion variability and bursts described in \\S \\ref{sec_intro}. In this paper we have coupled radiative transfer models with hydrodynamical simulations of collapsing cores predicting accretion rates that both decline with time and feature episodic accretion bursts caused by fragments torqued onto the protostar. We have calculated the time evolution of standard observational signatures (\\lbol, \\tbol, and \\lbolsmm) for cores collapsing following these simulations. We have compared our results to a database of 1024 YSOs containing 112 embedded protostars recently compiled by Evans et al.~(2009). We summarize our main conclusions as follows: \\begin{enumerate} \\item The hydrodynamical simulations presented by Vorobyov \\& Basu (2005, 2006, 2010) reproduce the full spread of observed embedded protostars in a diagram of \\lbol\\ vs.~\\tbol. The models resolve the luminosity problem and provide a reasonable match to the observed protostellar luminosity distribution (K-S value of 0.85). The models predict a large number of sources at very low ($\\la 0.1$ \\lsun) luminosities absent in the observations due to the observational sensitivity limit; removing such low luminosities from the comparison between models and observations increases the K-S value to 0.87. The models predict that only 0.2\\% of the total time is spent at \\lbol\\ $\\geq$ 100 \\lsun; a larger dataset than the 112 protostars considered here is necessary to test this prediction. Time-averaged models that filter out the accretion variability and bursts do not provide as good of a match to the observed luminosity problem, suggesting that the bursts are required. \\item The models do not provide a good match to the distribution of observed \\tbol\\ for embedded protostars (K-S value of 0.42). Instead, the models predict a subtantial population of embedded protostars at \\tbol\\ $\\ga$ 1000 K and thus classified as Class II or Class III sources. Most of this discrepancy arises from the method by which we adopted analytic disk and core density profiles and will be alleviated with future models that incorporate the exact physical structure from the hydro simulations, but some of the discrepancy may also be due to a population of embedded protostars with high values of \\tbol\\ missing from the current database of protostars due to the selection criteria applied to construct this database. The planned future models will not significantly alter the model luminosity distribution since this is primarily determined by the accretion rates and protostellar masses, not by the detailed physical structure adopted in the radiative transfer calculations. \\item The models reproduce the full spread of sources in a plot of \\lbol\\ vs.~\\mcore, including the existence of very low luminosity protostars with \\lbol\\ $\\sim 0.1-0.2$ \\lsun\\ but relatively high core masses of $1-2$ \\msun. \\item The duration of each model is approximately proportional to the final stellar mass produced, yet these models resolve the luminosity problem and provide an excellent match to the observed protostellar luminosity distribution. This is in contrast to recent results in the literature claiming that models that tend toward constant accretion time rather than constant accretion rate are necessary to match the observed distribution. \\item The IMF-weighted average duration of the embedded phase in our models is 0.12 Myr, whereas Evans et al.~(2009) recently determined the embedded phase duration to be 0.44 Myr. We have suggested a number of possible means by which these two estimates of the embedded phase duration may be reconciled. The IMF-weighted average model Stage 0 duration is 0.027 Myr, or 23\\% of the total embedded phase duration. Observationally determined values based on the ratio of Class 0 to Class I sources range from $16\\%-30\\%$ (Enoch et al.~2009a; Maury et al.~2011). Our models are consistent with this range. \\item On average, these models spend 1.3\\% of their total time in accretion bursts, during which time 5.3\\% of the final stellar mass accretes. In the most extreme models these values reach 11.8\\% and 35.5\\%, respectively. Thus accretion is not truly ``episodic'' since it actually occurs at all times rather than only in episodes. A better description is that accretion is ``variable with episodic bursts.'' \\end{enumerate} Future work must concentrate on improving the accuracy and self-consistency of the hydrodynamical simulations and radiative transfer models, and on compiling a more complete and more accurate database of protostars and their observational signatures. Nevertheless, we expect our primary conclusion that the Vorobyov \\& Basu (2005, 2006, 2010) simulations resolve the luminosity problem and match the observed protostellar luminosity distribution to remain unchanged to such future improvements." }, "1112/1112.4430_arXiv.txt": { "abstract": " ", "introduction": "With the advent of future wide-field surveys, weak gravitational lensing will become a premier probe of cosmology and an important tool to constrain dark energy, neutrinos and the initial conditions (see for instance \\cite{2010AnP...522..324D,Giannantonio:2011ya}). In order to fully exploit the potentiality of the convergence and shear fields, it will be important to use their whole statistics. In particular, the lensing bispectrum represents a complementary probe to the power spectrum, as it will provide constraints that are comparable to those obtained from the power spectrum alone \\cite{Berge:2009xj}. One of the primary interests of wide-field surveys is to look for primordial non-Gaussianities generated in the early universe and the lensing bispectrum represents a natural tool to capture such features \\cite{Bernardeau:2002fc,Takada:2003sv,2011MNRAS.411..595P,Jeong:2011rh,2011arXiv1107.1656S}. Given the current and forecasted constraints on $\\fNL$, there is a plethora of second-order effects intervening between the initial conditions and the observations that could be potentially relevant. Some of these effects, such as the second-order growth of matter fluctuations in Eulerian perturbation theory, the Born correction, the lens-lens coupling and the non-linear conversion between the galaxy shape distortion and the (observable) reduced shear, have been thoroughly studied in the past years and are expected to dominate the lensing 3-point statistics on small-angular scales \\cite{1997A&A...322....1B,Cooray:2000uu,2002A&A...389L..28B,Cooray:2002mj,2003MNRAS.344..857T,Dodelson:2005zj,2006JCAP...03..007S,2010A&A...523A..28K,Dodelson:2005rf,Schneider:1997ge}. When studying weak lensing one usually restricts the analysis to small angular separations (or large multipole moments), which is justified by the fact that so far cosmic shear surveys have covered only a limited portion of the sky. However, nearly full-sky surveys are currently under preparation and on large angular scales general relativistic second-order couplings will become relevant. These corrections are usually neglected because they are suppressed by the ratio between the scale probed and the Hubble scale, but are important on angular scales of the order of the angular diameter distance to the source. These are analogous to the second-order effects from general relativity affecting the CMB bispectrum when at least one of the scales probed is larger than the horizon at recombination \\cite{Bartolo:2004ty,Boubekeur:2008kn,Pitrou:2010sn}. In the CMB, these effects have been recently found to be negligible for the contamination of Planck searches for a primordial local signal in the squeezed limit \\cite{Creminelli:2011sq,Bartolo:2011wb}. In case of cosmic shear observations, these effects however cannot be a priori neglected. The aim here is to provide an exhaustive evaluation of the impact of these contributions to observations. Those results extend recent studies on the impact of relativistic corrections to the observations of large-scale galaxy clustering in \\cite{Yoo:2009au,Bonvin:2011bg,Challinor:2011bk,Baldauf:2011bh,Yoo:2011zc}. The complete study of all second-order effects in the cosmic shear, including the general relativistic ones, has been recently undertaken in \\cite{Bernardeau:2009bm} by solving at second order the Sachs equation \\cite{1961RSPSA.264..309S}, which describes the deformation of the cross-section of a light bundle, mapping galaxy shapes into their angular images. The main results of this paper are reviewed in \\sect~\\ref{sec:GReffects}. The goal of the current paper is to compute the bispectrum of the electric part of the cosmic shear from all second-order effects and compare the importance of the general relativistic corrections with the standard couplings. We review the computation of the shear power spectrum in \\sect~\\ref{sec:PS} where we also present one of the numerical shortcuts we will use throughout the paper, the second-order Limber approximation. The formal expression of the bispectrum, computed in \\sect~\\ref{sec:BS}, is determined by two types of non-linear contributions: the dynamical couplings, which depend on the particular metric solution of Einstein's equations at second order, and the geometrical couplings, so-called because they depend on the geometry of the solution of the Sachs equation at second order. Among the latter we also include the corrections coming from the inhomogeneities of a fixed-redshift source plane. The calculations of \\sect~\\ref{sec:BS} involve multiple integrals and complicated manipulations of spherical harmonics and Wigner symbols. (In appendix~\\ref{app:spart} one can find useful relations for these calculations.) We are afraid to say that those calculations are extremely lengthy and technical. Impatient readers can skip this section and go directly to \\sect~\\ref{sec:results}. In this section we present our results concentrating on the squeezed limit, where one of the multipole moments $l$ involved in the bispectrum is small and we discuss the functional forms of the resulting terms as well as their relative importance. In \\sect~\\ref{sec:primordial} we compare our results with the bispectrum generated by primordial non-Gaussianities of the local type and we compute their contamination to a primordial $f_{\\rm NL}^{\\rm loc}$. Finally, in \\sect~\\ref{sec:conclusion} we conclude and discuss the results of the paper. ", "conclusions": "\\label{sec:conclusion} In this paper we present an exhaustive calculation of the full-sky bispectrum of the shear field as it would be built by an observer measuring shapes of galaxies at a given redshift. The calculation is performed at tree order (i.e.~neglecting loop corrections), assuming that the initial curvature perturbation $\\zeta$ obeys Gaussian statistics, such as after single-field inflation. Furthermore, we study the relative importance of the different couplings and we compare them with primordial non-Gaussianities of the local type. Our derivation fully exploits and completes the recent study of Ref.~\\cite{Bernardeau:2009bm}, where the full-sky second-order expression of the shear field has been computed. This article derives all second-order terms in the metric fluctuations that contribute to the shear field and it is thus the natural starting point of the computation of the tree order bispectrum presented here. In this paper we focus our calculations on the bispectrum of the scalar, i.e.~electric, part of the shear field alone. This is somewhat restrictive since the second-order shear field contains a pseudo-scalar, i.e.~magnetic, mode. We leave the study of the $B$-mode and of the correlators where it is relevant for the future. The computation of the bispectrum from the formal second-order shear expression is an involved exercise, which includes two steps. The first consists in the formal derivation of the bispectrum in harmonic space. The electric part of the shear field itself is defined from its harmonic space expression, in a way similar to the mode decompositions of the CMB polarization. Its bispectrum is then computed with the help of the decomposition properties of the spherical harmonics and their spin weighted extensions. These results are presented in \\sect~\\ref{sec:BS} of the paper where the contribution of each term is fully computed. Then, the second part consists in numerically evaluating the terms that we have found, in the context of the $\\Lambda$CDM model. Two approximations are commonly employed for weak lensing calculations, namely the Limber approximation \\cite{1953ApJ...117..134L}, which consists in neglecting longitudinal modes, as they tend to average out along the line of sight, and the flat sky approximation, which consists in replacing a small portion of the spherical sky with a 2-dimensional plane.\\footnote{Note that these two approximations are not equivalent and one does not necessarily imply the other.} In order to be consistent with our full-sky treatment, it is a priori not possible to employ neither of these two approximations. Abandoning the Limber approximation makes the computations numerically challenging due to the large number of integrations involved and, as a consequence, only a limited number of configurations could be exactly computed. In order to circumvent this problem we have used the higher-order Limber approximation discussed in \\cite{1992ApJ...388..272K}. Apart from for a specific type of contributions, we have found that sufficiently high accuracy can be reached with the help of the second-order approximation. The procedure that we have used to implement this approximation to the angular power spectrum and the bispectrum is presented in appendix~\\ref{app:limber}. Our results are described in Sect. 5. In this section we give a precise account of the relative importance of the various contributing terms, as well as of their dependence with the multipole configuration. We distinguish the different contributing terms as geometrical and dynamical, depending on whether they originate from pure general relativistic effects on the line-of-sights or due to the dynamical evolution of the metric field. Note however that this distinction is somewhat arbitrary, since it depends on the gauge choice. Furthermore, in each case we distinguish the standard terms as those which survive at small angular scale, from the non-standard ones that are mainly relevant at large angular scales. We discuss the contribution to the bispectrum of the different nonlinear couplings focussing on the squeezed limit configuration, where the non-standard general relativistic corrections are more likely to play a significant role. In \\sect~\\ref{sec:results} we give a precise account of the relative importance of the various contributing terms as well as of their dependence with the multipole configuration. We find that the general relativistic effects, i.e. the non-standard coupling terms, are typically smaller than the standard non-linear couplings. However their relative importance increases at smaller source redshifts. Among the new couplings, the dynamical couplings are always much smaller than the others, while those due to the inhomogeneity in the redshift of the source dominate. The latter can even become of the same order of the standard couplings when the redshift of the sources is below $0.5$. These results are summarized in fig.~\\ref{fig:bz1} and \\ref{fig:bzp0}. In \\sect~\\ref{sec:primordial} we finally compute the corresponding level of contamination induced by the standard and non-standard terms on the amplitude of primordial non-Gaussianities of the local type. Note that, as in the squeezed limit the standard non-linear couplings depend on the angle between the short and long mode, they differ significantly from a local signal. For $z_S \\gtrsim0.8$, the contamination to $f_{\\rm NL}^{\\rm loc}$ of the standard geometrical couplings amounts to $f_{\\rm NL}^{\\rm loc} \\gtrsim 100$. On the other hand, the relativistic corrections induce a non-Gaussianity which is mainly of the local type and can contaminate the search for a primordial signal by $f_{\\rm NL}^{\\rm loc} \\sim 1$ to $f_{\\rm NL}^{\\rm loc}\\gtrsim 10$ when the source redshifts vary from $1$ to $0.3$. Our investigations did not go as far as computing the signal-to-noise ratio for the bispectrum induced by the general relativistic corrections. However, given the much larger amplitude of the standard contributions, we expect it to be rather small. This might not be the case when, besides the only $E$-mode three-point correlators we have investigated here, quantities involving $B$-modes are also included. Relativistic corrections indeed generically induce $B$-modes with an amplitude a priori comparable to the $E$-modes~\\cite{Bernardeau:2009bm}. This is at variance with the standard dynamical couplings that induce second-order $E$-modes only. It ensures that the correlations between $E$ and $B$ modes are only non vanishing in the presence of general relativistic corrections making such quantities more likely to be observable. We leave such calculations for a future study." }, "1112/1112.2600_arXiv.txt": { "abstract": "The AGILE scientific instrument has been calibrated with a tagged $\\gamma$-ray beam at the Beam Test Facility (BTF) of the INFN Laboratori Nazionali di Frascati (LNF). The goal of the calibration was the measure of the Point Spread Function (PSF) as a function of the photon energy and incident angle and the validation of the Monte Carlo (MC) simulation of the silicon tracker operation. The calibration setup is described and some preliminary results are presented. ", "introduction": "AGILE (Astro-rivelatore Gamma a Immagini LEggero) is a Small Scientific Mission of the Italian Space Agency (ASI) launched on April 2007 and dedicated to high-energy astrophysics \\cite{agimis}. The AGILE satellite is designed to detect and image photons in the 18 - 60 keV, 30 MeV - 50 GeV and 350 keV - 100 MeV energy bands with excellent spatial resolution, timing capability, and large field of view.\\\\ AGILE is the most compact ($\\approx 0.25 m^3$), light (120 kg for the instrument, 350 kg for the whole satellite) and low power ($\\approx 60 W$) scientific instrument ever developed for high-energy astrophysics.\\\\ The AGILE scientific payload (shown in Fig.\\ref{agidraw}) consists of three detectors with independent detection capability. The Gamma-Ray Imaging Detector (GRID) consists of a Si-W converter-tracker \\cite{st1} sensitive in the $\\gamma$-ray energy range 30 MeV - 50 GeV, a shallow ($1.5\\, X_0$ on-axis) CsI Calorimeter \\cite{minical} and a segmented AntiCoincidence system based on plastic scintillators \\cite{ac}.\\\\ In addition to the GRID, a coded-mask hard X-ray imaging system (SuperAGILE), made of a Si detector plane and a W mask, ensures coverage in the range $18-60 \\mathrm{keV}$ \\cite{superagile}.\\\\ The AGILE main feature is the combination of two co-aligned imaging detectors (SuperAGILE and GRID) sensitive in the hard X-ray and in the $\\gamma$-ray ranges with large field of view ($\\approx 1.0\\mathrm{sr}$ and $\\approx 2.5 \\mathrm{sr}$ respectively).\\\\ Moreover the CsI MiniCalorimeter (MCAL) can operate in stand alone \"burst mode\" in the $350\\,\\mathrm{keV}- 100\\,\\mathrm{MeV}$ range to detect GRB.\\\\ On ground and subsequently on flight calibrations of a detector are essential to the interpretation of its results. The purpose of the calibration of a scientific instrument is to reproduce, under controlled condition, the detector response in operation.\\\\ This paper describes the on-ground calibration of the silicon tracker and some results on the instrument performances derived by it. \\begin{figure}[!ht] \\begin{center} \\includegraphics[width=5cm]{AgileFigure.eps} \\end{center} \\caption{A schematic view of the AGILE scientific instrument.} \\label{agidraw} \\end{figure} ", "conclusions": "This paper presents some preliminary results of the calibration of the AGILE ST at the BTF of the LNF in 2005.\\\\ The setup is described in detail as well as the calibration requirements. We discussed the problems encountered in exploiting the PTS originally designed and a novel approach devised to circumvent those problems: the phase analysis.\\\\ We concentrated on the measurements of the PSF presenting two possible definitions: the Gaussian and the 68\\%\\ PSF.\\\\ The calibration results are compared with the MC simulations for a broad set of variables, showing good consistency with some poorer agreement for 68\\%\\ PSF mainly at low energies.\\\\ These results give confidence on the use of the MC simulation in the untested part of the $\\gamma$-ray parameters (e.g. higher $E_\\gamma$) especially in flight conditions, i.e. without low energy background, and in the measurement of detector parameters, like absolute efficiency and energy resolution, that are difficult to measure without exploiting the PTS information." }, "1112/1112.5216_arXiv.txt": { "abstract": "In this paper, a decay vacuum model $\\bar{\\rho}_\\Lambda=3\\sigma M_p^2H_0 H$ is revisited by detailed analysis of background evolution and perturbation equations. We show the imprints on CMB temperature and matter power spectrum from the effective coupling terms between dark sectors by comparing to the standard cosmological constant model and observational data points (WMAP7 and SDSS DR7). We find that the decay vacuum model can describe the expansion rate at late times as well as the standard cosmological constant model but it fails to simultaneously reproduce the observed CMB and matter power spectrum. Its generalization $\\bar{\\rho}_\\Lambda=3M_p^2(\\xi_1 H_0 H+\\xi_2 H^2)$ is also discussed. Detailed analysis of the background evolution shows that the dimensionless parameter $\\xi_{2}$ would be zero to avoid the unnatural 'fine tuning' and to keep the positivity of energy density of dark matter and dark energy in the early epoch. ", "introduction": "Since the accelerating expansion of the universe has been found from the measures of the luminosity-redshift relation $d_L(z)$ of type Ia supernovae (SN Ia) \\cite{Riess}, a cosmic component called dark energy was often introduced to explain the acceleration within the framework of general relativity. Now more and more evidence, such as cosmic microwave background (CMB) \\cite{Bennett,ref:wmap7}, baryon acoustic oscillations (BAO) \\cite{Eisenstein}, weak gravitational lensing \\cite{Jarvis} and x-ray clusters \\cite{Allen,Allen2}, indicated that the universe is spatially flat and dominated by dark energy at present. Apart from dark energy models, modified gravity \\cite{Carroll,Trodden} can also explain the acceleration of the present universe. However, we just focus on dark energy model in this paper. Among the various dark energy models \\cite{Copeland}, including scalar field \\cite{Ratra1}, vector field\\cite{Zhang94,Kiselev}, holographic dark energy \\cite{Holographic}, Chaplygin gas \\cite{Kamenshchik} and so on. The cosmological constant model ($\\Lambda$CDM) \\cite{Weinberg} is the simplest one. However, as is well known, the $\\Lambda$CDM suffers from a fine tuning problem: the observed vacuum energy density of order $\\sim10^{-47}\\text{GeV}^4$ is about $10^{121}$ orders of magnitude smaller than the value expected by quantum field theory for a cutoff scale that is the Plank scale, and is still about $10^{44}$ orders smaller even for a cutoff scale that is the QCD scale \\cite{Copeland}. As an extension to $\\Lambda$CDM, the decaying vacuum (DV) dark energy model was proposed \\cite{Borges1,Carneiro}, based on the incomplete quantum field theory in the curved 4-dimension space-time. In this model, the vacuum serves as dark energy, whose energy density decays with the expansion of the universe leading to an additional production of the matter component. In the late-time with a quasi-de Sitter background, the vacuum density is proportional to the Hubble rate, $\\rho_\\Lambda(t) \\propto H(t)$. However, the equation of state for the vacuum is a constant value $w=p_\\Lambda(t)/\\rho_\\Lambda(t)=-1$, the same as that in the $\\Lambda$CDM model. Moreover, as an interesting feature, the late-time dynamics of the DV model is similar to $\\Lambda$CDM \\cite{Borges1,Carneiro}. The quasar APM 08279+5255 at $z=3.91$ was used to examine the DV model \\cite{mltong2}, and it was found that the DV model can greatly alleviate the high redshift cosmic age problem existing in the $\\Lambda$CDM model. In order to distinguish the DV model from other dark energy models at the late-time Universe, the state finder and $Om$ diagnostics of the DV model were also presented in \\cite{mltong2}. Moreover, the DV model has been tested by $\\chi^2$ analysis using the observational data of SN Ia \\cite{Carneiro2}, a joint data from SN Ia, BAO and CMB\\cite{Carneiro3,Pigozzo}, and the joint data that the Gamma-ray bursts, Hubble rate and x rays in galaxy clusters were added \\cite{tongnoh}. It was found that, the DV model favors a relatively larger value of the matter density contrast, $\\Omega_m=(0.34\\sim0.43)$. In this paper, compared to previous work \\cite{Pigozzo,tongnoh}, we take the radiation component into account in a more reasonable way. We will demonstrate the temperature anisotropies of CMB induced by the matter perturbations in the DV model with various values of $\\Omega_m$. The matter power spectrum is investigated. This paper is structured as follows. In section \\ref{sec:bg}, the background evolution equations are given in a spatially flat FRW universe. We give a brief review of cosmological perturbation theory in section \\ref{sec:cp}. The main results of this paper are presented in section \\ref{sec:results}. In section \\ref{sec:tvv}, we give a brief discussion on a generalized form $\\bar{\\rho}_{\\Lambda}=3M_p^2(\\xi_1 H_0 H+\\xi_2 H^2)$ and point out that this model is not viable. Section \\ref{sec:con} is the conclusion. ", "conclusions": "\\label{sec:con} In this paper, a decay vacuum model $\\bar{\\rho}_\\Lambda=3\\sigma M_p^2H_0 H$ and its generalization $\\bar{\\rho}_\\Lambda=3M_p^2(\\xi_1 H_0 H+\\xi_2 H^2)$, a time variable cosmological constant model, are revisited. At first, the background evolution equation in a spatially flat FRW universe containing cold dark matter, radiation, baryon and time variable cosmological constant is given. The relative departure from $\\Lambda$CDM model is minor, please see the left panel of Fig. \\ref{fig:evolution}. So to discriminate the decay vacuum model from $\\Lambda$CDM model, high redshift observations are needed. In the decay vacuum model case, an effective interaction between cold dark matter and vacuum can be introduced. Then the evolution of cold dark matter will depart from the conventional power law $a^{-3}$. And the large scale structure formation will be strongly different from that of $\\Lambda$CDM model. Though the baryon component evolves in the scaling $a^{-3}$, the background evolves different from $\\Lambda$CDM model for the effective interaction between cold dark matter and decay vacuum. Then the dynamic evolution would be modified. So the cosmological perturbations are taken into account. As results, the angular power spectrum of CMB and matter power are presented with different parameter values of cold dark matter abundance $\\omega_c$, please see Fig. \\ref{fig:clspk}. From this figure, one can conclude that CMB observations and matter power spectrum can distinguish the decay vacuum model from $\\Lambda$CDM model markedly. When $\\Omega_{c} h^{2}=0.2158$, i.e. $\\Omega_{c0}=0.4404$, the purple dashed line in Fig. \\ref{fig:clspk} is close to observational data points. It means that increasing the abundance of cold dark matter will depress the acoustic peaks to cosmic observational data points in this model. However, in the right panel of Fig. \\ref{fig:clspk}, one sees that increasing the abundance of cold dark matter will enhance the matter power spectrum at small scale but depress that at large scale. That makes it difficult to match observational data points. With these observations, this model would be ruled out. But to know in what kind of levels to rule out this model, testing this model with current available cosmic observational data sets, for example type Ia supernovae, baryon acoustic oscillation, full CMB and SDSS DR7 etc, would be interesting. Furthermore, a generalized vacuum model $\\bar{\\rho}_{\\Lambda}=3M_p^2(\\xi_1 H_0 H+\\xi_2 H^2)$ was discussed. From a detailed analysis, one can find that the parameter space of $\\xi_{2}$ is a very small negative dimensionless parameter. To keep the positivity of energy density of dark matter and dark energy at early epoch, the parameter $\\xi_{2}$ suffers from the fine tuning problem. So to avoid this unnatural condition, the $\\xi_{2}$ would be set to zero. Then it reduces to the decay vacuum model. In this sense, it would not be a viable dark energy model." }, "1112/1112.0928_arXiv.txt": { "abstract": "{Many physical parameters change with time in star forming regions. Here we attempt to correlate changes in infall and outflow motions in high mass star forming regions with evolutionary stage using JCMT observations. } {From a sample of 45 high mass star forming regions in three phases of evolution, we investigate the presence of established infall and outflow tracers to determine whether there are any trends attributable to the age of the source.} {We obtained JCMT observations of HCO$^+$/H$^{13}$CO$^+$ J=4-3 to trace large scale infall, and SiO J=8-7 to trace recent outflow activity. We compare the infall and outflow detections to the evolutionary stage of the host source (high mass protostellar objects, hypercompact HII regions and ultracompact HII regions). We also note that the integrated intensity of SiO varies with the full width at half maximum of the H$^{13}$CO$^+$.} {We find a surprising lack of SiO detections in the middle stage (Hypercompact HII regions), which may be due to an observational bias. When SiO is detected, we find that the integrated intensity of the line increases with evolutionary stage. We also note that all of the sources with infall signatures onto Ultracompact HII regions have corresponding outflow signatures as well.} {} ", "introduction": "High mass stars are a vital player in the evolution of galaxies. During their formation they inject energy into their surroundings in the form of outflows and ionizing photons. As they evolve their stellar winds continue to stir turbulence in the surrounding medium, and when they explode as supernovae they additionally enhance the metallicity of their environs. The processes involved in this earliest stage, the formation of high mass stars is not nearly as well constrained as are the mechanisms responsible for lower mass stars. The pre main sequence evolution of low mass stars \\citep[i.e.][]{shu87,andre93} has been well constrained and statistically based lifetimes of the stages have come about, in part, due to observing large numbers of individual sources \\citep[i.e][]{Enoch09,Spezzi08}. Similar studies for the early evolution of high mass stars (M$\\gtrsim 8$ M$_\\odot$) have proven problematic due to limiting factors such as the large average distances to high mass star forming regions and the clustered nature of their formation. However studies such as \\citet{Beltran11} suggest high mass stars are likely to form though disk mediated accretion in a similar but scaled up version of lower mass star formation. An evolutionary sequence is developing, however progress is hampered by a lack of large sample surveys of the molecular gas dynamics in these region \\citep[see, for example][]{ZY07,Beuther07}. Evolution in low mass systems is often characterized by changes in the spectral energy distributions (SEDs). This is not as easily done in high mass systems \\citep{Molinari08}, which is why studying the gas dynamics is very important. In broad terms, the evolution of massive protostars begins in an infrared dark cloud \\citep[IRDC, i.e.][]{Egan98}. The central condensation then begins heating its environment, at which point the core starts emitting at IR wavelengths and it becomes a high mass protostellar object \\citep[HMPO, i.e.][]{Sridharan02} within a hot core. The protostar continues to gain mass and heat its environment eventually forming a hypercompact (HC) HII region \\citep[i.e.][]{Keto07} which then grows into an ultracompact (UC) HII region \\citep[i.e.][]{WC89}. This sequence has been shown graphically in \\citet{Zapata10}. The goal of this paper is to follow the larger scale (that observable with a single dish) outflow and infall structures surrounding sources in the later three of these four evolutionary stages (HMPO, HCHII and UCHII region), and determine whether any evolutionary trends are distinguishable. In the case of low mass star formation, \\citet{andre93} conclude that outflow motions can be seen quite early in the Class 0 phase, and \\citet{Bontemps96} state that the outflows from Class 0 sources are more powerful than those from Class I sources at the same luminosity. They also suggest that all deeply embedded low mass protostars should have outflows. As for infall in the low mass regime, inverse P-Cygni profiles indicative of infall have been seen for a decade now \\citep{difrancesco01,Zapata08,Furuya11}, and some observations even suggest that infall is seen perpendicular to the outflow direction \\citep{Arce04}. That these trends can be applied to higher mass protostars is starting to be explored \\citep[i.e.][]{LS11}. Here we present HCO$^+$ and H$^{13}$CO$^+$ (J=4-3) observations to study the bulk infall signatures in high mass star forming regions. We also present SiO (J=8-7) observations to trace recent outflow activity from the same sources. We have collected a sample of 45 high mass star forming regions in three evolutionary stages: HMPOs, and sources with HC and UCHII regions. In Section \\ref{sec:observations} we present our observations and in Section \\ref{sec:method} we present our methodology. In Section \\ref{sec:results} we present our results, which we discuss in Section \\ref{sec:discussion} where we also conclude. ", "conclusions": "" }, "1112/1112.0575_arXiv.txt": { "abstract": "We consider issues related to the conformal mapping between the Einstein and Jordan frames in $f(R)$ cosmology. We consider the impact of the conformal transformation on the gauge of a perturbed system and show that unless the system is written in a restricted set of gauges the mapping could produce an inconsistent result in the target frame. Newtonian gauge lies within the restricted group but synchronous gauge does not. If this is not treated carefully it could in principle contaminate numerical calculations. ", "introduction": "\\noindent Extended gravity theories, where the Einstein-Hilbert Lagrangian density $\\mathcal{L}_{\\rm EH}=R$ is replaced by a more general function including terms of higher-order in derivatives of the metric ($R^2$, $R_{\\mu\\nu}R^{\\mu\\nu}$, $R_{\\alpha\\beta\\mu\\nu}R^{\\alpha\\beta\\mu\\nu}\\ldots$) and couplings to new dynamical degrees of freedom, have long been of interest in relativity. The last few decades have seen increasing applications of these models to cosmology. See \\cite{Capozziello:2011et,Nojiri:2006ri} and their references for an overview and further details on these models and their applications to cosmology.\\footnote{Recently \\cite{Biswas:2011ar} presented the most general action guaranteed to be free of ghosts when quantising perturbations on a classical vacuum; a viable model of extended gravity will be constrained to conform with this action.} In the last decade attention has focused on exploiting extended gravity to model dark energy without the need to introduce exotic particle species. A model frequently employed in this context is $f(R)$ gravity (see for example \\cite{DeFelice:2010aj,Nojiri:2010wj} for recent reviews), where the Einstein-Hilbert Lagrangian density is replaced with an arbitrary function of the Ricci scalar, $\\mathcal{L}=f(R)$, while the matter couples minimally to the metric and follows its geodesics in free motion. This representation of an extended gravity model is known as the ``Jordan frame''. The action can be transformed to a variety of different forms. One of the most % common transformations is into the so-called ``Einstein frame'' where the action is manipulated to isolate a Ricci scalar of a new metric. The residual terms can be interpreted as an effective scalar field to which matter is non-minimally coupled and deflected from the geodesics of the metric. The field equations are otherwise those of standard general relativity. We review the model in the Jordan and Einstein frames in \\S\\ref{Section:Model}. Transforming from the Jordan frame to the Einstein frame is extremely useful in the study of $f(R)$ gravity. We employ the metric formulation, in which the action is varied with respect to the metric alone, and in which the field equations are fourth-order. In the Einstein frame, as in standard GR, the theory is second-order -- a significant simplification. Aspects of the transformation have been controversial for some time (see for example \\cite{Faraoni:1998qx,Deruelle:2010ht,Quiros:2011wb,Capozziello:2011et} and further references in \\S\\ref{Section:Equivalence}). The discussion has centred upon the nature of the equivalence, and authors can be separated \\cite{Faraoni:1998qx,Capozziello:2011et} into two camps: those who feel the equivalence is ``physical'' and observables can be calculated in either frame, and those who feel the equivalence is mathematical in nature and that observables should be calculated in a chosen ``physical frame''. We briefly discuss this issue in \\S\\ref{Section:Equivalence}. Modern cosmology is the study of Friedmann-Lema\\^itre-Robertson-Walker (FLRW) spacetimes, perturbed to linear or higher orders, and this provides an illustration of the major motivation for our paper. Numerical studies targeting the cosmic microwave background (CMB) and the matter power spectrum (such as \\cite{Bean:2006up,MGCAMB11,Gu:2011kk,Li:2011vk,Dossett:2011tn}) often employ modifications to Boltzmann codes such as CAMB \\cite{CAMB}. While it is common to employ a general parameterisation \\cite{MGCAMB11,Gu:2011kk,Li:2011vk,Dossett:2011tn} in the Jordan frame, working in the Einstein frame (e.g. \\cite{Schimd:2004nq,Bean:2006up}) provides a flexible alternative. Exploiting the Einstein frame allows us to consider any modified gravity model which possesses an Einstein frame and for which background solutions in the Jordan frame might be extremely difficult to find, either analytically or numerically. A recent example is the study of the string gas cosmology (see e.g. \\cite{Brandenberger:2011et}), in which the primordial power spectrum is frequently calculated in the Einstein frame as opposed to the string frame, as in \\cite{Brandenberger:2006pr,Kaloper:2006xw}. The use of the Einstein frame allows us to import our intuition and understanding from standard gravity -- at least while we remain within the frame -- and, perhaps more importantly, it allows us to directly leverage well-tested codes developed for general relativity and which contain a wealth of physics and arbitrary collections of fluids. Codes of comparable breadth would be extremely lengthy to implement in the Jordan frame, and their fourth-order nature has consequences for speed and stability. Considering the nature of the transformation for a perturbed model reveals an issue which to the best of our knowledge is as yet unappreciated. Perturbed systems in relativity exhibit the gauge issue -- the mapping between the fictitious background spacetime and the physical perturbed spacetime is non-unique. In cosmology, gauge freedoms allow one to eliminate four perturbative degrees of freedom (two scalar and two vector, where the classification is based on how the modes transform on a 3-surface), which specifies the slicing and threading we have chosen for our foliation. The transformation between the Einstein and the Jordan frames tangles this choice. We consider the impact of this and how one can resolve the resulting gauge ambiguities. We illustrate with a simple $f(R)$ model in a vacuum FLRW spacetime, but it should be emphasised that this issue will in principle occur in any perturbed spacetime and a wide range of extended theories of gravity. We choose to work in the standard (``metric'') approach to cosmological perturbations (e.g. \\cite{MukhanovFeldmanBrandenberger,MaAndBertschinger}) and leave potential issues in the 3+1 gauge-invariant and covariant (GIC) approach \\cite{EllisBruni,Li:2007xn,Abebe:2011ry} for future study.\\footnote{The issue of perturbed models of modified gravity in the 3+1 GIC approach has been considered in \\cite{Carloni:2009gp}. The authors wish to thank Sergei Odintsov for bringing this work to our attention.} While many authors working on perturbed spacetimes in modified gravity choose to employ ``gauge-invariant'' variables (for example, \\cite{Hwang1990CQG,MukhanovFeldmanBrandenberger}) or work in the Jordan frame (such as \\cite{Hwang1996PRD,Song:2006ej,Carroll:2006jn,Tsujikawa:2007tg,delaCruzDombriz:2008cp,DeFelice:2010gb,Bertacca:2011wu}), the analysis of gauge problems is necessary for three reasons. Firstly, fixing a gauge (as in \\cite{Hwang:1996np,Bean:2006up,Zhang:2007nk,Tsujikawa:2007gd}), is a valid approach, and any problems when transforming to the Einstein frame must be understood. Secondly, gauge-invariant variables are themselves based on variables in a particular gauge \\cite{Malik:2008im}. The frequently-employed Bardeen potentials are, for example, gauge-invariant generalisations of the potentials in Newtonian gauge. If there is an issue with the choice of the ``natural'' gauge of a gauge-invariant perturbation, the perturbation itself cannot necessarily be trusted. Thirdly, our study is motivated not least by the use of Boltzmann codes in the Einstein frame. By far the widest used Boltzmann code is CAMB \\cite{CAMB}, which is written in a frame equivalent to synchronous gauge.\\footnote{Of the other modern Boltzmann codes, cmbeasy \\cite{CMBEasy} can be used in synchronous or Newtonian gauges, and CLASS \\cite{CLASS} is best tested in synchronous gauge although it was updated with support for Newtonian gauge after the release of the first version of this paper. Very recently, CosmoLib \\cite{CosmoLib} was released, programmed entirely in Newtonian gauge. COMICS \\cite{COSMICs} could be employed in both Newtonian and synchronous gauges, but was rapidly superseded by CMBFast \\cite{CMBFast} which could only be used in synchronous gauge. Both of these codes are now obsolete, with CMBFast in turn superseded by CAMB and cmbeasy. Our comments concerning CAMB and synchronous gauge equally apply to other Boltzmann codes in other gauges.} If there is an issue with the frame transformation and synchronous gauge, then the initial conditions for modifications of CAMB could be suspect, as could the final results even after being transformed back into the Jordan frame. Worse, if there is an issue the error is compounded: each transformation between frames will induce errors. Given that the use of CAMB (or similar codes) is of interest, the study of transformations of systems in synchronous gauge is forced upon us, despite its known shortcomings: synchronous gauge is not fully fixed \\cite{Malik:2008im,Christopherson:2011ra} and ambiguities in the choice of threading remain. The alternative formulation of $f(R)$ gravity employs the Palatini approach, where the metric and the connections are treated as independent objects. The Palatini approach is popular in cosmology (e.g. \\cite{Capozziello:2010ef,Capozziello:2011et,Olmo:2011uz} and their references) and often employed in studies of cosmological perturbations (e.g. \\cite{Koivisto:2005yc,Li:2006vi,Sotiriou:2008rp}). The nature of the transformation between the frames implies that the issues we highlight in this paper are also relevant in the Palatini approach. Since the field equations in the Palatini approach are intrinsically second-order the usefulness of the transformation is somewhat lessened, but if a system is more easily evaluated in the Einstein frame then the potential ambiguities should still be considered. The gauge problem is discussed in \\S\\ref{Section:GaugeAmbiguity}A and illustrated in \\S\\ref{Section:GaugeAmbiguity}B. We close the paper in \\S\\ref{Section:Discussion} with a brief discussion. We employ a signature $(-+++)$ and the Ricci tensor $R_{\\mu\\nu}=R^\\alpha_{\\phantom{\\alpha}\\mu\\alpha\\nu}=\\Gamma^\\alpha_{\\mu\\nu,\\alpha}+\\ldots$ (as in for example \\cite{Gravitation,Maeda,DeFelice:2010aj}) and generally follow the notation of \\cite{DeFelice:2010aj}. An overdot will represent differentiation with respect to conformal time $d\\eta=dt/a$, and the conformal Hubble rate is written $\\mathcal{H}=\\dot{a}/a$. ", "conclusions": "\\label{Section:Discussion} In this paper we have discussed a technical subtlety of the conformal transformation between the Jordan and Einstein frames that to the best of our knowledge has not been highlighted before: na\\\"ively transferring a perturbed system between the frames tangles your choice of gauge. It should be emphasised that this effect is \\emph{calculational}, not physical, and that if sufficient care has been taken no previous results are changed by this. It is also extremely important to note that this issue is not restricted to $f(R)$ gravity or to the study of cosmological perturbations, and that it potentially arises whenever a perturbed system is conformally transformed regardless of the background metric or the model of gravity. $f(R)$ theory and vacuum cosmology provide a useful, straightforward illustration. While we have concretely demonstrated the issue only for this vacuum $f(R)$ system, in principle it arises generically, although demonstrating this explicitly is beyond the scope of this paper. The properties of the conformal transformation restrict us to a particular set of gauges which we refer to as the ``restricted group'': gauges that possess a lapse, a spatial curvature, and a density perturbation. This restriction only applies when the transformation is performed. At all other times one is free to work in any convenient gauge. However, we additionally argued that there are alternatives to working in the restricted group -- one can instead choose to ``refix'' the gauge, or to evolve the system using the redundant Einstein field equations. We have illustrated our arguments with a simple vacuum system, showing that there are no ambiguities introduced by equations (\\ref{BardeenTransformations}, \\ref{FieldTransformations}), so long as one takes such care. If this is not done and the results of a transformation from a gauge outside the restricted group are interpreted as if they themselves lie in that gauge, the system will be inaccurately specified. Perhaps the most important place where this issue will arise is in the treatment of initial conditions. For instance, Boltzmann codes are frequently \\cite{COSMICs,CMBFast,CAMB,CLASS} written in synchronous gauge, which lies outside of the restricted group. Interpreting the Jordan frame as physical, the initial conditions must be set in the Jordan frame. However, regardless of the gauge chosen, the transformation into the Einstein frame induces a spurious lapse function which must be reabsorbed if the results of the calculation are to make any sense. This itself introduces an additional problem: the transformation to synchronous gauge is not fully specified, and introduces an arbitrary function of the 3-coordinates which must itself be removed with care. The alternative is to include the redundant degrees of freedom in the Boltzmann codes, which requires much additional effort. A careless study without attention to these issues would either leave the system with the spurious lapse, or risk introducing a gauge mode. Some Boltzmann codes \\cite{COSMICs,CMBEasy,CLASS} include modules written in Newtonian gauge, which lies within the restricted group, and studies of extended gravity that employ these codes remain consistent. However, COSMICS is very outdated, while cmbeasy and CLASS are programmed in C++ and C respectively; given the widespread use of Fortran in cosmology, its large codebase and the extensive testing it has undergone, CAMB remains the dominant Boltzmann code and the study of synchronous gauge is necessary.\\footnote{CosmoLib \\cite{CosmoLib} is also written in Fortran but in Newtonian gauge; however it has appeared too recently to yet have much of an impact. However, it could prove extremely useful for studies of modified gravity.} Similar arguments apply to the initial conditions employed in n-body simulations of modified gravity performed in the Einstein frame. The other time at which the gauge ambiguity becomes important is at the end of a calculation. Taking the Jordan frame to be physical, the results of a simulation in the Einstein frame must be transformed back. Given that calculations are frequently performed in synchronous gauge then one must either transform into a safer gauge before performing the conformal transformation, or deal with the redundant system in the Jordan frame. Otherwise one runs a serious risk of contamination. At this stage it seems easier to refix the gauge in the Jordan frame to a fully-specified gauge (such as Newtonian or uniform curvature), or to leave the system unspecified. So long as one does so consistently, observables will not be affected, since gauge transformations cannot change an observation. It is interesting to note that the lensing potential $\\Phi+\\Psi$ is uncontaminated by the transformation. As such, if one wishes to calculate the weak lensing signal on a constant-time hypersurface, this can be achieved transforming from the Jordan frame into the Einstein frame in any gauge and not worrying about refixing the system \\cite{Schimd:2004nq}. For a static (or quasi-static) system, this will be a good approximation. However, carelessly evolving the system without due care will in principle introduce errors, so that the frame invariance does not imply such a straightforward estimate of the integrated Sachs-Wolfe effect. In summary, we recommend that authors exploiting the Einstein frame to study extended theories of gravity work in the restricted group. This removes the need for additional gauge transformations to reabsorb the extra apparent degrees of freedom, or the need to employ redundant Einstein equations. If one chooses the latter options, care should be taken to ensure that the result remains consistent." }, "1112/1112.1745_arXiv.txt": { "abstract": "*{I discuss the effects of measurement error on regression and density estimation. I review the statistical methods that have been developed to correct for measurement error that are most popular in astronomical data analysis, discussing their advantages and disadvantages. I describe functional models for accounting for measurement error in regression, with emphasis on the methods of moments approach and the modified loss function approach. I then describe structural models for accounting for measurement error in regression and density estimation, with emphasis on maximum-likelihood and Bayesian methods. As an example of a Bayesian application, I analyze an astronomical data set subject to large measurement errors and a non-linear dependence between the response and covariate. I conclude with some directions for future research.} \\vspace{-1.5 in} \\abstract{I discuss the effects of measurement error on regression and density estimation. I review the statistical methods that have been developed to correct for measurement error that are most popular in astronomical data analysis, discussing their advantages and disadvantages. I describe functional models for accounting for measurement error in regression, with emphasis on the methods of moments approach and the modified loss function approach. I then describe structural models for accounting for measurement error in regression and density estimation, with emphasis on maximum-likelihood and Bayesian methods. As an example of a Bayesian application, I analyze an astronomical data set subject to large measurement errors and a non-linear dependence between the response and covariate. I conclude with some directions for future research.} ", "introduction": "\\label{intro} Measurement error is ubiquitous in astronomy. Astronomical data consists of passive observations of objects, whereby astronomers are able to directly measure the flux of an object as a function of wavelength, its location on the sky, and the time of the observation. Because the number of photons detected from an astronomical object follows a Poisson process, this makes the measurement of a source's intensity intrinsically subject to measurement error, even if none is introduced from the detector. Therefore, the very nature of astronomical data makes measurement error unavoidable. Moreoever, quantities that are derived from an object's observed emission, either by fitting a model to the spectral energy distribution (SED) or by employing scaling relationships, are also `measured' (derived) with error. Examples include mass, metallicity, and distance. Often the measurement error on the derived quantities is significant. This is unfortunate as inference on the derived quantities is often the goal of astronomical data analysis. Therefore, there has been considerable interest in how to perform statistical inference in the presence of measurement error. Measurement error is a problem that affects, at various levels, all scientific research. Because of this, numerous methods for handling measurement errors have been developed (Fuller 1987, Cheng \\& Van Ness 1999, and Carroll et al. 2006 are good references). In this contribution, I will present a survey of methods for handling measurement error that have been developed and used in astronomical data analysis. Because astronomical measurement errors are, in general, heteroskedastic (having different variances), I will limit my discussion to methods developed for heteroskedasticity. I will focus on situations where a deterministic relationship is not assumed between the variables, but where all variables of interest are random and are measured with error. Because of this, I will ignore situations where the measurement error is the only source of randomness in one's data. An example of this type of situation is fitting a model to an observed spectrum, where the measurement error is the only source of randomness; i.e., in the absence of measurement error a deterministic relationship is assumed between, say, flux density and wavelength. Methods for handling measurement error in this case are relatively well-established, and typically one simply minimizes the usual $\\chi^2$ statistic (e.g., Bevington 2003). However, it is worth pointing out that many complications may still exist, and more sophisticated methods may be needed, especially when dealing with low-count X-ray and $\\gamma$-ray data (e.g., van Dyk et al. 2001) or when incorporating calibration unceratinties (Lee et al. 2011). Instead, I will focus on methods for analyzing data from astronomical samples, where the variables are a random sample from an underlying distribution. Within the context of regression, this implies that intrinsic scatter (referred to as equation error in the statistics literature) exists in the relationship among the variables, and thus a deterministic relationship is not assumed between the variables even without the presence of measurement error. Most of the techniques I will discuss focus on accounting for measurement error in regression. The goal of regression is often to understand how one variable changes with another. For example, how does the mass of a black hole change as a function of the stellar velocity dispersion of the host galaxy's bulge? Typically one simply estimates how the average value and dispersion of one variable depends on another. Measurement error statistical models are typically divided into two types: `functional' and `structural' models. In functional modeling, one assumes that the unknown true values of the variables are fixed, whereas in structural modeling the unknown true values of the variables have their own intrinsic distribution. As a result, in structural modeling one must parameterically model the distribution of the true values of the variables, whereas in functional modeling one does not. Density estimation is another important technique in astronomical data analysis, being the foundation for luminosity and mass function estimation. The methods I will discuss for handling measurement error in structural models are also applicable to density estimation, as in this case regression and density estimation are based on the same formalism. When discussing regression methods, I will refer to the `dependent' variable as the response, and the `independent variables' as the covariates. ", "conclusions": "" }, "1112/1112.2948_arXiv.txt": { "abstract": "M87 is the first detected non-blazar extragalactic Tera-Electron-Volt (TeV) source with rapid variation and very flat spectrum in the TeV band. To explain the two-peaks in the spectral energy distribution (SED) of the nucleus of M87 which is similar to those of blazars, the most commonly adopted models are the synchrotron self-Compton scattering (SSC) models and the external inverse Compton scattering (EIC) models. Considering that there is no correlated variation in the soft band (from radio to X-ray) matching the TeV variation, and the TeV sources should not suffer from the $\\g\\g$ absorption due to the flat TeV spectrum, the EIC models are advantageous in modeling the TeV emission from M87. In this paper, we propose a self-consistent EIC model to explain the flat TeV spectrum of M87 within the framework of fully general relativity, where the background soft photons are from the advection-dominated accretion flow (ADAF) around the central black hole, and the high energy electrons are from the mini-jets which are powered by the magnetic reconnection in the main jet \\citep{2010MNRAS.402.1649G}. In our model, both the TeV flares observed in the years of 2005 and 2008 could be well explained: the $\\g\\g$ absorption for TeV photons is very low, even inside the region very close to the black hole $20R_g\\sim50R_g$; at the same region, the average EIC cooling time ($\\sim 10^2\\sim10^3s$) is short, which is consistent with the observed time scale of TeV variation. Furthermore, we also discuss the possibility that the accompanying X-ray flare in 2008 is due to the direct synchrotron radiation of the mini-jets. ", "introduction": "Unlike blazars and the Galactic sources (e.g. SNRs), which composite the most population of TeV sources, M87 is the first discovered radio galaxy with TeV radiation. Recently three other radio galaxies Centaurus A, 3C66B and IC310 were also identified as TeV sources \\citep{ 2009ApJ...695L..40A,2009ApJ...692L..29A, 2010ApJ...723L.207A}. Comparing to blazars, M87 has much milder variations in both optical and X-ray band; most importantly, the prominent kpc-scale jet of M87 has a very large viewing angle $\\sim30^\\circ$ with respect to the line of sight \\citep{bick96,ahar06}. In 2005, rapid TeV variation ($1-2$days) was discovered with flat spectrum for the first time, but without correlated X-ray variation from the nuclear core \\citep{ahar06}. While in 2008, the radio, X-ray, TeV joint observation discovered that TeV flares lasts 2 weeks accompanied with an X-ray flare and a radio flare inside the unresolved nuclear core ($30\\times60R_s$) as well as with a radio blob moving out of the unresolved core \\citep{2009Sci...325..444A}. Here $R_s$ used in the VLBA observations are based on the black hole (BH) mass of M87 $M_{BH}=6\\times10^9M_{\\sun}$ \\citep{2009ApJ...700.1690G}. Recently, a TeV flare in 2010 which is very similar to the previous TeV flares in 2005 and 2008 was reported. Further observations at X-rays and radio find the correlated X-ray flare, but no enhanced radio flux from the core \\citep{Abramowski2011,Harris2011}. In this paper, we adopt $M_{BH}=3\\times10^9M_{\\sun}$ \\citep{1997ApJ...489..579M} to be consistent with the advection-dominated accretion flow (ADAF) models and the mini-jet models in the previous studies \\citep{2009ApJ...699..513L,2010MNRAS.402.1649G}. Although M87 has a large viewing angle, it is very near to the Earth with a distance of $\\sim16.7$Mpc. Owing to its proximity, it is proposed that M87 might be a misaligned blazar \\citep{1998ApJ...493L..83T}. Whereas blazars are believed to have jets beaming towards us, TeV sources in M87 could be shocks with relativistic bulk velocity inside the jets. Recent observed minute-scale variation from galaxies like Mrk 421 (Fossati et al. 2008) and PKS 2155-304 \\citep{2007ApJ...664L..71A} indicates that TeV sources should be compact and very close to the BH. Comparing to the TeV flares of blazars, the slower variation ($1-2$ day) detected in M87 could be the results of too few observed TeV photons (each data point requires integration of photons for the whole night) or the much lower Doppler factor of the bulk velocity of the TeV sources due to the large inclination. Furthermore, unlike the steeper TeV spectrum of blazars, the much flatter TeV spectrum of M87 could be mainly due to the lack of $\\g\\g$ absorption. The large viewing angle may play an important role in avoiding $\\g\\g$ absorption within the jet; thanks to its proximity and the dimness of its host galaxy, the absorption to TeV photon from M87 caused by the galactic and the intergalactic background soft photons, as well as the cosmic background photons is rather weak \\citep{nero07}. To explain the spectral energy distribution (SED) of the nucleus of M87, the one zone synchrotron self-Compton(SSC) models, which have been applied to the TeV flares in blazars \\citep[e.g.][]{1998ApJ...509..608T, 2001A&A...367..809K}, face difficulties in fitting the two peaks in the SED, however, in the multi-zone SSC models, the TeV source is separated from the soft ones, therefore the whole SED could be easily fitted as long as there are enough SSC blobs with specific locations, electron distributions and bulk velocity \\citep[e.g.][]{2008MNRAS.385L..98T,2007sf2a.conf..196L,2005ApJ...634L..33G}. The multi-zone SSC models could also provide better explanation for the rapid variation, and the orphan TeV flares (the variations in the soft band can not connect with the TeV variation). Furthermore, considering the very flat power-law SED of M87 in the TeV band with index $p\\simeq-2.2\\sim-2.6$, the $\\g\\g$ absorption provides crucial constrains on the SSC models. So far, most of the SSC models still have problems in explaining the very flat TeV spectrum due to the certain $\\g\\g$ absorption, therefore the external inverse Compton (EIC) process could be more likely responsible for the TeV flare \\citep{2008MNRAS.384L..19B,2009MNRAS.395L..29G,2010MNRAS.402.1649G}. In the EIC models, the TeV source is far away from the soft one, or has the relativistic bulk velocity with respect to the soft one, thus the $\\g\\g$ absorption can be reduced. In those models mentioned above, how the very high energy (VHE) particles are produced is still an open question. It is generally believed that VHE particles are accelerated by the shocks in jets. There are some other possibilities include the mini-jets powered via the magnetic reconnection in the main jet \\citep[][and references therein]{2010MNRAS.402.1649G}; the magnetic centrifugal acceleration in the vicinity of BH \\citep{nero07,rieg08}, and so on. In the mini-jets model \\citep{2010MNRAS.402.1649G}, magnetic reconnection within the main jet produces two oppositely directed (in the rest frame of the main jet) mini-jets, in the laboratory frame, one of them always points within the angle of the main jet and is observable in blazars because their jets point at us. The other mini-jet (its counterpart) points outside the opening angle of the main jet and is potentially observable to off-axis observers in case of the misaligned jets, such as those of M87 and Centaurus A. In a word, the great advantage of the mini-jets model is that even at large inclination, mini-jets with a high bulk speed can still be detected, which is helpful for explanation of the fast TeV variation from M87. After all, in order to avoid the $\\g\\g$ absorption, the energy density of the soft synchrotron photons in the TeV source must be limited, therefore, the minimal Lorentz factor of the VHE particles is generally assumed to be $10^{3-4}$ and the strength of the magnetic field in the TeV source below several Gauss \\citep{2010MNRAS.402.1649G}. Beside the direct inverse Compton process of the VHE electrons, there are also some alternative models to produce the SED of the TeV flares, such as the hadronic models including, among which, the interaction between the VHE protons and soft photons \\citep{2004A&A...419...89R} and the proton-proton collision process when a red giant was passing the base of the jet \\citep{2010ApJ...724.1517B}. In this paper, we develop a fully general relativity EIC model for explaining the TeV emission from M87, in which the soft photons emanate from the ADAF around the BH. In \\S2, we investigate the safe zone of TeV photons (the $\\g\\g$ optical depth is below unity). The technical details of our fully general relativity EIC model can be found in \\S3. The numerical results and the discussions are given in \\S4 and \\S5, respectively. \\begin{figure}[t!] \\centering \\includegraphics[width=0.6\\textwidth, trim=0pt 80pt 0pt 0pt]{fig1.eps} \\caption{ The contour of $\\tau_{\\g\\g}=0.1$ or $1$ in the $\\alpha,\\beta$ plane. Where $r$ represents the distance to the BH in the Boyer-Lindquist frame. The viewing angle is taken to be $30^\\circ$. The blank region in the lower panels is the region with $\\tau_{\\g\\g} \\leq 1$. } \\label{fig1} \\end{figure} ", "conclusions": "In this paper, by taking into account the fully general relativity effects, we propose a disk-dominating external Compton-scattering model for explaining the flat TeV radiation from M87. The external Compton-scattering model suffers much less self-$\\g\\g$ absorption, comparing to the one-zone self-synchrotron Compton model, in which the soft photons and the VHE electrons are from the same blobs. The advantage of the EIC model in which the soft radiation is from the accretion disk is that the soft radiation source could be more compact and further away from the TeV source than ones in the SSC models, in which the $\\g\\g$ absorption can be reduced significantly. Besides, our EIC model is also supported by the observed mild IR-UV-X flux variations of the nucleus and the large viewing angle of the jet of M87 (the nuclei flux with viewing angle $30^\\circ$ is unlikely to be contaminated by the jet, neither to be blocked by the disk itself). It should be noticed that the VHE electrons and TeV photons may be influenced by some outflows with the anisotropic radiation at the soft band which are not beaming towards us, such as the jet itself. Unfortunately, the base of the jet of M87 appears to be quite chaotic, thus it is hard to draw any conclusions about the detailed structure of jet/outflows \\citep{2011arXiv1109.6252P}. In our model, the TeV sources are the mini-jets which are easily able to beam towards us with high Doppler factor and give rise to the observed VHE radiation. How to distinguish between SSC and EIC models by the future observations? The main difference between these two models is that the IR-UV variation and TeV flares are not necessarily correlated in the EIC models, but they are in the SSC models, especially the multi-zone SSC model. Therefore, we can distinguish between the SSC and EIC models by the simultaneous observations of the IR-UV emission during the TeV flare. Unfortunately, during the 2005 TeV flares, there is no accompanied HST observation of M87. With the better sensitivity of CTA (it is about $>$10 times better than those of the present Cerenkov telescopes), we will be able to detect the fainter flares and obtain the minimal variation time scale of the TeV flares, which increases the opportunity of finding the correlated variation between IR-UV and TeV emission. In our models, with little $\\g\\g$ absorption, all the EIC spectra at the observational band have inherited the original power-law index of the VHE electrons. After fitting those spectra, we calculate the cooling time of the VHE electrons in the mini-jets (see section 4) to make sure that they would not travel longer than $ct_{\\rm var}\\delta_D$. It turns out when TeV source is located at $H_{\\rm TeV}<50R_{\\rm g}$, the IC cooling time is very much shorter than the upper-limits of the observed TeV variation of 1-2 days. Therefore, if possibly, the TeV source is located further away from the disk ($H_{\\rm TeV}>50R_{\\rm g}$), the cooling efficiency of the VHE electrons in the mini-jets is much lower, the mini-jets could move out of the main jet. As a result, the variation of the soft radiation background at the different locations above the disk should be considered. We also discuss the probability that the direct synchrotron radiation from the mini-jets may cause the correlated X-ray flare in 2008. As expected, according to our model, if the magnetic field is about several Gauss, the direct synchrotron flux from these mini-jets which lies at the X-ray band can explain the X-ray flare very well but has no influence on the observed radiation at the GeV-TeV band. However, the magnetic field $B$ can not be constrained by any direct observation, as discussed above, and we only have a rough estimate of the strength of the magnetic field supposed to be consistent with the mini-jets model. Considering that the synchrotron flux is roughly about $F_{\\rm syn}\\sim N_0B^2$, it is obvious that the $N_0$ and $B$ are degenerated. If the TeV source is confined inside the region $20R_g1$; $F(\\nu) \\propto \\nu^{-\\alpha}$] after the peak. This leaves the accretion disk contribution unhidden. By modelling the disk component we can derive the black hole mass and accretion rate. } \\label{0227} \\end{figure} ", "conclusions": "" }, "1112/1112.1666_arXiv.txt": { "abstract": "I briefly review the flatness problem within the context of classical cosmology and examine some of the debate in the literature with regard to its definition and even the question whether it exists. I then present some new calculations for cosmological models which will collapse in the future; together with previous work by others for models which will expand forever, this allows one to examine the flatness problem quantitatively for all cosmological models. This leads to the conclusion that the flatness problem does not exist, not only for the cosmological models corresponding to the currently popular values of \\lnull\\ and \\onull\\ but indeed for all Friedmann\\ndash Lema\\^{\\i}tre models. ", "introduction": "The flatness problem has been called one of the outstanding puzzles in cosmology \\citep[\\eg][]{RDickePPeebles79a}. This in itself is rather puzzling in view of the fact that the arguments in favour of it being a problem are rather vague and heuristic, while quantitative arguments have been presented against the claim that it is a problem, at least for some classes of cosmological models \\citep[\\eg][]{PColesGEllis97a,KLake05a}. The flatness problem is one of the main motivations for the inflationary scenario \\citep{AGuth81a}. Of course, if there is no flatness problem (or, indeed, even if there were no motivation at all for inflation), this does not mean that inflation could not have occurred. However, it does mean that inflation should not be taken as given based on the belief that it explains away the flatness problem and thus without it classical cosmology leads to absurd conclusions. The plan of this paper is as follows. In \\Sect~\\ref{cosmo} I present the basic equations needed in the rest of the paper, mainly to define my notation (unfortunately, there is not a uniform notation in the literature) and give an overview of the entire cosmological parameter space relevant to the discussion. Section~\\ref{history} gives a brief historical overview of the flatness problem and some qualitative arguments against it. In \\Sect~\\ref{collapse} I discuss a new quantitative argument regarding cosmological models which will collapse in the future. Sections~\\ref{Lake} and~\\ref{anthro} discuss previous quantitative results by others for other classes of cosmological models. Section~\\ref{summary} summarizes the results for all cosmological models. ", "conclusions": "\\label{summary} The qualitative flatness problem, \\ie~the puzzle why the universe was arbitrarily close to the Einstein\\ndash de~Sitter model\\footnote{Or, for an empty universe, the Milne or de~Sitter model.} at early times, does not exist. It is merely a consequence of the way $\\lambda$ and $\\Omega$ are defined. Neither does the quantitative flatness problem exist: although the cosmological parameters in general evolve with time, it is not puzzling that we don't observe extreme values for them today. In the case of models which will collapse in the future this is because large (absolute) values of $\\lambda$ and $\\Omega$ occur only during a relatively short time in the lifetime of such a universe, namely near the time of maximum expansion. $\\lambda$ and $\\Omega$ can become large only when $H$ becomes small, and this happens only during the time when the universe is at or near its maximum size. [Arbitrarily small (absolute) values, if they occur at all, also occur for only a relatively short time]. For models which will expand forever, large values are possible only for $k=+1$. However, this occurs only for $\\alpha \\approx 1$. In this case, the fine-tuning argument is reversed; only in the case of fine-tuning do $\\lambda$ and $\\Omega$ become arbitrarily large. Since all models which will expand forever asymptotically approach $\\Omega=0$, arbitrarily small values of $\\Omega$ can occur. Those with $\\lambda=0$ (and hence $k=-1$) approach the Milne model with $\\Omega=0$; models with $\\lambda>0$, whatever the value of $k$, approach the de~Sitter model with $\\lambda=1$ (the Milne and de~Sitter models themselves are of course stationary points). (If $\\lambda=0$ at any time then $\\lambda=0$ at all times. Otherwise, arbitrarily small values of $\\lambda$, if they occur at all, occur only for a relatively short time.) However, if \\hnull\\ has a value similar to or smaller than the observed value, small values of $\\Omega$ will occur only in the far future when anthropic arguments probably make the observation of such a low value of $\\Omega$ unlikely. While (for $\\lambda>0$) a higher value of \\hnull\\ would allow a low value of $\\Omega$ even for an age near the observed age, such a universe would have spent only a very short time during which $\\Omega$ was not very small, so structure formation would have been strongly suppressed. It is interesting to note that the three arguments presented here make it unlikely that we would observe extreme values of \\lnull\\ or \\onull. This automatically solves the so-called coincidence problem, which has been called deeply puzzling \\citep*[\\eg][]{MTegmarkAVilenkinLPogosian05a}. Also interesting is that in the appendix to his seminal paper, \\citet{AGuth81a} anticipates much of the subsequent discussion. He points out that even though essentially all cosmological models begin arbitrarily close to the Einstein\\ndash de~Sitter universe (the qualitative flatness problem), this still leaves the question as to why the universe is so close to the Einstein\\ndash de~Sitter universe today (the quantitative flatness problem). He also rejects the idea that some basic principle must force the universe to conform exactly to the Einstein\\ndash de~Sitter model on the grounds that this is obviously only an approximation in the case of the real universe (see the discussion of this above). He basically recasts the flatness problem as the longevity problem: fine tuning is required in order that the universe does not recollapse or thin out to extremely low density within a very short time. However, this argument relies on an assumption for the value of \\hnull, while our arguments do not need this assumption except in the third category. Also, as I have shown here, even in an extremely short-lived universe (which of course recollapses), extreme values of $\\lambda$ or $\\Omega$ are observed only during a relatively small fraction of the lifetime of the universe. If there is no flatness problem, what does this mean for inflation? If there is no flatness problem (or, indeed, even if there were no motivation at all for inflation), this does not mean that inflation could not have occurred. However, it does mean that inflation should not be taken as a given based on the belief that it explains away the flatness problem and thus without it classical cosmology leads to absurd conclusions. Inflation also solves the monopole and isotropy problems. However, the monopole problem seems to be more a problem with theories of particle physics than with cosmology \\citep{JNarlikarTPadmanabhan91a} while \\citet{JBarrow95a} claims that there is no isotropy problem.\\footnote{It is, however, debatable whether Barrow's conclusion is as general as he claims or depends too strongly on his assumptions.} If all of these claims are true, then this still does not prove that inflation didn't happen, but the necessity for inflation or something like it is weakened if not destroyed altogether." }, "1112/1112.6270_arXiv.txt": { "abstract": "{} {We construct a theoretical model to predict the number of orphan afterglows (OA) from gamma-ray bursts (GRBs) triggered by primordial metal-free (Pop III) stars expected to be observed by the Gaia mission. In particular, we consider primordial metal-free stars that were affected by radiation from other stars (Pop III.2) as a possible target. } {We use a semi-analytical approach that includes all relevant feedback effects to construct cosmic star formation history and its connection with the cumulative number of GRBs. The OA events are generated using the Monte Carlo method, and realistic simulations of Gaia's scanning law are performed to derive the observation probability expectation.} {We show that Gaia can observe up to 2.28 $\\pm$ 0.88 off-axis afterglows and 2.78 $\\pm$ 1.41 on-axis during the five-year nominal mission. This implies that a nonnegligible percentage of afterglows that may be observed by Gaia ($\\sim 10\\%$) could have Pop III stars as progenitors. } {} ", "introduction": "The first stars (hereafter, Pop~III-primordial metal-free) in the Universe are thought to have played a crucial role in the early cosmic evolution by emitting the first light and producing the first heavy elements \\citep{bromm09}. Understanding such objects is very important since their detection would permit the pristine regions of the Universe to be probed. However, there has been no direct observation of the so-called Pop~III stars up to now. Pop III stars may produce collapsar gamma-ray bursts (GRBs) whose total isotropic energy could be $\\approx 2$ orders of magnitude larger than average \\citep{barkov2010,komissarov2010,meszaros2010,suwa2011,toma2011}. Even if the Pop III star has a supergiant hydrogen envelope, the GRB jet can break out of it because of the long-lasting accretion of the envelope itself \\citep{nagakura2011,suwa2011}. It is of great importance to study the rate and detectability of Pop~III GRB prompt emissions and afterglows in current and future surveys. We explore here the possibility to observe these objects through their afterglows \\citep{toma2011}. Observations of GRB afterglows make it possible to derive physical properties of the explosion mechanism and the circumburst medium. It is intriguing to search for signatures of metal-poor stars in the GRB afterglows at low and high redshifts. GRB optical afterglows are one of the possible transients to be detected by the Gaia\\footnote{http://www.rssd.esa.int/GAIA/} mission. Recently \\citet{Japelj2011} explored the detectability of such afterglows with Gaia using a Monte Carlo approach that inspired us. As the GRB jet sweeps the interstellar medium, the Lorentz factor of the jet is decelerated and the jet starts to expand sideways, eventually becoming detectable by off-axis observers. These afterglows are not associated with the prompt GRB emission and are called orphan afterglows (OA) \\citep{nakar2002, rossi2008}. De Souza et al. (2011) showed that, considering EXIST\\footnote{http://exist.gsfc.nasa.gov/design/} specifications, we can expect to observe a maximum of $\\approx 0.08$ GRBs with $z>10$ per year originating from primordial metal-free stars (Pop III.1) and $\\approx 20$ GRBs with $z>6$ per year coming from primordial metal-free stars that were affected by the radiation from other stars (Pop III.2). In the context of the current \\textit{Swift}\\footnote{http://swift.gsfc.nasa.gov/docs/swift/swiftsc.html} satellite, $\\approx 0.2$ GRBs with $z>6$ per year from Pop III.2 stars are expected. These numbers reflect the fact that, compared to Pop~III.1 stars, Pop III.2 stars are more abundant and can be observed in a lower redshift range, which makes them more suitable targets. In the light of such results, the calculations presented here will focus on Pop III.2 stars alone. Searches have been made of OAs by both X-ray surveys \\citep{grindlay1999,greiner2000} and optical searches \\citep{becker2004,rykoff2005,rau2006,malacrino2007}. The purpose of the present paper is to calculate the Pop~III.2 GRB OA rate that might be detected by the Gaia mission \\citep[for more details about Gaia, see, e.g.,][]{Perryman:2001p3838, Lindegren:2009p8828}. The Gaia mission is one of the most ambitious projects of modern astronomy. It aims to create a very precise tridimensional, dynamical, and chemical census of our Galaxy from astrometric, spectrophotometric, and spectroscopic data. In order to do this, the Gaia satellite will perform observations of the entire sky in a continuous scanning created from the coupling of rotations and precession movements called the scanning law. For point sources, these observations will be unbiased and the data of all the objects bellow a certain limiting magnitude (G=20) will be transferred to the ground. Certainly, galactic and extragalactic sources will be among those objects. Typically, Pop~III.2 stars are formed in an initially ionized gas \\citep{Johnson06,Yoshida07}. They are thought to be less massive than Pop~III.1 stars but still massive enough to produce GRBs. Recent results from \\citet{greif2011} show that, instead of forming a single object, the gas in mini-halos fragments vigorously into a number of protostars with a range of different masses. It is not clear up to now how this initial range of mass will be mapped into the final mass function of Pop III stars. The most likely conclusion is that Pop III stars are less likely to reach masses in excess of $\\sim 140 M_{\\odot}$, which consequently affect the estimated number of GRBs from Pop III.1. \\citet{hosokawa2011}, performing state-of-the-art radiation-hydrodynamics simulations, showed that the typical mass of Pop III stars could be $\\sim 43 M_{\\odot}$. Here we assume that this will not affect significantly the mass range assumed for Pop III.2 ($\\sim 10-100 M_{\\odot}$). The paper is organized as follows. In Sect. 2, we calculate the formation rate of primordial GRBs. In Sect. 3, we calculate the OA light curves and their redshift distribution. In Sect. 4, we discuss the details of the Gaia mission and derive the probability of a given event to be observed by Gaia. In Sect. 5, we discuss the results, and finally, in Sect. 6, we give our concluding remarks. Throughout the paper, we adopt the standard $\\Lambda$ cold dark matter model with the best-fit cosmological parameters from \\citet{jarosik2010} (WMAP-Yr7\\footnote{http://lambda.gsfc.nasa.gov/product/map/current/}), $\\Omega_{\\rm m} = 0.267, \\Omega_{\\Lambda} = 0.734$, and $H_0 = 71 {\\rm km}~{\\rm s}^{-1}{\\rm Mpc}^{-1}$. ", "conclusions": "It is important to emphasize that our knowledge concerning first stars and their GRBs is still quite incomplete. Many of their properties (e.g., characteristic mass, SFR and efficiency to trigger GRBs) are still very uncertain, and more reliable information can only come once a detection is confirmed. Recently, \\citet{hosokawa2011}, performing state-of-the-art radiation-hydrodynamics simulations, showed that the typical mass of primordial stars could be $\\sim 43 M_{\\odot}$, i.e., less massive than originally expected by theoretical models. Their results, though, are affected by assumptions on the initial conditions. This confirms that we are far away from understanding all characteristics of these objects and any observation would be of paramount importance to improve theoretical models. In this work, we estimated the average number of OAs events originating from Pop III stars that the Gaia mission may observe to be up to 2.28 $\\pm$ 0.88 off-axis afterglows and 2.78 $\\pm$ 1.41 on-axis ones. In case such events are found among Gaia data, valuable physical properties associated with the primordial stars of our Universe and their environment could be constrained." }, "1112/1112.5981_arXiv.txt": { "abstract": "This article reviews the present knowledge on oscillating main sequence A/F stars that belong to more than one class of pulsator. Due to recent results of asteroseismic space missions, we now know of many $\\delta$~Scuti/$\\gamma$~Doradus stars. However, $\\gamma$~Doradus variability was also detected in a rapidly oscillating Ap star, and solar-like oscillations were discovered in a $\\delta$~Scuti star. The astrophysical information that can be gained from these pulsators is discussed, and confronted with what is believed to be known about pulsational driving. ", "introduction": "Over the last 40 years, our knowledge about pulsating stars has greatly expanded. The ever increasing accuracy in stellar observations and advances in theory have revealed a large number of previously unknown classes of variable star. Figure 1 compares the knowledge on the positions of pulsating stars in the HR diagram over this time span. \\begin{figure} \\plotfiddle{hakone1.ps}{8.3cm}{00}{72}{72}{-203}{-40} \\caption{Left: classes of pulsating star known in the late 1960's. Right: a selection of classes of pulsating star known to date. Areas hatched from lower right to upper left depict domains of g-mode pulsators, areas hatched from lower left to upper right delineate domains of p-mode pulsators; overlapping areas may contain \"hybrid\" pulsators. Parts of model evolutionary tracks for main sequence, horizontal branch and post-AGB stars are shown as dashed-dotted lines for orientation.} \\end{figure} The most striking difference between the two panels in this figure is that the HR Diagram nowadays is well filled with pulsating variables. This has the important consequence that stellar oscillations can be utilized to sound the interiors of stars in many different evolutionary stages, and over wide ranges of mass and chemical composition. Besides better defined locations of the pulsating stars over the last 40 years, another new feature is obvious: several of the different instability strips overlap. These instability strips contain stars with different pulsational behaviour, or in other words, with different types of mode excited. Consequently, it can be suspected that stars possessing more than a single set of mode spectra exist. Ground based observations have shown that they do, and the history of their study has been reviewed elsewhere (e.g., Handler 2009). The most complicated region in the HR Diagram in terms of overlap of different classes of pulsating star is where the lower classical instability strip intersects the main sequence, at spectral types of A and F. There one finds: \\begin{itemize} \\item $\\delta$ Scuti stars, pulsating in radial and nonradial pressure and mixed modes of low radial order with periods between 20 minutes and 6 hours, driven by the classical $\\kappa$ mechanism in the He II ionization zone (Baker \\& Kippenhahn 1962) \\item rapidly oscillating Ap (roAp) stars, high-order pressure modes pulsators governed by a global magnetic field, with periods of 5 - 20 minutes, believed to be driven by the $\\kappa$ mechanism in the He I/H ionization zone (Balmforth et al.\\ 2001) \\item $\\gamma$ Doradus stars that oscillate in high-order g modes with periods between 0.3 - 3 days, and likely driven by blocking of flux at the base of the envelope convection zone (Guzik et al.\\ 2000) \\item solar-like oscillations with periods around 10 - 40 minutes are theoretically predicted to be present and driven by surface convection (Houdek et al.\\ 1999) \\end{itemize} The present article reviews the present knowledge on hybrid pulsators in this domain in the light of recent results of asteroseismic space missions. There will be a strong focus on Kepler data, not only because they have the best quality, but for the practical reason that the author is involved in their analysis. ", "conclusions": "The $\\delta$ Scuti stars, the roAp stars, the $\\gamma$ Dor stars and solar like oscillators all overlap at the intersection of the classical instability strip and the main sequence. Each of these classes of pulsating star are driven in different regions of the star. To date, we know hybrid pulsators between: $\\delta$ Scuti and $\\gamma$ Dor stars, roAp stars and $\\gamma$ Dor stars as well as $\\delta$ Scuti stars and solar-like oscillators. The other three possible combinations of two classes of pulsator have not yet been observed. This status makes sense from the point of view of pulsational driving (as we believe to understand it to date), because the mechanisms operating in the known hybrids, or their basic physical characteristics, do not affect each other negatively. On the other hand, the absence of roAp/solar-like hybrids is also not a surprise, as the strong magnetic fields of Ap stars are believed to suppress surface convection. Similar arguments may be invoked to explain the non-detection of roAp/$\\delta$ Scuti hybrids: the diffusion and settling of chemical elements is supposed to deplete the driving zone for $\\delta$ Scuti oscillations of the required Helium, and slow wave leakage due to strong magnetic fields is believed to provide additional damping. What about $\\gamma$ Doradus/solar-like hybrids? Such stars have not been reported to date, but the author believes this is only a question of time. Along those lines, given the large fraction of Delgam Scudor stars, one may also expect the existence of $\\delta$ Scuti/$\\gamma$ Doradus/solar-like hybrids, which would be another goldmine for asteroseismology of main sequence stars." }, "1112/1112.5723_arXiv.txt": { "abstract": "We calculate the angular power spectrum of galaxies selected from the Sloan Digital Sky Survey (SDSS) Data Release 7 (DR7) by using a quadratic estimation method with KL-compression. The primary data sample includes over 18 million galaxies covering more than 5,700 square degrees after masking areas with bright objects, reddening greater than 0.2 magnitudes, and seeing of more than 1.5 arcseconds. We test for systematic effects by calculating the angular power spectrum by SDSS stripe and find that these measurements are minimally affected by seeing and reddening. We calculate the angular power spectrum for $\\ell \\le 200$ multipoles by using 40 bandpowers for the full sample, and $\\ell \\le 1000$ multipoles using 50 bandpowers for individual stripes. We also calculate the angular power spectrum for this sample separated into 3 magnitude bins with mean redshifts of $z = 0.171$, $z = 0.217$, and $z = 0.261$ to examine the evolution of the angular power spectrum. We determine the theoretical linear angular power spectrum by projecting the 3D power spectrum to two dimensions for a basic comparison to our observational results. By minimizing the $\\chi^2$ fit between these data and the theoretical linear angular power spectrum we measure a loosely-constrained fit of $\\Omega_m = 0.31^{+0.18}_{-0.11}$ with a linear bias of $b = 0.94 \\pm 0.04$. ", "introduction": "\\label{Introduction} The angular power spectrum, $C_{\\ell}$, is a statistical measure that quantitatively characterizes the large scale angular distribution of matter \\citep{peebles73}. Therefore, calculating the angular power spectrum of galaxies is useful as both a method of data compression, reducing clustering information of an arbitrary number of galaxy positions down to a set of $C_{\\ell}$ and their corresponding window functions, and also since the $C_{\\ell}$ values derived from the observations can be easily compared to theoretical predictions. Calculations of angular power spectra are well known to cosmologists for their usefulness in studying the Cosmic Microwave Background (CMB), as the CMB provides a detailed and precise measurement of the density variations in the early universe (e.g., \\citealt{smoot92,netterfield02,spergel07}). However, to study large scale structure in other eras, it is necessary to analyze how mass clusters by using galaxies as a tracer of the underlying dark matter distribution. Angular power spectra of galaxies have been calculated for galaxy surveys with various depth and survey area (e.g., \\citealt{huterer01,blake04,frith05}) including the SDSS (\\citealt{tegmark02}, hereafter T02; \\citealt{blake07,thomas10}). By using angular power spectra to calculate galaxy clustering, we study the Fourier modes of the galaxy distribution; this method is most sensitive to large scale effects. Recent galaxy surveys such as the APM Galaxy Survey \\citep{maddox90}, the Two Micron All Sky Survey \\citep{skrutskie06}, and the SDSS \\citep{abazajian09} have cataloged large areas of the sky, thereby providing enormous numbers of galaxies for which we can measure angular clustering. However, to date the galaxy angular power spectrum has not been calculated for the full SDSS main galaxy sample. In this paper, we address this deficiency. The angular power spectrum is useful for large scale clustering, while it is complemented by the two-point angular correlation function on small scales. The two-point angular correlation function (e.g., \\citealt{brunner00,myers07,ross10}), which is related to the angular power spectrum by the Legendre transform (T02), is more sensitive to smaller scale clustering because the calculation is done in configuration space where the distances between nearby pairs of galaxies can be calculated faster. This makes the two-point angular correlation function advantageous to use on scales where non-linear evolution is important. This regime is also where the angular power spectrum at large $\\ell$ is more difficult to measure and model, partly due to correlations introduced between the $C_{\\ell}$. To calculate the angular power spectrum, we want to find the most probable parameters $C_{\\ell}$ that could produce the data we observe. To do this, we need the likelihood function of the angular power spectrum, which is proportional to the probability of the data given the $C_{\\ell}$. Though in theory we would like to know the entire likelihood function, calculating this $\\ell_{max}$-dimensional function is difficult \\citep{oh98}. Fortunately, since we are only interested in the most probable $C_{\\ell}$, we only really need to know the maximum of this function. To determine the $C_{\\ell}$ that maximize the likelihood function, we use the quadratic estimation method (\\citealt{tegmark97a,bond98}, hereafter BJK98). This technique fits a quadratic function to the shape of the likelihood function for some initial angular power spectrum, finds the $C_{\\ell}$ that maximize this quadratic, and uses these $C_{\\ell}$ for a new quadratic fit to iteratively converge to the true maximum of the likelihood function. Once we have found the angular power spectrum of galaxies, we can use the results to infer what cosmological parameters are consistent with the measurement (e.g., \\citealt{jaffe99}). In this paper, we discuss the SDSS DR7 data, our selected sample and subsamples, and our systematic tests and masks in Section \\ref{Data}. In Section \\ref{Method}, we discuss our pixelization scheme, KL-compression, and the quadratic angular power spectrum estimation method of BJK98 in detail. In Section \\ref{Results}, we apply this estimator to the complete SDSS DR7, selected subsamples, and individual SDSS stripes, and present the results. We construct a theoretical linear angular power spectrum to compare with the observational results, and we extract cosmological matter density and linear bias from this computation in Section \\ref{Theory}. Finally, we discuss our results in Section \\ref{Discussion}, and conclude the paper in Section \\ref{Conclusion}. ", "conclusions": "\\label{Conclusion} We have used the quadratic estimation method with KL-compression to determine the SDSS DR7 angular power spectrum, first as a means of radical compression of the angular clustering information, and second to match these observed angular power spectra with theoretical angular power spectra to extract the linear bias and cosmological matter density. We masked for observational effects and applied this method to over 18 million SDSS DR7 galaxies and three magnitude subsamples out to $\\ell \\le 200$. We also measured the angular power spectrum for each individual stripe out to $\\ell \\le 1000$ for stripes 9--37. We have used the photometric redshift distribution of these galaxies to project the 3D power spectrum to two dimensions to obtain theoretical linear angular power spectrum, and used $\\chi^2$ minimization to determine the best fit parameters given the observations. As the linear angular power spectrum approximation is not valid for the entire range of our estimated angular power spectrum, these parameter constraints have a large allowed range of values. We found that the linear bias of our samples was $b = 1.09 \\pm 0.05$ in the 18--19 magnitude range, $b = 1.03 \\pm 0.04$ for 19--20, and $b = 0.92 \\pm 0.04$ for 20--21, with an overall bias of $b = 0.94 \\pm 0.04$ for our combined 18--21 magnitude sample. We have also calculated the cosmological density of matter as $\\Omega_m = 0.31^{+0.18}_{-0.11}$ from our entire sample." }, "1112/1112.0330_arXiv.txt": { "abstract": "The dark energy dominated warm dark matter (WDM) model is a promising alternative cosmological scenario. We explore large-scale structure formation in this paradigm. We do this in two different ways: with the halo model approach and with the help of an ensemble of high resolution $N$-body simulations. Combining these quasi-independent approaches, leads to a physical understanding of the important processes which shape the formation of structures. We take a detailed look at the halo mass function, the concentrations and the linear halo bias of WDM. In all cases we find interesting deviations with respect to CDM. In particular, the concentration-mass relation displays a turnover for group scale dark matter haloes, for the case of WDM particles with masses of the order $m_{\\rm WDM}\\sim0.25 \\keV$. This may be interpreted as a hint for top-down structure formation on small scales. We implement our results into the halo model and find much better agreement with simulations. On small scales the WDM halo model now performs as well as its CDM counterpart. ", "introduction": "\\label{sec:Intro} Over the last decade the vacuum energy dominated cold dark matter (hereafter $\\Lambda$CDM) scenario, has emerged as a standard model for cosmology. This owes largely to the combination of information from galaxy clustering surveys such as the 2dFGRS and SDSS with WMAP measurements of the temperature anisotropies in the microwave background \\citep{Coleetal2005short,Tegmarketal2006short,Komatsuetal2011short}. However, the nature of the two dark components in the $\\Lambda$CDM model are still completely unknown and it is therefore important to keep exploring alternative models and test their compatibility with observations. In the $\\Lambda$CDM model the dark matter is assumed to be composed of heavy, cold thermal relic particles that decoupled from normal matter very early in the history of the Universe \\citep{Peebles1982,Blumenthaletal1984,KolbTurner1990,Jungmanetal1996}. Whilst there is a large body of indirect astrophysical evidence that strongly supports CDM, there are some hints that it has shortcomings. Firstly, CDM galaxy haloes contain a huge number of subhaloes \\citep{Mooreetal1999c,DiemandKuhlen2008,Springeletal2008,Stadeletal2009}, while observations indicate that only relatively few satellite galaxies exist around the Milky Way and M31 \\citep{Mooreetal1999c,Klypinetal1999}. Secondly, the highest resolution halo simulations show that the slope of the inner density profile decreases linearly at smaller radii \\citep{Navarroetal1997,Mooreetal1999c,Diemandetal2004,Springeletal2008,Stadeletal2009}, whereas the density profiles inferred from galaxy rotation curves are significantly shallower \\citep{Mooreetal1999c} \\citep[and for recent studies see][and references there in]{Swatersetal2003,Saluccietal2007,deBloketal2008,Gentileetal2009}. Thirdly, the observed number of dwarf galaxies in the voids appears to be far smaller than expected from CDM \\citep{Peebles2001,Tikhonovetal2009,PeeblesNusser2010}. Another example is the excess in the prediction of dwarf galaxy concentrations \\citep{Lovelletal2011}. Whilst, it has become clear that some of these discrepancies might be resolved through an improved understanding of galaxy formation, they have led some to consider changes to the $\\Lambda$CDM paradigm. One possible solution might be warm dark matter (WDM) \\citep{BondSzalay1983,Bardeenetal1986,Bodeetal2001}. In this scenario, the dark particle is considered to be lighter than its CDM counterpart, and so remains relativistic longer and also retains a thermal velocity. Since WDM particles are collisionless and decouple early, they may `free-stream' or diffuse out of perturbations whose size is smaller than the Jeans' length\\footnote{Although originally defined in the context of gas dynamics, the Jeans length can be generalized to collisionless systems by replacing the sound speed with the velocity dispersion. The reason for this tight analogy lies in the linearized equation of perturbations, which has the same structure for gas and collisionless fluids \\citep[see][for more details]{Peebles1982}.} in the radiation dominated Universe \\citep{KolbTurner1990}. This free-streaming of the WDM particles acts to damp structure formation on small scales. Two potential candidates are the sterile neutrino \\citep{DodelsonWidrow1994,ShaposhnikovTkachev2006}, and the gravitino \\citep{Ellisetal1984,Moroietal1993,Kawasakietal1997,Gorbunovetal2008}, both of which require extensions of the standard model of particle physics. Recent observational constraints have suggested that sterile neutrinos can not be the dark matter: the Lyman alpha forest \\citep{Seljaketal2006b,Boyarskyetal2009a} and QSO lensing \\citep{MirandaMaccio2007} bounds are $m_{\\nu_s}>8\\keV$, whilst those from the X-ray background are $m_{\\nu_s}<4\\keV$ \\citep{Boyarskyetal2008}\\footnote{Lower bounds on the mass of a fully thermalized WDM particle can be obtained using \\Eqn{eq:massrescale} \\citep[see][]{Vieletal2005}.}. However, a more recent assessment has suggested that a better motivated particle physics model based on resonant production of the sterile neutrino, may evade these constraints: the Lyman alpha forest bound is brought down to $m_{\\nu_s}\\gtrsim 2\\keV$ and the X-ray background is pushed to $m_{\\nu_s}<50\\keV$ (for very low mixing angles) \\citep{Boyarskyetal2009b}. It therefore seems that additional, independent methods for constraining the $\\Lambda$WDM scenario would be valuable. In \\citet{Markovicetal2010} and \\citet{SmithMarkovic2011}, it was proposed that the $\\Lambda$WDM scenario could be tested through weak lensing by large-scale structure. The advantage of such a probe is that it is only sensitive to the total mass distribution projected along the line of sight. However, to obtain constraints on the WDM particle mass, an accurate model for the nonlinear matter clustering is required. In these papers, an approach based on the halo model was developed. Accurate predictions from this model require: detailed knowledge of the abundance of dark matter haloes, their spatial large-scale bias, and their density profiles. In these studies, it was assumed that the semi-analytic methods, which were developed for CDM, would also apply to WDM. In this paper we perform a series of very high resolution CDM and WDM $N$-body simulations with the specific aim of exploring the halo model ingredients in the $\\Lambda$WDM scenario. Over the past decade, there have been a limited number of numerical simulation studies of nonlinear structure formation in the WDM model \\citep{Colombietal1996,Mooreetal1999c,Colinetal2000,WhiteCroft2000,Avila-Reeseetal2001, Bodeetal2001,Bullocketal2002,ZentnerBullock2003,Colinetal2008,Zavalaetal2009,MaccioFontanot2010,Lovelletal2011,Vieletal2011,Dunstanetal2011}. In most of these previous studies, conclusions have been drawn from object-by-object comparison of a relatively small number of haloes simulated in boxes of typical size $L=25\\Mpc$. In this work we are more interested in the overall impact that the WDM hypothesis has on the statistical properties of large-scale structures. We therefore simulate boxes that are 10 times larger than have been typically studied before, hence having roughly $\\sim1000$ times larger sampling volume. This means, that our conclusions will have greater statistical weight, than those from previous studies. Furthermore, our results should be less susceptible to finite volume effects, which can lead to underestimates of the nonlinear growth. The paper is structured as follows: In \\S\\ref{sec:WDMtheory} we provide a brief overview of the salient features of linear theory structure formation in the WDM model and we review the halo model approach. In \\S\\ref{sec:Simulations} we describe the $N$-body simulations. In \\S\\ref{sec:Ingredients} we explore the main ingredients of the halo model: the halo mass function, bias and density profiles. In \\S\\ref{sec:Comparison} we compare the halo model predictions for the matter power with our measurements from the simulations. Finally, in \\S\\ref{sec:Conclusion} we summarize our findings. ", "conclusions": "\\label{sec:Conclusion} In this paper we have explored nonlinear structure formation in the WDM cosmological model, through a large suite of cosmological $N$-body simulations and through the halo model. The study was done for a set of fully thermalized WDM models with particle masses in the set $m=\\{0.25,0.5,0.75,1.0,1.25\\} \\keV$. These masses range from purely pedagogical models, towards more realistic scenarios for the dark particle. For the simulations we chose a box size ${L=256\\Mpc}$, which was small enough to resolve both the small scales, where WDM effects play an important role, and the large scales, which are required for correct linear evolution of the box-modes. All models were simulated with $N=\\{256^3, 512^3, 1024^3\\}$ particles. This was done in order to disentangle physical effects from numerical ones. In the original halo model calculation for WDM by \\citet{SmithMarkovic2011}, it was shown that in order to make robust predictions, one requires good understanding of dark matter halo profiles, the mass function and halo bias. In this work we performed a detailed study of all of these ingredients. Our findings can be summarized as follows: \\begin{enumerate}{\\leftmargin = 1.0em} \\item Mass function: Below a certain scale, the WDM mass function is suppressed with respect to CDM. This suppression is considerably stronger than that obtained by simply applying the Sheth-Tormen approach together with the linear power spectrum of WDM. In agreement with \\citet{SmithMarkovic2011}, we found that the mass functions for the different WDM models could be transformed into a single locus of points. This was achieved by taking the ratio of the WDM mass function with that for CDM, and then rescaling the masses by $M_{\\rm hm}$ (or equivalently $M_{\\rm fs}$). We used a fitting function similar to that proposed in \\citet{Dunstanetal2011} to link the Sheth-Tormen mass function to the measured one. The fitting function, which has only one free parameter, was able to reproduce all of the data with an accuracy of a few percent. We also found a strong boost in the mass function at very small mass scales. We showed that this was consistent with artificial halo formation around the initial particle lattice \\citep[cf.][]{WangWhite2007}. \\item Halo bias: We measured the linear halo bias, using the four largest modes in our simulations. For smaller mass haloes, we found a small enhancement of the bias in WDM simulations, which was qualitatively consistent with the predictions of \\citet{SmithMarkovic2011}. However, owing to the simulation box being too small, we were unable to quantify this more robustly. At very small masses we found a prominent boost in the bias. We found that this was again a sign of artificial halo formation. \\item Density profiles: In the CDM model, the density profiles of dark matter haloes can be characterized by an NFW profile, with a monotonically decreasing concentration-mass relation. In the WDM scenario, we have shown that the NFW profile remains valid for the models and resolution limits of our simulations, and we saw no evidence for a central density core. A simple adaption of the CDM concentration-mass relation, would suggest a strong flattening towards small masses. Whilst, we found such a flattening, the measurements in fact revealed a turnover towards smaller masses. This somewhat surprising result may be interpreted as a sign of top-down structure formation. We modelled the mean relation by adapting a fitting formula similar to that for the mass function. Our fit to the $c(M)$ data was good to an accuracy of $\\sim10\\%$. Interestingly, we found that the deviations from CDM in the WDM model, appear in the $c(M)$ relation for halo masses one order of magnitude larger than for the mass function. \\end{enumerate} After analyzing these ingredients in detail and developing new fitting functions for them, we were able to improve the small-scale performance of the WDM halo model. We found that for $k\\gtrsim3\\kMpc$, we could predict the absolute amplitude of the power spectrum to better than $\\sim10\\%$. However, we were able to predict the ratio of the WDM to CDM spectra, at better than $\\lesssim2\\%$. This was competitive with the latest fitting formulae \\citep{Vieletal2011}. One of the many advantages of the halo model based approach, is that we may more confidently extrapolate our power spectra predictions to smaller scales than can be done from a fitting formula, since the model is built on physical quantities. Furthermore, we may also use the model to study the clustering of galaxies \\citep{Zehavietal2005}. It is hoped that this may lead to a method for constraining WDM models from galaxy clustering studies. Lastly, one further issue for future study, is to establish a better theoretical understanding of what shapes the mass function and halo concentrations in WDM. In particular, in finding the turnover in the concentration mass relation, have we really seen the reversal of bottom-up structure formation. This promises to be an interesting future challenge." }, "1112/1112.2103_arXiv.txt": { "abstract": "In the current-driven, kink-type Tayler instability (TI) a sufficiently strong azimuthal magnetic field becomes unstable against non-axisymmetric perturbations. The TI has been discussed as a possible ingredient of the solar dynamo mechanism and a source of the helical structures in cosmic jets. It is also considered as a size limiting factor for liquid metal batteries. We report on a liquid metal TI experiment using a cylindrical column of the eutectic alloy GaInSn to which electrical currents of up to 8 kA are applied. We present results of external magnetic field measurements that indicate the occurrence of the TI in good agreement with numerical predictions. The interference of TI with the competing large scale convection, resulting from Joule heating, is also discussed. ", "introduction": " ", "conclusions": "" }, "1112/1112.0106_arXiv.txt": { "abstract": "A molecular line survey has been carried out toward the carbon-rich asymptotic giant branch star CW~Leo employing the HIFI instrument on board of the {\\it Herschel} satellite. Numerous features from 480~GHz to beyond 1100~GHz could be assigned unambiguously to the fairly floppy SiC$_2$ molecule. However, predictions from laboratory data exhibited large deviations from the observed frequencies even after some lower frequency data from this survey were incorporated into a fit. Therefore, we present a combined fit of all available laboratory data together with data from radio-astronomical observations. ", "introduction": "\\label{introduction} Silacyclopropynylidene, SiC$_2$, somewhat better known as silicon dicarbide, is a fascinating molecule for spectroscopists, structural and quantum chemists as well as astronomers. In 1926, uncataloged bands near 500~nm were discovered in the spectra of several carbon-rich asymptotic giant branch (AGB) stars~\\cite{SiC2_A-X_astro_1926a,SiC2_A-X_astro_1926b}. These are late-type stars which produce elements heavier than helium and which eject large quantities of gaseous material as well as dust which form a circumstellar envelope (CSE). Thirty years later, laboratory spectroscopy established that the molecule SiC$_2$ is the carrier of these bands~\\cite{SiC2_A-X_lab-ident_1956}. It was assumed that the molecule has a linear SiCC structure (silapropadienediylidene) in analogy with the isoelectronic propadienediylidene, C$_3$. Although some later studies cast doubt on the linear structure of the molecule, it took almost another 30 years until the analysis of the rotational structure of the electronic origin band unmistakably determined the structure as silacyclopropynylidene~\\cite{SiC2_A-X_lab-non-lin_1984}. In the course of their analysis, the authors instigated quantum chemical calculations which provided evidence that the cyclic isomer of SiC$_2$ may be lower in energy than the linear form~\\cite{SiC2_ai_1984}. A plethora of quantum chemical calculations on various properties of SiC$_2$ have been published later, yielding energy differences between the linear and the cyclic form which depended strongly on the level of the calculation and the size of the basis set. A high level ab initio calculation concluded that the cyclic isomer of SiC$_2$ is the only minimum on the potential energy surface and that the linear transition state is 24.3~kJ/mol higher in energy~\\cite{SiC2_ai_1997}. However, the authors attached a caveat to this value as an anharmonic force field calculation provided a much too small value for the vibrational energy of $\\varv_3 = 1$ and much too large anharmonicity constants. Therefore, some of the authors revisited the problem of the energy difference between the two SiC$_2$ structures, the last time in 2003 when very high level calculations combined with very large basis sets, basis set extrapolation to infinite size as well as additional corrections yielded a value of 26.5~kJ/mol~\\cite{SiC2_ai_2003}. The SiC$_2$ molecular parameters obtained in Ref.~\\cite{SiC2_A-X_lab-non-lin_1984} laid the foundation for progress in laboratory spectroscopy. The $J = 1 - 0$ rotational transition frequencies of the three isotopologs SiC$_2$\\footnote{Unlabeled atoms refer to $^{12}$C and $^{28}$Si.}, $^{29}$SiC$_2$, and $^{30}$SiC$_2$, as well as the permanent electric dipole moment were measured using Fourier transform microwave spectroscopy~\\cite{SiC2_1-0_dip_1989}. Subsequently, 34 additional transition frequencies were measured for the main isotopic species between 93 and 370~GHz~\\cite{SiC2_rot_1989}. Even though a comparatively large number of 15 spectroscopic parameters of a standard Watson-type Hamiltonian in the $A$-reduction, all parameters up to sixth order, were employed in the fit, the transition frequencies were reproduced on average to only four times the experimental uncertainties. Similarly large bodies of laboratory transition frequencies were obtained for SiC$^{13}$C between 339 and 405~MHz~\\cite{SiCC-13_isos_astro_1991} and, only very recently, for $^{29}$SiC$_2$ and $^{30}$SiC$_2$ between 140 and 360~GHz~\\cite{Si-29_30-C2_rot_2011}. Higher excited vibrational states have been studied for the main isotopolog. Rotational transitions in its low-lying $\\varv_3 = 1$ vibrational state between 186 and 399~GHz~\\cite{SiC2_001_rot_1991} and in $\\varv_3 = 1$ and 2 were obtained between 140 and 400~GHz~\\cite{SiC2_001_002_rot_1994}. The symmetry of $\\varv_3 = 1$ is $b_2$ and can be viewed as an asymmetric bending state which facilitates internal rotation of the C$_2$ unit with respect to the Si atom. Its vibrational energy has been determined as 196.37~cm$^{-1}$ from an investigation into the laser-induced and the dispersed fluorescence of a jet-cooled sample of SiC$_2$~\\cite{SiC2_LiF_DF_1991}. Several attempts have been made to model rotational and sometimes also rovibrational data to a varying extent and accuracy. A semirigid bender (SRB) Hamiltonian was employed~\\cite{SiC2_modeling_1994} to reproduce $\\varv_3 = 0$ and 1 rotational transition frequencies~\\cite{SiC2_rot_1989,SiC2_001_rot_1991} as well as rovibrational data from their previous~\\cite{SiC2_LiF_DF_1991} and present work~\\cite{SiC2_modeling_1994}. The analysis suggested that the linear configuration is not a local minimum, and the cyclic form is 22.5~kJ/mol lower than the linear form; the estimated uncertainty was 2.4~kJ/mol. They found the energy difference to be in good agreement with results from their own ab initio calculations; their highest level value being 21.8~kJ/mol. The reproduction of the vibrational data was reasonable, that of the rotational data was only qualitative and thus of no use for radio-astronomical purposes. The $\\varv_3 = 0$ and 1 state rotational data~\\cite{SiC2_rot_1989,SiC2_001_rot_1991} were also modeled with a dedicated internal rotation Hamiltonian~\\cite{SiC2_modeling_1993}. A greatly improved reproduction, within 1.5~times the experimental uncertainties, was achieved for the ground vibrational state, albeit at the expense of 16 spectroscopic parameters compared to 15 previously~\\cite{SiC2_rot_1989}. The barrier to linearity was derived as $\\sim$54~kJ/mol, more than twice the value from ab initio calculations of that time. A reasonably successful result has been obtained by employing a conventional Watson-type Hamiltonian in the $S$-reduction~\\cite{IRC_10216_2008}. The reproduction of the ground state rotational data~\\cite{SiC2_1-0_dip_1989,SiC2_rot_1989} was converged after varying 17 spectroscopic and keeping one fixed. The data were reproduced on average as well as in the fit which employed an internal rotation Hamiltonian~\\cite{SiC2_modeling_1993}; the number of varied parameters, however, was larger by still another one. After scaling of the parameters, the transition frequencies of $^{29}$SiC$_2$ and $^{30}$SiC$_2$ available at that time, mostly from astronomical observation~\\cite{SiCC-13_isos_astro_1991,IRC_10216_2008}, could be reproduced well after releasing only 5 spectroscopic parameters~\\cite{IRC_10216_2008}; 13 parameter were released in the fit for SiC$^{13}$C because of the large amount of accurate laboratory data~\\cite{SiCC-13_isos_astro_1991}. The unambiguous assignment of a spectroscopic feature observed in space relies on reliable predictions which are usually based on laboratory data which, for the most part, have been obtained in approximately the same frequency domain. The predictions may also be based in part on data from astronomical observations, recent examples include H$^{13}$CO$^+$~\\cite{HC-13-O+_2004}, DCO$^+$~\\cite{DCO+_2005,DCO+_etc_2009}, DNC and HN$^{13}$C~\\cite{DCO+_etc_2009}, C$_2$H~\\cite{C2H_2009}, and C$^{13}$CH~\\cite{CC-13-H_2010}. Sometimes, identifications are even possible in the absence of laboratory data, as demonstrated recently by the detection of C$_5$N$^-$ in the CSE of CW~Leo~\\cite{det_C5N-}. Frequencies have to be predicted with accuracies better than around one tenth of the line width to permit extraction of dynamical information. Some type of intensity information, such as that at a certain temperature, the line strength, or the Einstein $A$-value is needed as is additional auxiliary information such as quantum numbers, lower or upper state energies etc. Pickett's {\\scriptsize SPCAT} and {\\scriptsize SPFIT} programs \\cite{Herb} have been developed for that purpose and have evolved over time, see e.\\,g. Ref.~\\cite{editorial_Herb-Ed}. These programs are routinely used in the Cologne Database for Molecular Spectroscopy\\footnote{Internet address: http://www.astro.uni-koeln.de/cdms}, CDMS~\\cite{CDMS_1,CDMS_2}, to provide in its catalog section\\footnote{Internet address: http://www.astro.uni-koeln.de/cdms/catalog} predictions of (mostly) rotational spectra of molecules which may be found in various environments in space. The spectroscopic parameters from the first rotational analysis of the electronic spectrum of SiC$_2$~\\cite{SiC2_A-X_lab-non-lin_1984} were accurate enough to identify nine unassigned emission features, previously observed between 93 and 171~GHz toward the carbon-rich AGB star CW~Leo, also known as IRC~+10216, as belonging to SiC$_2$ and improved the SiC$_2$ structural parameters~\\cite{SiC2_det-IRC_1984}. Using the structural parameters from that work, the three $J = 4 - 3$, $\\Delta K_a = 0$ transitions of $^{29}$SiC$_2$ and $^{30}$SiC$_2$ were detected toward the same source very soon thereafter~\\cite{Si-29_30-C2_det-IRC_1986}; two of these transitions were marginally detected for SiC$^{13}$C. Extensive sets of transition frequencies of $^{29}$SiC$_2$, $^{30}$SiC$_2$, and SiC$^{13}$C were obtained from radio-astronomical observation between 90 and 241~GHz together with laboratory rest frequencies for the latter isotopolog~\\cite{SiCC-13_isos_astro_1991}. Radio lines of SiC$_2$ have also been detected in the CSEs of other C-rich AGB stars, e.\\,g. toward II~Lup, which is also known as IRAS 15194-5115~\\cite{SiC2_II_Lup_1993}. However, SiC$_2$ features are particularly strong toward CW~Leo, which, to a large extent, is due to its proximity to our Solar system. CW~Leo has been studied extensively, and many molecular species, such as CN$^-$~\\cite{det_CN-} and FeCN~\\cite{det_FeCN}, have been detected toward its CSE exclusively or for the first time, including many Si-containing ones, such as SiCN and SiNC~\\cite{det_SiNC}. The source is a subject of observations in several key projects carried out with the recently launched {\\it Herschel} satellite~\\cite{Herschel}. One of these projects is a molecular line survey carried out with the Heterodyne Instrument for the Far-Infrared (HIFI)~\\cite{HIFI}. This high-resolution instrument covers, in several bands, the 480$-$1250 and 1410$-$1910~GHz regions. A preliminary analysis of a lower frequency region (554$-$637~GHz) revealed that a rather small number of molecules account for a large fraction of the emission features~\\cite{SiC2_HIFI_2010}. These molecules are CS, SiO, SiS, HCN, and, in particular, SiC$_2$. Predictions of the rotational spectrum based on Ref.~\\cite{IRC_10216_2008} turned out to be very good for transitions with higher values of $K_a$, but showed increasing deviations of up to 10~MHz for transitions with decreasing $K_a$, contrary to common expectations. A combined fit of these transitions frequencies together with those from laboratory spectra~\\cite{SiC2_1-0_dip_1989,SiC2_rot_1989} required only one additional parameter~\\cite{SiC2_HIFI_2010} to reproduce the astronomical data within uncertainties, and the laboratory data about as well as before~\\cite{IRC_10216_2008}. Predictions of the SiC$_2$ rotational spectrum based on these results are currently available as version~2 in the CDMS\\footnote{http://www.astro.uni-koeln.de/cgi-bin/cdmssearch?file=c052527.cat; see http://www.astro.uni-koeln.de/cgi-bin/cdmsinfo?file=e052527.cat for the documentation}. Since they were still not appropriate to predict the observed emission features satisfactorily to high frequencies (beyond 1100~GHz), we present here a combined fit using these as well as laboratory data, supplemented with data from additional astronomical observations. \\begin{figure*} \\begin{center} \\includegraphics[width=15.0cm]{sic2_fig_vs1.eps} \\end{center} \\caption{Sections of the molecular line survey of CW~Leo displaying selected SiC$_2$ transitions. Their quantum numbers $J'_{K'_a,K'_c} - J''_{K''_a,K''_c}$ are given in red (grey). The formulae and quantum numbers of other lines are given in blue (black).} \\label{spectrum} \\end{figure*} ", "conclusions": "\\label{Conclusions} Astronomical observations with {\\it Herschel} have been used to improve spectroscopic parameters of SiC$_2$ in its ground vibrational state greatly. The transition frequencies obtained from laboratory data as well as from astronomical observations have been reproduced to within the estimated uncertainties employing a conventional Watson-type Hamiltonian. These data should be useful to test alternative models to describe the rotational or even rovibrational energy levels of this floppy molecule. Predictions of the rotational spectrum as well as line, parameter and other auxiliary files, both from present fits as well as previous ones, will be available in the CDMS~\\cite{CDMS_1,CDMS_2}." }, "1112/1112.1046_arXiv.txt": { "abstract": "*{New instrumentation is providing new insights into intermediate mass pulsating Cepheids, particularly about their formation and history. Three approaches are discussed, using space (Hubble and Chandra) and ground-based studies (radial velocities). First, we are conducting a survey of Cepheids with the Hubble Space Telescope Wide Field Camera 3 (WFC3) to identify possible resolved companions (for example Eta Aql) and thus provide constraints on star formation. Followup X-ray observations (Chandra and XMM-Newton) can confirm whether possible low mass companions are young enough to be physical companions of Cepheids. In a related study of intermediate mass stars, Chandra X-ray observations of late B stars in Tr 16 have been used to determine the fraction which have X-ray active low mass companions. Finally, the Tennessee State Automatic Spectroscopic Telescope AST and the Moscow University group have obtained velocities of a number of Cepheids. As an example, the orbit of V350 Sgr has been redetermined, providing a new level of accuracy to the orbital velocity amplitude, which is needed for mass determination. } \\abstract{New instrumentation is providing new insights into intermediate mass pulsating Cepheids, particularly about their formation and history. Three approaches are discussed, using space (Hubble and Chandra) and ground-based studies (radial velocities). First, we are conducting a survey of Cepheids with the Hubble Space Telescope Wide Field Camera 3 (WFC3) to identify possible resolved companions (for example Eta Aql) and thus provide constraints on star formation. Followup X-ray observations (Chandra and XMM-Newton) can confirm whether possible low mass companions are young enough to be physical companions of Cepheids. In a related study of intermediate mass stars, Chandra X-ray observations of late B stars in Tr 16 have been used to determine the fraction which have X-ray active low mass companions. Finally, the Tennessee State Automatic Spectroscopic Telescope AST and the Moscow University group have obtained velocities of a number of Cepheids. As an example, the orbit of V350 Sgr has been redetermined, providing a new level of accuracy to the orbital velocity amplitude, which is needed for mass determination. } \\vskip .1truein ", "introduction": "This contribution focuses on two aspects of binary Cepheids: information they provide about star formation and about stellar evolution (masses). \\begin{figure} \\centering \\includegraphics[scale=.20]{evans_fig1a.eps} \\hspace{0.5cm} \\includegraphics[scale=.25]{evans_fig1b.eps} \\caption{Left: The center of the HST WFC3 V image of $\\eta$ Aql. The image has a log scale and is approximately 10$''$ wide. Right: The difference image with the Cepheid T Mon subtracted from the $\\eta$ Aql image. The companion is clearly visible 0.7$''$ from the Cepheid. Only the vertical column bleeding remains uncorrected. } \\label{fig:1} % \\end{figure} ", "conclusions": "" }, "1112/1112.3569_arXiv.txt": { "abstract": "Simple models of magnetic field generation by convection in rotating spherical shells exhibit properties resembling those observed on the sun. The {assumption of the Boussinesq approximation made in these models} prevents a realistic description of the solar cycle, but through a physically motivated change in the boundary condition for the differential rotation the propagation of dynamo waves towards higher latitudes can be reversed at least at low latitudes. ", "introduction": "A well known difficulty in modeling the solar dynamo is the fact that the dynamo waves which describe the nearly time periodic dynamics of the magnetic field tend to propagate from lower to higher latitudes instead in the opposite sense as observed on the sun. This effect is well known in mean field models of the solar cycle, see for example Stix (1976), but it is also observed in numerical solutions {of convection-driven} dynamos in rotating spherical fluid shells which are supposed to model processes in the solar convection zone. This shortcoming and others are caused by the inadequate representation of the compressibility of the solar atmosphere. Even with the huge power of modern supercomputers it is not yet possible to resolve adequately the dynamics of convection in the presence of the large density variation between bottom and top of the solar convection zone and to model appropriately the {dependence of density on pressure.} {In their early} direct numerical models of solar convection and magnetic field generation, Gilman and Miller (1981) assumed the Boussinesq approximation in which the fluid is regarded as incompressible except in connection with the gravity term where the temperature dependence of the density is taken into account. This approximation eliminates the need for a separate equation of state and leads to a system of equations describing long period processes while the short period acoustic modes no longer enter the analysis. The same effect is obtained in the anelastic approximation in which the horizontally averaged density variation is taken into account, but the fluctuating component of the density is still only represented in the gravity term. For applications of the anelastic approximation in models of the solar convection zone see Gilman and Glatzmaier (1981) and Elliot et al. (2000). Convection in rotating spherical fluid shells heated from below is always associated with a differential rotation generated by the Reynolds stresses of convection. An analytical model demonstrating the preference of banana shaped convection cells girdling the equator and the associated solar like differential rotation was presented by Busse (1970, 1973). While the analytical solution for stress-free boundaries exhibits a depth independent differential rotation, a differential rotation decreasing with depth is always found in fully nonlinear numerical models. This property together with the fact that the differential rotation reaches its maximum at the equator is responsible for the propagation of the dynamo waves towards higher latitudes (Yoshimura, 1975). \\begin{figure*} \\psfrag{Omg}{$\\Omega$} \\begin{center} \\epsfig{file=Fig01.eps,width=1.5\\columnwidth,clip=} \\end{center} \\caption[]{{(color online)} Differential rotation of the solar convection zone according to helioseismology (right, image courtesy NSF's National Solar Observatory) and differential rotation of the numerical model in the case $\\eta=0.65$, $P=1$, $\\tau=2$, $R=120000$, $P_m=4$, $\\beta=1.5$ and mixed velocity boundary conditions {(left)}. } \\label{fig010} \\end{figure*} The solar differential rotation also decreases with depth throughout most of the convection zone with the exception of the tachocline region near its bottom and a region near its surface as indicated in figure \\ref{fig010}(b). There is no general agreement about the origin of {the} upper 30Mm deep layer in which the differential rotation increases with depth. Here we assume that it is caused by supergranular convection that is characterized by a strong asymmetry between rising hot plasma and descending cool plasma. This type of convection has been modeled by hexagonal convection cells in the presence of rotation (Busse, {2007}). As shown in this paper the asymmetry between rising and descending flow in hexagonal convection cells does indeed generate a differential rotation that increases with depth. We use this dynamical property of convection as motivation to modify the usually assumed stress-free boundary condition. Solely for the differential rotation we apply the condition given in expression (A11) of the Appendix. The resulting profiles shown in the example of figure \\ref{fig010}(a) are still not very solar like, but they show an increase with depth of the differential rotation in the upper layer and tend to generate a more solar like behavior in the numerical simulations, {as discussed below.} ", "conclusions": "In this report an attempt is described to explore the extent to which the operation of the solar dynamo can be understood on the basis of a minimal, but physically consistent, {convection-driven} dynamo model. Although the Boussinesq approximation is a highly unrealistic assumption in the case of the Sun, the convection columns are similar to those found in numerical simulations based on {anelastic} models, see, for instance, Brun et al.~(2004) and Ghizaru et al.~(2010). Although there is little solar evidence for this type {of} convection, there are generally believed to exist as \"giant cells\" in the deeper region of the solar convection zone. The structure of the magnetic field found in our simulations differs substantially from the commonly assumed structure of the solar magnetic field dominated by a strong axisymmetric toroidal component. When the evidence for this traditional view is examined, however, it is found that observations do not contradict fields dominated by $m=1$ or $m=2$ components such as those shown in figures \\ref{fig020} and \\ref{fig050}. On the contrary, the presence of active longitudes on the sun indicates at least that components with the azimuthal wavenumbers $m=1$ and $m=2$ play a significant role. {A more detailed exploration of the parameter space of our minimal model is likely to reveal an even better correspondence with solar observations.} \\appendix" }, "1112/1112.4349_arXiv.txt": { "abstract": "{We present the analysis of the bright X-ray binary 4U~1820-30, based mainly on \\xmm-RGS data, but using complementary data from \\xmm-\\epi, \\inte, and \\ch-HETG, to investigate different aspects of the source. The broad band continuum is well fitted by a classical combination of black body and Comptonized emission. The continuum shape and the high flux of the source ($L/L_{\\rm Edd}\\sim0.16$) are consistent with a \"high state\" of the source. We do not find significant evidence of iron emission at energies $\\geq6.4$\\,keV. The soft X-ray spectrum contain a number of absorption features. Here we focus on the cold-mildly ionized gas. The neutral gas column density is $N_{\\rm H}\\sim 1.63\\times10^{21}$\\,cm$^{-2}$. The detailed study of the oxygen and iron edge reveals that those elements are depleted, defined here as the ratio between dust and the total ISM cold phase, by a factor $0.20\\pm0.02$ and $0.87\\pm0.14$, respectively. Using the available dust models, the best fit points to a major contribution of Mg-rich silicates, with metallic iron inclusion. Although we find that a large fraction of Fe is in dust form, the fit shows that Fe-rich silicates are disfavored. The measured Mg:Fe ratio is $2.0\\pm0.3$. Interestingly, this modeling may point to a well studied dust constituent (GEMS), sometimes proposed as a silicate constituent in our Galaxy. Oxygen and iron are found to be slightly over- and under-abundant, respectively (1.23 and 0.85 times the solar value) along this line of sight. We also report the detection of two absorption lines, tentatively identified as part of an outflow of mildly ionized gas ($\\xi\\sim-0.5$) at a velocity of $\\sim1200$\\,\\kms.} {} ", "introduction": "\\label{par:intro} The interstellar medium (ISM) in the plane of our Galaxy is a dynamic and complex environment, composed of mainly neutral matter in both gas and dust form and by a warmer gas phase in the form of diffuse emission in, and above, the Galactic plane. The properties of the cold phase in the diffuse ISM have been extensively studied at long wavelengths, from the far-infrared to the far-UV \\citep[e.g.][ for a review]{draine03}. A sizeable fraction of the cold phase is locked up in dust grains \\citep[e.g.][ and references therein]{ss96,jenkins09}. Amorphous silicate materials together with graphite and policyclic aromatic carbon should account for the majority of the depleted elements measured in the ISM: C, O, Fe, Mg, Si \\citep[e.g.][]{wd01,wooden08}. One of the major spectroscopic signatures of the presence of Fe- and Mg- rich amorphous silicates is the 10\\,$\\mu$m emission feature. Extensive studies of this feature lead to the conclusion that the Fe:Mg ratio should be approximatively 1 \\citep[e.g.][]{lidraine01}. Main sources of both Mg and Fe silicates are O-rich asymptotic giant branch stars (AGB) and supernovae. However, the process of amorphization of the dust agglomerates, for instance by rapid cooling of the gas phase \\citep{wooden05} or cosmic rays bombardment \\citep{carrez02}, strongly favors the survival of Mg silicates. In fact, a recent analysis successfully models the 10\\,$\\mu$m feature in terms of Mg-rich silicates, if non-spherical grain shapes are used \\citep{min07}. Glassy material, consisting of Mg-rich silicates with metallic iron and sulfide inclusion \\citep[called GEMS,][]{bradley94} have also been commonly found during the {\\em Stardust} mission. Their origin is mostly from the interplanetary environment \\citep[e.g.][]{kel_mes04}, but a fraction have a composition compatible with an ISM origin \\citep{kel_mes08}. Iron is however a highly depleted element \\citep[70--99\\% of Fe is in dust,][]{wilms00,whi03} whose inclusion into solid grains is not completely understood \\citep[e.g.][]{whi03}. This is mainly due to the difficulty of modeling iron emission, which does not display any sharp feature in long-wavelength spectra. The abundances of the most important metals in the Galactic disk smoothly decrease with the galactocentric distance. The average slope of the distribution is $\\sim0.06$\\,dex\\,kpc$^{-1}$ \\citep[][ and references therein]{chen03}. However, a large scatter in the abundance measurements as a function of the Galactic radius is reported. This is attributed to different factors which contaminate the smooth mixing due to the pure stellar evolution process. Indeed the medium can be locally influenced by e.g. supernovae ejecta, and in-falling metal-poor material into the disk \\citep{lugaro99,nittler05}. In recent years it has become clear that the X-ray band could provide an excellent laboratory to study the silicate content of the diffuse ISM, as the absorption K edges of O ($E=0.538$\\,keV), Mg ($E=1.30$\\,keV) and Si ($E=1.84$\\,keV) together with the Fe LII and LIII edges ($0.71$ and $0.72$\\,keV, respectively) fall in the low-energy X-ray band. The method used is to study the absorbed spectrum of bright X-ray binaries, located in different regions in the disk, observed with high-energy resolution instruments. This allows to probe the interstellar dust (ID) content in a variety of environments, with different extinction and with possibly different dust formation history. Previous studies of X-ray spectra taken along different lines of sight led first to the recognition that not only gas but also dust plays a role in shaping the iron and oxygen edges \\citep{takei02,kaastra09} and later led to the quantitative modeling of those edges \\citep{lee09,devc09,ciro}. There is not yet a clear picture of the chemical composition of the ID as seen in X-rays. Silicates containing andratite (iron-rich silicates) were reported studying the oxygen edge of \\object{GS~1826-238} \\citep{ciro}, while iron oxides, rather than iron silicates were reported along the line of sight of \\object{Cyg~X-1} \\citep{lee09}. This may point to a chemically inhomogeneous distribution of ID.\\\\ \\object{4U~1820-30} is an extensively studied source, by virtue of its extraordinary intrinsic properties. It is an ultracompact \\citep[orbital period 11.4\\,minutes,][]{1987ApJ...312L..17S} X-ray binary consisting of a neutron star and a He white-dwarf \\citep{1987ApJ...322..842R}. The presence of X-ray bursts associated with the neutron star has been early recognized \\citep[first by][]{grindlay76}. \\uu\\ is classified as an atoll source \\citep{1989A&A...225...79H}, showing kilohertz quasi-periodic oscillation at different frequencies in its power spectrum \\citep[e.g.][]{smale97,zhang98}.\\\\ The broad band spectrum has been studied with several instruments. The source displays a Comptonized continuum and a soft black body component \\citep[e.g.][]{sidoli01}. With the advent of high resolution spectroscopy, the low-energy X-ray spectrum also revealed interesting features. Absorption by ionized ions has been reported in several studies \\citep{futamoto04,yw05,juett06,cackett08}. The absence of blueshift in the lines of this gas and lack of variability generally points to an interstellar origin \\citep[but see][]{cackett08}. The hypothesis of a gas intrinsic to the source is intriguing as other X-ray binaries often display absorption by ionized gas either at the rest frame of the source \\citep{vanpeet} or outflowing \\citep[e.g.][]{jmiller06,neilsenlee09}. The source has also been used as a backlight to illuminate the interstellar dust along the line of sight. This results in a dust scattering halo, which in this source is moderate \\citep[$\\sim3.2$\\% at 1\\,keV,][]{peter95}, given the relatively low Galactic column density ($N_{\\rm H}\\sim1.63\\times10^{21}$\\,cm$^{-2}$, this study). Absorption by cold interstellar dust has been never studied in detail in this source. In this paper we aim for a comprehensive view of the cold absorbing medium along the line of sight to this source. This study benefits from the combined information provided by the \\ch\\ and \\xmm\\ high resolution instruments, which allows us to meaningfully study both the Fe\\,L and O\\,K edges. In addition, we make use of dust models updated with the latest laboratory measurements \\citep[e.g.][]{lee08,lee09}. We also present the broad band continuum behavior underlying the absorption features.\\\\ The adopted protosolar abundances follow the prescription given by \\citet{lodders09} and discussed in \\citet{lodders10}. The broad band energy spectrum (\\epi\\ and \\inte) is fitted using the $\\chi^2$ minimization method, taking care that at least 20 counts per bin are present in each data set. For the high resolution spectra (RGS and \\ch-HETG) the Cash statistic has been used\\footnote{http://heasarc.gsfc.nasa.gov/docs/xanadu/xspec/manual/XSappendixCash.html}. The errors quoted are for 68\\% confidence level, corresponding to $\\Delta\\chi^2=1$ or $\\Delta C^2=1$. The spectral fitting package used in this paper is SPEX\\footnote{www.sron.nl/spex} \\citep{kaastra96}. The adopted distance is 7.6\\,kpc \\citep{kuulkers03}. The nominal hydrogen column density toward the source is $1.29\\times10^{21}$\\,cm$^{-2}$ \\citep{kal05}, to be compared with $1.52\\times10^{21}$\\,cm$^{-2}$ \\citep{dl90}. This paper is organized as follows: in Sect.~\\ref{par:data}, we illustrate the data handling for the different instruments used. Sect.~\\ref{par:cont} is devoted to the continuum determination, using \\epi\\ and \\inte\\ data. In Sect.~\\ref{par:high} we describe in detail the modeling of the different absorption components using RGS and \\ch-MEG. The discussion can be found in Sect.~\\ref{par:discussion} and the conclusions in Sect.~\\ref{par:conclusions}.\\\\ ", "conclusions": "\\label{par:conclusions} In this paper we present the X-ray analysis of the continuum of \\uu\\ (using a quasi-simultaneous observation of \\xmm\\ and \\inte) and of the absorption features due to the cold matter in the line of sight (using \\xmm-RGS and \\ch-MEG data).\\\\ The continuum shape and the Eddington ratio show that the source was caught in a high-state. The continuum is well fitted by black body emission plus a Comptonization component which extends up to 40\\,keV. We do not find evidence of iron emission, either from neutral or ionized matter. This may be naturally explained by the metal-poor accretion stream expected from the white-dwarf companion.\\\\ The absorption spectrum shows the presence of many components with different ionization. We focused on the cold and mildly ionized phase only. Oxygen has been found slightly overabundant by a factor 1.23 times the solar value. Iron is on the contrary slightly underabundant ($\\sim$0.85 times solar). The abundance values are not dramatically deviating from the solar ones and do not allow us to assign a precise location of the absorbing gas. However, a location close to the observer seems likely. Thanks to the simultaneous study of absorption by dust and gas we measured also the element depletion. Oxygen is mildly depleted by a factor about 0.20. The depletion of iron is more evident, as the depletion factor is 0.87. The depletion of Mg and Si are more difficult to determine. We find that they are depleted of a factor $>0.97$ and $>0.86$, respectively\\\\ We modeled the dust contribution with the currently availables absorption profiles of dust compounds. Our conclusions bear the uncertainty due to the still limited dust data-base and a lower sensitivity in selected spectral regions. However we clearly find that both the oxygen and iron edges cannot be fitted by iron-rich silicates. On the contrary, the oxygen edge is consistent to be mostly absorbed by enstatite (MgSiO$_3$, possibly in a glass-form). Metallic iron should be the main absorber in the iron edge. This leads to the interesting possibility that a GEMS-like form (Mg-rich silicates with metallic iron inclusion) of grain may be absorbing along this line of sight. A fraction of the studied GEMS, in particular the sulfur-poor grains, are believed to be of ISM origin and have also been proposed as constituents of ISM. For the first time an X-ray absorption analysis provides a tentative confirmation of this scenario.\\\\ Finally, we report the tentative detection of a mildly ionized outflow ($v_{\\rm out}\\sim1200$\\,\\kms), highlighted by the \\oiv\\ and \\ov\\ absorption lines. Both a photo- or collisional- ionizing process could fit the lines, leaving open the interpretation on the nature of this gas." }, "1112/1112.4663_arXiv.txt": { "abstract": "The super-massive objects in galactic nuclei are thought to be the Kerr black holes predicted by General Relativity, although a definite proof of their actual nature is still lacking. The most massive objects in AGN ($M \\sim 10^9$~$M_\\odot$) seem to have a high radiative efficiency ($\\eta \\sim 0.4$) and a moderate mass accretion rate ($L_{\\rm bol}/L_{\\rm Edd} \\sim 0.3$). The high radiative efficiency could suggest they are very rapidly-rotating black holes. The moderate luminosity could indicate that their accretion disk is geometrically thin. If so, these objects could be excellent candidates to test the Kerr black hole hypothesis. An accurate measurement of the radiative efficiency of an individual AGN may probe the geometry of the space-time around the black hole candidate with a precision comparable to the one achievable with future space-based gravitational-wave detectors like LISA. A robust evidence of the existence of a black hole candidate with $\\eta > 0.32$ and accreting from a thin disk may be interpreted as an indication of new physics. For the time being, there are several issues to address before using AGN to test the Kerr paradigm, but the approach seems to be promising and capable of providing interesting results before the advent of gravitational wave astronomy. ", "introduction": "Gravity has been tested and verified for distances in the range $\\sim$1~mm to $\\sim$1~pc (mainly within its Newtonian limit) and for weak gravitational fields~\\cite{w,5th}. The research is now moving to check the validity of the theory at cosmological scales, sub-millimeter distances, and for strong gravitational fields. One of the most intriguing predictions of General Relativity (GR) is that the collapsing matter produces singularities in the space-time. According to the weak cosmic censorship conjecture, singularities of gravitational collapse must be hidden within black holes (BHs)~\\cite{p}. In 4-dimensional GR, uncharged BHs are described by the Kerr solution, which is completely specified by two parameters, the mass, $M$, and the spin angular momentum, $J$~\\citep{c}. The condition for the existence of the event horizon is $a_* \\le 1$, where $a_* = |J/M^2|$ is the spin parameter\\footnote{Throughout the paper I use units in which $G_N = c = 1$, unless stated otherwise.}. When $a_* > 1$, there is no horizon and the central singularity is naked, violating the weak cosmic censorship conjecture. Astronomers have discovered at least two classes of BH candidates (for a review, see e.g. Ref.~\\cite{n}): stellar-mass objects in X-ray binary systems ($M \\sim 5 - 20$~$M_\\odot$) and super-massive objects in galactic nuclei ($M \\sim 10^5 - 10^9$~$M_\\odot$). The estimates of the masses of these objects are robust, because determined via dynamical measurements and without any assumption about the geometry of the space-time. The key-point is that the stellar-mass objects in X-ray binary systems are too heavy to be neutron or quark stars for any reasonable matter equation of state~\\cite{kb}, while the super-massive objects at the centers of galaxies are too heavy, compact, and old to be clusters of non-luminous bodies, as the cluster lifetime would be shorter than the age of these systems~\\cite{m}. All these objects are therefore thought to be the BHs predicted by GR, as they cannot be explained otherwise without introducing new physics. There are also some observations interpreted as an indirect evidence for the existence of the event horizon~\\cite{bln} (but see~\\cite{a}). On the contrary, there is no indication that the geometry around these objects is described by the Kerr metric. Testing the Kerr BH hypothesis is thus the next step to progress in this research field and several authors have indeed suggested possible ways to do it using present and future data (for a review, see e.g. Ref.~\\cite{b4}). A very promising approach is the detection of extreme mass ratio inspirals (EMRIs, i.e. systems consisting of a stellar-mass compact object orbiting a super-massive BH candidate) with future space-based gravitational-wave antennas. Missions like LISA will be able to follow the stellar-mass compact object for millions of orbits around the central super-massive BH candidate, and therefore deviations from the Kerr geometry will lead to a phase difference in the gravitational waveforms that grows with the number of observed cycles~\\cite{r}. However, these data will not be available shortly, as the first mission will be at best in the early 2020s. The nature of BH candidates can also be tested by extending the methods currently used to estimate the spin of these objects, such as X-ray continuum~\\cite{bb1} and K$\\alpha$-iron measurements~\\cite{pj}, observations of quasi-periodic oscillations~\\cite{jpp}, and measurements of the cosmic X-ray background~\\cite{b2,b5}. These methods can in principle be applied even with present data, provided that the systematic errors are properly understood. Future observations of the shadow of nearby super-massive BH candidates are another exciting possibility to test the Kerr BH paradigm~\\cite{bf}. Previous studies have clearly pointed out that ``rapidly-rotating'' objects are the best candidates to test the Kerr BH hypothesis: if the object rotates fast, even a small deviation from the Kerr background can cause significant differences in the properties of the electromagnetic radiation emitted by the gas of the accretion disk and peculiar features, otherwise absent in the Kerr geometry, may show up~\\cite{bb2,bm}. The aim of this paper is to investigate the potentialities of the most massive BH candidates in AGN, through the measurement of their radiative efficiency $\\eta$, defined by $L_{\\rm bol} = \\eta \\dot{M}$, where $L_{\\rm bol}$ is the bolometric luminosity of the source and $\\dot{M}$ is the mass accretion rate of the BH candidate. The estimate of the mean radiative efficiency of AGN through the Soltan's argument~\\cite{s} already suggests the presence of rapidly-rotating BHs~\\cite{erz,ho}. Recently, Davis and Laor have proposed a way to measure the radiative efficiency of individual AGN~\\cite{dl}. The authors found that the most massive BH candidates (with a mass $M \\sim 10^9$~$M_\\odot$) would have a high radiative efficiency, up to $\\eta \\sim 0.4$, and a moderate mass accretion rate, $L_{\\rm bol}/L_{\\rm Edd} \\sim 0.3$, where $L_{\\rm Edd}$ is the Eddington luminosity of the source. The standard accretion disk model in a Kerr background would predict a high value of the spin parameter $a_*$ for these objects, extremely close to 1. At the same time, the moderate luminosity (in Eddington units) may indicate a thin accretion disk and the applicability of the standard accretion disk model. If these estimates and these considerations are correct, the most massive compact objects in AGN would be excellent candidates to test the Kerr paradigm. The sole measurement of $\\eta$ can potentially constrain either $a_*$ and a deviation from the Kerr geometry. The paper is organized as follows. In Section~\\ref{s-2}, I review the standard accretion disk model, its assumptions and properties, its effects on the evolution of the spin parameter of the central object, and its applicability. In Section~\\ref{s-k}, I consider an accretion disk in the Kerr background, summarizing well-known results that should be expected if the BH candidates are the BHs of GR. In Sections~\\ref{s-jp} and \\ref{s-mn}, I discuss accretion disks respectively in the Johannsen-Psaltis (JP)~\\cite{jp} and in the Manko-Novikov (MN)~\\cite{mn} space-times. These are two metrics that can be conveniently used to describe a background deviating from the Kerr geometry. The nature of the two metrics is definitively different: the JP metric describes non-Kerr BHs in a putative alternative theory of gravity, while the MN one is an exact solution of the Einstein's vacuum equations and can describe the exterior gravitational field of generic compact objects. In both cases, the body is characterized by a mass, a spin angular momentum, and an infinite number of ``deformation parameters'', even if here, for the sake of simplicity, I will consider only a single deformation parameter at a time. I will show that the two metrics present common features. In particular, a high radiative efficiency necessarily requires a very rigid compact object, much stiffer than a self-gravitating fluid with ``normal'' equations of state. The confirmation of the existence of individual AGN with high radiative efficiency ($\\eta > 0.3$) can potentially either be used to put strong constraints on the Kerr nature of astrophysical BH candidates and to discover new physics, as in the framework of the standard accretion disk model $\\eta$ cannot exceed 0.32. In Section~\\ref{s-d}, I discuss the findings of this paper in relation with current estimates of the radiative efficiency of AGN. The conclusions and the issues that need to be addressed before using the most massive objects in AGN to really test the Kerr BH hypothesis are reported in Section~\\ref{s-c}. In Appendices~\\ref{a-1} and \\ref{a-2}, the reader can find the non-vanishing metric coefficients respectively of the JP and of the MN metric. ", "conclusions": "\\label{s-c} There is some evidence that the most massive BH candidates in AGN have a high radiative efficiency and a thin accretion disk. In this case, they could be excellent candidates to test GR in the strong field regime and, in particular, the Kerr BH paradigm. A robust observation of a high radiative efficiency together with the confirmation of the validity of the standard accretion disk model for the accretion process onto these objects would constrain possible deviations from the Kerr geometry of the space-time around astrophysical BH candidates. If we restrict our attention only to objects more oblate than a Kerr BH with the same spin parameter, a radiative efficiency $\\eta > 0.30$ requires that the quadrupole moment of the compact object deviates not more than $\\sim 10^{-3}$ with respect to the one of a Kerr BH. If we allow for the existence of compact objects more prolate than a Kerr BH (whose existence, however, may be questionable, see the footnote~3), the constraints is much weaker, at the level of 20\\%. On the other hand, the observation of an accreting BH candidate with $\\eta > 0.32$ might indicate that the super-massive BH candidates in galactic nuclei are not the Kerr BH of GR, as there is no realistic astrophysical mechanism capable to spin up them to $a_* > 0.998$. For the time being, there are at least two main issues to address before using the measurement of the radiative efficiency of individual AGN to test GR: \\begin{enumerate} \\item We do not really know if the NT model can describe thin accretion disks around astrophysical BH candidates. If it can or it is easy to estimate the necessary corrections, the measurement of $\\eta$ could provide an interesting way to probe the geometry of the space-time around BH candidates. If it cannot, it seems to be impossible to test GR (with this or other approaches) from the properties of the radiation emitted by the gas in the accretion disk and we should wait for the advent of gravitational wave astronomy. \\item We need robust and more precise measurements of $\\eta$. At present, there are several sources with $\\eta > 0.32$ at high masses, but we cannot really exclude they actually have $\\eta < 0.32$~\\cite{dl}. There are some sources of uncertainty in the final estimate of the radiative efficiency and the most important one is the determination of the mass of the BH candidate. Moreover, the results presented in~\\cite{dl} are based on a preliminary study and it is necessary to further investigate and verify the validity of the method proposed by these authors. \\end{enumerate}" }, "1112/1112.4488_arXiv.txt": { "abstract": "{ We propose a general scenario for moduli stabilization where low-energy supersymmetry can be accommodated with a high scale of inflation. The key ingredient is that the stabilization of the modulus field \\emph{during} and \\emph{after} inflation is not associated with a single, common scale, but relies on two different mechanisms. We illustrate this general scenario in a simple example, where during inflation the modulus is stabilized with a large mass by a K\\\"ahler potential coupling to the field which provides the inflationary vacuum energy via its F-term. After inflation, the modulus is stabilized, for instance, by a KKLT superpotential. } ", "introduction": "\\label{Sec:Introduction} Low-energy supersymmetry (SUSY) at the TeV-scale provides an attractive extension of the Standard Model (SM) for various reasons, in particular, since it can solve the gauge hierarchy problem. Furthermore, it is going to be tested in the ongoing LHC experiments. Regarding cosmology, the paradigm of cosmic inflation has emerged as the prime candidate for early universe physics to resolve the flatness and horizon problems associated with the hot big bang scenario. Typically, models of inflation require a comparatively high scale of inflation, close to the energies where the gauge interactions of the SM can be unified. The scale of inflation might soon be tested by an observation of tensor modes with the PLANCK satellite. It is therefore an interesting question whether a high scale of inflation can be realized together with TeV-scale SUSY. A priori, the two topics of low-energy SUSY and high-scale inflation seem to be unrelated. However, they turn out to be connected due to the issue of ``moduli stabilization'', in particular in the context of string theory. String compactifications generically lead to many light scalar fields, aka moduli, which for example parametrize the size and shape of the compactified internal manifolds. Tracing the physics of extradimensions, the couplings in the 4D effective low-energy theory become functions of the moduli fields. To make sure that the internal dimensions do not decompactify, and also to avoid observational constraints on the space-time variability of the low-energy coupling constants, the moduli fields must be stabilized (i.e. must acquire a suitable mass) both during and after inflation. In the context of type IIB string theory, the issue of moduli stabilization is most well understood in the well-known KKLT scenario \\cite{Kachru:2003aw}. It assumes that the dilaton and the complex structure moduli are stabilized via fluxes \\cite{Giddings:2001yu,Dasgupta:1999ss,Taylor:1999ii}.\\footnote{For reviews and an extensive list of references on moduli stabilization with fluxes and nonperturbative effects in type IIB string theory cf. e.g. \\cite{Grana:2005jc,Douglas:2006es,Blumenhagen:2006ci,Denef:2007pq}.} Therefore, only the dynamics of the volume moduli are important for the low-energy physics. The volume moduli are stabilized by the contributions from nonperturbative effects such as gaugino condensates. Moreover, moduli stabilization also nicely combines with dynamical supersymmetry breaking with the moduli comprising the hidden sector. When the issue of moduli stabilization is discussed in conjunction with inflation, however, a severe problem emerges, namely adding the inflationary sector may destabilize the moduli, which was pointed out by Buchm\\\"uller, Hamaguchi, Lebedev and Ratz and by Kallosh and Linde in \\cite{Buchmuller:2004xr,Buchmuller:2004tz,Kallosh:2004yh}. Here, we are concerned with a particular version of this problem sometimes referred to as the Kallosh-Linde (KL) problem \\cite{Kallosh:2004yh}: One often finds an upper bound on the inflationary energy scale in terms of the present-day gravitino mass, \\begin{equation} \\Hinf \\leq m_{3/2}^{\\text{today}} \\, , \\end{equation} to avoid the destabilization of the volume modulus. For TeV-scale SUSY breaking, one has $m_{3/2} \\sim \\text{TeV}$ and thus the scale of inflation is bound to be very small, much below the scale required for many model building approaches (and also very much below observational sensitivities for gravitational waves). Basically, the problem appears since there is effectively only one scale in the problem, which sets both the gravitino mass today and the height of the barrier towards decompactification. A SUSY breaking uplifting term turns the AdS minimum into a nearly Minkowski minimum, and also sets the height of the barrier separating the metastable vacuum from the vacuum at infinity. Typically, inflation in such a setup can be viewed as an additional uplifting which induces a runaway potential for the modulus. Thus, if its contribution becomes too large, the barrier and hence the minimum disappear. In summary, the inflationary scale is constrained to be smaller than the height of the barrier, and therefore the gravitino mass. We will review the problem in more detail in the next section. As a solution to the problem, KL \\cite{Kallosh:2004yh} suggested a form of the superpotential for which the gravitino mass in the present vacuum is unrelated to the height of the barrier. Therefore, choosing the gravitino mass at the TeV-scale, we can always independently increase the barrier height such that high-scale inflation models do not destabilize the modulus. This corresponds to a fine-tuning of the terms in the superpotential. Following the approach of KL, in the context of volume modulus inflation in a racetrack setup, the problem has been addressed when one of the exponents of the nonperturbative terms is positive \\cite{Abe:2005rx,Badziak:2008yg,Abe:2008xu,Badziak:2008gv,Badziak:2009eh}. However, this can be done only at the expense of introducing more parameters in the theory. In the large volume scenario, attempts have been made to accommodate a small gravitino mass with a high scale of inflation, when inflation happens exponentially far away from the present Minkowski vacua. Other than some inevitable fine tuning, the working models have several phenomenological difficulties \\cite{Conlon:2008cj}. Dynamical avenues for the case of chaotic inflation \\cite{He:2010uk} and hybrid inflation \\cite{Kobayashi:2010rx} have also been explored, where the gravitino mass becomes inflaton-dependent in a suitable way. Difficulties related to the realizations of high-scale inflation together with low-energy SUSY breaking have been discussed in \\cite{Davis:2008sa,Davis:2008fv} and some resolutions have been proposed in \\cite{Mooij:2010cs,Davis:2008sa}. However, this has been successfully achieved only for the superpotential of \\cite{Kallosh:2004yh}. Combining chaotic inflation and supersymmetry breaking within the KL scheme has recently been discussed in \\cite{Kallosh:2011qk} for the general chaotic inflation models of \\cite{Kallosh:2010ug,Kallosh:2010xz,Demozzi:2010aj}. For other approaches to the KL problem see e.g. \\cite{Misra:2009ei}. Here, we propose a different resolution of the problem, namely to use two different mechanisms to stabilize the moduli during and after inflation. During inflation, the modulus field receives a large mass proportional to the inflationary vacuum energy. This is achieved, for example, by a suitable moduli-dependence of the K\\\"ahler metric of the field driving supersymmetry breaking during inflation. The role of such a type of coupling for moduli stabilization was first noted in \\cite{Antusch:2008pn}. At the end of inflation, the vacuum energy goes away and we invoke a stabilization mechanism like KKLT relying as usual on nonperturbative terms in the superpotential. We illustrate our general idea in a simple example: chaotic inflation protected by a shift symmetry combined with a KKLT-type superpotential. However, we stress that our framework can be applied to more general setups. We begin with a brief review of the Kallosh-Linde Problem in section~\\ref{Sec:ReviewKalloshLindeProblem}. Afterwards, we describe our general framework for resolving the problem in section~\\ref{Sec:ResolutionGeneralFramework} and illustrate it in a simple example with shift symmetric chaotic inflation and a KKLT superpotential in section~\\ref{Sec:AnExplicitExample}. Some results for a generic choice of $W_{\\mod}(T)$ are presented in appendix \\ref{Sec:AppendixGeneralWmod}. Finally, we give our conclusions in section~\\ref{Sec:Conclusion}. ", "conclusions": "\\label{Sec:Conclusion} In this article, we have proposed a general scenario for moduli stabilization where low-energy supersymmetry breaking can be accommodated together with a high scale of inflation. In our proposal the KL problem (which is reviewed in section \\ref{Sec:ReviewKalloshLindeProblem}) is resolved because the stabilization of the modulus field \\emph{during} and \\emph{after} inflation is not associated with a single, common scale, but instead relies on two different mechanism to stabilize the modulus during and after inflation. More explicitly (c.f.\\ section \\ref{Sec:ResolutionGeneralFramework}), we suggest to consider a K\\\"ahler potential which features a coupling between the modulus field and the field whose F-term drives inflation in such a way that the term $V_{\\inf} \\sim e^{K} \\, K^{X \\bar{X}} \\lvert D_{X} W \\rvert^{2}$ creates a minimum for the modulus which stablizes it with large mass during inflation. After inflation, when $D_{X} W$ vanishes, a ``standard'' mechanism involving nonperturbative terms in the superpotential can take over to stabilize the modulus as usual. The way we avoid the KL problem in this setup works essentially as follows: The gravitino mass $m_{3/2}$ now only sets the scale for moduli stabilization after inflation and therefore remains small all the time. During inflation, the scale for moduli stabilization is set by the inflationary energy scale $\\Hinf$ itself and no longer also by $m_{3/2}^{\\text{today}}$. This allows to consistently combine low scale SUSY and high scale inflation. There is of course a price to pay: Since the two minima for the modulus during and after inflation generically do not coincide, we have to make sure that there is no overshoot problem after inflation. This requires, for instance, that the difference between the two minima for the modulus lie not too far apart (c.f.\\ section \\ref{Sec:ExampleOvershoot}). Without a dynamical mechanism to guarantee a smooth transition between the two minima, achieving $\\sigma_{0} \\approx \\sigma_{\\inf}$ may require some amount of tuning of the model parameters. However, notice that also the KL solution requires some tuning to disentangle the height of the barrier from the gravitino mass today. Also note that, for a KKLT-type stabilisation mechanism after inflation, the mass of the modulus amounts about $m_T \\sim 16\\pi^2 \\, m_{3/2}$, which means it can be haevy enough to decay before BBN, avoiding the standard cosmological moduli problem. We have illustrated our general strategy in a simple model of chaotic inflation with a shift symmetry supplemented by a KKLT-type superpotential and uplifting term (c.f.\\ section \\ref{Sec:AnExplicitExample}). Moduli stabilization during inflation is achieved considering a rather general K\\\"ahler potential coupling between the field which provides the inflationary vacuum energy by its F-term and the modulus. We also showed that in the limit of high-scale inflation and low-energy supersymmetry breaking, i.e.~for $W, D_{T} W \\ll W_{X}$, the corrections to the inflationary observables from the modulus sector become negligible. The appendix contains results for a general moduli stabilizing superpotential $W_{\\mod}(T)$ combined with the simple chaotic inflation model. Results for hybrid (or tribrid) inflation models and models based on a Heisenberg symmetry instead of a shift symmetry will appear elsewhere \\cite{HeisenbergTribrid}. Finally, emphasize that our general strategy may work for more general scenarios. Even though many ingredients of our scenario are motivated from a string theory point of view, we did not consider a particular embedding in a string theory compactification here and defer this discussion to the future. \\subsection*{Acknowledgements} We thank P.~M.~Kostka for collaboration on the initial stage of this project. We would like to thank Alexander Westphal for several illuminating comments and discussions. S.H.~is supported by the German Science Foundation (DFG) Cluster of Excellence ``Origin and Structure of the Universe''. K.D.~is supported by the German Science Foundation (DFG) within the Collaborative Research Center 676 ``Particles, Strings and the Early Universe''. S.A.~acknowledges support by the Swiss National Science Foundation." }, "1112/1112.3057_arXiv.txt": { "abstract": "We study the impact of the spectral dependence of the linear polarization rotation induced by an achromatic half-wave plate on measurements of cosmic microwave background polarization in the presence of astrophysical foregrounds. We focus on the systematic effects induced on the measurement of inflationary gravitational waves by uncertainties in the polarization and spectral index of Galactic dust. We find that for the experimental configuration and noise levels of the balloon-borne EBEX experiment, which has three frequency bands centered at 150, 250, and 410 GHz, a crude dust subtraction process mitigates systematic effects to below detectable levels for 10\\% polarized dust and tensor to scalar ratio of as low as $r=0.01$. We also study the impact of uncertainties in the spectral response of the instrument. With a top-hat model of the spectral response for each band, characterized by band-center and band-width, and with the same crude dust subtraction process, we find that these parameters need to be determined to within 1 and 0.8~GHz at 150 GHz; 9 and 2.0~GHz at 250 GHz; and 20 and 14~GHz at 410 GHz, respectively. The approach presented in this paper is applicable to other optical elements that exhibit polarization rotation as a function of frequency. ", "introduction": "The Cosmic Microwave Background (CMB) polarization field can be decomposed into two orthogonal E and B modes. On large angular scales the B-mode signal encodes information about inflation, a period of rapid expansion in the early universe \\citep{Kamionkowski97,seljak97}. The signal is characterized by the tensor-to-scalar ratio $r$ which quantifies the relative strength of inflationary gravitational waves (IGW) and density perturbations generated by inflation. The level of the IGW signal encodes information about the energy scale at which inflation occurred. The current upper limit is $r~<~0.2$ \\citep{WMAP7yr_Komatsu}. A number of experimental efforts are ongoing to search for the signal at levels as low as $r \\sim 0.01$ over the coming years. On small angular scales, the B-mode signal is dominated by the `lensing signal', which results from gravitational lensing of the CMB photons by the large scale structure of the universe. The lensing converts E-mode to B-mode polarization \\citep{Zaldarriaga1998}. Galactic foregrounds are expected to be a source of confusion for measurements of the B-mode signal. Above 70~GHz the polarized emission from Galactic dust is predicted to dominate over much of the sky and be comparable to the IGW signal with an $r$ value of 0.1 or less even for the cleanest regions of the sky \\citep{Page2007_WMAP3yrpol, Gold2009, Fraisse2011}. Therefore many experimental efforts plan to employ multiple frequencies which will enable foreground identification and subtraction. Some CMB polarimeters use a half-wave plate (HWP) to modulate the observed linear polarization, such as EBEX \\citep{BRKspie}, SPIDER \\citep{SpiderSPIE} and POLARBEAR \\citep{PolarbearSPIE}. When observing at multiple frequency bands simultaneously, an achromatic half-wave plate (AHWP) can be used. An AHWP is a stack of monochromatic HWPs with a particular set of orientation angles relative to each other \\citep{Pancharatnam1955}. While an AHWP has a higher modulation efficiency across a broad frequency range compared to a single HWP, it rotates the polarization angle of the incident light by an amount that depends on frequency \\citep{AHWP_original,Matsumura09}. The amount of rotation depends on the construction parameters of the AHWP and on the spectrum and polarized intensity of the constituent signals, which for this paper are CMB and Galactic dust. With knowledge of the spectrum and relative polarization intensity, the amount of rotation can be calculated and corrected. However, while the spectrum of the CMB component is well known, that of dust is not. The polarized intensities of dust and CMB are also not well known. These uncertainties may pose challenges in the extraction of the underlying IGW signal. Various authors studied the impact of HWP non-idealities on measurements of CMB polarization \\citep{Brown2009, Brian2010}. However, this particular frequency dependent rotation effect has not been studied in the context of B-mode measurements. The goal of this paper is to quantitatively assess this effect. For concreteness we adopt the AHWP model, frequency bands and approximate noise information that are applicable to the E and B experiment (EBEX) \\citep{BRKspie}, a balloon-borne CMB polarimeter targeting the IGW signal at the $r \\sim 0.04$ level. In Section 2 we describe the basic components of the simulation. Section 3 focuses on quantifying the effect of rotation due to the AHWP in the 150 GHz band. In Section 4 we use multiple frequency information to account for rotation due to Galactic dust. In Section 5, we study the additional effects of uncertainties in the spectral response of the instrument, and in Section 6 we make concluding remarks. ", "conclusions": "The spectral response of an achromatic half-wave plate may induce bias in the estimation of polarized signals. We analyze the level of such bias in the context of measurements of the B-mode signal of the CMB in the presence of Galactic dust, the dominant source of foreground emission in cases of interest here. For concreteness, we use the specific experimental configuration corresponding to the EBEX balloon-borne experiment. For the area of sky considered we find that with reasonable assumptions about the magnitude and spectral shape of dust, the effects of rotation induced by the AHWP are only appreciable when dust is polarized at a level of about 5\\% and above and the tensor-to-scalar ratio $r$ is less than $\\sim0.05$. In the regime when the effects of rotation are appreciable, even a crude process of dust estimation and subtraction mitigates the effects of AHWP rotation to below detectable levels. For example, using the crude dust subtraction process we find no bias in the estimation of the B-mode power spectrum for dust polarization fraction as large as 10\\% and $r$ as low as $0.01$. For 2\\% dust polarization fraction, $r$ of 0.02 or higher is recovered without bias. We also find that the dust subtraction causes the power spectrum error bars to increase by a modest 30\\% on average. Employing the same dust estimation and subtraction process, but now assuming errors in knowledge of the experiment's detection band-center and band-width, we find the accuracy with which these need to be measured. For example, for the particular experimental configuration considered, we found that band-center and band-width of the 150 GHz band need to be determined to better than 1 and 0.8~GHz, respectively. It is possible that this requirement may not need to be as stringent if a more sophisticated foreground estimation and subtraction process is used. This research is ongoing. We explore the sensitivity of the particular experimental configuration to high frequency spectral leaks. Using a rejection level that is readily achievable experimentally we show that spectral leaks are not expected to pose challenges for the operation with an AHWP. The analysis and subtraction approach discussed in this paper are applicable to other optical elements for which polarization rotation is a function of frequency. For example, ~\\citet{OBrient_thesis} describes a broadband, mm-wave detection technique that is based on sinuous antenna. It is well documented that such antennas change the phase response of polarized signals, and that this effect is frequency dependent. Thus they exhibit fundamentally the same behavior as an AHWP. Our methods and approach apply to such cases." }, "1112/1112.1578_arXiv.txt": { "abstract": "{The daily sunspot numbers of the whole disk as well as the northern and southern hemispheres from January 1, 1945 to December 31, 2010 are used to investigate the temporal variation of the rotational cycle length through the continuous wavelet transformation analysis method. The auto-correlation function analysis of daily hemispheric sunspot numbers shows that the southern hemisphere rotates faster than the northern hemisphere. The results obtained from the wavelet transformation analysis are: there exists no direct relationship between the variation trend of the rotational cycle length and the variation trend of solar activity in the two hemispheres; the rotational cycle length of both hemispheres has no significant period appearing at the 11 years, but has significant period of about 7.6 years. Analysis concerning the solar cycle dependence of the rotational cycle length shows that in the whole disk and the northern hemisphere acceleration seems to appear before the minimum time of solar activity. Furthermore, the cross-correlation study indicates that the rotational cycle length of the two hemispheres has different phases, and the rotational cycle length of the whole disk as well as the northern and southern hemispheres also has phase shifts with the corresponding solar activity. What's more, the temporal variation of North-South (N-S) asymmetry of the rotational cycle length is also studied; it displays the same variation trend as the N-S asymmetry of solar activity in a solar cycle as well as in the considered time interval, and it has two significant periods of 7.7 and 17.5 years. Moreover, the N-S asymmetry of the rotational cycle length and the N-S asymmetry of solar activity are highly correlated. It's inferred that the northern hemisphere should rotate faster at the beginning of solar cycle 24. ", "introduction": "% \\label{sect:intro} There are two main methods used in investigating the solar rotation rate: the trace method and the spectroscopic method \\cite{Lucio}. And it is found that the Sun has a higher rotation rate in the equatorial region: 26 days at the equator while 30~days at $60^\\circ$ latitude \\cite{Lawrence2008, Le2007}. More details about different measures of the Sun's rotation rate can be found in the review papers \\cite{ Howard1984a, Schroeter1985, Snodgrass1992, Beck2000, Lucio}. Hoping to reach a more synthetic view of solar rotation, Heristchi \\& Mouradian \\cite{Heristchi2009} suggested a method called global rotation applied to structures of solar activity. Using this method, they indicated that individual structures, local proper motions, meridian drift or differential rotation could be analyzed together in the considered time. How the solar differential rotation varies in a solar cycle as well as in a long time is still an unsolved problem. Li et al. \\cite{li2011a, li2011b} used a continuous complex Morlet wavelet transformation to investigate the temporal variations of the rotational cycle length of daily sunspot areas and daily sunspot numbers from a global point of view, and indicated that the rotational cycle length of the Sun had a secular trend, and the rotational period had no relation with the Schwabe cycle. Li et al. \\cite{li2011b} pointed out that a lower than average rotation velocity should statistically appear around the maximum time of solar activity, while around the minimum time the rotation velocity was very close to the average. But Gilman \\& Howard \\cite{Gilman1984}, Zuccarello \\& Zappal\\'{a} \\cite{Zuccarello} and Braj\\v{s}a et al. \\cite{ Brajsa2006} claimed a higher than average rotation velocity appear in the minimum time of solar activity. The North-South (N-S) asymmetry in solar activity is an important part of solar physics. A lot of research has been done based on various solar activity indices on the solar surface. More details about the N-S asymmetry can be found in Vizoso \\& Ballester \\cite{Vizoso1990}, Verma \\cite{Verma1993}, Carbonell et al. \\cite{Carbonell1993, Carbonell2007}, Li et al. \\cite{li2001, li2002,li2010}, and S\\'{y}kora \\& Ryb\\'{a}k \\cite{Sykora2010}. Besides, the rotational periods are also subjected to a N-S asymmetry \\cite{Temmer2003}. Gilman \\& Howard \\cite{Gilman1984} found that in the northern hemisphere the rotation was more solid-body-like. Javaraiah \\& Ulrich \\cite{Jav2006} indicated that there existed difference in the hemispheric rotation rates. Howard et al. \\cite{Howard1984b} analyzed the large spots data and found that the rotation rate increased less in the northern hemisphere. Antonucci et al. \\cite{Antonucci1990} investigated the rotational period of the photospheric magnetic field during cycle 21 and their results showed that the two hemispheres had different dominant periods--- 26.9 days for the northern hemisphere and 28.1 days for the southern hemisphere. Also, the result of Temmer et al. \\cite{Temmer2002a, Temmer2002b} concerning the rotational periods of $H\\alpha$ flare and sunspot numbers accorded with the periods found by Antonucci et al. \\cite{Antonucci1990}. However, the observational result of Balthasar et al. \\cite{Balthasar1986} indicated that the sunspots had a little higher rotation rate in the southern hemisphere by analyzing sunspot groups of all types in the period 1874-1976. Georgieva \\& kirov \\cite{Georgieva2003} indicated that the two hemispheres not only rotated differently but aslo had different periodicities in the variations of the rotation parameters. The N-S asymmetry in hydrogen filament rotation has been studied by Gigolashvili \\cite{Gigolashvili2001} and Gigolashvili et al. \\cite{Gigolashvili2003}. They found that the sign of asymmetry changed with the hale period, and they suggested that the N-S asymmetry of the solar rotation might be connected with the N-S asymmetry of solar activity. This work follows the previous study of Li et al. \\cite{li2011a,li2011b}. We still use the continuous complex Morlet wavelet transformation to obtain the rotational signals reflected in the daily hemispheric sunspots' wavelet power spectrum from a global point of view, and then conduct further research on temporal variation of the solar rotation separately into the northern and southern hemispheres and on their relationship with the hemispheric solar activity. In addition, we investigate the N-S asymmetry of the solar rotational cycle length, including its time-variation, its periodicity, and also its relationship with the N-S asymmetry of solar activity. ", "conclusions": "\\label{sect:discussion} The long-time variations of the solar rotation rate are studied in the northern and southern hemispheres respectively through a continuous wavelet transformation method from a global point of view, and the main results are listed as follows: 1. The autocorrelation function indicates that the southern hemisphere rotates faster than the northern hemisphere in the considered time interval. Lusting \\cite{Lusting} studied the solar differential rotation by using positions of sunspots of the years from 1947 to 1981, and found that the southern hemisphere had a smaller gradient of the differential rotation than the northern hemisphere had, in other words, the southern hemisphere rotated faster than the northern hemisphere. And from the Table 1 of Javaraiah et al. \\cite{Jav2005}, we may find that the southern hemisphere indeed had a smaller gradient of differential rotation. To answer why the southern hemisphere rotates faster, one need to make further study of the reasons for the difference of the long-time variations of the gradient of differential rotation in two hemispheres. 2. The rotational cycle length of the northern and southern hemispheres has different variation trends, while solar activity in the two hemispheres has the same variation trend. It means that the long-time variation trend of hemispheric rotation rate has no direct relation with the variation trend of solar activity in the considered time. Li et al. \\cite{li2011c} pointed out that secular trends of solar rotation on an average of latitudes or at a certain latitude should change with latitudes. Thus it may be one possible reason for the different trends of the two hemispheres' rotational cycle length that the sunspots of the northern and southern hemispheres form in different average latitudes in the considered time. Further research is needed. 3. The rotational cycle length of both the hemispheres has no significant period (scale) appearing at the 11-year Schwabe cycle, in accordance with Li et al. \\cite{li2011a}, but both has significant period of about 7.6 years. And the period of about 7.6 years has been found in the periodicity of the surface equatorial rotation rate by Javaraiah et al. \\cite{Jav2009}. 4. In the whole disk and the northern hemisphere, a higher than average velocity appear before the minimum time of solar activity. This may be caused by the phase difference and periodic difference between the hemispheric rotational cycle length and the hemispheric solar activity, and may be also influenced by spatial-temporal distribution of the sunspots. The solar-cycle dependence of the two hemispheres' rotational cycle length is also different, and this may be the result of the phase shift between the northern and southern rotational cycle length, as well as the phase shift between the northern and southern solar activity. 5. The rotational cycle length of the northern and southern hemispheres shows difference in their phases. Additionally, the phase shifts between the rotational cycle length and the sunspot numbers in the north, south and the whole disk are different from one another. Since the relation between the hemispheric rotation and the hemispheric sunspot activity is complex, in-depth research is needed. 6. The N-S asymmetry of the rotational cycle length has the same variation trend as the N-S asymmetry of solar activity in a solar cycle as well as in the considered time interval. The N-S asymmetry of the rotational cycle length has two significant periods --- 7.7 and 17.5 years. Moreover, it has high correlation with the N-S asymmetry of sunspot activity. On the basis of the aforementioned characteristic and the regularity advanced by Vizoso \\& Ballester \\cite{Vizoso1990} and Li et al. \\cite{li2002}, it's inferred that the northern hemisphere should rotate faster at the beginning of solar cycle 24. \\normalem" }, "1112/1112.3327_arXiv.txt": { "abstract": "Large H\\,\\textsc{ii} regions, with angular dimensions exceeding 10 pc, usually enclose numerous massive O-stars. Stellar winds from such stars are expected to play a sizeable role in the dynamical, morphological and chemical evolution of the targeted nebula. Kinematically, stellar winds remain hardly observable i.e., the typical expansion velocities of wind-blown bubbles being often confused with other dynamical processes also regularly found H\\,\\textsc{ii} regions. However, supersonic shock waves, developed by stellar winds, should favor shock excitation and leave a well-defined spectral signature in the ionized nebular content. In this work, the presence of stellar winds, observed through shock excitation, is investigated in the brightest portions of the Galactic IC\\,1805 nebula, a giant H\\,\\textsc{ii} region encompassing at least 10 O-stars from main-sequence O9 to giant and supergiant O4. The use of the imaging Fourier transform spectrometer SpIOMM enabled the simultaneous acquisition of the spectral information associated to the H{$\\alpha$}\\,$\\lambda$6563 $\\mbox{\\AA}$, $[$N\\,\\textsc{ii}$]$\\,$\\lambda$$\\lambda$6548, 6584 $\\mbox{\\AA}$, and $[$S\\,\\textsc{ii}$]$\\,$\\lambda$$\\lambda$6716, 6731 $\\mbox{\\AA}$ ionic lines. Diagnostic diagrams, first introduced by Sabbadin and collaborators, were used to circumscribe portions of the nebula likely subject to shock excitation from other areas dominated by photoionization. The gas compression, expected from supersonic shocks, is investigated by comparing the pre- and post-shocked material's densities computed from the $\\frac{[\\textnormal{S}\\,\\textsc{ii}]\\,\\lambda6716}{[\\textnormal{S}\\,\\textsc{ii}]\\,\\lambda6731}$ line ratio. The typical $\\frac{[\\textnormal{N}\\,\\textsc{ii}]\\,\\lambda6584}{[\\textnormal{N}\\,\\textsc{ii}]\\,\\lambda6548}$ line ratio slightly exceeds the theoretical value of 3 expected in low-density regimes. To explain such behavior, a scenario based on collisional de-excitations affecting the $[$N\\,\\textsc{ii}$]$\\,$\\lambda$6548 $\\mbox{\\AA}$ line is proposed. ", "introduction": "The presence of ionized material in the interstellar medium (hereafter, ISM) can be attributed to two distinctive mechanisms. First, photoionization of the surrounding neutral gas by the strong, energetic ultra-violet (UV) flux of nearby massive stars is largely responsible for the detection of the standard hydrogen Balmer series, typically used for the morphological and kinematical description of H\\,\\textsc{ii} regions. Secondly, stellar winds with high terminal velocities and violent supernova blasts are commonly associated with the propagation of transonic and supersonic shock waves in the surrounding medium. The important increase of the post-shocked gas' temperature favors its ionization through shock excitation. Pioneering work by \\citet{Sab1977} has compared specific flux ratios for a variety of ionic transitions in the ISM optical gas. This allowed the authors to approximately separate standard H\\,\\textsc{ii} regions and planetary nebulae (hereafter, PNe) mostly governed by photoionization from shock-dominated supernova remnants (hereafter, SNRs). These diagnostic diagrams were often used, in literature, to classify large amount of ionized targets in large-scale objects, for example more-or-less distant galaxies (e.g., \\citealt{Mag2003,Rie2006}). This has led to emission-line ratio plots in which each object is usually statistically represented by a single point. Obviously, intrinsic variations, within a given object, can be investigated by targeting Galactic objects individually (e.g., \\citealt{Phi1998,Phi1999,Phi2010}). This allows to spatially resolve much smaller ($\\ll$\\,0.1 pc) structures and artifacts characterized by peculiar line ratios that are, otherwise, unrevealed or statistically negligible in observations using poorer angular resolutions. The investigation of close ionized objects has already revealed that photoionization and shock excitation can both be found in individual regions. Kinematically speaking, standard H\\,\\textsc{ii} regions encompassing massive stars earlier than O6 have shown little indication of a strong impact on the surrounding gas attributed to stellar winds, referred earlier as a potential source for shock excitation in nebulae. This remains usually true even when using observations with high spectral resolutions that allow to measure velocity fluctuations down to a few km s$^{-1}$. The main reason for this resides in the double-shock model said to accurately describe the dynamical evolution of wind-blown bubbles \\citep{Wea1977}. The reverse shock quickly (i.e., often within less than a spatial element of resolution) converts high-velocity ($>$\\,1\\,000 km s$^{-1}$) terminal winds into low-velocity, high-temperature gas. The thermal energy of this hot, pressurized material then initiates and fuels the expansion of the dense shell of post-shock ISM material found at much greater distances with respect to the central star. The typical expansion velocity of the shocked shell is usually a few km s$^{-1}$ above 10 km s$^{-1}$, roughly the speed of sound for warm ($\\sim$10\\,000 K) H$^{+}$ gas. Unfortunately, in complex tridimensional geometries, this kind of velocities could be easily confused with standard accelerated outflows in H\\,\\textsc{ii} regions, such as Champagne \\citep{Ten1979,Bod1979} or photoevaporated \\citep{Art2006,Mel2006} flows, turbulent motions \\citep{Jon1984,Ars1988,Lag2011} or fluid instabilities. This work mainly explores the possibility of detecting shock waves associated, in particular, to stellar winds using the emissive properties of the ionized gas rather than a more typical, and not always successful, approach based on the information retrieved from radial velocities and non-thermal line widths. The IC\\,1805 nebula is located in the Perseus arm of our Galaxy. The most massive stars of the Melotte 15 star cluster are currently responsible for the energetic support of the large H\\,\\textsc{ii} region. The south-central portion of the nebula (see Figure 1), in the direct vicinity of the star cluster, is gas-rich and will be used in this work to fill the \\citet{Sab1977} diagrams. Our goals are to (1) obtain reliable, high signal-to-noise ratios (hereafter, S/N), spectra of the optical emission in the brightest portions of the IC\\,1805 star-forming complex, (2) provide the corresponding series of line-ratio diagnostic diagrams, and (3) investigate the impact of photoionization and shock excitation in the targeted gas volume. We present, in $\\S$~2, the Galactic H\\,\\textsc{ii} region IC\\,1805 and its associated star cluster, Melotte 15. Information related to our spectrometric observations and methods used for data reduction are detailed in $\\S$~3. Results of our study and diagnostic diagrams are provided in $\\S$~4. Interpretation and discussion follow in $\\S$~5. Summarized results and general conclusions will finally be provided in $\\S$~6. ", "conclusions": "The use of the imaging Fourier transform spectrometer SpIOMM allowed us to obtain series of emission-line profiles of the optical gas in the brightest, central portions of the Galactic IC\\,1805 H\\,\\textsc{ii} region. The bandwidth used at data acquisition allowed the simultaneous observations of the H{$\\alpha$}\\,$\\lambda$6563 $\\mbox{\\AA}$, $[$N\\,\\textsc{ii}$]$\\,$\\lambda$$\\lambda$6548, 6584 $\\mbox{\\AA}$, and $[$S\\,\\textsc{ii}$]$\\,$\\lambda$$\\lambda$6716, 6731 $\\mbox{\\AA}$ ionic lines (see $\\S$~3). The main goal of this work was to investigate on the presence of supersonic shock waves attributed to stellar winds in the vicinity of Melotte 15, the current star cluster actually fueling up the expansion of the IC\\,1805 nebula. Literature has long suggested that a kinematical detection of stellar winds, in H\\,\\textsc{ii} region, might represent a difficult task since the typical expansion velocity of shocked shells could be commonly confused with other dynamical processes revealing similar kinematical behaviors (see $\\S$~1). On the other hand, specific line ratios retrieved from the optical gas may indicate if the presence of ionized material is attributed to standard photoionization or to shock excitation. These line ratios therefore provide a non-kinematical tool for identifying shocks in the nebular volume. Our results are summarized as follow: \\begin{enumerate} \\renewcommand{\\labelenumi}{\\Roman{enumi}.} \\item The $\\frac{[\\textnormal{N}\\,\\textsc{ii}]\\,\\lambda6584}{[\\textnormal{N}\\,\\textsc{ii}]\\,\\lambda6548}$ line ratio genuinely deviates from the theoretical value of 3 (see $\\S$~5.1). Values varying between 2.5 and 4 were commonly found. The distribution of the density measurements has not allowed to demonstrate that the $[$N\\,\\textsc{ii}$]$\\,$\\lambda$6548 $\\mbox{\\AA}$ line could be affected by collisional de-excitations, hence reducing its peak intensity. This scenario remains nonetheless plausible and could be verified if densities, in the N$^{+}$ volume, could be measured precisely (see $\\S$~5.1). Densities extracted from the $\\frac{[\\textnormal{S}\\,\\textsc{ii}]\\,\\lambda6716}{[\\textnormal{S}\\,\\textsc{ii}]\\,\\lambda6731}$ ratio may not perfectly reflect the conditions prevailing in the N$^{+}$ volume if both nitrogen and sulfur are not co-spatial in central IC\\,1805. \\item The diagnostic diagrams, introduced by \\citet{Sab1977}, indicate, in first approximation, that photoionization most likely dominate in IC\\,1805 (see $\\S$~4.2.5). This initially holds even for the densest, most emissive structures in our field-of-view, directly exposed to the radiation fields and stellar winds of the nearby, most massive stars (see $\\S$~5.2). \\item Evidences for shock excitation appear only following the subtraction of the diffuse foreground/background material. This component likely occupies a very large fraction of the nebular volume in IC\\,1805; a volume sufficiently large that, even though being tenuous in nature, the foreground/background material strongly dilutes the signal emanating from shocked condensations along the line-of-sight. Its mean density is estimated at 25 cm$^{-3}$ (see $\\S$~5.2.1). \\item Oriented on a south-north axis and surrounded by numerous O-stars, a bright, large ionized feature occupies the central area of our field-of-view. The last, tenuous fragments of an old molecular cloud can be found near its southern portion. Shocks may have contributed to ionize material found in the direct vicinity of the molecular clump while the northern parts of the ionized feature, deprived of molecular emission, appears to be largely dominated by photoionization (see $\\S$~5.2.1). \\item The shock-excited ionized gas has a mean density of 175 cm$^{-3}$ and the compression factor, between pre- and post-shocked gas, is typically between 2 and 10. For isothermal shocks, this suggests shock velocities between 15 and 30 km s$^{-1}$, in agreement with models describing the expansion of wind-blown bubbles (see $\\S$~5.2.1). Geometrically speaking, given the apparent proximity between the central, ionized structure and the most massive stars of Melotte 15, winds seem to have had sufficient time to reach the structure within a timescale corresponding to the age of the star cluster (see $\\S$~5.2.1). This gives credence to our assumption that shock excitation can be found in the south-central portions of our field-of-view. \\item Points identified with a high probability of shock excitation reveal compression factors typically greater than those points where shocks may be present although less certain (see $\\S$~5.2.1). \\item Shocks did not seem to have played a major role in the ionization of a molecular cloud's envelope located in the south-eastern portion of our field-of-view. This most likely results from the molecular fragment being located too far from the ionizing sources so that (1) shocks induced by stellar winds have not reached yet the cloud or (2) shocks have reached the cloud with velocities too low to initiate shock excitation (see $\\S$~5.2.2.1). \\item Shock development was clearly detected on the outskirts of an apparently weak, but well-defined elephant trunk located in the eastern portion of our field-of-view (see $\\S$~5.2.2.2). This is in agreement with theoretical works developed on such feature typically found in H\\,\\textsc{ii} regions. \\end{enumerate}" }, "1112/1112.2926_arXiv.txt": { "abstract": "We present a mechanism for the dark matter stability in the framework of a non-Abelian flavour symmetry renormalizable model. The same non-abelian discrete flavor symmetry which accounts for the observed pattern of neutrino oscillations, spontaneously breaks to a $Z_2$ subgroup which renders DM stable. The simplest scheme leads to a scalar doublet DM potentially detectable in nuclear recoil experiments, inverse neutrino mass hierarchy, hence a neutrinoless double beta decay rate accessible to upcoming searches. ", "introduction": "Non-baryonic Dark Matter (DM) is one of the most compelling problems of modern cosmology. Despite the fact that its existence is well established by cosmological and astrophysical probes, its nature remains elusive. Still, observations can constraint the properties of dark matter and give some hints about its identity. For instance, a fundamental requirement for a viable dark matter candidate is the stability over cosmological times. This suggests the existence of an exact or slightly-broken symmetry protecting or suppresing its decay. It has been shown recently that such symmetry can also be related to the flavor structure of the Standard Model~\\cite{arXiv:1007.0871}. The model proposed in ~\\cite{arXiv:1007.0871} is based on a $A_4$ symmetry with four $SU(2)$ Higgs doublets. After the electroweak symmetry breaking, the $A_4$ (even permutation of four objects) group is spontaneously broken into a $Z_2$ subgroup which is responsible for the DM stability. The leptonic sector is also extended. It consist in four right handed neutrinos and the light neutrino masses are generated through the type-I seesaw mechanism and obey an inverted hierarchy mass spectrum with $m_{\\nu_3}=0$ and vanishing reactor angle $\\theta_{13}=0$. We have also been consider models with a different matter content for the right handed neutrinos with the same DM stability mechanism but with different neutrino phenomenology~\\cite{arXiv:1011.1371} or a model based on the dihedral group $D_4$ where the some flavour changing neutral currents are present and constraint the DM sector~\\cite{arXiv:1104.0178}. Other models with flavor symmetries but with decaying DM have also been considered, see for instance~\\cite{arXiv:1010.4963,arXiv:1011.5753}. For a model with stable DM but with the same behavior of the DM, in the sense that instead of being inert totally, the DM couples to some right handed neutrinos in a similar way in our model\\footnote{This afther the $A_4$ is broken into the $Z_2$ symmetry.} see~\\cite{arXiv:1111.4938}. ", "conclusions": "We have studied a model where the stability of the dark matter particle arises from a flavor symmetry. The $A_4$ non-abelian discrete group accounts for both the observed pattern of neutrino mixing and for DM stability. We have analyzed the constraints that follow from electroweak precision tests, collider searches and perturbativity. We have also analyzed the prospects for direct and indirect dark matter detection and found that, although the former already excludes a large region in parameter space, we cannot constrain the mass of the DM candidate. In contrast, indirect DM detection is not yet sensitive enough to probe our predictions. However, forecasted sensitivities indicate that Fermi-LAT should start probing them in the near future. All of the above relies mainly on the properties of the scalar sector responsible for the breaking of the gauge and flavor symmetries. The motivation of our approach is to link the origin of dark matter to the origin of neutrino mass and the understanding of the pattern of neutrino mixing, two of the most outstanding challenges in particle physics today. At this level one may ask what are the possible tests of this idea in the neutrino sector. We found an inverted neutrino mass hierarchy, hence a neutrinoless double beta decay rate accessible to upcoming searches, while $\\theta_{13}=0$ giving no CP violation in neutrino oscillations. Note however that the connection of dark matter to neutrino properties depends strongly on how the symmetry breaking sector couples to the leptons." }, "1112/1112.0438_arXiv.txt": { "abstract": "We review basic properties of the population of dwarf galaxies in the Local Group focusing on dwarf spheroidal galaxies found in the immediate vicinity of the Milky Way. The evidence for dark matter in these objects is critically assessed. We describe the methods of dynamical modelling of such objects, using a few examples of the best-studied dwarfs and discuss the sources of uncertainties in mass estimates. We conclude with perspectives for dwarf galaxies as targets for dark matter detection experiments. ", "introduction": "The population of dwarf galaxies in the Local Group offers a unique opportunity to test our theories of structure formation in the Universe. Starting from the Magellanic Clouds which were known since antiquity, the census of the dwarf galaxies in our immediate vicinity still grows. The sample of dwarf galaxies in the Local Group can be divided using the morphological criteria into classes of dwarf irregulars (dIrr) that are flattened, rotating, bright and still forming stars and dwarf spheroidals (dSph) which are rounder, faint and contain mostly old stellar populations dominated by random motions. Dwarf galaxies of the Local Group tend to cluster around the two main hosts, the Milky Way and Andromeda, exhibiting a pronounced morphology-density relation: while dSphs are typically found close to one of the big galaxies, dIrrs occupy more isolated regions. The origin of this relation is an interesting issue that the theories of structure formation must address. Another question in which dwarf galaxies played a role is the so-called problem of missing satellites: the number of observed satellites of the Milky Way is much smaller than predicted by theories based on cold dark matter. Dwarf galaxies, especially nearby dSphs, have also drawn attention once a significant number of stellar velocities could be measured and their masses could be determined. These masses turned out to be much larger than could be explained by the stellar content indicating large amounts of dark matter present. \\begin{figure} \\epsfig{file=lokas_fig1.eps, height=10cm} \\caption{Observational parameters of dwarf galaxies in the Local Group: the central surface brightness $\\mu_V$ (upper left panel), the half-light radius $r_{1/2}$ (lower left panel), the ratio of the rotation velocity to the velocity dispersion $V/\\sigma$ (upper right panel) and the ellipticity $e=1-b/a$ (lower right panel) as a function of the total visual magnitude $M_V$. The open circles show the data for dIrr galaxies, filled black circles the data for the dSph and dSph/dE galaxies and gray circles for transitory dIrr/dSph dwarfs.} \\label{lokas_fig1} \\end{figure} Observational properties of dwarf galaxies \\cite{mateo98} are usually characterized by the total visual magnitude $M_V$, the central surface brightness $\\mu_V$, the half-light radius (containing half the total luminosity) $r_{1/2}$, the ellipticity $e=1-b/a$ and the ratio of the rotation velocity to the velocity dispersion $V/\\sigma$. The latter four quantities are plotted in Figure~\\ref{lokas_fig1} as a function of magnitude for the dwarf galaxies in the Local Group with $-16$ mag $< M_V < -8$ mag using data compiled in \\cite{lokas11}. The sample includes the eight ``classical\" best-known dSph galaxies, satellites of the Milky Way: Carina, Draco, Fornax, Leo I, Leo II, Sculptor, Sextans and Ursa Minor with magnitudes in the range $-13$ mag $< M_V < -8$ mag which all have low central surface brightness $\\mu_V$ ($>22.4$ mag arcsec$^{-2}$) and small half-light radii ($< 0.7$ kpc). The properties of these eight dwarfs are listed in Table~\\ref{properties}. All classical dwarfs are believed to be characterized by large mass-to-light ratios between 10 and a few hundred solar units. \\begin{table} \\begin{center} \\caption{Properties of the classical dSph galaxies, satellites of the Milky Way.} \\begin{tabular}{lrcccc} \\hline Dwarf\t & $M_V \\ $ & $r_{1/2}$ & $\\mu_V$ & $V/\\sigma$ & $e=1-b/a$ \\\\ galaxy & [mag] & [kpc] & [mag arcsec$^{-2}$] & & \\\\ \\hline Carina &$\t-8.62 $ & 0.241 & 25.5\t& 0.43\t&\t0.33 \\\\ Draco &$\t-8.74 $ & 0.196 & 25.3 & 0.21\t&\t0.29 \\\\ Fornax &$\t-13.03$ & 0.668 & 23.4 & 0.18\t&\t0.31 \\\\ Leo I &$\t-11.49$ & 0.246 & 22.4 & 0.33 & 0.21 \\\\ Leo II &$\t-9.60 $ & 0.151 & 24.0\t& 0.28\t&\t0.13 \\\\ Sculptor &$\t-10.53$ & 0.260 & 23.7 & 0.30\t&\t0.32 \\\\ Sextans &$\t-9.20 $ & 0.682 & 26.2 & 0.48\t&\t0.35 \\\\ Ursa Minor &$\t-8.42 $ & 0.280 & 25.5\t& 0.49\t&\t0.56 \\\\ \\hline \\label{properties} \\end{tabular} \\end{center} \\end{table} In recent years we have witnessed discoveries of new dSph galaxies in the Local Group, mainly in the northern part of the Sloan Digital Sky Survey \\cite{belokurov07}. These new dwarfs are generally fainter and more irregular in shape than the classical ones but not more distant. If analogous discoveries follow in the southern hemisphere, they may significantly help to solve the problem of missing satellites. Spectroscopic studies of the stellar populations \\cite{simongeha} of the faint dwarfs revealed quite large velocity dispersions and thus large masses for these dwarfs were estimated suggesting even higher dark matter content than in well-known dSphs. It has been speculated \\cite{strigari} that the characteristic masses of all dwarfs are of the order of $10^7$ solar masses, independently of their luminosity. A more acceptable proposal \\cite{walker} suggests instead a universal relation between the half-light radius and the mass contained within it. ", "conclusions": "The case for a high dark matter content in dSph galaxies of the Local Group is quite strong. The objects are found to contain significant amounts of dark matter even if the modelling is performed with very conservative assumptions such as the one that mass follows light and on samples which were subject to restrictive algorithms of interloper rejection. It has been demonstrated that the classical dwarfs, even if strongly affected by tidal force from the Milky Way can be reliably modelled by standard methods assuming virial equilibrium. The high mass-to-light ratios of the classical dwarfs may turn out to be even higher in the case of the newly discovered population of ultra-faint satellites of the Milky Way. If the trend is confirmed, the fainter dwarfs may become excellent targets for experiments aiming for the direct detection of dark matter in the Universe via self-annihilation of dark matter particles. Dwarf galaxies are good candidates for such targets not only because of the high dark matter content but also because they have low astrophysical gamma ray backgrounds and they are small and localized so may easily fit in the field of view of many instruments. Recently, observations of dwarf galaxies with the Fermi Large Area Telescope began to yield interesting constraints on the parameters of a variety of supersymmetric models that provide candidates for weakly interacting massive particles as dark matter \\cite{abdo10}. Although no clear signal of annihilation has been detected from dwarf galaxies so far, upper limits on the annihilation cross-section can be obtained. The upper limits do not yet reach the interesting region of density parameter values compatible with the standard cosmological model, but are already able to exclude a significant number of supersymmetric models of particle physics." }, "1112/1112.0324_arXiv.txt": { "abstract": "We present our \\emph{XMM-Newton} observation of the fastest rotating spiral galaxy UGC~12591. We detect hot gas halo emission out to 110~kpc from the galaxy center, and constrain the halo gas mass to be smaller than $3.5\\times10^{11} \\msun$. We also measure the temperature of the hot gas as $T=0.64\\pm0.03$~keV. Combining our X-ray constraints and the near-infrared and radio measurements in the literature, we find a baryon mass fraction of 0.03--0.04 in UGC~12591, suggesting a missing baryon mass of 75\\% compared with the cosmological mean value. Combined with another recent measurement in NGC~1961, the result strongly argues that the majority of missing baryons in spiral galaxies does not reside in their hot halos. We also find that UGC~12591 lies significantly below the baryonic Tully-Fisher relationship. Finally, we find that the baryon fractions of massive spiral galaxies are similar to those of galaxy groups with similar masses, indicating that the baryon loss is ultimately controlled by the gravitational potential well. The cooling radius of this gas halo is small, similar to NGC~1961, which argues that the majority of stellar mass of this galaxy is not assembled as a result of cooling of this gas halo. ", "introduction": "Observations show that nearby galaxies are missing most of their baryons (e.g., Hoekstra et al.\\ 2005; Heymans et al.\\ 2006; Mandelbaum et al.\\ 2006; Gavazzi et al.\\ 2007; Jiang \\& Kochanek 2007; Bregman 2007) when compared to the cosmological baryon to matter ratio (e.g., $f_b = 0.171\\pm0.009$ from WMAP, Dunkley et al.\\ 2009). For example, the Milky Way is missing two thirds of its baryon allotment (e.g., Sakamoto et al.\\ 2003) and less massive galaxies have retained less than 10\\% of their baryons (e.g., Corbelli 2003; Walker et al.\\ 2007). This situation is confirmed in other galaxies through a variety of methods (e.g., Hoekstra et al. 2005; McGaugh 2007). However, most of these studies do not include the baryon mass in the hot gas halo of galaxies, and it is possible that the majority of the missing baryons in galaxies actually resides in their hot has halos based on theoretical predictions (e.g., White \\& Frenk 1991; Sommer-Larsen 2006; Fukugita \\& Peebles 2006; also see the review by Benson 2010). This component can be difficult to detect for spiral galaxies due to its faintness, especially if the gas density profile is flat. As another possibility, the missing baryons from galaxies may have escaped from the potential wells of the galaxies but reside in their parent groups or clusters (e.g., Humphrey et al.\\ 2011). Finally, the missing baryons can be in the form of warm-hot intergalactic medium (Cen \\& Okstiker 1999; 2006). Deep X-ray observations are needed to distinguish these different scenarios. The situation is different for rich galaxy clusters, where the gas mass dominates the baryon content, and the baryon fractions in these massive systems are close to the cosmological value after combining the gas and stellar baryon contributions (e.g., Vikhlinin et al.\\ 2006; Allen et al.\\ 2008). The measurements of the baryon fraction in different systems suggest that the fraction depends on the dynamical mass of the systems: rich clusters retain their cosmological allotment of baryons, while galaxies are baryon-poor. We summarize the situation in Dai et al.\\ (2010) by combining the archival data points reported in the literature and our stacking analysis result using the ROSAT All-Sky Survey data of 4,000 nearby galaxy groups and clusters (Dai et al.\\ 2007). We find that the baryon fractions from dwarf galaxies to rich galaxy clusters can be fit by a broken power-law model with the break at the circular velocity of $V_c \\sim 440~\\kms$. The scatter of the fractions about the mean relation is small considering the huge dynamic range of the systems. Further examining the relation, we find that the baryon fractions are similar for different systems with similar total masses but different compositions. For example, the baryon fractions of poor galaxy groups, where the baryon mass is still dominated by the gas mass, are close to those of massive galaxies, where the baryon mass is dominated by the stellar mass. Such a coincidence is puzzling considering the differences between their mass compositions and energy feedback mechanisms. To test whether the missing baryons reside in galaxy halos and further constrain the baryon loss in different mass scales, we focus on the massive galaxy \\ugc. \\ugc\\ is a spiral galaxy with the largest measured rotational velocity to date (466-500 km/sec, Giovanelli et al. 1986; Paturel 2003). To appreciate this galaxy, its optical-IR luminosity is nine times that of M31. Nine big spirals were in a single group is richer than Hickson groups and is about half of the Fornax cluster, but much more compact. In this paper, we combine our \\xmm\\ observation of \\ugc\\ with 2MASS and other data to determine the baryon fraction and the composition of this massive spiral galaxy. Throughout the paper, we use the cosmological parameters from WMAP with $H_0 = 72~\\rm{km~s^{-1}~Mpc^{-1}}$, $\\Omega_{\\rm m} = 0.26$, and $\\Omega_{\\Lambda}= 0.74$ (Dunkley et al.\\ 2009). ", "conclusions": "\\begin{deluxetable}{lccccccccc} \\tabletypesize{\\scriptsize} \\tablecolumns{10} \\tablewidth{0pt} \\tablecaption{Gravitational and Baryon Mass Components in \\ugc \\label{tab:mass}} \\tablehead{ \\colhead{Component} & \\colhead{within 50~kpc radius} & \\colhead{within 500~kpc radius} } \\startdata Stellar & $45\\pm10$ & $45\\pm10$ \\\\ Cold Gas & 0.7 & 0.7 \\\\ Hot Gas & $0.41\\pm0.03$ & 13 \\\\ Including Flattened Hot Gas & 0.55 & $\\ls 35$ \\\\ Gravitational & 270 & 1900 \\\\ Stellar-to-gas ratio $r_{sg}$ & 39 & 3.3 \\\\ $r_{sg}$ including the flattened gas & 33 & 3.3--1.3 \\\\ Baryon Fraction $f_b$ & 0.17 & 0.03 \\\\ $f_b$ including the flattened gas & 0.17 & 0.03--0.04 \\\\ \\enddata \\tablecomments{The masses are in the unit of $10^{10} \\msun$. The \\rtwo\\ is at $\\simeq 550$~kpc.} \\end{deluxetable} \\subsection{Baryon Mass Components in \\ugc} We list the various baryon mass components in \\ugc\\ in Table~\\ref{tab:mass}. We have constrained the hot gas mass of \\ugc\\ using the \\xmm\\ observation as $(4.1\\pm0.3)\\times10^{9} \\msun$ within 50~kpc with an average temperature of $T=0.64\\pm0.03$~keV. We have also constrained the hot gas mass of $1.3\\times10^{11} \\msun$ within 500~kpc regions using our best fit $\\beta$ model, and the hot gas mass is below $3.5\\times10^{11} \\msun$ within 500~kpc even if we add another flatter $\\beta$ model component in our fits. Beside the hot gas mass, there are other baryon mass components in the galaxy including the stellar mass and cold gas mass components. For the cold gas mass component, \\citet{gi86} measure the \\HI\\ mass of $5.3\\times10^{9} h_{72}^{-2} \\msun$ from the radio data. Assuming the \\HI\\ mass is 75\\% of the total cold gas mass, we find that the total cold gas mass is $M_{cg} = 7.1\\times10^9 \\msun$, where we ignore the contribution from the molecular gas component. We estimate the stellar mass of \\ugc\\ as $(4.5\\pm1.0)\\times10^{11} \\msun$ within 29~kpc radius using its $K$ band total magnitude ($K = 8.89$~mag) from the 2MASS Extended Source Catalog and a range of mass-to-light ratio from 0.6 to 0.95 (e.g., Bell et al.\\ 2003). The 2MASS team calculates the total magnitude by integrating the surface brightness profile out to $\\sim 4$ disk scale lengths from the isophotal aperture well below the $1\\sigma$ noise level. For the mass-to-light ratio, Bell et al.\\ (2003) measure a value of 0.95 as the cosmic mean value. However, since \\ugc\\ is a late-type galaxy and could have a lower mass-to-light ratio, we choose to use $0.78\\pm0.18$ in our calculation. We find the largest uncertainties in the baryon mass are from the systematical uncertainty in the stellar mass-to-light ratio and the flattened gas halo. Combining the two effects, we find an uncertainty of $2.4\\times10^{11} \\msun$ for the total baryon mass within 500~kpc. In the central region within $\\sim50$~kpc, the baryon mass is clearly dominated by the stellar mass, and the stellar-to-gas mass ratio is $r_{sg} \\simeq 39$. Using the rotational velocity of 466--500~\\kms, we measure a total mass of $m_{tot} = 2.7\\times10^{12} \\msun$ within 50~kpc and a baryon mass fraction of $f_b \\simeq 0.17$, consistent with the cosmological baryon fraction. Out to the 500~kpc region (\\rtwo $\\simeq 550$~kpc), we use the gravitational mass $m_{tot} = 1.9\\times10^{13} \\msun$ estimated from the X-ray data, because the rotational curve is only constrained within 28$h_{72}^{-1}$~kpc \\citep{gi86}, smaller than the total mass $m_{tot} = 2.7\\times10^{13} \\msun$ estimated using the rotational curve. The baryon mass within 500~kpc is $m_b = 5.9\\times10^{11} \\msun$, and we measure a baryon fraction of $f_b \\simeq 0.03$. Considering a second flattened gas component, the baryon fraction within 500~kpc can reach to $f_b \\ls 0.04$. Since we use the smaller total mass estimate in the calculation, the baryon fraction quoted should be treated as a conservative upper limit. The stellar mass component is still more important with $r_{sg} = 3.3$ within 500~kpc, or $r_{sg} \\gs 1.3$ with the additional flatter gas component. To summarize, combining our \\xmm\\ observation and the 2MASS and radio data in the literature, we have constrained that \\ugc\\ has lost at least 75\\% of the baryons compared to the cosmological value. The missing baryons do not reside in the hot halos for spiral galaxies. Our result confirms the recent measurements in another giant spiral NGC~1961 using \\chandra\\ by Anderson \\& Bregman (2011), who find that NGC~1961 has also lost 75\\% of its baryon content. \\subsection{Baryonic Tully-Fisher Relationship} \\begin{figure} \\epsscale{1} \\plotone{f9.eps} \\caption{The baryonic Tully-Fisher relationship of McGaugh (2005) adding NGC~1961 (Anderson \\& Bregman 2011) and \\ugc. \\label{fig:btf}} \\end{figure} Using the fastest rotating galaxy \\ugc, we are able to extend the baryonic Tully-Fisher relationship (BTF), a correlation between the baryon mass and rotational velocity of galaxies (McGaugh 2005; 2011) to the high rotational velocity regime of 500~\\kms. We plot \\ugc\\ in the BTF diagram together with the galaxies in McGaugh (2005) and the other massive galaxy NGC1961 (Anderson \\& Bregman 2011) with a rotational velocity of 402~\\kms\\ in Figure~\\ref{fig:btf}. Anderson \\& Bregman (2011) find that NGC1961 deviates slightly from the linear fits to BTF. However, the authors caution that the offset can be caused by systematic uncertainties. Indeed, assuming a $K$ band mass-to-light of 0.95, NGC~1961 would be on the BTF relation. Here, \\ugc\\ provides another challenge to the linear BTF relation, which predicts a baryon mass of $2.2\\times10^{12} \\msun$ for \\ugc, whereas we measure a baryon mass in the range of 5.9--8.1$\\times10^{11} \\msun$ with an uncertainty of $2.4\\times10^{11} \\msun$. Thus, \\ugc\\ is $6\\sigma$ below the BTF relation. If the offset from the BTF relation is caused by the uncertainties in the $K$ band mass-to-light ratio, a ratio of 3.1 is needed to put \\ugc\\ on the BTF relation, which is extremely unlikely (e.g., Bell et al.\\ 2003). Thus, it is possible that the BTF relation turns over for massive galaxies with $v_c \\gs 400 \\kms$. However, measurements from a larger sample of massive spiral galaxies are needed to confirm this result. \\subsection{Overall Relationship of Baryon Fractions with Total Mass} \\begin{figure} \\epsscale{1} \\plotone{f10.eps} \\caption{The baryon fraction as a function of the gravitational potential well indicated by the circular velocity at \\rtwo. We plot the new measurements from massive spiral galaxies \\ugc\\ and NGC~1961 (Anderson \\& Bregman 2011) over the archival data from Sakamoto et al.\\ (2003), McGaugh (2005), Flynn et al.\\ (2006), Vikhlinin et al.\\ (2006), Gavazzi et al.\\ (2007), Walker et al.\\ (2007), Stark et al.\\ (2009), Sun et al.\\ (2009), Dai et al.\\ (2010). The dotted line is the cosmological baryon fraction measured from CMB, and the dashed line is our best fit broken power law model for baryon losses. For massive spiral galaxies like \\ugc\\ and NGC~1961, the baryon loss is at least 75\\%. \\label{fig:bf}} \\end{figure} We plot the baryon fractions of \\ugc\\ and NGC~1961 against their rotational velocities, a proxy for the depth of the gravitational potential well, in Figure~\\ref{fig:bf}. We include the data composed by Dai et al.\\ (2010) including the stacked results from optically-selected clusters (Dai et al.\\ 2007; 2010), individual galaxy clusters (Vikhlinin et al.\\ 2006; Sun et al.\\ 2008) corrected for the stellar component (Dai et al.\\ 2010), individual galaxies (McGaugh 2005; Walker et al.\\ 2007; Stark et al.\\ 2009), elliptical galaxies (Gavazzi et al.\\ 2007) and the Milky Way (Sakamoto et al.\\ 2003; Flynn et al.\\ 2006), and the recent addition of gas-rich late-type galaxies (Begun et al.\\ 2008; Trachternach et al.\\ 2009) composed by McGaugh (2011). This enables us to compare the baryon loss across a large range of systems from dwarf galaxies to rich galaxy clusters. For massive spiral galaxies such as \\ugc\\ and NGC~1961, their dark matter halo masses are close to that of a medium galaxy group, and we find that their baryon losses are comparable to the stacked results for galaxy groups in Figure~\\ref{fig:bf}. For different systems as spiral galaxies and galaxy groups, the consistency between their baryon fractions suggests that the baryon loss is ultimately controlled by the potential well of the dark matter halo. However, currently we still lack constraints from individual galaxy groups to confirm the stacking results of Dai et al.\\ (2010). The overall baryon fractions for all systems can be fit by a broken power law model (Dai et al.\\ 2010). With the addition of new measurements, especially from those gas-rich late-type galaxies (Begun et al.\\ 2008; Trachternach et al.\\ 2009), we can better fit the power law slope for baryon fractions in less massive systems. Thus, we re-fit the data with a broken power law model to find that \\begin{equation} f_{b} = \\frac{0.16 (v_c/643~{\\rm km/s})^{a}}{{(1+(v_c/643~{\\rm km/s})^{c})}^{b/c}}, \\end{equation} where $a=1.26$, $b=1.24$, and $c=2$. The baryon fraction, $f_{b}$, scales as $f_b\\propto v_c^{a-b=0.02}$ above the break and \u00a0$f_{b} \\propto v_c^{a=1.26}$ below the break, and the parameter $c$ in the equation is the smoothness of the broken power law model, which is fixed in our fit. Comparing with the fit in Dai et al.\\ (2010), we find the major difference lies in the shallower slope $f_b \\propto v_c^{1.26}$ for the baryon loss in galaxies. We also find a larger break location and a shallower slope for galaxy clusters. \\subsection{Cooling of the Gas Halo} The gas halo is predicted to play an important role in galaxy formation and evolution. With our detection of the gas halo emission in \\ugc, we can estimate the cooling time of this hot halo and the implied accretion rate onto the galaxy, which can provide constraints on the gas available for new star formation. We define the cooling radius as the radius where the cooling time is 10~Gyr, using the expression of the cooling time (Fukugita \\& Peebles 2006), \\begin{equation} \\tau(r) = \\frac{1.5 nkT}{\\Lambda n_e \\left(n-n_e\\right)} \\approx \\frac{1.5kT\\times1.92}{\\Lambda n_e \\times 0.92}, \\end{equation} where the latter expression assumes a primeval He abundance resulting in a total particle density of $n = 1.92 n_e$. For $T = 10^{6.85}$~K, $Z/ Z_{\\odot} = 0.5$, and $\\Lambda = 10^{-22.85}$ erg cm$^{3}$ s$^{-1}$ (Sutherland \\& Dopita 1993), the cooling radius is at $n_e = 6.8\\times10^{-4}$ cm$^{-3}$. For the range of best-fit $\\beta$-model profiles constrained in this paper, this corresponds to a cooling radius between 15.6 and 18.0 kpc, and a hot halo mass of $6.2-9.2\\times10^8 M_{\\odot}$ within that radius. We can roughly estimate the cooling time and rate by dividing the thermal energy in the hot gas within the cooling radius by the luminosity within that radius, and this yields a wide range in cooling time of 2.8-6.3 Gyr for material within the cooling radius, but a fairly narrow range in the effective cooling rate of $0.15-0.21 M_{\\odot}$ year$^{-1}$. This halo accretion rate is two orders of magnitude too low to assemble the stellar mass of this galaxy within a Hubble time. Therefore, significant accretion must have occurred via some other mode, such as cold flows or mergers, to produce the stellar mass seen in this galaxy today, confirming the conclusion drew in NGC~1961 (Anderson \\& Bregman 2011)." }, "1112/1112.5447_arXiv.txt": { "abstract": "We discuss the combined constraints on a Finite Scale Factor Singularity (FSF) universe evolution scenario, which come from the shift parameter $\\mathcal{R}$, baryon acoustic oscillations (BAO) $\\mathcal{A}$, and from the type Ia supernovae. We show that observations allow existence of such singularities in the $2\\times10^9$ years in future (at $1\\sigma$ CL) which is much farther than a Sudden Future Singularity (SFS), and that at the present moment of the cosmic evolution, one cannot differentiate between cosmological scenario which allow finite scale factor singularities and the standard {\\bf $\\Lambda$CDM} dark energy models. We also show that there is an allowed value of $m=2/3$ within $1\\sigma$ CL, which corresponds to a dust-filled Einstein-de-Sitter universe limit of the early time evolution and so it is pasted ito a standard early-time scenario. ", "introduction": "\\setcounter{equation}{0} Finite scale factor singularities (FSF) are one of types of new and exotic singularities which were found first in Ref. \\cite{BGT}. The inspiration to search these new types of singularities was due to the observations of high-redshift type Ia supernovae (SNIa) which provided strong evidence that the expansion of the universe is accelerating due to an unknown form of dark energy\\cite{supernovaeold} phenomenologically behaving as the cosmological constant. Further observational data \\cite{supernovaenew} made cosmologists think of more accelerating universe filled with phantom \\cite{phantom} which violated all energy conditions: the null ($\\varrho c^2 + p \\geq 0$), weak ($\\varrho c^2 \\geq 0$ and $\\varrho c^2 + p \\geq 0$), strong ($\\varrho c^2 + p \\geq 0$ and $\\varrho c^2 + 3p \\geq 0$), and dominant energy ($\\varrho c^2 \\geq 0$, $-\\varrho c^2 \\leq p \\leq \\varrho c^2$) ($c$ is the speed of light, $\\varrho$ is the mass density in $kg m^{-3}$ and $p$ is the pressure). Phantom-driven dark energy leads to a big-rip singularity (BR or type I according to \\cite{nojiri}) in which the infinite values of the energy density and pressure ($\\rho$, $p\\to\\infty$) are accompanied by the infinite value of the scale factor ($a\\to\\infty$) \\cite{caldwellPRL}. The list of new types of singularities contains: a big-rip (BR), a sudden future singularity (SFS) \\cite{barrow04}, which can appear in inhomogeneous and anisotropic models too \\cite{barrow042, sfs1}, a generalized sudden future singularity (GSFS), a finite scale factor singularity (FSF) \\cite{aps}, a big-separation singularity (BS). They are characterized by violation of all or some of the energy conditions which results in a blow-up of all or some of the physical quantities: the scale factor, the energy density and the pressure. The finite scale factor singularity, which is the subject of this paper, is a weak singularity according to Tipler's definition, but strong according to Kr\\'{o}lak's definition \\cite{lazkoz}. Apart from mentioned above there are also $w$-singularities \\cite{wsin}, which are not physical singularities, but are singularities of a barotropic index $w$ which are present in different cosmological models of $f(R)$ gravity \\cite{star1980}, in scalar field models \\cite{Setare}, and in brane cosmologies \\cite{Sahni}. For a finite scale factor singularity (FSF) $\\rho \\to \\infty$ and $p \\to \\infty $ diverge, while the scale factor $a$ remains constant. This means that it is similar to a big-bang singularity with only one exception - the scale factor is a constant $a = a_s=$ const instead of zero $a_{BB} = 0$. Besides, FSF singularities are stronger than a sudden future singularities (SFS) [10], so they placed themselves in between the big-bang (which is strong and geodesically incomplete) and the SFS. It has been shown that SFS appear in physical theories such as $f(R)$ gravity, scalar field cosmologies, brane cosmologies \\cite{yuri} and, in particular, they plague loop quantum cosmology \\cite{LQC}. It seems then that after an appropriate choice of the scalar field, $f(R)$ function or brane, FSF will be easily deduced to appear in such physical theories as well. \\indent The paper is organized as follows. In section \\ref{s2} we present FSF scenario. In section \\ref{oc} we derive the expressions for the observables: baryon acoustic oscillations, distance to the last scattering surface, type Ia supernovae, used to test an FSF scenario. In section \\ref{rac} we give the results and discussion. ", "conclusions": "\\label{rac} \\setcounter{equation}{0} Overall $\\chi^2$ that has been used is \\be \\chi^2=\\chi^2_{SN}+\\chi^2_{R}+\\chi^2_{A} \\ee In Fig. \\ref{fig1} we present contour plots showing the joint marginal posterior distribution for each pair of FSF model parameters. Each sub-panel shows three contours, denoting roughly 68\\%, 95\\% and 99\\% (from light gray to dark grey respectively) credible regions. In Fig. \\ref{fig1} we see that there are two disjoined regions in the parameter space for which FSF singularities are allowed in the universe. $\\delta$ is always positive which is not surprising since only for positive $\\delta$ there can be an accelerated expansion in the investigated model. Characteristic feature of the picture is that there are two qualitatively different regions divided by the value of the parameter $m$. There are two branches, the first for $m>1$, and the second for $m<1$. In the following tables we present ranges of the values of the parameters for three confidence levels. For the case when $m>1$: \\begin{center} \\begin{tabular}{cccc } {\\bf $1\\sigma$ CL} & {\\bf $2\\sigma$ CL} & {\\bf $3\\sigma$ CL} \\\\\\hline $m\\in(1.00,1.14)$ & $m\\in(1.00,1.21)$ & $m\\in(1.00,1.28)$ \\\\\\hline $\\delta > 1.00$ & $\\delta > 1.00$ & $\\delta > 1.00 $ \\\\\\hline $n\\in(0.0,0.97)$ & $n\\in(0.0,0.98)$ & $n\\in(0.0,0.99)$ \\\\\\hline $y_0\\in(0.0,0.50)$ & $y_0\\in(0.00,0.66)$ & $y_0\\in(0.00,0.76)$ \\\\ \\end{tabular} \\end{center} Below the case of $m<1$: \\begin{center} \\begin{tabular}{ccc} {\\bf $1\\sigma$ CL} & {\\bf $2\\sigma$ CL} & {\\bf $3\\sigma$ CL} \\\\\\hline $m\\in(0.41,0.66)$ & $m\\in(0.39,0.67)$ & $m\\in(0.39,0.72)$ \\\\\\hline $\\delta\\in(0.52,0.99)$ & $\\delta\\in(0.36,0.99)$ & $\\delta\\in(0.26,0.99)$ \\\\\\hline $n\\in(0.0,0.85)$ & $n\\in(0.0,0.91)$ & $n\\in(0.0,0.94)$ \\\\\\hline $y_0\\in(0.45,0.88)$ & $y_0\\in(0.43,0.94)$ & $y_0\\in(0.41,0.97)$ \\\\ \\end{tabular} \\end{center} \\newpage For the case of the branch $m>1$ the maximum value at $3\\sigma \\ y_0$ parameter is $0.76$, which means that such a singularity can happen later than for the second branch $m<1$. While $y_0$ approaches $0.76$, $\\delta$ grows stronger and as the value of $\\delta$ more grows, $m$ becomes better constrained. What seems to be important is that there is an allowed value of $m=2/3$ within $1\\sigma$ CL, which could correspond to the dust filled Einstein-de-Sitter universe in a close to big-bang limiting case. What is also interesting for this branch is that while the allowed values of the nonstandarcity parameter $\\delta$ are small, i.e. $\\delta<1$, the parameter $y_0$ approaches unity which means that the singularity may happen in the nearest future for this case. For $1\\sigma$ CL, $y_0=0.885$ corresponds to $\\sim 2\\times 10^9$ years to the time of the singularity. For $3\\sigma$ CL, $y_0=0.97$ corresponds to the present time at $\\sim 0.37 \\times 10^9$ years before the time of the singularity.\\\\ \\indent In conclusion, we have shown that for a finite scale factor singularity there is an allowed value of $m = 2/3$ within $1\\sigma$ CL, which corresponds to the dust-filled Einstein-de-Sitter universe for the close to big-bang limiting case. The finite scale factor singularity may happen within $2\\times 10^9$ years in future for $1\\sigma$ CL and its prediction at the present moment of cosmic evolution cannot be distinguished, with current observational data, from the prediction given by the standard quintessence scenario of future evolution in the Concordance Model \\cite{chevallier00, linder02, linder05, koivisto05, caldwell07, zhang07, amendola07, diporto07, hu07b, linder09}." }, "1112/1112.2784_arXiv.txt": { "abstract": "VO-KOREL is a web service exploiting the technology of Virtual Observatory for providing the astronomers with the intuitive graphical front-end and distributed computing back-end running the most recent version of Fourier disentangling code KOREL. The system integrates the ideas of the e-shop basket, conserving the privacy of every user by transfer encryption and access authentication, with features of laboratory notebook, allowing the easy housekeeping of both input parameters and final results, as well as it explores a newly emerging technology of cloud computing. While the web-based front-end allows the user to submit data and parameter files, edit parameters, manage a job list, resubmit or cancel running jobs and mainly watching the text and graphical results of a disentangling process, the main part of the back-end is a simple job queue submission system executing in parallel multiple instances of FORTRAN code KOREL. This may be easily extended for GRID-based deployment on massively parallel computing clusters. The short introduction into underlying technologies is given, briefly mentioning advantages as well as bottlenecks of the design used. ", "introduction": "The Virtual observatory (hereafter VO) is a global infrastructure of distributed astronomical archives and data processing services enabling the standardised discovery and access to the astronomical data worldwide as well as a large set of powerful tools for scientific analysis and visualisation (\\cite[Solano 2006]{solano}). VO supports mainly the multi-wavelength research or discovery of rare objects by cross-matching huge catalogues. For interaction with the user as well as other computers the VO uses the modern technology of Web Services (WS). The WS is typically complex processing application using the web technology to transfer input data to the main processing back-end and the results (after intensive number crunching) back to user. All this can be done using only an ordinary web browser (and in principle the science may be done on the fast palmtop or advanced mobile phone). An example of WS is the e-shop or ticket reservation system. ", "conclusions": "The VO-KOREL service is not only giving comfortable environment for Fourier disentangling of spectra, but it is a test-bed of general cloud infrastructure for execution of most scientific computationally intensive codes, like models of stellar atmospheres or special processing of complex data sets." }, "1112/1112.1672_arXiv.txt": { "abstract": "{} {A thorough multiwavelength, multiheight study of the vector magnetic field in a compact active region filament (NOAA 10781) on 2005 July 3rd and 5th is presented. We suggest an evolutionary scenario for this filament. } {Two different inversion codes were used to analyze the full Stokes vectors acquired with the Tenerife Infrared Polarimeter (TIP-II) in a spectral range which comprises the chromospheric \\ion{He}{i} 10830\\,\\AA\\ multiplet and the photospheric \\ion{Si}{i} 10827\\,\\AA\\ line. In addition, we used \\textit{SOHO}/MDI magnetograms as well as BBSO and \\textit{TRACE} images to study the evolution of the filament and its active region (AR). High resolution images of the Dutch Open Telescope were also used. } {An active region filament (that was formed before our observing run) was detected in the chromospheric Helium absorption images on July 3rd. The chromospheric vector magnetic field in this portion of the filament was strongly sheared (parallel to the filament axis) whereas the photospheric field lines underneath had an inverse polarity configuration. From July 3rd to July 5th, an opening and closing of the polarities at either side of the polarity inversion line (PIL) was recorded, resembling the recently discovered process of the sliding door effect seen by Hinode. This is confirmed both with TIP-II and \\textit{SOHO}/MDI data. During this time, a newly created region that contained pores and orphan penumbrae at the PIL was observed. On July 5th, a normal polarity configuration was inferred from the chromospheric spectra, while strongly sheared field lines aligned with the PIL were found in the photosphere. In this same data set, the spine of the filament is also observed in a different portion of the FOV and is clearly mapped by the Silicon line core. } {The inferred vector magnetic fields of the filament suggest a flux rope topology. Furthermore, the observations indicate that the filament is divided in two parts, one which lies in the chromosphere and another one that stays trapped in the photosphere. Therefore, only the top of the helical structure is seen by the Helium lines. The pores and orphan penumbrae at the PIL appear to be the photospheric counterpart of the extremely low-lying filament. We suggest that orphan penumbrae are formed in very narrow PILs of compact ARs and are the photospheric manifestation of flux ropes in the photosphere.} ", "introduction": "Solar filaments, also called prominences when observed in emission outside the disk, are elongated structures formed of dense plasma which lies above polarity inversion lines (PILs) of the photospheric magnetic field \\citep{babcock55} and has a lower temperature than its surroundings. A functional definition of the PIL would be an imaginary line that separates two close areas of opposite polarity. On disk filaments are readily identifiable in the quiet Sun (quiescent filaments) and in active regions (active region filaments) when observed using common chromospheric wavelengths, e.g., the \\ion{He}{i} 10830\\,\\AA\\ multiplet and the H$\\alpha$ 6563\\,\\AA\\ line. The magnetic field plays a major role in their formation, stability and evolution. Hence, spectropolarimetric observations at different heights of the atmosphere combined with high angular resolution images are needed to obtain an overall picture of the physical processes which take place in filaments. The magnetic field strength and its orientation can be inferred using appropiate diagnostic techniques that are able to interpret the Zeeman effect as well as scattering polarization and its modification through the Hanle effect \\citep[][and references therein]{tandberg95}. Although many line-of-sight observations have been carried out in the past decades, it is important to emphasize the need for full-Stokes measurements in order to have complete information of the vector magnetic field. According to observational studies, the magnetic field topology under active region (AR) filaments has sheared or twisted field lines along the PIL which can support dense plasma in magnetic dips \\citep{lites05, lopez06}. Previous findings about prominence magnetic structure have shown models with dipped field lines in a normal \\citep{kippen57} or inverse\\footnote{Throughout this paper, by {\\em inverse configuration} we mean a magnetic field vector, perpendicular to the filament's axis, that points from the negative to the positive polarity. This is the opposite to what one would expect in a potential field solution - where the field points from the positive to the negative side (and is referred to as normal configuration).} \\citep{kuperus74, pneuman83} polarity configuration, the latter having a helical structure. From a large sample of quiescent prominence observations, \\citet{leroy84} found both types of configurations and a strong correlation between the magnetic field topology and the height of the prominence. On the other hand, \\citet{bommier94} found predominantly inverse polarity configurations. However, more recent photospheric spectropolarimetric observations of AR filaments have revealed that both configurations can coexist in the same filament \\citep{guo10} or even evolve with time from one type to the other, as presented by \\citet{okamoto08} from the analysis of vector magnetogram sequences. The formation process of filaments is still a controversial issue in solar physics and has been widely debated. On the one hand, there is the \\emph{sheared arcade model} which, by large scale photospheric footpoint motions (such as shear at, and convergence towards the PIL, and subsequent reconnection processes) is able to form dips, and even helical structures, where plasma can be gravitationally confined \\citep[e.g.,][]{pneuman83,balle89, antiochos94,aulanier98, devore00,martens01,welsch05,karpen07}. This model is capable of reproducing both inverse and normal polarity configurations in the same filament \\citep{aulanier02}. However, recent observational works in a quiescent filament \\citep{hindman06} and along an AR filament channel \\citep{lites10} did not find evidence for these systematic photospheric flows that converged at the PIL and triggered reconnection processes. In our opinion, more observations are needed to support this important result. On the other hand, the \\emph{flux rope emergence model} assumes that the twist in the field lines is already present in the convection zone before emerging into the atmosphere \\citep[eg.,][]{kurokawa87, low94, low95, lites05}. The rise of twisted magnetic fields has been studied by various authors through observations \\citep{lites95, leka96, lites10} as well as three-dimensional (3D) numerical simulations \\citep[e.g.,][]{fan01,archontis04,magara04,sykora08,fan09,yelles09}. Such an emerging helical flux rope scenario, combined with the presence of granular flows, is suggested to be the main process of formation and maintenance of the AR filament studied by \\citet{okamoto08, okamoto09}. Recent nonlinear force-free field (NLFFF) extrapolations of photospheric magnetic fields underneath AR filaments \\citep{guo10, canou10, jing10} agreed that the main structure has to be a flux rope, although \\citet{guo10} also found dipped arcade field lines in the same filament. Flux rope emergence from below the photosphere encounters severe difficulties to reach chromospheric layers and above, as it is loaded with mass \\citep[e.g.,][]{archontis08}. Indeed, the previous work suggests that a second flux rope formed from instabilities and atmospheric reconnection is what lifts up the mass and forms the observed active region filaments. The main difference between the sheared arcade and the emerging flux rope models is the existence of a flux concentration stuck at photospheric levels in the latter scenario. The reader is referred to the paper of \\citet{mackay10} for a recent review on the magnetic structure of filaments. In the past, only a few measurements of the magnetic field strength in AR filaments have been done. For instance, \\citet{kuckein09}, using the same data set that is described in this paper, found a predominance of Zeeman-like signatures in the Stokes profiles. Using three different methods, they inferred very strong magnetic fields in the filament (600--700\\,Gauss), with a dominant transverse component (with respect to the line-of-sight). The aim of this paper is to study the strength and topology of the magnetic field in an active region filament at photospheric and chromospheric heights simultaneously. Recently, several studies have presented analyses of the vector magnetic field in filaments or prominences from observations either in the photosphere \\citep[e.g.,][]{lites05,lopez06,okamoto08,guo10,lites10} or the chromosphere \\citep[e.g.,][]{lin98,casini03,merenda06,kuckein09}, but non of them have inferred the field at both heights at the same time. The 10830\\,\\AA\\ spectral region, which includes a chromospheric \\ion{He}{i} triplet and a photospheric \\ion{Si}{i} line, offers a unique spectral window to understand the physical processes that take place in AR filaments as already shown by \\citet{sasso11}. In this work, we focus on the overall magnetic configuration observed simultaneously in the photosphere and the chromosphere. ", "conclusions": "The main conclusions of the present work are: \\begin{enumerate} \\item The ``sliding doors'' effect described by Okamoto et al. (2008) was seen in our observations during the period between July 3rd and July 5th. This has been identified in \\textit{SOHO}/MDI data as well as in the TIP-II \\ion{Si}{i} magnetograms for those days. \\item The observed AR filament can be separated into two areas. The first one is well observed on July 3rd and shows the filament axis, or spine, in the Helium absorption images. The \\ion{He}{i} data show the magnetic field (with horizontal field strengths of 400--500\\,G) aligned with the filament, indicating a strong sheared configuration. This region is also observed in the FOV of July 5th, when portions of the spine are even seen in the Silicon line core images. This would indicate that the part of the spine observed this day lies in lower layers than the portion observed two days earlier. \\item The photospheric fields for this area of the filament show an inverse configuration of the magnetic field lines that we interpret to be the bottom of a flux rope structure. Similarly, we interpret the spine seen in Helium as the signature of the flux rope axis. \\item The second area corresponds to the orphan penumbrae that appear during the sliding door timeframe and clearly shows a different magnetic configuration. Helium observations exhibit a normal polarity configuration whereas the Silicon data show a strongly sheared region with very intense horizontal fields. This magnetic topology is interpreted as a scenario in which the chromospheric magnetic fields trace the top part of the flux rope, while the spine -or flux rope axis- is still lying in the photosphere. \\item The observed orphan penumbrae is thus the photospheric counterpart of this active region filament. Since the spine is also seen in the Silicon line, we emphasize that active region filaments can have a clear photospheric signature. In particular, it would be interesting to find out if all orphan penumbral regions observed inside active regions are related to their filaments. \\item The whole emergence process of these orphan penumbrae does not leave a clear imprint in the flux history of the active region. This is interpreted as due to the flux system being mostly transverse, with a low enough twist to not generate identifiable signals in \\textit{SOHO}/MDI longitudinal magnetograms. \\item We do not observe flux loops with a normal polarity configuration in the photosphere, as suggested by some filament generation models based on footpoint motions and reconnection \\citep[see][]{balle89, vanballe07}. While the preexisting filament observed on July 1st in the BBSO data could have been created in the way described in these works, the configuration and the evolution described here for the period from July 3rd to July 5th suggests otherwise. \\end{enumerate} The current study was limited, not only by the small FOV of TIP-II, which nowadays has a slit twice as long as during our observing campaign in 2005, but also by the observational gap on July 4th. Magnetic field extrapolations could undoubtedly shed more light on clarifying the magnetic structure of this filament. It is now crucial to carry out more multiwavelength measurements, as the one presented in here, with higher cadence and bigger FOVs to fit the pieces of the puzzle together, in order to fully understand the origin, evolution and magnetic topology of AR filaments. In particular, continuous vector magnetograms of active regions together with simultaneous imaging of the corona should be able to prove/disprove the proposed scenario for the last stages of their evolution. The instrument suite on board the NASA/SDO satellite is the best candidate for such an study." }, "1112/1112.6052_arXiv.txt": { "abstract": "We investigate the constraints on the scalar, vector and spin-3/2 dark matter interaction with the standard model particles, from the observations of dark matter relic density, the direct detection experiments of CDMS and XENON, and the indirect detection of the $\\bar p/p$ ratio by PAMELA. A model independent way is adopted by constructing general 4-particle operators up to dimension 6 for the effective interaction between dark matter and standard model particles. We find that the constraints from different experiments are complementary with each other. Comparison among these constraints may exclude some effective models of dark matter and limit some parameters of others. The spin-independent direct detection gives strong constraints for some operators, while the indirect detection of $\\bar p/p$ data can be more sensitive than direct detection or relic density for light dark matter (whose mass $\\lesssim 70$~GeV) in some cases. The constraints on some operators for spin-3/2 dark matter are shown to be similar to those on their analogous operators for Dirac fermionic dark matter. There are still some operators not sensitive to the current dark matter direct and indirect search experiments. ", "introduction": "It is by now established~\\cite{Tegmark:2006az,Komatsu:2008hk,Komatsu:2010fb} that about 23\\% of the constituents of the Universe are composed of dark matter (DM)~\\cite{kolb-turner,Jungman:1995df,Bertone:2004pz, Murayama:2007ek,Feng:2010gw}. However, its nature remains unclear. A well-motivated candidate for DM is the weakly interacting massive particle (WIMP), which must be stable, nonrelativistic, electrically neutral and colorless. If the WIMP mass is from a few GeV to TeV while their interaction strength is of the weak scale, they can naturally yield the observed relic density of DM~\\cite{Feng:2010gw}. Although there is no stable WIMP in the Standard Model (SM), WIMP candidates exist in various theoretical models trying to solve the SM problems at the weak scale, such as supersymmetric models \\cite{Jungman:1995df,Goldberg:1983nd,Ellis:1983ew,Arnowitt:1992aq,Nath:1992ty, Kane:1993td}, extra dimensional models~\\cite{Kolb:1983fm,Cheng:2002ej, Hooper:2007qk,Servant:2002aq,Servant:2002hb,Agashe:2004ci, Agashe:2004bm,Agashe:2007jb}, little Higgs models \\cite{Cheng:2004yc,Low:2004xc,Birkedal:2006fz,Freitas:2009jq,Kim:2009dr}, left-right symmetric models~\\cite{Dolle:2007ce,Guo:2008hy,Guo:2008si, Guo:2010vy}, and some other models (e.g.~\\cite{Khlopov:2008ty,Li:2010rz, Cui:2011wk,Kanemura:2011mw}). The specific models mentioned above are very attractive, but still lack experimental support. The well-running LHC experiment may find some important signals of these models in the near future. However, if other new particle species are all so heavy that the DM particle is the only new particle within the reach of LHC, it will be very difficult to know which model the DM particle belongs to. In addition, it is possible that the DM particle may be first observed by direct or indirect detection experiments. These early observations may only provide information about some general properties of the DM particle, and may not be able to distinguish the underlying theories. Therefore, the model-independent studies of the DM phenomenology can play an important role as they may avoid theoretical bias \\cite{Birkedal:2004xn,Giuliani:2004uk,Kurylov:2003ra,Beltran:2008xg, Cirelli:2008pk,Shepherd:2009sa}. Recently there have been quite a few papers to study various phenomenologies related with DM in the model-independent way~\\cite{Cao:2009uv,Cao:2009uw, Beltran:2010ww,Fitzpatrick:2010em,Goodman:2010yf,Bai:2010hh, Goodman:2010ku,Goodman:2010qn,Bell:2010ei,Zheng:2010js,Cheung:2010ua, Cheung:2011nt,Mambrini:2011pw,Rajaraman:2011wf,Fox:2011pm,Goodman:2011jq, Kamenik:2011vy,Shoemaker:2011vi}. Especially the relic density measured by WMAP~\\cite{Komatsu:2010fb}, direct detection from CDMS~\\cite{Ahmed:2009zw,Akerib:2005za} and XENON \\cite{Aprile:2011hi,Angle:2008we}, and possible collider signals from Tevatron (e.g.~\\cite{Aaltonen:2008hh,CDF:monojet}) and from LHC (e.g. \\cite{Chatrchyan:2011nd,ATLAS:monojet}) are considered in these studies. In our previous work~\\cite{Zheng:2010js}, we investigated a general set of 4-fermion operators for the effective interaction between the spin-1/2 fermionic DM and the SM particles, and gave the phenomenological constraints from the observed DM relic density, the direct detection experiments by CDMS and XENON, and the indirect detection of the $\\bar{p}/p$ ratio by PAMELA~\\cite{Adriani:2010rc}. It was found that the constraints from different observations are quite complementary. Besides the possibility of spin-1/2 WIMPs, it is also possible that DM is composed of scalar, vector, or spin-3/2 WIMPs, which belong to different representations of Lorentz group. In these cases, the forms of possible effective operators are different. These differences may lead to distinguishable phenomenological results. In this work we extend our previous analysis to the cases of scalar, vector and spin-3/2 DM. We will consider a general set of 4-particle operators up to dimension 6 for the effective interaction between the WIMPs and the SM fermions and compute their phenomenological constraints. We will use the updated limit of XENON100 SI direct detection \\cite{Aprile:2011hi}, which is stronger than that adopted in our previous work~\\cite{Zheng:2010js}. This paper is organized as follows. In Sections \\ref{sec-scal} and \\ref{sec-vect}, the effective models of scalar and vector DM are discussed, respectively. In the subsections of these two sections, we explore the constraints on these models from the DM relic density, direct and indirect detection searches and the validity of effective theory, and then present the combined constraints on the effective coupling constants of these models. In Section \\ref{sec-3/2}, the study on the effective models of spin-3/2 DM are carried out briefly. The conclusions are given in Section \\ref{sec-con}. ", "conclusions": "} In this work, we give a general analysis of the 4-particle interaction between SM particles and DM which consists of scalar, vector or spin-3/2 WIMPs. The most general forms of the 4-particle operators up to dimension 6 have been considered. We find that for scalar, vector and spin-3/2 DM, the constraints from DM relic density, DM direct and indirect detection are complementary to each other. Thus the comparison among different kinds of experimental results gives us a complete picture about the current DM searches. In general, the constraints from SI direct detection are the most stringent, while those from SD direct detection are quite weak. On the other hand, for light DM (whose mass $\\lesssim 70$~GeV) the cosmic-ray $\\bar p/p$ data can be more sensitive than direct detection or relic density in some cases. Assuming one operator dominates the effective interaction between DM and SM fermions, we find that in some cases the constraints are so strong that the Universe will be overclosed by DM thermal production. If the standard cosmology is still retained, the DM models in such cases should be excluded, which are indicated in Tables \\ref{tab:scalar_sum}, \\ref{tab:vector_sum} and \\ref{tab:3/2_sum}. In the case of scalar DM, recent direct detection experiments exclude some $M_\\phi$ regions for S and V interactions, while the PAMELA $\\bar p/p$ ratio excludes some small $M_\\phi$ regions ($\\lesssim 70$~GeV) for S and SP interactions. In the case of vector DM, recent direct detection experiments exclude some $M_X$ regions for S and V interactions, and for $\\widetilde{\\mathrm{VA}}$ interaction only with universal couplings. The PAMELA $\\bar p/p$ ratio, however, excludes some small $M_X$ regions ($\\lesssim 70$~GeV) for S, T, SP and $\\widetilde{\\mathrm{T}}$ interactions, and most of these interactions cannot be excluded by direct detection. In the case of spin-3/2 DM, the constraints on most interactions (S, P, V, A, T1, T4, SP, PS, VA and AV) are much similar to those on their analogous interactions of Dirac fermionic DM in our previous work \\cite{Zheng:2010js}. Besides, the constraints on T2 and T3 interactions are so weak that only direct detection can exclude a little region of $M_\\chi$ in the case of T2 interaction with universal couplings. For T5 and T6 interactions, some small $M_\\chi$ regions ($\\lesssim 70$~GeV) are excluded by the PAMELA $\\bar p/p$ ratio. Among the scalar, vector and spin-3/2 DM effective models, there are still some effective interactions to which the recent DM direct and indirect search experiments are not sensitive at all." }, "1112/1112.4866_arXiv.txt": { "abstract": "We have recently proposed a special class of scalar tensor theories known as {\\it the Fab Four}. These arose from attempts to analyse the cosmological constant problem within the context of Horndeski's most general scalar tensor theory. {\\it The Fab Four} together give rise to a model of self-tuning, with the relevant solutions evading Weinberg's no-go theorem by relaxing the condition of Poincar\\'e invariance in the scalar sector. {\\it The Fab Four} are made up of four geometric terms in the action with each term containing a free potential function of the scalar field. In this paper we rigorously derive this model from the general model of Horndeski, proving that {\\it the Fab Four} represents the {\\it only} classical scalar tensor theory of this type that has any hope of tackling the cosmological constant problem. We present the full equations of motion for this theory, and give an heuristic argument to suggest that one might be able to keep radiative corrections under control. We also give {\\it the Fab Four} in terms of the potentials presented in Deffayet {\\it et al}'s version of Horndeski. ", "introduction": "The cosmological constant problem has been % described as the most embarrassing fine-tuning problem in Physics today. According to our current understanding of particle physics and effective {quantum} field theory, the vacuum receives zero point energy contributions from each particle species right up to the {UV} cut-off, which may be as high as the Planck scale. The trouble is that in General Relativity, any matter, including vacuum energy, gravitates and the only way to make it compatible with observation is to demand considerable fine-tuning between the vacuum energy and the bare cosmological constant. The situation is exacerbated by phase transitions in the early universe that can give rise to constant shifts in the vacuum energy contribution. To date, particle physicists have failed to come up with a satisfactory solution to this problem, so some recent attempts have instead focussed on gravitational physics. This alternative approach requires a non-trivial modification of Einstein's theory at large distances (see \\cite{review} for a detailed review of modified gravity). One particularly interesting direction involves scalar-tensor theories of gravity. It seems sensible to require that any theory maintains second order field equations in order to avoid an Ostrogradski instability \\cite{ostro}, and the most general scalar-tensor theory satisfying that criteria in four dimensions was written down back in 1974 by Horndeski \\cite{horndeski:1974} (it has recently been rediscovered independently in \\cite{general}). Such theories of modified gravity cover a wide range of models, ranging from Brans-Dicke gravity \\cite{bdgravity} to the recent models \\cite{covgal,galmodels} inspired by galileon theory \\cite{galileon}. Galileon models are examples of higher order scalar tensor Lagrangians with second order field equations, and, as a result, they are closely related to Kaluza-Klein compactifications of higher dimensional Lovelock theories \\cite{kkl, VanAcoleyen:2011mj}. Of course all of these scalar-tensor models can be considered as special cases of Horndeski's original action. In \\cite{Charmousis:2011bf} we obtained a new class of solutions arising out of Horndeski's theory on FLRW backgrounds. The new solutions gave a viable self-tuning mechanism for solving the (old) cosmological constant problem, at least at the classical level, by completely screening the spacetime curvature from the net cosmological constant. This would seem to be in violation of Weinberg's famous no-go theorem \\cite{nogo} that forbids precisely this kind of self-adjustment mechanism. However, Weinberg assumes Poincar\\'e invariance to hold universally across all fields whereas we allow it to be broken in the scalar field sector. In other words, we continue to require Poincar\\'e invariance at the level of spacetime curvature, but not at the level of the self-adjusting scalar field. A similar approach was adopted in the context of bigalileon theory \\cite{bigalileon} where only a small vacuum energy could be successfully screened away. In \\cite{Charmousis:2011bf}, we provided a brief sketch of how the system works for scalar tensor theories where matter is only minimally coupled to the metric (required to ensure compatibility with Einstein's Equivalence Principle (EEP)). By demanding the presence of a viable self-tuning mechanism we were able to place powerful restrictions on the allowed form of Horndeski's original Lagrangian. Whereas the original model is complicated, with many arbitrary functions of both the scalar and its derivatives, we showed that once the model is passed through our self-tuning filter (to be defined shortly), it reduces in form to just four base Lagrangians each depending on an arbitrary function of the scalar only, coupled to a curvature term. We called these base Lagrangians {\\it the Fab Four}: John, Paul, George and Ringo. Together, {\\it the Fab Four} make up the most general scalar-tensor theory capable of self-tuning. Individually they are given by the following \\begin{eqnarray} \\label{eq:john} {\\cal L}_{john} &=& \\sqrt{-g} V_{john}(\\phi)G^{\\mu\\nu} \\nabla_\\mu\\phi \\nabla_\\nu \\phi \\\\ \\label{eq:paul} {\\cal L}_{paul} &=&\\sqrt{-g}V_{paul}(\\phi) P^{\\mu\\nu\\alpha \\beta} \\nabla_\\mu \\phi \\nabla_\\alpha \\phi \\nabla_\\nu \\nabla_\\beta \\phi \\\\ \\label{eq:george} {\\cal L}_{george} &=&\\sqrt{-g}V_{george}(\\phi) R \\\\ \\label{eq:ringo} {\\cal L}_{ringo} &=& \\sqrt{-g}V_{ringo}(\\phi) \\GB \\end{eqnarray} where $R$ is the Ricci scalar, $G_{\\mu\\nu}$ is the Einstein tensor, $P_{\\mu\\nu\\alpha \\beta}$ is the double dual of the Riemann tensor \\cite{mtw}, $\\GB=R^{\\mu\\nu \\alpha \\beta} R_{\\mu\\nu \\alpha \\beta}-4R^{\\mu\\nu}R_{\\mu\\nu}+R^2$ is the Gauss-Bonnet combination, and in what follows the Greek indices $\\mu,\\nu =0..3$. The purpose of this paper is to rigorously derive the conditions that lead to these four base Lagrangians, showing how they naturally lead to self-tuning solutions, provided that $\\{V_{john}, V_{paul}, V_{george}\\} \\neq \\{0,0, constant\\}$. Note that this constraint means that GR is {\\it not} a {\\it Fab Four} theory, consistent with the fact that it does not have self-tuning solutions. To be clear as to what is meant by ``self-tuning\", let us define our self-tuning filter. We require that \\begin{itemize} \\item the theory should admit a Minkowski vacuum\\footnote{For simplicity throughout the introductory part of the text we have simply written, ``Minkowski vacuum\" to stand for``a patch of Minkowski vacuum\". This technical issue will be made clear later on in section \\ref{self-tune}} for any value of the net cosmological constant \\item this should remain true before and after any phase transition where the cosmological constant jumps instantaneously by a finite amount. \\item the theory should permit a non-trivial cosmology \\end{itemize} The last condition ensures that Minkowski space is not the only cosmological solution available, something that is certainly required by observation. The idea is that the cosmological field equations should be dynamical, with the Minkowski solution corresponding to some sort of fixed point. In other words, once we are on a Minkowski solution, we stay there -- otherwise we evolve to it dynamically. This last statement would indicate that the self-tuning vacuum is an attractive fixed point. We do not prove this here, but in our companion paper on cosmology \\cite{fab4cos} we will see plenty of examples where it is indeed the case. The first two conditions are the basic requirements of any successful self-tuning mechanism. There are many examples in the literature which pass the first condition, but fall down at the second. This includes the much explored co-dimension two braneworld models in which the compact extra dimensions are shaped like a rugby ball \\cite{Carroll}. The brane tension controls the deficit angle, while the brane geometry is completely determined by the bulk cosmological constant and the magnetic flux. Therefore, this passes our first condition. However, when the brane tension changes after a phase transition it affects the brane curvature via the backdoor, by altering magnetic flux and the theory falls foul of our second condition \\cite{cline}. It is interesting to note that any diffeomorphism invariant theory that passes both the first and second condition will admit a Minkowski solution in the presence of {\\it any} cosmological fluid, not just a cosmological constant. The point is that our vacuum energy density corresponds to a piecewise constant function, with discontinuities at the phase transitions. In principle these transitions can occur at any given time, so a Minkowski solution can be returned for all piecewise constant energy densities. The energy density of an arbitrary cosmological fluid can be well approximated by a piecewise constant function, and so it follows that it must also admit a Minkowski solution. Like we said, this property must hold for {\\it any} diffeomorphism invariant theory passing our first two conditions, and not just {\\it the Fab Four}. We might worry that this prevents any hope of a sensible matter dominated cosmology. However, this is where the third condition comes into play, and we once again refer the reader to our companion paper \\cite{fab4cos} for evidence that sensible cosmologies are indeed possible within this theory. Even so, the main aim of this paper is not to extoll the virtues of {\\it the Fab Four} but to push a very general class of modified gravity theories through our self-tuning filter and to see what happens. In a sense we are testing the scope of Weinberg's theorem, relaxing one of his assumptions and seeing how far we can go. It turns out that our filter is very efficient -- it removes most of Horndeski's original theory-- but it is not $100\\%$ efficient. We are left with {\\it the Fab Four}. The layout of the paper is as follows: in section \\ref{horndeski-action} we present the original action of Horndeski \\cite{horndeski:1974}, minimally coupled to matter, and derive the Hamiltonian and scalar field equations of motion for the system. In section \\ref{self-tune} we demonstrate how a self-tuning solution can in principle be obtained by relaxing Weinberg's no-go theorem to allow the scalar field to evolve in time. This is followed in section \\ref{self-tune-horndeski} with a derivation of the self-tuned Horndeski action, where we show how the initial complicated Lagrangian reduces to four simple terms each one being an arbitrary function of the scalar field alone coupled to a curvature term. Of particular note is that any dependence on the kinetic energy of the scalar field drops out. In section \\ref{conc} we bring everything together and discuss further demands we may wish to make on our theory, over and above our original filter, ranging from cosmological and solar system tests, to issues of stability. We also elucidate the elegant geometrical structure possessed by {\\it the Fab Four} and present their equations of motion in full. We have a number of appendices, most of which are technical additions to the main text. The exceptions are appendices \\ref{app:dgsz} and \\ref{radiative}. In appendix \\ref{app:dgsz} we present {\\it the Fab Four} in the language of the potentials of Deffayet {\\it et al}'s version of Horndeski \\cite{general}. In appendix \\ref{radiative} we discuss the issue of radiative corrections to {\\it the Fab Four}. This is an important question, because radiative corrections are at the heart of the cosmological constant problem. We do not attempt a detailed analysis -- that is certainly beyond the scope of the current paper --- but we do perform some heuristic calculations. It seems that radiative corrections can be kept under control given some not too restrictive conditions. ", "conclusions": "" }, "1112/1112.2501_arXiv.txt": { "abstract": "Over the past decade advancements in the understanding of several astrophysical phenomena have allowed us to infer a concordance cosmological model that successfully accounts for most of the observations of our universe. This has opened up the way to studies that aim to better determine the constants of the model and confront its predictions with those of competing scenarios. Here, we use strong gravitational lenses as cosmological probes. Strong lensing, as opposed to weak lensing, produces multiple images of a single source. Extracting cosmologically relevant information requires accurate modeling of the lens mass distribution, the latter being a galaxy or a cluster. To this purpose a variety of models are available, but it is hard to distinguish between them, as the choice is mostly guided by the quality of the fit to the data without accounting for the number of additional parameters introduced. However, this is a model selection problem rather than one of parameter fitting that we address in the Bayesian framework. Using simple test cases, we show that the assumption of more complicate lens models may not be justified given the level of accuracy of the data. ", "introduction": "Over the past years, our understanding of the universe has greatly improved. The concordance model explains most of the cosmological observations. We have now entered a phase where finding new observational ways of measuring the constants of this model as well as confronting predictions with those of competing scenarios is crucial to further advance the research in cosmology. Strongly lensed quasars constitute such a cosmological probe. In a strong gravitational lens, each image is the result of a different light-path. As a result, if the source behind the lens has a variable luminosity, as quasars do, it will manifest with a time delay between the two images. This time delay $\\Delta t$ depends on the gravitational potential of the lens, and the underlying cosmological model. The time delay between two images A and B is: \\begin{equation} \\Delta t_{A,B} = (1 + z_l) \\frac{d_l d_s}{d_{ls}} \\left( \\frac{1}{2}((\\theta_A - \\beta)^2 -(\\theta_B -\\beta)^2) + \\psi(\\theta_A) - \\psi(\\theta_B) \\right) \\end{equation} where $\\Delta t_{A,B}$, $z_l$, $\\theta_A$ and $\\theta_B$ are observables, $\\beta$, $\\psi(\\theta_A)$ and $\\psi(\\theta_B)$ depend on the lens model and $d_l$, $d_s$ and $d_{ls}$ depend on cosmology. Using the above relation, we can derive constraints on cosmological parameters, provided we assume a lens model. Time delays are particularly sensitive to the value of the Hubble constant $H_0$. Unfortunately, a change in the lens model can shift the inferred value of $H_0$ by a factor of two. Hence, the modeling of the lens, as well as a robust discrimination between lens models, is critical to the study of time delays. Here, we first discuss the lens models used in our study, then describe our methodology based on Bayesian statistical analysis, and finally present our results. ", "conclusions": "Bayesian techniques are a good way to discriminate between lens models and allow us to decide which double lenses can be accurately modeled by a simple power-law model. With the result from this analysis, we now have a good sample of double lenses, together with lens models, to determine cosmological parameters more accurately. %" }, "1112/1112.2737_arXiv.txt": { "abstract": "We present absorption line indices measured in the integrated spectra of globular clusters both from the Galaxy and from M~31. Our samples include 41 Galactic globular clusters, and more than 300 clusters in M~31. The conversion of instrumental equivalent widths into the Lick system is described, and zero-point uncertainties are provided. Comparison of line indices of old M~31 clusters and Galactic globular clusters suggests an absence of important differences in chemical composition between the two cluster systems. In particular, CN indices in the spectra of M~31 and Galactic clusters are essentially consistent with each other, in disagreement with several previous works. We reanalyze some of the previous data, and conclude that reported CN differences between M~31 and Galactic clusters were mostly due to data calibration uncertainties. Our data support the conclusion that the chemical compositions of Milky Way and M~31 globular clusters are not substantially different, and that there is no need to resort to enhanced nitrogen abundances to account for the optical spectra of M~31 globular clusters. ", "introduction": "The history of star formation and chemical enrichment of galaxies is encoded in the ages and chemical compositions of their stellar populations. In particular, powerful insights on the processes leading to the assembly of the Galactic halo are gained by studies of the chemical abundances of their constituent populations of field stars and globular clusters (GCs). It is only natural to extend such studies to the nearest giant spiral galaxy, M~31. While detailed abundances of individual stars in the Galactic halo, disk, and bulge, have been obtained and extensively analyzed in the past several decades, similar studies of comparable samples of M~31 halo stars await the advent of 20-30~m class telescopes, equipped with efficient high-resolution spectrographs. In the meantime, integrated light studies of M~31 GCs have historically provided valuable qualitative information about the chemical composition of the halo of M~31, leading to important clues on its formation history. As stellar population synthesis models and instruments both grow in sophistication, quantitative information on elemental abundances from integrated-light studies of M~31 GCs is also becoming available \\citep[e.g.,][]{co09,ca10}. The history of this field has been punctuated by heroic observational efforts, based on optical spectra painstakingly collected with, according to today's standards, relatively small telescopes, rather inefficient spectrographs, and low-quantum efficiency (and often difficult to calibrate) detectors \\citep[e.g.,][]{bu84,bh90}. The main results emerging from these efforts are: 1) M~31 GCs are on average slightly more metal rich than their Galactic counterparts, while spanning roughly the same range of metallicities \\citep{bh91}; 2) M~31 GCs are $\\alpha$-enhanced, just like those in the MW \\citep[e.g.,][]{pu05}; 3) M~31 GCs are CN-enhanced compared to MW GCs \\citep{bu84,bh90,be04,pu05}. In particular, \\cite{bu84} found that the CN band at $\\lambda$ 4170 ${\\rm\\AA}$ is stronger by $\\sim$ 0.05 mag in the spectra of moderately metal-poor M31 GCs than in those of their MW counterparts of same metallicity; and 4) M~31 GCs are possibly younger/older than those in the MW. \\cite{bu84} found that $H\\beta$ is stronger in the spectra of M~31 GCs than in those of their MW counterparts by about 0.5 ${\\rm\\AA}$, indicating a younger age or differences in horizontal branch morphology. The latter two results are contradicted by the findings presented in this paper. While \\cite{ca10} discuss the Balmer line strengths in the spectra of the two GC systems, comparing their ages, CN is the main focus of this paper. Those significant conclusions may teach us about important aspects of the formation of the two galaxies. For instance, the comparable mean metallicities may be indicative of similar overall chemical enrichment, suggesting that the early star formation efficiency has been similar in the two systems---or perhaps in the sub-components that eventually assembled to form them. The similar $\\alpha$-enhancement, assessed by measurements of Mg and Fe-sensitive absorption lines, suggests that either the time scale for star formation, the IMF, or a combination thereof, were similar in both galaxies. The difference in CN strength, which has been ascribed to a difference in nitrogen abundance, is difficult to interpret, owing mostly to uncertainties in the models for the nucleosynthesis of that element. The issue is further complicated by the fact that CNO elements are seen to present strong star-to-star variations in Galactic GCs \\citep[e.g.,][and references therein]{gr04}, which may be associated with the presence of multiple stellar populations in those GCs \\citep[][and references therein]{cs10,pi09,ms09,can98}, and by inference, in their M~31 counterparts. Regardless, any scenario for the formation of the MW and M~31 haloes, and their GC systems, will be challenged by the large nitrogen abundance differences between systems that look otherwise very similar. This is the fourth of a series of papers dedicated to analyzing the kinematics and chemistry of a large sample of M~31 GCs, based on high-quality integrated spectra for several hundred M~31 clusters, obtained with MMT/Hectospec. In Paper I \\citep{ca09a} we characterized the population of young ($\\simless$ 2 Gyr) M~31 clusters in terms of their ages, metallicities, masses, and kinematics. In Paper II \\citep{ca10}, the ages and metallicities of M~31 old GCs were studied. In particular, we found no differences between the ages of the two old GC systems, in disagreement with previous claims. Moreover, we found that the M~31 GC system does not have a bimodal metallicity distribution, in agreement with recent findings \\citep{yo06}. In Paper III \\citep{mo10}, we suggest that the old bulge GCs in M~31 are characterized by a bar-like kinematics. In this paper, we present absorption line indices measured in integrated spectra of a large sample of M~31 and MW GCs. The data for M~31 GCs come from spectra described by \\cite{ca09a}, and those for MW GCs are based on the integrated GC spectra from \\cite{s05}. While the focus of this paper is on the comparison of CN strengths between the two GC systems, \\cite{s11} will present an analysis of abundance ratios, based on application of stellar population synthesis models to these index measurements. This paper is organized as follows: in Section~\\ref{data}, the measurement of Lick absorption line indices is described, and a comparison between CN indices in M~31 and MW GCs is presented in Section~\\ref{analysis}. Our conclusions are summarized in Section~\\ref{epilogue}. ", "conclusions": "\\label{epilogue} We present absorption line index measurements taken in integrated spectra of a large number of M~31 and MW GCs, from \\cite{ca09a} and \\cite{s05}, respectively. We discuss the conversion of instrumental measurements to a common equivalent system \\citep[defined in][]{s07}, as well as the uncertainties in the zero points of these systems. By comparing measurements of CN indices in old GCs belonging to both data sets, we conclude that, in disagreement with previous work by \\cite{bu84}, \\cite{bh90}, \\cite{da90}, \\cite{bu04}, \\cite{be04}, and \\cite{pu05}, among others, M~31 and MW GCs of the same [Fe/H] {\\it generally} have similar CN-band strengths. Because this result disagrees with conclusions by many different groups, we have reanalyzed the data by some of these different authors and suggest that in most cases their conclusions were a result of calibration problems. In particular, the blue Lick CN indices are relatively weak and are very sensitive to flux calibration uncertainties. We also find that the two MW GCs with [Fe/H] $\\simgreater -0.4$, NGC~6528 and 6553, have weaker CN bands than their M~31 counterparts at the same [Fe/H]. It is not entirely clear whether this latter difference is real or a by-product of sky-subtraction errors, or of issues with the sampling of the GCs' brightest stars. Other data sets \\citep[e.g.,][]{pu05} do not show similar CN discrepancies between these GCs and their M~31 metal-rich counterparts. Even if our results are correct, and these two GCs indeed have weaker CN indices than metal-rich M~31 GCs, the fact that they seem to depart significantly from a $\\langle$Fe$\\rangle$--$CN_1$ trend that is present in {\\it both} samples suggests that NGC~6528 and 6553 are the exception rather than the rule. The result by \\cite{bu84} that M~31 GCs have stronger CN bands than their MW counterparts of same metallicity, even though confirmed by several later studies, has always been difficult to understand because it is very hard to explain how nitrogen abundances could differ substantially between otherwise very similar systems. \\cite{lb03} proposed a scenario where GCs were formed from zero-metallicity material, pre-enriched by hypernovae explosions in the center of $\\sim 10^6 M_\\odot$ gas clouds. Nitrogen abundance differences between the two GC systems was the one piece of evidence that could not be explained in that scenario. We hope that the task of devising a mechanism for the formation of the haloes of the Milky Way and Andromeda galaxies is made easier by the findings presented in this paper. In a forthcoming paper, we contrast measurements of the abundance patterns of M~31 and Galactic GCs, and discuss their implications to our understanding of the formation of the two galaxy haloes." }, "1112/1112.2995_arXiv.txt": { "abstract": "{We address the issue of cosmological backreaction from non-linear structure formation by constructing an approximation for the time evolved metric of a dust dominated universe based on a gradient expansion. Our metric begins as a perturbation of a flat Friedmann-Robertson-Walker state described by a nearly scale invariant, Gaussian, power-law distribution, and evolves in time until non-linear structures have formed. After describing and attempting to control for certain complications in the implementation of this approach, this metric then forms a working model of the universe. We numerically calculate the evolution of the average scale factor in this model and hence the backreaction. We argue that, despite its limitations, this model is more realistic than previous models that have confronted the issue of backreaction. We find that the \\emph{instantaneous} effects of backreaction in this model could be as large as $\\sim10\\%$ of the background. This suggests that a proper understanding of the \\emph{cumulative} effects of backreaction could be crucial for precision cosmology and any future exploration of the dark sector.} \\begin{document} ", "introduction": "The issue of cosmological backreaction \\cite{Buchert:1999er} has fueled a lively debate in the literature. The question is: can inhomogeneities in our universe backreact and affect the average dynamics within our causal horizon such that the observed acceleration can be attributed to their influence? If the answer is positive then the mystery of the cosmological constant will have found a solution requiring no new physics but which will nevertheless demonstrate the subtlety of gravitational phenomena. Even if the answer is negative, backreaction will be operative at some level due to the non-linear nature of gravity and its effects may be visible in the new generation of cosmological observations. Either way, calculating its magnitude is an important cosmological question. The problem with assessing the magnitude of this backreaction lies with the complexities of the non-linear structures forming under gravity in the late universe. For example, using second order perturbation theory falls short of providing an answer since the effect is expected to emerge in the non-linear regime where perturbation theory breaks down. Attempts to model the non-linear structures involving toy models of voids and overdensities may be useful for developing intuition but are nevertheless simplistic. N-body simulations imply that the effect is small but they are Newtonian with zero global backreaction by construction; the backreaction is a purely relativistic effect \\cite{Buchert:1995fz}. A strong argument for the smallness of the effect uses the fact that even with non-linear overdensities, the local metric perturbations and peculiar velocities are still much smaller than unity (away from black holes) and that a perturbative FRW framework should therefore hold. However, counterarguments have been put forward involving subtleties in the choice of background employed in these treatments. For references finding a a small backreaction see e.g. \\cite{Ishibashi:2005sj, Paranjape:2008jc, Baumann:2010tm, Alonso:2010zv, Green:2010qy, Mattsson:2010ky} while arguments for significant backreaction can be found in e.g. \\cite{Rasanen:2006kp, Rasanen:2008it, Clarkson:2009hr, estimate, Collins:2010ea, Bochner:2011dr}. For recent articles and reviews on the subject with more extensive references see \\cite{Buchert:2011sx, arXiv:1103.2335, arXiv:1102.1449,arXiv:1103.5974, arXiv:1102.1015, arXiv:1105.0909, arXiv:1106.1693, arXiv:1103.2016, arXiv:1102.0408, Kolb, arXiv:1105.1886, arXiv:1005.0788}. In this paper we discuss cosmological backreaction in a novel manner, with a fully relativistic framework for a universe that begins as a perturbed Friedmann-Robertson-Walker, $\\Omega=1$, CDM universe. To be precise, we employ a gradient expansion to express the metric as a series of terms with an increasing number of spatial gradients and coefficients which are functions of proper time. The initial conditions are the standard adiabatic and Gaussian post-inflationary primordial perturbations. We use the synchronous gauge: our coordinate lines comove with CDM particles and our time hypersurfaces are labeled by their proper time. Of course the gradient series has to be truncated and thus does not capture the developing non-linearities entirely realistically. However, we argue that even if truncated this series can still provide a well motivated \\emph{model} for the true geometry which can be made increasingly accurate in principle. Furthermore, it extends into the non-linear regime, describing the collapse of initial over-densities and the rarefaction of initial under-densities which go on to form the voids dominating the cosmological volume.\\footnote{The synchronous gauge and a gradient expansion to study backreaction was first used in \\cite{Kolb:2005da} to argue that a non-perturbative approach is really needed to settle the issue. In the present work we use a truncated gradient series but also consider it in the non-perturbative regime, beyond its apparent radius of convergence. As explained in the text we argue that this can be used as a model for the metric of the universe which captures the evolution of over- and under-dense regions.} Here is the outline of the rest of the paper: In the following section we obtain the series solution for the metric in a gradient expansion using a Hamilton-Jacobi formulation. This approach, first developed in e.g. \\cite{Croudace:1993yt} and \\cite{Stewart:1994wq}, simplifies the calculation significantly compared to a more straightforward expansion of the standard Einstein equations. Our treatment follows a slightly different logical development. Then, in section 3, we apply these results to the backreaction problem by numerically evaluating the evolution of the average scale factor and the backreaction parameter $Q$. We find that, up to certain qualifications which we explain, backreaction leads to non-negligible deviations from the unperturbed background model. Our results indicate that the effect might be relevant, indeed crucial, for precision cosmology and the exploration of the dark sector. We summarize and discuss our findings in section 4 where future directions are also laid out. ", "conclusions": "\\subsection{Potential improvements to the model}\\label{sec:extensions} While our model is interesting and internally consistent and perhaps even an improvement in sophistication compared to earlier models, what we want is to know is the magnitude and nature of the backreaction in the real universe. The most obvious next step in improving our model is to add a cosmological constant term, $\\Lambda$. With that achieved, it would be possible to do a full Markov Chain Monte Carlo over the parameters of the model to see which set of parameters fits the observed data best. If our model, with backreaction, fitted the data better than the concordance model, without backreaction, then this would indicate that the assumptions we made to solve the various complications were well-motivated. However, another complication that is less easily solved would need to be addressed before comparisons to the real universe can be meaningfully made. In this paper we have calculated the backreaction that arises as structures form over a small range of scales. We can vary where this range lies, but we cannot make it arbitrarily wide. To compare the model to observation we need a reliable calculation of the \\emph{total} backreaction that arises over \\emph{all} scales. In principle this problem can be solved by considering a larger box; however, in practical terms, this can never be fully achieved. Given the hierarchical nature of structure formation, an ideal solution to this problem is a time-dependent grid spacing and box size. The time-dependence could be global, with the grid spacing $R$ set as a function of time $R(t)$. $R(t)$ would be chosen based on which scales start forming structures at which times and at which times structure growth freezes out. Alternatively, the time-dependence could be set locally. There could be a progressive ``zooming out'' as individual gridpoints expand into neighbouring gridpoints. Whenever two gridpoints collide it would be possible to stop considering them separately and to recalculate the metric from that time onwards as if they were one gridpoint. Precisely what metric to use for this combined gridpoint would need to be determined. Either a global or a local time-dependence to the grid spacing would face this difficulty. That is, the question of precisely what initial conditions, or background model, to use for the zoomed out metric. Should $t_0$ (which was set as a function of $h$) become a time-dependent quantity? If so, should the $h$ used in the transfer function also be dependent on the grid spacing and thus time? If the zooming out problem could be solved it would remove any problems that relate to how to fix the over expansion of gridpoints. However, if the zooming out problem is not solved, there are some improvements that can be made to our fixing methods. Firstly, it should be possible to pick $g_{\\rm rat}$ from observations, rather than to treat it as a free parameter, as we have here. Secondly, we chose to fix our expanding regions to the initial background model. This is not ideal. When any region reaches the point where we wish to fix its expansion, the \\emph{average} expansion rate, and more importantly, the average volume and density, are not the same as those of the initial background, extrapolated to this time. If the universe is to reach a new asymptotic FRW state it should be related to the new average density, not the one extrapolated from the initial background model. However, doing this is not easy. The deviation of the average density from the initial background model changes with time. Therefore, the correct background model to fix the expanding regions to also changes with time. When the total backreaction is $<10\\%$ these issues will be less important than the many other features of the model. However, when comparing the model to precise measurements, perhaps $10\\%$ of $10\\%$ is no longer an ignorably small quantity. \\subsection{Conclusions} Let us close by summarizing our methodology and the conclusions we can draw from it. We have approximated the evolving inhomogeneous synchronous gauge metric of a CDM $\\Omega=1$ universe with realistic initial conditions using a gradient expansion, and studied the backreaction on the average dynamics. This gradient expansion is an approximation to the true metric of such a universe which goes beyond the description of standard perturbation theory. Instead of being limited by the smallness of the inhomogeneities, it has a temporal range of validity set by the initial local 3-curvature. For timescales $t>t_{\\rm con}$ the gradient expansion does not converge anymore. This does not pose a real problem for studying regions with positive initial curvature which collapse at around that timescale and eventually virialize. However, it does present limitations in following the late time evolution of under-dense regions with negative initial curvature. Nevertheless, if the late time evolution of such regions is fixed in some prescribed manner, this approach offers a way to construct a realistic \\emph{model} even for the late-time metric of such a universe. Our methodology improves over previous approaches in that it is fully relativistic, goes beyond perturbation theory and does not treat the universe as a collection of unconnected under-dense and over-dense structures with high symmetry. Of course it does require modeling of the late time behaviour of the under-dense regions which is not naturally captured by the gradient expansion but such modeling can be simply parameterized and possibly informed by observation. We have attempted to achieve this in a simple manner in this paper but it can still be made more realistic in a number of ways. Given the approximations mentioned, we computed the backreaction of inhomogeneities in this model metric. Our qualitative results indicate that backreaction may constitute more than a percent effect and is thus highly relevant in future considerations of precision cosmology. Of course more realism is needed before definite statements can be made regarding the backreaction in our universe. There is another aspect in which such a model would aid in understanding the effects of inhomogeneities for cosmology. We have focused on the average dynamics but what we actually see is light propagating through the inhomogeneous universe and not the average scale factor. One could question whether the latter is at all relevant for our observations. To really assess the impact of inhomogeneities on observations one should trace light rays through the inhomogeneous spacetime and determine what observers would see \\cite{arXiv:1002.1232,arXiv:1109.2484}. This is easy to do in a box described by our metric. In particular we can compute what happens to the redshift of photons or the travel time though over-dense and under-dense regions. The great advantage of our approach is that this is easily calculable within our model. Therefore, even irrespective of whether our model describes the universe entirely accurately, we can answer the question of whether a large backreaction in the synchronous gauge corresponds to a large shift in the time it takes a photon to traverse the universe and/or the redshift it experiences. We will return to this issue and further improvements to our model in future work." }, "1112/1112.0281.txt": { "abstract": "We present the initial-final mass relation derived from 10 white dwarfs in wide binaries that consist of a main sequence star and a white dwarf. The temperature and gravity of each white dwarf was measured by fitting theoretical model atmospheres to the observed spectrum using a $\\chi^{2}$ fitting algorithm. The cooling time and mass was obtained using theoretical cooling tracks. The total age of each binary was estimated from the chromospheric activity of its main sequence component to an uncertainty of about 0.17 dex in log \\textit{t} The difference between the total age and white dwarf cooling time is taken as the main sequence lifetime of each white dwarf. The initial mass of each white dwarf was then determined using stellar evolution tracks with a corresponding metallicity derived from spectra of their main sequence companions, thus yielding the initial-final mass relation. Most of the initial masses of the white dwarf components are between 1 - 2 M$_{\\odot}$. Our results suggest a correlation between the metallicity of a white dwarf's progenitor and the amount of post-main-sequence mass loss it experiences - at least among progenitors with masses in the range of 1 - 2 M$_{\\odot}$. A comparison of our observations to theoretical models suggests that low mass stars preferentially lose mass on the red giant branch. %Wide binaries tend to be much older than nearby clusters. Thus, they contribute valuable limits to the low mass end of the initial-final mass relation where clusters provide sparse information. %The lack of a similar correlation among cluster stars suggests that the mechanism for mass loss in wide binaries differs from that in open clusters or that not enough low mass white dwarfs are observed in clusters to see such a relation. %Willson's model when the age-metallicity correction for progenitors of white dwarfs is taken into account. %% what keyword punctuation is appropriate. ", "introduction": "Over 90 percent of all stars shed at least half their mass as they evolve towards their final state - a white dwarf (WD). The initial-final mass relation (IFMR) represents a mapping between the mass of a WD remnant and the mass of its hydrogen-burning main-sequence (MS) progenitor. It also characterizes the amount of material stars with primordial masses M $\\lesssim$ 8 M$_{\\odot}$ recycle to the interstellar medium. Thus, it is of paramount importance to understanding the chemical enrichment and the efficiency of star formation in galaxies. This relation is also a key constraint on stellar evolution theory. %Unfortunately, currently it is not well constrained either by theory or by observation. One of the first attempts to empirically determine the IFMR was undertaken by Weidemann (1977). Reimers $\\&$ Koester (1982) and Koester $\\&$ Reimers (1996) presented observations of WDs in the open cluster NGC 2516 and obtained an IFMR using these WDs. Weidemann (2000; W00 hereafter) updated the IFMR by incorporating new theoretical and observational data. Claver et al. (2001) observed six WDs in the Praesepe open cluster and determined a monotonic IFMR. Williams et al. (2004; 2009) presented an empirical determination of the IFMR based on a spectroscopic analysis of massive white dwarfs in NGC 2168 (M35). They showed that the resultant white dwarf mass increases monotonically with progenitor mass for masses greater than 4 M$_{\\odot}$. Ferrario et al. (2005) re-evaluated the ensemble of data that has been used to determine the IFMR and characterized a mean IFMR about which there is an intrinsic scatter. They showed that a linear IFMR predicts a mass distribution in reasonable agreement with the Palomar-Green survey. Kalirai et al. (2005) determined the IFMR from observations of very faint WDs in the rich open cluster NGC2099 (M37). They found stars with initial masses between 2.8 and 3.4 M$_{\\odot}$ lose 70 - 75\\% of their mass during post-MS evolution. Dobbie et al. (2006) also constructed a new IFMR based on 11 WDs in the Praesepe open cluster. Rubin et al. (2008) constructed an IFMR based on 19 spectroscopically identified WDs in NGC 1039 (M34). Catal\\'{a}n et al. (2008a) studied the IFMR using six WDs in wide binaries and suggested the IFMR may not be a single-valued function. Kalirai et al. (2008) presented constraints on the low-mass end of the IFMR using older open clusters NGC7789 (\\textit{t} = 1.4 Gyr ), NGC 6819 (\\textit{t} = 2.5 Gyr) and NGC6791 (\\textit{t} = 8.6 Gyr). Later, Kalirai et al. (2009) extended the IFMR to lower masses using the globular cluster M4. Since most nearby clusters are relatively young, the initial masses of those WDs tend to be high. %So far, WDs with low initial masses of only a few clusters have been included in a determination of the IFMR (Kalirai et al. 2008; 2009). In this paper we investigate the IFMR using wide ``fragile\" binary systems containing a WD with a MS companion. The systems have relatively large orbital separations ($$ $\\sim$ 10$^{3}$ AU; Oswalt et al. 1993; Silvestri et al. 2001, 2002, 2005). Thus, one can safely assume that each component has evolved independently, unaffected by mass exchange or tidal coupling that complicate the evolution of closer pairs. Components of a given binary are coeval (Greenstein 1986). Essentially, each may be regarded as an open cluster with only two components. Although it is difficult to obtain ages for wide binaries as accurate as ages for open clusters, they tend to be nearer, brighter and are far more numerous than nearby clusters. Moreover, they span a much more continuous range in age. Catal\\'{a}n et al. (2008a; C08a hereafter) used six wide binaries to investigate the IFMR; three of the WDs had low initial mass ($<$ 2M$_{\\odot}$). Our sample contains additional WDs at the low initial mass limit. Previous research has indicated a large scatter in the empirical IFMR. What is the source of this scatter? Kalirai et al. (2005) found some weak evidence of a metallicity dependence in the IFMR. Kalirai et al. (2007) found evidence for enhanced mass loss at extremely high metallicities by studying the white-dwarf mass distribution in the supersolar-metallicity star cluster NGC 6791 ([Fe/H] = +0.4). Kalirai et al. (2009) found a relatively flat relation between mass loss and metallicity ([Fe/H] between -1.1 to solar metallicity) by extending these studies to WDs in the globular cluster M4. However, a clear relation between the metallicity and scatter in the IFMR has not been demonstrated (cf. Williams 2007; Catal\\'{a}n et al. 2008b, C08b hereafter). We investigated whether there is a metallicity dependence on the IFMR in our wide binary sample using the spectra of their MS companions. Section 2 provides an overview of the observations and reductions for our sample. The astrophysical properties of the WDs are discussed in section 3. The MS companions are discussed in section 4. In section 5, we present and discuss our IFMR. Section 6 compares our observations to theoretical models of post-MS mass loss. A discussion of the implications of our findings is given in section 7. ", "conclusions": "%Eight are DA type and two are DB type. The $T$$\\rm_{eff}$ and log $g$ of DA WDs were derived from the H line fits between the observed spectra and %theoretical spectra, while those of two DB WDs are obtained from the literature. The mass and cooling times of the WDs %were determined from model cooling tracks. The total ages of the binaries were estimated from the CA of their MS components. First, S$\\rm_{HK}$ is measured. Then we construct the calibration between our $S\\rm_{HK}$ and $S\\rm_{MW}$. After $S\\rm_{HK}$ transformed into $S\\rm_{MW}$, $R\\arcmin\\rm_{HK}$ could be calculated by Equation Noyes et al. (1984). Finally, the ages of each binary were determined by the relation between $R\\arcmin\\rm_{HK}$ and age (Soderblom et al. 1991). The metallicities of wide binaries were derived by template matches between the MS spectra and theoretical spectra. The difference between the total age and the cooling time of each WD was taken as the lifetime of its progenitor. The initial masses of 10 WDs were then determined by the evolution track. Our IFMR contains six WDs whose M$\\rm_{i}$ are lower than 2 M$_{\\odot}$. They contribute to the low initial mass limit not well sampled by clusters. Our WDs in wide binaries suggest a linear IFMR over the initial mass range 1.1 M$_{\\odot}$ to 4.1 M$_{\\odot}$: In this study, we constructed an empirical IFMR using 10 WDs in wide binaries. Our IFMR contains six WDs whose M$\\rm_{i}$ are lower than 2 M$_{\\odot}$. They contribute to the low initial mass limit that is not well-sampled by clusters. Our WDs in wide binaries suggest a linear IFMR over the initial mass range 1.1 M$_{\\odot}$ to 4.1 M$_{\\odot}$ (Equation 5). We compared our mass loss vs. metallicity relation to theoretical models for evolving lower MS stars ($<$ 2 M$_{\\odot}$). In general, the models predict a net mass loss and IFMR that agree with the values found from our observation within the current uncertainty of measurement. Kalirai et al. (2007) and Kalirai et al. (2009) tentatively found a metallicity dependence on the IFMR. We find that at least part of the scatter seen in the IFMR is correlated with metalllcity. Stars with lower metallicity apparently shed less mass when they become WDs. %{\\bf Except NGC 6791 (Kalirai et al. 2007) and M4 (Kalirai et al. 2009), IFMRs from most of open clusters have shown no such dependance on metallicity, }it is curious that evolution for stars in wide binaries seems to be different from that in open clusters. Perhaps this is because the Z dependance of mass loss is most pronounced in low mass progenitors, which are poorly sampled by observation of cluster. %Although only two DBs are in our current sample, we suspect the metallicity effect on the IFMR is different for DA and DB stars. More observational data will be needed to prove this. % %We made a comparison between our mass loss vs. [Fe/H] relation and theoretical model for lower mass WDs ($<$ 2 M$_{\\odot}$). %Overall, the mass loss models predict net mass loss and M$\\rm_{f}$ vs. M$\\rm_{i}$ that are very close to the values found from the white dwarf pairs. This will be true for other mass loss formulae with similar death-lines, as discussed by Willson (2007, 2008a, 2008b, 2009). In addition, the Z-dependence of our models is consistent with the data from these white dwarf stars. The bulk of the Z-dependence in the mass loss rates comes from the dependence of R on Z, with a smaller effect from the efficiency of forming dust (the gas/dust ratio) in the models, it is important to note. %% The displaymath environment will produce the same sort of equation as %% the equation environment, except that the equation will not be numbered %% by LaTeX. %% If you wish to include an acknowledgments section in your paper, %% separate it off from the body of the text using the" }, "1112/1112.3595_arXiv.txt": { "abstract": "We have conducted a series of numerical experiments with the spherically symmetric, general relativistic, neutrino radiation hydrodynamics code \\aboltz\\ to examine the effects of several approximations used in multidimensional core-collapse supernova simulations. Our code permits us to examine the effects of these approximations quantitatively by removing, or substituting for, the pieces of supernova physics of interest. These approximations include: (1) using Newtonian versus general relativistic gravity, hydrodynamics, and transport; (2) using a reduced set of weak interactions, including the omission of non-isoenergetic neutrino scattering, versus the current state-of-the-art; and (3) omitting the velocity-dependent terms, or observer corrections, from the neutrino Boltzmann kinetic equation. We demonstrate that each of these changes has noticeable effects on the outcomes of our simulations. Of these, we find that the omission of observer corrections is particularly detrimental to the potential for neutrino-driven explosions and exhibits a failure to conserve lepton number. Finally, we discuss the impact of these results on our understanding of current, and the requirements for future, multidimensional models. ", "introduction": "Colgate and White \\citeyearpar{CoWh66} were the first to propose that core-collapse supernovae may be neutrino-driven and performed the first numerical simulations of such events, launching more than four decades of research that continues to this day. A significant milestone occurred nearly two decades later with Wilson's discovery that delayed neutrino-driven explosions could be obtained. Based on his models, Wilson concluded \\citep{Wils85,BeWi85} that the stalled supernova shock wave could be revived via neutrino absorption on a time scale of several hundred milliseconds given the intense flux of neutrinos emerging from the proto-neutron star liberating the star's gravitational binding energy. Observations of the neutrinos from SN1987A, the first such observations of supernova neutrinos \\citep{BiBlBr87,HiKaKo87}, provided support for the central role of neutrinos in the explosion mechanism. State-of-the-art simulations today continue to explore Wilson's neutrino-driven explosion mechanism in the context of two- and three-dimensional models \\citep[e.g., see][]{BuLiDe07,MaJa09,BrMeHi09b,SuKoTa10}. Neutrinos are weakly interacting particles whose cross sections are energy dependent. Thus, unlike all other components in a supernova model, they are not well described as a fluid, except in the deepest layers, and their transition in space to non-fluid-like behavior depends on their energy. Instead, the evolution of the neutrino radiation field, particularly in the semi-transparent regime, is far better characterized by classical kinetics---specifically, the general relativistic Boltzmann kinetic equation \\citep[e.g., see][]{CaMe03}, \\notetoeditor{Superscripts should be placed to the left of subscripts in the following equation:} \\begin{equation} p^{\\hat \\mu} \\left( {\\Lambda^{\\bar\\mu}}_{\\hat\\mu} {e^\\mu}_{\\bar\\mu} \\frac{\\partial f}{\\partial x^\\mu} - {\\Gamma^{\\hat\\nu}}_{\\hat\\rho \\hat\\mu} p^{\\hat\\rho} \\frac{\\partial f}{\\partial p^{\\hat\\nu}}\\right) ={C}[f], \\label{eq:boltz} \\end{equation} where, for spherically symmetry, \\begin{eqnarray} \\frac{1}{E}{C}[f] &=& \\label{eq:collision} (1-f)j-\\chi f \\\\ \\nonumber &+&\\frac{1}{c}\\frac{1}{h^{3}c^{3}}E^{2}\\int d\\mu' R_{{\\rm IS}}(\\mu, \\mu',E)f \\\\ \\nonumber &-&\\frac{1}{c}\\frac{1}{h^{3}c^{3}}E^{2}f\\int d\\mu' R_{{\\rm IS}} (\\mu, \\mu',E)\\\\ \\nonumber &+&\\frac{1}{h^{3}c^{4}} (1-f) \\int dE'E'^{2}d\\mu' R_{{\\rm NIS}}^{{\\rm in}}(\\mu, \\mu',E,E') f \\\\ \\nonumber &-&\\frac{1}{h^{3}c^{4}} f \\int dE'E'^{2}d\\mu' R_{{\\rm NIS}}^{{\\rm out}}(\\mu, \\mu',E,E') (1-f) \\\\ \\nonumber &+&\\frac{1}{h^{3}c^{4}} (1-f) \\int dE'E'^{2}d\\mu' R_{{\\rm PR}}^{{\\rm in}}(\\mu, \\mu',E,E') (1-\\bar{f}) \\\\ \\nonumber &-&\\frac{1}{h^{3}c^{4}} f \\int dE'E'^{2}d\\mu' R_{{\\rm PR}}^{{\\rm out}}(\\mu, \\mu',E,E') \\bar{f}. \\end{eqnarray} Equation (\\ref{eq:boltz}) describes the evolution of the neutrino distribution function $f(t,x_1,x_2,x_3,\\mu_1,\\mu_2,E)$, which at time $t$ and spatial location $(x_1,x_2,x_3)$ supplies the distribution of neutrinos in direction cosines $(\\mu_1,\\mu_2)$ and energy $E$--i.e., the angular and spectral distribution of neutrinos. One such Boltzmann equation is solved for each flavor of neutrino---electron, muon, and tau (\\nue, \\numu, and \\nutau, respectively)---and for their antineutrino partners (\\nuebar, \\numubar, and \\nutaubar). The invariant collision term, ${C}[f]$, in equation~(\\ref{eq:collision}) is written using emission, $j$, absorption, $\\chi$, and scattering and pair kernels, $R$, following the forms often used for neutrino transport \\citep[e.g., see][]{MeMe99}, where $\\bar{f}$ is the distribution function for the partner antineutrino and $\\mu$ is the neutrino direction cosine. In equations (\\ref{eq:boltz}) and (\\ref{eq:collision}), $f$ is a function of $(\\mu,E)$, as well as position and time. The $(\\mu',E')$ dependence of $f$ and $\\bar{f}$ inside the integrals illustrates the physical coupling of all energies and angles for each neutrino species and of neutrino and antineutrino partners. The first term on the left-hand side of equation (\\ref{eq:boltz}) describes the time evolution of the local neutrino distribution owing to spatial transport through the volume of interest. The second, far more complex term on the left-hand side (\\momspder) describes the evolution of the local neutrino distribution in angle and energy as the result of (A) the coordinate system chosen, (B) special relativistic effects, and (C) general relativistic effects. In what follows, we will refer to (B) as ``observer corrections.'' Terms describing (A) depend on the choice of coordinate system. For example, in spherical-polar coordinates, the neutrino direction cosine relative to the outwardly pointing radial vector changes as the neutrino propagates through a local volume. This coordinate-system effect is included in \\momspder\\ and is present even in the absence of fluid motion or general relativity. For Cartesian coordinates, the neutrino direction cosines do not change as a result of the coordinate system choice alone and, consequently, such a term is absent. Terms describing (B) depend on the frame of reference chosen to measure the neutrino direction cosines and energies. The comoving frame, with neutrino direction cosines and energies measured in an inertial frame of reference instantaneously comoving with the stellar core fluid with which the neutrinos interact, is often used. Neutrino--matter interactions are naturally expressed in this frame. Given this choice, the terms in \\momspder\\ present a significant numerical challenge. Finding discrete representations that guarantee conservation of lepton number and energy is one of the most difficult aspects of modeling neutrino transport in stellar cores. This has been achieved for general relativistic, spherically symmetric flows \\citep{LiMeMe04}, providing the conceptual and implementation groundwork for achieving the same in axisymmetric (2D) and non-symmetric (3D) flows. Further theoretical foundations have been laid \\citep{CaMe03,CaLeMe05}; steps toward the development of a 2D Boltzmann solver have been taken \\citep{OtBuDe08}; and the challenge now is to fully implement lepton energy and number conserving discretizations in 2D and 3D models. For another choice of reference frame---measuring neutrino angles and energies relative to the inertial, ``lab\" frame of a distant observer---the terms encapsulating the special relativistic effects in \\momspder\\ are absent, simplifying the left-hand side of the Boltzmann equation. In such a frame of reference, the neutrino direction cosines and energies do not change from observer to observer in the frame. However, this simplification comes at a price because the neutrino--matter interactions are naturally described in the comoving frame. In the lab frame, a Lorentz transformation is required in order to express the comoving-frame neutrino--matter interactions in terms of the lab-frame direction cosines and energies, which introduces non-trivial velocity dependencies into the lab-frame collision term. One approach to the complexity of the lab-frame collision term is the ``mixed frame'' approach, which uses the lab-frame 4-momenta and an $\\mathcal{O}(v/c)$ Taylor-series expansion in energy of the comoving-frame emissivities and opacities \\citep{MiKl82}. \\citet{HuBu07} have proposed to use the mixed-frame approach for core-collapse simulations with extensions for non-isotropic and non-isoenergetic scattering. The mixed-frame approach may be difficult to extend to arbitrarily relativistic flows, and has not yet been used in the context of a full-physics core-collapse supernova simulation. In a general relativistic setting, such as core-collapse supernovae, we must contend with (B) and/or (C) regardless of the frame of reference chosen to describe the neutrino direction cosines and energies. Even for static general relativistic environments, angular aberration, gravitational red shift, and other effects occur, and the resulting terms in (C) are always present. Regardless of approach, comoving- or lab-frame, it is problematic to adapt the simplicity of both approaches, simultaneously simplifying the left- and right-hand sides of the Boltzmann equation, as has been done in \\citet{BuLiDe06,BuLiDe07}, \\citet{OtBuDe08}, and other models using the \\vulcan\\ code, although one can view the implementation in these works as steps toward a more complete description. They deploy a lab-frame approach for terms describing angular aberration and energy shift on the left-hand side (or assume such terms are unimportant in a comoving-frame approach), while simultaneously deploying a comoving-frame approach for the collision term describing the neutrino--matter interactions on the right-hand side. This is not a mixed-frame approach in the sense described above. It is an approach not based in any reference frame, and it is physical only for static cases in which there is no distinction between lab and comoving frames. One of the goals of this study is to investigate the importance of the terms in \\momspder\\ in a comoving-frame approach, and whether they can be ignored while using a comoving-frame approach for the collision term. Modeling general relativistic Boltzmann kinetics is also challenging because of the complexity of the collision term on the right-hand side of the Boltzmann equation, even in a comoving-frame formulation. Looking at equation (\\ref{eq:collision}), we see that the collision term describes the full, direct coupling of all neutrino angles and energies for each neutrino species, owing to neutrino isoenergetic (IS) scattering on nuclei and non-isoenergetic (NIS) scattering on electrons and nucleons. The pair creation and annihilation processes (PR) such as electron--positron annihilation and nucleon--nucleon bremsstrahlung also couple the angles and energies of the neutrino and antineutrino species of each flavor together. The coupling of all neutrino angles and energies through the relevant set of weak interactions dominates the computation associated with the solution of the neutrino Boltzmann equations. It has been argued \\citep{BuLiDe06,BuLiDe07,NoBuAl10} these couplings are subdominant and can be ignored, greatly simplifying the neutrino Boltzmann equations and significantly reducing the computational cost associated with their solution. A second goal of this study is to investigate whether or not such approximations to the collision term are realistic. The complete general relativistic Boltzmann equation was solved in spherically symmetric models of core-collapse supernovae by the Oak Ridge-Basel collaboration \\citep{LiMeTh01,LiMeMe04} and by Sumiyoshi and collaborators \\citep{SuYaSu05}. Achieving this in three-dimensional models of core-collapse supernovae presents a major challenge, one that will likely require sustained exascale resources to meet. The overarching goal of this study is to use general relativistic, spherically symmetric Boltzmann simulations to guide and, more importantly, set minimum requirements for accurate 2D and 3D simulations. We use the Oak Ridge-Basel code \\aboltz\\ in these studies to compare general relativistic--full weak interaction physics (\\grfull), Newtonian--full weak interaction physics (\\nfull), Newtonian--reduced weak interaction physics (\\nreduc), and Newtonian--reduced weak interaction physics--no observer correction (\\noc) models. These models will demonstrate the importance of general relativity, a complete weak interaction set and treatment, and the terms in \\momspder\\ to stellar core collapse and the post-core-bounce evolution. Current multidimensional models suggest that spherical symmetry is a reasonable approximation for the first 100--150 ms after bounce \\citep{MaJa09,BrMeHi09b,SuKoTa10}. Thus, the simulations presented here are relevant for discussing the initial conditions present for all multidimensional phenomena that might ensue; e.g., neutrino-driven convection and the standing accretion shock instability (SASI). ", "conclusions": "We have examined the consequences of removing (1) GR effects, (2) non-isoenergetic scattering and detailed nuclear EC opacities, and (3) observer corrections from spherically symmetric models of core-collapse supernovae. We have found that all of these changes, individually and especially when taken together, affect the progress of stellar collapse and the post-bounce evolution of the shock and core thermodynamic properties in significant ways, in constrast to the assessments made by \\citet{BuLiDe06,BuLiDe07} and \\citet{NoBuAl10}. We have computed variations in the shock radius, neutrino luminosities, and neutrino RMS energies as large as 60~km, 35~\\Bethes, and 10~\\mev, respectively, across the four models considered here. Omission of GR results in a less compact core and an unrealistically more favorable shock progression after bounce. Eliminating non-isoenergetic scatterings and simplifying electron capture on nuclei drastically reduces the core deleptonization and expands the homologous core at bounce. Omission of the observer corrections dramatically alters core deleptionization, the shock position, and neutrino luminosities after bounce, in part resulting from a complete breakdown of lepton number conservation. The lepton non-conservation and non-promotion of neutrino energy resulting from omitting observer corrections in our \\noc\\ model results in a compact, low-$Y_L$ core and a shock trajectory that is the least favorable of our models. The artificial loss of lepton number, lower neutrino luminosities, and the consequent lower neutrino heating rate and shallower shock trajectory may explain the lack of neutrino-driven explosions in models computed with \\vulcan\\ \\citep[see][]{BuLiDe07}, in contrast to the results reported by others \\citep{MaJa09,BrMeHi09b,SuKoTa10,TaKoSu11}. Moreover, the changes in $Y_{e}$ and $Y_{L}$, their gradients, and the entropy gradients that we see as we traverse the models shown here will change the location and strength of convectively unstable regions in the proto-NS and between the proto-NS and the shock. The lepton and entropy gradients in the proto-NS drive prompt convection, the entropy gradients between the proto-NS and the shock drive neutrino-driven convection, and these in turn seed and are seeded by the SASI. That is, the changes we have documented in this transport study have implications for all of the multidimensional phenomena we know to be important in multidimensional supernova models once spherical symmetry is broken. All of the ingredients (1)--(3) above must be included in multidimensional simulations of core-collapse supernovae to ensure physical fidelity. Their omission is not the only approximation used in current multidimensional simulations, some of which (like the ray-by-ray approximation) are inadequately understood and need to be better understood or phased out. Certainly, further examination of these approximations is required within the context of multidimensional simulations." }, "1112/1112.3240_arXiv.txt": { "abstract": "{} % {This work aims to study the unexplained sulfur depletion observed toward dense clouds and protostars.} {We made simulation experiments of the UV-photoprocessing and sublimation of H$_2$S and H$_2$S:H$_2$O ice in dense clouds and circumstellar regions, using the Interstellar Astrochemistry Chamber (ISAC), a state-of-the-art ultra-high-vacuum setup. The ice was monitored in situ by mid-infrared spectroscopy in transmittance. Temperature-programmed desorption (TPD) of the ice was performed using a quadrupole mass spectrometer (QMS) to detect the volatiles desorbing from the ice.} {Comparing our laboratory data to infrared observations of protostars we obtained a more accurate upper limit of the abundance of H$_2$S ice toward these objects. We determined the desorption temperature of H$_2$S ice, which depends on the initial H$_2$S:H$_2$O ratio. UV-photoprocessing of H$_2$S:H$_2$O ice led to the formation of several species. Among them, H$_2$S$_2$ was found to photodissociate forming S$_2$ and, by elongation, other species up to S$_8$, which are refractory at room temperature. A large fraction of the missing sulfur in dense clouds and circumstellar regions could thus be polymeric sulfur residing in dust grains.} {} ", "introduction": "\\label{intro} Sulfur is depleted in molecular clouds by a factor of 1000 compared to its estimated cosmic abundance (Tieftrunk et al. 1994), while in the diffuse interstellar medium the abundance of sulfur in the gas phase is comparable to the cosmic abundance. This suggests that there is a form of sulfur in the gas phase that was not observed in molecular clouds, which could be atomic sulfur, or alternatively that sulfur chemistry on icy grain mantles, present in dense clouds and regions around YSOs but not in the diffuse interstellar medium, plays an important role. There is compelling evidence that supports the role of dust grains on sulfur chemistry. Several gas-phase S-containing molecules were observed in hot cores, such as OCS, H$_2$S, H$_2$CS, SO, SO$_2$, HCS$^+$, and NS (van der Tak et al. 2003). Current gas-phase chemical models are unable to explain the abundances of S-species like HCS$^+$ and OCS measured toward protostars (Doty et al. 2004). The depletion of sulfur is observed not only in dense clouds, but also toward Class 0 and Class I sources (Buckle \\& Fuller 2003) and toward hot cores (Wakelam et al. 2004). The abundances of S-bearing species, including H$_2$S, suggest that these molecules are formed on grain surfaces and subsequently released to the gas phase. Several S-bearing molecules have been detected in comets. Among them H$_2$S has the largest abundance, from 0.2 to 1.5\\% relative to H$_2$O (Irvine et al. 2000). There seems to be a general agreement between the molecular abundances observed in circumstellar ices and in comets (Bockel\\'ee-Morvan et al. 2000). That suggests that H$_2$S is expected to be present in circumstellar ice mantles. The H$_2$S ice will be strongly processed by UV and ion irradiation (Garozzo et al. 2010; Grim \\& Greenberg 1987; Moore et al. 2007). We explore here the possibility that the missing sulfur atoms may be present in icy grain mantles. The cosmic abundance of sulfur is 1.23 $\\times$ 10$^{-5}$ N$_{\\rm H}$, or 37 times less abundant than oxygen (Snow \\& Witt 1996), and like oxygen, sulfur belongs to group 16 of the periodic table. The electronegativity values of S and O are 2.58 and 3.44 on the Pauling scale, and the electron affinities are 200 and 141 kJ mol$^{-1}$. The bond energy of an S-H bond is 363 kJ mol$^{-1}$, while that of an O-H bond is 458.9 kJ mol$^{-1}$, and therefore S-H bond formation is favored over O-H bond formation. Given that H is the most abundant element, S atoms will tend to form H$_2$S molecules because they impinge on icy grain mantles. The feature due to the \\ stretching mode of H$_2$S at 3.925 $\\mu$m (2548 cm$^{-1}$) has a band strength of $A$ $\\approx$ 2.9 $\\times$ 10$^{-17}$ cm molecule$^{-1}$ (Smith 1991), and $A$(H$_2$O) = 2.0 $\\times$ 10$^{-16}$ cm molecule$^{-1}$ (Hagen et al. 1981). Therefore, if we make the crude assumption that the H$_2$S abundance in the ice is roughly about 1/37 that of H$_2$O ice, based on the S/O = 1/37 cosmic abundance ratio, the absorbance area of H$_2$S relative to that of H$_2$O might just be 0.4\\%, which is close to the detection limit of most observations. With the exception of a weak CH$_3$OH absorption, the feature falls on a relatively {\\it clean} part of the mid-infrared spectrum. If present, it might be observable in the spectra of circumstellar or dense interstellar icy grains. Solid H$_2$S has not been detected in the interstellar medium. The presence of H$_2$S ice was inferred in W33A, a high-mass protostar (Geballe 1985; 1991), and a band at 4.9 $\\mu$m (2040 cm$^{-1}$) was attributed to OCS (Geballe 1985, Palumbo et al. 1995). The detection of OCS is possible since the 4.9 $\\mu$m band has a considerable band strength, $A$ = 1.5 $\\times$10$^{-16}$ cm molecule$^{-1}$ (Hudgins et al. 1993). In addition to OCS, SO$_2$ was detected in ice mantles (Boogert et al. 1997). The H$_2$S detection by Geballe is not fully supported in the literature. Van der Tak et al. (2003) argue that infrared observations do not support the assumption that H$_2$S is the main S reservoir in grain mantles, and they provide the ISO-SWS observations of W33A as an example. One of the problems for identifying the 3.925 $\\mu$m (2548 cm$^{-1}$) band of H$_2$S in H$_2$O-rich ice mantles was the lack of laboratory spectra of H$_2$S embedded in an H$_2$O matrix, which is expected to affect this band significantly. Ultraviolet emission spectra of the coma of comet IRAS-Araki-Alcock showed the presence of S$_2$, and the spatial profiles indicate a release of this species directly from, or very close to, the nucleus (A'Hearn et al. 1983). Diatomic sulfur was later found in the comet Hyakutake (Laffont et al. 1996). Based on the formation of S$_2$ at 12 K from irradiation of dirty ice containing H$_2$S, Grim \\& Greenberg (1987) suggested that S$_2$ was formed in interstellar ice mantles that ultimately aggregate into comets. Subsequently, fast reactions between OCS and metastable S, produced during the dissociation of CS$_2$, were proposed. This supports the formation of S$_2$ in the coma (e.g., A'Hearn et al. 2000). The formation of polymeric sulfur in interstellar environments involves the dissociation of H$_2$S and might, therefore, serve as a probe of energetic processing of the precometary ice. Ion irradiation experiments of ice analogs containing CO, CH$_3$OH, and S-bearing species lead to formation of OCS and CS$_2$ molecules (Ferrante et al. 2008; Garozzo et al. 2010). Previously, we performed an experiment consisting of the photoprocessing of H$_2$O:CO:NH$_3$:H$_2$S ice followed by warm-up to room temperature. This residue was analyzed by means of gas chromatography coupled to mass spectroscopy. Among the residue products several N-heterocycles and a number of S-bearing molecules were detected. S-polymers, S$_6$ through S$_8$, were formed, resulting from the S atoms released after photodissociation of H$_2$S ice. But also pentathian (S$_5$CH$_2$), hexathiepan (S$_6$CH$_2$), and c-(S-CH$_2$-NH-CH$_2$-NH-CH$_2$) were detected (Mu\\~noz Caro 2002). The presence of S-containing refractory molecules in icy grain mantles, like sulfur polymers, might be the reservoir of the missing sulfur in dense clouds and circumstellar environments (Wakelam et al. 2005). To approach the S-depletion dilemma in dense clouds and YSOs, we carried out a series of experiments on the deposition of pure H$_2$S or H$_2$S in an H$_2$O-matrix under UHV conditions at 7 K, to mimic interstellar/circumstellar conditions followed by warm-up. We measured the mid-infrared spectra of both pure H$_2$S ice and H$_2$S in an H$_2$O ice matrix at different temperatures from 7 K to sublimation. Similar spectra were reported by Moore et al. (2007) at higher temperatures. Using a quadrupole mass spectrometer (QMS), we obtained the temperature programmed desorption (TPD) plots showing the abundances of the molecules released to the gas phase as a function of temperature during warm-up. These experiments were repeated, including UV irradiation of the ice. The infrared spectra of H$_2$S ice measured in the laboratory were compared with spectroscopic observations performed by ISO, providing new upper limits on the H$_2$S abundance toward protostars. The layout of this paper is as follows. In Sect.~2 we describe the experimental protocol. The experimental results are presented in Sect.~3. The astrophysical implications are discussed in Sect.~4, and the main conclusions summarized in Sect.~5. ", "conclusions": "\\label{conc} We determined the desorption temperature of H$_2$S ice, which depends on the initial H$_2$S/H$_2$O ratio. Pure H$_2$S ice desorbs around 82 K. When H$_2$S is present in an H$_2$O ice matrix, a fraction of the H$_2$S co-desorbs with H$_2$O in the 130--170 K temperature range, showing two maxima around 145 and 163 K. These results agree with Collings et al. (2004). Comparison of the laboratory infrared spectra of H$_2$S, pure or in an H$_2$O matrix, with ISO observations of protostars W33A and IRAS18316-0602 provided upper limits of 0.7 and 0.13\\% on the solid H$_2$S abundance relative to H$_2$O. These values are too low to explain the S depletion observed toward dense clouds and circumstellar regions. Another reservoir of S in these regions could be the products of H$_2$S ice photoprocessing. It was found that solid H$_2$S, which expected in icy grain mantles, photolyzes very readily, leading to the formation of several species, including H$_2$S$_2$, HS$^{\\cdot}$, HS$_2^{\\cdot}$, S$_2$, and H$_2$SS. The set of reactions leading to these species is known, and it was possible to determine the values of some of the rate constants involved. If H$_2$S was present in an H$_2$O ice matrix the SO$_2$, SO$_4^=$, HSO$_3^-$, HSO$_4^-$, H$_2$SO$_2$, H$_2$SO$_4$, and H$_2$S$_2$ photoprocessing products were formed in our experiments. Proton bombardment of H$_2$S:H$_2$O ice led to formation of H$_2$S$_2$ and SO$_2$, while a similar processing of SO$_2$:H$_2$O ice formed H$_2$O$_2$, H$_3$O$^+$, HSO$_3^-$, HSO$_4^-$, and SO$_4^=$ (Moore et al. 2007). With the exception of H$_2$SO$_4$, all the above products of H$_2$S:H$_2$O ice photoprocessing desorbed between 100-200 K. S-polymers from S$_3$ to S$_8$ are also products of H$_2$S photoprocessing; these species can stick on grains even after sublimation of the ice mantle (Mu\\~noz Caro 2002, this work). In general, given the expected relative low abundances of S-polymers in ice mantles, they would be difficult to detect by infrared spectroscopy, but some of them will be observable with ALMA in the gas phase (Wakelam et al. 2005). The UV-irradiation of H$_2$O:CO:CH$_3$OH:NH$_3$:H$_2$S ices is left for future work. It was found that S$_8$ was by far the most abundant refractory product in these experiments, even for low initial H$_2$S ice abundances on the order of 2\\% relative to H$_2$O ice (Mu\\~noz Caro 2002). Including C in H$_2$S-containing ices, in the form of abundant ice components such as CH$_3$OH or CO, may lead to species like OCS, which is observed in circumstellar ice mantles." }, "1112/1112.0902_arXiv.txt": { "abstract": "{ We review our knowledge of the origin and phenomenology of low states (also known as VY Scl states) in cataclysmic variables, making special emphasis on the cataclysmic variable population found just above the period gap. This orbital period range between approximately 3 and 4 hours is the preferred land for the elusive SW Sextantis stars, which are believed to be about to enter the gap when their donor stars become fully convective. Despite their main role in our understanding of the whole picture of cataclysmic variable evolution, the study of their component stars is almost impossible due to the extreme veiling brightness of the accretion disc during the high state. Here we present the first steps toward the characterisation of the white dwarfs and the donor stars in these VY Scl systems in the low state, and the discovery of sporadic accretion events and satellite emission lines like those observed in the AM Her systems in the low state, which are likely related to magnetic activity of the fast rotating secondary stars. ", "introduction": "Cataclysmic variables (CVs) are occasionally caught fading toward states of greatly diminished brightness, or `low states' (also known as VY Scl states). During these unusual quiescent stages CVs shine $\\sim$3-5 mag fainter than they do in the high state, and can stay at that level for days, months or even years before returning to the high state. Low states seem to occur independently of the magnetic field of the white dwarf (WD): they are observed in practically all AM Her stars (polars), some intermediate polars, and many weakly-magnetic CVs, including a large fraction of nova-likes and a number of dwarf novae and Z Cam stars. The causes and effects of low states have been previously reviewed by \\cite{king+cannizzo98}, \\cite{warner99}, and \\cite{hessman00}. The main driver of the low-state phenomenology came from the study of the discless AM Her stars. Without an accretion disc any large brightness drop must be caused by a significant decrease in the magnetic accretion luminosity, which is proportional to the mass transfer rate from the donor star. Therefore, the typical photometric changes seen during low states are directly related to variations in the rate at which the donor star provides matter. The picture, however, is not as simple as a brightness drop followed by a relatively long quiescence phase and a return to the high state. As Fig.~\\ref{fig1} shows CVs in the low state are always {\\em eager} to recover, displaying quick brightness changes between the high and the low state, and even getting stuck at intermediate states for a while. As mentioned, there is broad agreement that mass transfer from the donor star through the L1 point temporarily stops or greatly reduces during a low state, but the exact mechanism still remains unknown. \\cite{livio+pringle94} proposed the accumulation of starspots close to the L1 point as a way of inhibiting Roche-lobe overflow. Only gas provided by the magnetic activity or stellar wind of the donor star would then be available for accretion \\citep[e.g.][]{hessmanetal00}. In fact, entanglement of the magnetic fields of both stars has been proposed to explain the line emission patterns observed in AM Her during the low state \\citep[][and references therein]{kafkaetal08}. The observation of low states is therefore a unique opportunity to study the WDs and donor stars in CVs, especially the magnetic activity of the fast rotating donors, their solar-like activity cycles and the interplay with the strong WD magnetic field if present. On the other hand, the large mass accretion rate of nova-like CVs in the high state make their discs the dominating source of light, thus veiling the two stars and preventing dynamical measurements of their masses and other fundamental parameters such as the donor spectral type or the mean magnetic field of the WD from being done. It is only during deep low states that the WDs and the donor stars become visible for study. ", "conclusions": "Although we still don't exactly know what causes the low states in CVs, the observation with 8--10m class telescopes is providing new insights to the problem. Obtaining accurate stellar masses and radii in CVs thought to be close to enter the period gap is fundamental to test the current CV evolution theories. In addition, time-resolved, multifrequency observations during low states will help address open questions like the possible presence of a magnetic white dwarf or a remnant accretion disc in the CVs populating the 3--4 h orbital period region. \\begin{figure} \\includegraphics[width=7cm,angle=-90]{PabloRodriguezGil_CVs_2011_01_fig08.ps} \\caption{ \\footnotesize H$\\alpha$ trailed spectra diagram showing the satellite lines and their orbital variation. Black corresponds to maximum emission.} \\label{fig5} \\end{figure}" }, "1112/1112.2136_arXiv.txt": { "abstract": "A new three-body method is used to compute the rate of the triple-alpha capture reaction which is the primary source of $^{12}$C in stars. In this work, we combine the Faddeev hyperspherical harmonics and the R-matrix method to obtain a full solution to the three-body $\\alpha+\\alpha+\\alpha$ continuum. Particular attention is paid to the long range effects caused by the pairwise Coulomb interactions. The new rate agrees with the NACRE rate for temperatures greater than $0.07$ GK, but a large enhancement at lower temperature is found ($\\approx 10^{12}$ at $0.02$ GK). Our results are compared to previous calculations where additional approximations were made. We show that the new rate does not significantly change the evolution of stars around one solar mass. In particular, such stars still undergo a red-giant phase consistent with observations, and no significant differences are found in the final white dwarfs. ", "introduction": " ", "conclusions": "" }, "1112/1112.5185_arXiv.txt": { "abstract": "We investigate neutrino processes for conditions reached in simulations of core-collapse supernovae. Where neutrino-matter interactions play an important role, matter is partially degenerate, and we extend earlier work that addressed the degenerate regime. We derive expressions for the spin structure factor in neutron matter, which is a key quantity required for evaluating rates of neutrino processes. We show that, for essentially all conditions encountered in the post-bounce phase of core-collapse supernovae, it is a very good approximation to calculate the spin relaxation rates in the nondegenerate limit. We calculate spin relaxation rates based on chiral effective field theory interactions and find that they are typically a factor of two smaller than those obtained using the standard one-pion-exchange interaction alone. ", "introduction": "Understanding the mechanisms responsible for supernovae and neutron star formation requires a knowledge of the equation of state and transport mechanisms in matter at densities of the order of that in atomic nuclei and at temperatures ranging up to $10^{11}$K. Despite almost half a century of work on the subject, the question of how a fraction of the large thermal energy from the collapse of the core is transferred to the outer stellar layers, thereby causing a supernova explosion, is one that has yet to find a convincing answer. Because of the high matter density in the new-born protoneutron star, most macroscopic transport processes are ineffective, and a variety of other mechanisms have been considered. These include energy transfer by neutrinos that interact with matter via weak interactions \\citep{Colgate.White:1966,Bethe.Wilson:1985}, magnetic fields in combination with rotation \\citep{Ardeljan.Bisnovatyi.ea:2004,Kotake.Sato.Takahashi:2006}, convection \\citep{Herant.Benz.ea:1994}, or pressure waves and Alfv\\'en waves that may be emitted by the protoneutron star \\citep{Burrows.Livne.ea:2007,Sagert.Fischer.ea:2009}. Some of these mechanisms lead to very characteristic features in the gravitational wave \\citep{Ott:2009} and neutrino signatures \\citep{Dasgupta.Fischer.ea:2010}, and these will be constrained by observations of the next galactic supernova. Other mechanisms have been tested in simulations, which demonstrate that, with current microphysical input, neutrino transport alone is insufficient to generate explosions \\citep{Liebendoerfer.Mezzacappa.ea:2001,% Rampp.Janka:2002,Thompson.Burrows.Pinto:2003}. However, it is possible to obtain explosions when neutrino heating and fluid instabilities are combined in axisymmetric models of stellar core collapse \\citep{Buras.Rampp.ea:2006,Marek.Janka:2009,Bruenn2009,% Suwa:2009py,Brandt:2010xa}. In all the above mechanisms, transport by neutrinos is crucial: it is the dominant process for energy emission, it contributes to the transport of energy and lepton number, it influences the radial entropy and lepton fraction gradients that determine the stellar structure and fluid instabilities, and it gives rise to the neutrino luminosities and spectra, which will become the main observables for probing the properties of matter at high density in the next nearby supernova event. Calculations of rates of neutrino processes in dense matter shortly after the discovery of weak neutral currents were reviewed by \\citet{Freedman}. In the early work, the particles participating in the reactions were taken to be free, but, subsequently, effects of strong interactions were taken into account \\citep{sawyer, Iwamoto}. More recently, detailed calculations of rates of neutrino processes have been performed within a mean-field approach (the random phase approximation) by \\citet{Reddy,BurrowsSawyer,Reddy2,% BurrowsReddyThompson}. One effect not included in these calculations is that excitations can decay due to interactions in the dense medium. As stressed in \\citet{RaffeltSeckel,RaffeltSeckelSigl,HR}, this can lead to an energy transfer in neutrino processes considerably greater than that predicted on the basis of excitations with infinitely long lifetimes (i.e., from recoil effects alone). \\citet{LOP} showed how to include these damping effects in a mean-field approach, and a unified approach to structure factors was described by \\citet{LPS}. Detailed calculations were performed based on chiral effective field theory (EFT) interactions in \\citet{Bacca} for degenerate neutrons. The prime purpose of the present paper is to extend these studies to partially degenerate and nondegenerate conditions. The plan of the paper is as follows. In Section~\\ref{conditions}, we analyze the results of simulations of core-collapse supernovae and determine the conditions for which it is important to know rates of neutrino processes. These results point clearly to the need for a better understanding of neutrino properties in regions where nucleons are partially degenerate or nondegenerate. In Section~\\ref{formalism}, we develop a general formalism based on Landau's theory of normal Fermi liquids for calculating structure factors of strongly interacting matter. The spin relaxation rate for partially degenerate conditions is derived in Section~\\ref{time} and the nondegenerate limit is studied in Section~\\ref{nondegen}. A key quantity is the spin relaxation rate in partially degenerate neutron matter, and we calculate this in Section~\\ref{results} based on the one-pion-exchange approximation for nucleon-nucleon interactions, which is the standard one used in simulations \\citep{HR}, and from chiral EFT interactions. Particular attention is paid to conditions of importance in supernova simulations. Finally, we summarize and give future perspectives in Section~\\ref{summary}. ", "conclusions": "\\label{summary} We have analyzed simulations of core-collapse supernovae and have shown that in the post-bounce phase the relevant conditions for neutrino processes are partially degenerate or nondegenerate. We then developed a formalism for calculating neutrino rates in strongly interacting matter by generalizing to the partially degenerate regime the approach used for degenerate matter in \\citet{LPS}. The resulting spin dynamical structure factor takes into account both mean-field effects and collisions between excitations in neutron matter. We then calculated the spin relaxation rate, a key ingredient in the structure factor which enters expressions for the rates of neutrino processes. This was done at two levels of NN interactions, the OPE approximation, which is commonly used in simulations \\citep{HR}, and chiral EFT interactions. We have found that chiral NN interactions lead to a reduction of the spin relaxation rate typically by a factor of two for a broad range of conditions. This reduction is similar to what previously has been found in the degenerate regime \\citep{Bacca}. We have also found that our OPE rate, where the width is obtained consistently by solving the Boltzmann equation, differs from the OPE results of \\citet{HR}. This may be a consequence of the imposition of a normalization condition in \\citet{HR}. Moreover, our results demonstrate that the nondegenerate limit is an excellent approximation for the conditions encountered in the post-bounce phase of matter at subnuclear densities in supernova simulations. Future directions include the study of many-body contributions and of many-body forces and electroweak currents in chiral EFT, and the extension of the calculations to mixtures of neutrons and protons, where the central parts of nuclear interactions can cause relaxation of the axial charge, because of the different axial charges of the neutron and proton. These extensions will be greatly simplified by the finding that the nondegenerate limit provides an excellent approximation for the relevant supernova conditions. Throughout the paper we have assumed that the basic coupling of the weak neutral field (that of the $Z$ boson) to nucleons is via a one-body operator. However, the spin relaxation effect that we have considered amounts to coupling a single quasiparticle-quasihole pair to two quasiparticle-quasihole pairs. In other words, strong interactions have generated a two-body contribution to the weak charge operator. However, not all two-body contributions to the operator are included in calculating relaxation effects by this procedure. The approximation we have made includes only contributions with a single quasiparticle in an intermediate state, which are enhanced by $1/\\omega$. Consideration of other two-body contributions to the weak charge operator is left for future studies." }, "1112/1112.1009_arXiv.txt": { "abstract": "{} {A high density portion of the Orion Molecular Cloud 1 (OMC-1) contains the prominent, warm Kleinmann-Low (KL) nebula % plus a farther region in which intermediate to high mass stars are forming. Its outside is affected by ultraviolet radiation from the neighboring Orion Nebula Cluster and forms the archetypical photon-dominated region (PDR) with the prominent bar feature. Its nearness makes the OMC-1 core region a touchstone for research on the dense molecular interstellar medium and PDRs.} {Using the Atacama Pathfinder Experiment telescope (APEX), we have imaged the line emission from the multiple transitions of several carbon monoxide (CO) isotopologues over the OMC-1 core region. Our observations employed the $2\\times7$ pixel submillimeter \\champp\\ array to produce maps ($\\sim300\\arcsec\\times350\\arcsec$) of \\TWCO, \\THCO, and \\CXVIIIO\\ from mid-$J$ transitions ($J=6-5$ to $8-7$). We also obtained the \\THCO\\ and \\CXVIIIO\\ $J=3-2$ images toward this region.} {The \\TWCO\\ line emission shows a well-defined structure which is shaped and excited by a variety of phenomena, including the energetic photons from hot, massive stars in the nearby Orion Nebula's central Trapezium cluster, active high- and intermediate-mass star formation, and a past energetic event that excites the KL nebula. Radiative transfer modeling of the various isotopologic CO lines implies typical \\HH\\ densities in the OMC-1 core region of $\\sim 10^{4}-10^{6}$ cm$^{-3}$ and generally elevated temperatures ($\\sim 50-250$ K). We estimate a warm gas mass in the OMC-1 core region of 86--285 \\Msol.} {} ", "introduction": "\\begin{table*}% \\centering \\caption{\\label{table1}Observational parameters} \\begin{tabular}{lcrrcccr} \\hline \\hline Molecule/Line & Frequency\\tablefootmark{a} & $E_{\\rm up}/k$ & $\\theta_{\\rm MB}$ & $\\eta_{\\rm MB}$ & Receiver & PWV & $T_{\\rm{sys}}$\\\\ &(MHz) & (K) & (\\arcsec) & & & (mm) & (K)\\\\ \\hline \\CXVIIIO\\ $J=3-2$& 329330.553 & 32 &18.0 & 0.75 & APEX-2a & $\\sim1.0$ & 220 \\\\ \\THCO\\ $J=3-2$ & 330587.965 & 32 &18.0 & 0.75 & APEX-2a & $\\sim1.0$ & 210 \\\\ \\CXVIIIO\\ $J=6-5$& 658553.278 & 111 & 9.0 & 0.47 & \\champp\\ & $\\lesssim0.7$ & 1600 \\\\ \\TWCO\\ $J=6-5$ & 691473.076 & 116 & 8.6 & 0.47 & \\champp\\ & $\\lesssim0.5$ & 1800 \\\\ \\TWCO\\ $J=7-6$ & 806651.806 & 155 & 7.4 & 0.45 & \\champp\\ & $\\lesssim0.5$ & 3900 \\\\ \\THCO\\ $J=8-7$ & 881272.808 & 190 & 6.7 & 0.45 & \\champp\\ & $\\lesssim0.7$ & 3700 \\\\ \\hline \\end{tabular} \\tablefoot{Columns are, from left to right, CO isotopologue and transition, frequency, energy of the transition's upper energy level above the ground-state, the HPBW beam size, the main beam efficiency, the receiver used, the average precipitable water vapor column and the average system temperature during the observations. \\tablefoottext{a}{Taken from the Cologne Database for Molecular Spectroscopy \\citep[CDMS,][]{Mueller2005}\\footnote{http://www.astro.uni-koeln.de/cdms/}.} } \\end{table*} The Orion Molecular Cloud 1 (OMC-1) is a complex region of the interstellar medium (ISM) stretching over more than 2.4 pc ($20\\arcmin$, roughly north-south) on the sky \\citep{Kutner1976}. Its densest part, toward which the Great Orion Nebula (M42), a classical compact ''blister'' \\HII\\ region, and its associated Orion Nebular Cluster (ONC) appear in projection, is one of the best-studied regions in astronomy. It is a test bed for studies of (proto)stars and clusters and the formation of low-, intermediate-, and high-mass stars \\citep[for a review of this region, see][]{O'Dell2001,O'Dell2008}. Much of this region's prominence is due to its distance of just $414\\pm7$ pc \\citep{Menten2007}, which makes the ONC and OMC-1 the closest regions of recent (few million years old) and ongoing high-mass star formation. In the following, we shall use the term ``OMC-1 core'' or even just OMC-1 for the roughly $8\\arcmin\\times8\\arcmin$- or 1 pc$^{2}$-sized dense molecular cloud region, which is located closely (0.1--0.2 pc) behind M42 and most of the stars in the ONC \\citep[see][the latter give a comprehensive overview of this region and its phenomena]{Zuckerman1973,Genzel1989}. The OMC-1 core region may be divided into three main zones, all of which show bright (sub)millimeter wavelength emission from warm dust and molecular gas: the Becklin-Neugebauer/Kleinmann-Low (BN/KL) region, Orion South (OMC-1S or Orion-S) and the Orion Bar. The Orion BN/KL and Orion South regions are considered to be ``twin'' high-mass star forming regions because they have similar masses ($\\lesssim$ 100 \\Msol) and bolometric luminosities \\citep[$10^4-10^5$ \\Lsol,][]{Mezger1990,Drapatz1983} and show comparable levels of star-forming activity \\citep{O'Dell2008}. The BN/KL region harbors the eponymous ``hot core'' \\citep{Masson1984}, which was (and commonly still is) taken to be the prototype of the hot dense regions observed around many newly formed stars \\citep[see, e.g.,][]{Kurtz2000}. An interesting alternative explanation for this region's energetics (other than being powered by an embedded central heating source) is a protostellar merger event that released a few times $10^{47}$ erg of energy about 500 years ago \\citep{Bally2005,Zapata2011,Bally2011}. The Orion Bar is a well-described photon-dominated region (PDR) located at the side of the OMC-1 core, facing M42, which is heated and partially ionized by far-ultraviolet (FUV) photons from the young massive stars (most of them from $\\theta^1$ C, a spectral type O5--O7 star) that form the ``Trapezium'' at the center of the ONC \\citep[see, e.g.,][]{Hollenbach1997,Walmsley2000}. In addition, the Orion Bar appears to be located at the edge of the \\HII\\ blister tangential to the line of sight. On giant molecular cloud (GMC) scales, low-rotational level ($J$) \\TWCO\\ emission (commonly from the $J=1-0$ line) of the ambient, low density gas is usually used to trace the mass of the molecular ISM under a range of assumptions \\citep{Bloemen1984,Dame1985}. However, in massive star forming regions with much higher densities and temperatures, observations of the submillimeter and far-infrared (FIR) wavelength mid- or high-$J$ transitions of \\TWCO\\ and its $^{13}$C and $^{18}$O isotopologues are required for determinations of the gas temperature and density, usually in conjunction with Large Velocity Gradient (LVG) radiative transfer and PDR modelings. Much of the \\TWCO\\ emission from the OMC-1 core has been proposed to arise from the neutral and partially ionized back side of the \\HII\\ blister, i.e., the PDR \\citep{Genzel1989}. The pioneering \\textit{submillimeter} observations of the \\TWCO\\ $J=7-6$, $6-5$ and \\THCO\\ $J=7-6$ transitions by \\citet{Schmid-Burgk1989} and \\citet{Graf1990} revealed high density gas ($n\\geq$ $10^{4}$ cm$^{-3}$) and elevated temperatures ($T\\geq$ 50 K) in all parts of the OMC-1 core region. Observations of the even more highly excited \\TWCO\\ $J=9-8$ line by \\citet{Marrone2004} with the Receiver Lab Telescope (84\\arcsec\\ resolution) showed that the hot and broad velocity emission in the line profiles arises mainly from the BN/KL region, while much of the narrower ($\\sim3$--6 \\kms\\ wide) emission arises from the PDR. This was also observed in the spectra of the \\TWCO\\ $J=9-8$ transition taken by \\citet{Kawamura2002}, which indicated warm molecular gas from the extended ``quiescent ridge'' region, north of Orion BN/KL. Furthermore, the observations of \\TWCO\\ $J=7-6$ and $J=4-3$ made by \\citet{Wilson2001} together with LVG modeling suggest that the broad line widths of these lines from the BN/KL region are probably due to shock heating, while most of the narrow line extended emission can be explained by PDR models. Recently, \\citet{Furuya2009} presented the BN/KL region images in the \\TWCO\\ $J=7-6$ line. Clearly, multi-transition, high angular resolution \\textit{imaging} studies are needed to disentangle the relative contributions of shock and radiative excitation to the CO emission, and to identify possible energy sources. All of the spectral line observations described above report either single pointing toward selected positions in OMC-1 or limited mapping of special areas such as the Bar \\citep{Lis1998} or the BN/KL region \\citep{Furuya2009}. The only exceptions are the \\TWCO\\ $J=7-6$ map by \\citet{Schmid-Burgk1989} and the \\TWCO\\ $J=4-3$ and $J=7-6$ maps by \\citet{Wilson2001}. However, the first of these, which covers the whole OMC-1 core, was taken with the poor resolution of the Kuiper Airborne Observatory. Here we present the first high (better than $10\\arcsec$) resolution large scale maps of the OMC-1 core using the $2\\times7$ pixel submillimeter \\champp\\ array receiver on the APEX telescope. Seven of the receiver units cover the 620 to 720 GHz frequency range and seven the 780 to 950 GHz range, allowing imaging of two lines simultaneously. As we shall see in \\S\\ref{obs}, we imaged the OMC-1 core in \\TWCO\\ $J=6-5$ plus $J=7-6$ in one run and \\THCO\\ $J=8-7$ plus \\CXVIIIO\\ $J=6-5$ in another. The fast mapping afforded by the 14 unit instrument resulted in a superior image consistency. The mid-$J$ CO data were complemented by the maps of the \\THCO\\ and \\CXVIIIO\\ $J=3-2$ lines. ", "conclusions": "\\subsection{Large scale maps -- overall morphology} Figure \\ref{map1} shows $\\approx6\\arcmin\\times8\\arcmin$-size maps of the OMC-1 core in the intensity of the \\TWCO\\ $J=6-5$ and $J=7-6$ lines integrated over the local standard of rest (LSR) velocity interval from $-$25 to $+$30~\\kms. The emission distributions of both lines, clearly resolved in our observations, resemble each other closely and are consistent with the JCMT/SCUBA 850 \\micron\\ image obtained by \\citet{Johnstone1999}, where several regions are discernible. The most prominent of which are: (a) the emission maximum centered on the BN/KL region, (b) a secondary maximum coinciding with Orion South (OMC-1S), (c) a region $\\approx3\\arcmin$ east of the Trapezium, which is known as Orion East (OMC-1 E; see below), and (d), the prominent straight Orion Bar PDR. The bar actually appears connected to the Orion South region, forming a mirrored letter L shape. Furthermore, there are two minor maxima at $1\\farcm8$ N and $3\\farcm3$ NNE of the BN/KL peak position. Apart from their spatial distributions, the regions above can be distinguished by their centroid LSR velocities, $V_{\\rm LSR}$, and their line widths, $\\Delta V$. The Orion BN/KL region displays a spectacularly broad line up to $200$~\\kms\\ wide, centered around 6~\\kms\\ with a narrow ($\\Delta V \\approx 5$~\\kms) feature at $V_{\\rm LSR}\\approx8$~\\kms\\ superposed. Orion South is characterized by $\\sim6$~\\kms\\ wide lines, with higher values at positions with outflow activities. The \\TWCO\\ $J=6-5$ and $J=7-6$ lines observed toward the Orion Bar have FWHM widths of $\\sim4$~\\kms\\ and are centered around $V_{\\rm LSR}=10$~\\kms. The Orion East region has previously been mapped by \\citet{Houde2004} and \\citet{Herrmann1997}, but otherwise has received relatively little attention. The observed line widths and LSR velocities of Orion East, as well the high intensity of its CO emission ($T_{\\rm{MB}}\\gtrsim 160$ K) clearly demonstrate its PDR nature. All our measurements are in good agreement with the results of \\citet{Schmid-Burgk1989}, \\citet{Graf1990}, \\citet{Wilson2001}, \\citet{Kawamura2002}, and \\citet{Marrone2004}. The measurements of the CO isotopologues in the OMC-1 core are summarized in Table \\ref{table2}. Figure \\ref{map2} shows the integrated intensity of \\THCO\\ $J=8-7$ and \\CXVIIIO\\ $J=6-5$ maps in a similar area as \\TWCO\\ but in a smaller range of radial velocities (from $+$5 to $+$15 km s$^{-1}$ for \\THCO\\ and from $+$4 to $+$12 km s$^{-1}$ for \\CXVIIIO). Both maps also show a good correspondence but with the integrated emission from \\CXVIIIO\\ $J=6-5$ fainter than \\THCO\\ $J=8-7$. The two isotopologues have fainter and less broader lines compared with the \\TWCO\\ line emission which is expected for optically thick lines (see the spectra in Figs. \\ref{map1} and \\ref{map2}). The Orion Bar is well resolved in both lines and also shows a sharp edge. In the mid-$J$ images (Figs. \\ref{map1} and \\ref{map2}), the strong peak in the center of the Bar is also seen in the recent {\\it Herschel}/SPIRE high-$J$ \\TWCO\\ and \\THCO\\ data obtained by \\citet{Habart2010}. In addition, the line widths of \\THCO\\ $J=8-7$ and \\CXVIIIO\\ $J=6-5$ toward all four main regions are narrow, even in the outflow zones, i.e., Orion BN/KL and South. The \\THCO\\ $J=3-2$ and \\CXVIIIO\\ $J=3-2$ integrated intensity maps are shown in Figure \\ref{map3}. These images were made in a similar area as higher-$J$ \\TWCO\\ and in a range of radial velocities equal to \\THCO\\ $J=8-7$ and \\CXVIIIO\\ $J=6-5$. The \\CXVIIIO\\ $J=3-2$ emission is only present where the \\THCO\\ $J=3-2$ emission is strong. The \\THCO\\ and \\CXVIIIO\\ line profiles at this rotational transition toward the Orion BN/KL and South regions are very broad and show well-defined wings. The southwestern part of the Bar is especially strong in this lower-$J$ transition for both isotopologues, which is also shown in the {\\it Herschel}/SPIRE high-$J$ \\TWCO\\ and \\THCO\\ images reported by \\citet{Habart2010}. \\subsection{Velocity structure} The \\TWCO\\ $J=6-5$ and $J=7-6$ velocity channel maps are shown in Figure \\ref{chmap}. These maps reveal a well-resolved and complex structure of the OMC-1 core at ambient velocities as well as higher radial velocities. The outflows that are located mainly at Orion BN/KL, Orion South, and toward the north of Orion BN/KL are clearly seen in these images \\citep[see][]{Zapata2006,Zapata2009}, e.g., \\vlsr=0--6 \\kms\\ for the BN/KL region and \\vlsr=12--14 \\kms\\ for the Orion South region. Besides, no clear north-south velocity gradients across the OMC-1 core region are observed. The compact and warm structures toward the BN/KL region traced by the \\TWCO\\ $J=6-5$ and $J=7-6$ lines, which display a very broad range of velocities (up to about $\\pm100$ \\kms), are part of the enigmatic molecular outflow that seems to be produced by a violent explosion during the disruption of a massive young stellar system \\citep{Bally2005,Zapata2009,Bally2011}. Our observations reveal that some faint filamentary structures are very likely associated with the high-velocity \\TWCO\\ bullets (Peng et al, in prep.) reported by \\citet{Zapata2009}. % \\subsection{Line ratios} In Figures \\ref{ratio3}--\\ref{ratio2}, we show the line intensity ratio maps between the CO isotopologues and the different rotational transitions. From the image of the \\TWCO\\ to \\CXVIIIO\\ $J=6-5$ ratio (Fig. \\ref{ratio3} upper panel), we can see the spatial distribution of high optical depth regions at different velocities. The north-south dense ridge from Orion BN/KL to Orion South is clearly seen at 9 \\kms. This high optical depth ridge is similar to the filamentary structures of the NH$_{3}$ emission \\citep{Wiseman1996,Wiseman1998}, but has some morphological differences in the north of the Orion BN/KL region. Besides, the straight shape of the Orion Bar is pronounced at \\vlsr=10--11 \\kms, where three high optical depth regions are seen at the two ends and center of the Bar at \\vlsr=10 and 11 \\kms, respectively. In the lower panel of Figure \\ref{ratio3}, the \\CXVIIIO\\ $J=6-5$ to $J=3-2$ ratios around 10--11 \\kms\\ show a gradient in the Orion Bar which goes in the direction from the Trapezium stars. It indicates that \\CXVIIIO\\ $J=6-5$ is strongly excited at the edge of the Bar by UV photons from the Trapezium stars. \\begin{figure} \\centering \\includegraphics[angle=270,width=0.48\\textwidth]{Orion-A-12CO_7-6_6-5-ratio-chmap-LowRes.eps} \\caption{Velocity channel maps (5, 10, and 15 km s$^{-1}$) of the \\TWCO\\ $J=7-6$ to $J=6-5$ ratio ($\\theta_{\\rm HPBW}=10\\arcsec$) in the OMC-1 core. The 10 \\kms\\ image is overlaid with the \\TWCO\\ $J=6-5$ contours as shown in Fig. \\ref{map1}. The empty area are due to data blanking (3 $\\sigma$ level of the \\TWCO\\ $J=7-6$ temperature). The dashed-lines indicate that these ratio gradients are likely caused by outflows or filaments. Black triangles mark the Orion Hot Core position. \\label{ratio1}} \\end{figure} \\begin{figure} \\centering \\includegraphics[angle=270,width=0.45\\textwidth]{all-inte-ratio-multi-maps.eps} \\caption{(a) The OMC-1 core \\CXVIIIO\\ $J=3-2$ to \\THCO\\ $J=3-2$ integrated intensity ratio map ($\\theta_{\\rm HPBW}=20\\arcsec$) overlaid with the \\TWCO\\ $J=6-5$ contours (grey) as shown in Fig. \\ref{map1}. The black contour represents the \\CXVIIIO/\\THCO\\ abundance ratio of 0.122 (60/490). (b) The OMC-1 core \\CXVIIIO\\ $J=3-2$ to $J=6-5$ integrated intensity ratio map ($\\theta_{\\rm HPBW}=20\\arcsec$) overlaid with the \\TWCO\\ $J=6-5$ contours as shown in Fig. \\ref{map1}. Black triangles mark the Orion Hot Core position. \\label{ratio2}} \\end{figure} Figure \\ref{ratio1} shows the ratio between \\TWCO\\ $J=7-6$ and $J=6-5$ at three different radial velocities of 5, 10, and 15 \\kms. The three panels show clear variations across the integrated line emission. It is interesting to note that at the cloud velocity of 10 km s$^{-1}$, there are high 7--6/6--5 ratios located very close to the position of the Trapezium stars toward Orion BN/KL. These gradients are likely produced by those massive stars that heat the molecular cloud, causing the stronger \\TWCO\\ $J=7-6$ line emission compared with the $J=6-5$ line. A gradient is also seen in the Orion Bar which goes in the perpendicular direction to the Trapezium stars (Fig. \\ref{ratio1} at $V_{\\rm LSR}=10$ \\kms), and some patches with higher \\TWCO\\ $J=7-6$ brightness temperatures are seen inside the Bar or behind the ionization front. Some horizontal and vertical strips, artifacts from the OTF mapping, are also seen in the Orion Bar and South regions. These artifacts can affect the intensity ratio by $\\sim 14 \\%$ given the calibration errors of $10 \\%$ for both \\TWCO\\ $J=6-5$ and $J=7-6$. However, some structures seen in the \\TWCO\\ $J=7-6$ to $J=6-5$ ratio maps, e.g., the dashed-lines in Figure \\ref{ratio1}, are unlikely due to the calibration error or scanning strips during the observation, and seem to be footprints of filaments or outflows. Some similar filament structures have been noticed in the dust continuum emission observed by \\citet{Johnstone1999}. In the ratio map of \\CXVIIIO\\ $J=3-2$ to \\THCO\\ $J=3-2$ shown in Figure \\ref{ratio2}, a clear elongated and dense ridge is seen in the north-south direction, where the \\THCO\\ $J=3-2$ intensity is weaker in the Orion BN/KL region than at Orion South and the north of BN/KL. Therefore, the lower CO column density derived from the lower \\CXVIIIO\\ $J=3-2$ to \\THCO\\ $J=3-2$ ratio toward Orion BN/KL may be misleading. Even though the \\THCO\\ $J=3-2$ line is optically thick and may be self-absorbed, this will result in a higher \\CXVIIIO\\ $J=3-2$ to \\THCO\\ $J=3-2$ ratio instead of a lower ratio in Orion BN/KL. As \\citet{Goldsmith1997} pointed out, the column density calculated using low-$J$ \\CXVIIIO\\ lines is accurate except for the temperature $\\geq150$ K, which is the case in the OMC-1 core. Hence, higher-$J$ CO observations are critical to determine the CO column density in the OMC-1 core region. In addition, already shown in Figure \\ref{ratio1}, the low \\CXVIIIO\\ $J=3-2$ to $J=6-5$ ratios indicate that the \\CXVIIIO\\ $J=6-5$ intensity is strongly enhanced toward the Orion BN/KL, South, and Bar regions, and is likely due to their PDR nature. Nevertheless, shocks/outflows may also play an important role in the Orion BN/KL and South regions, leading to an extra heating in the cloud. \\subsection{Excitation temperature and density estimates} The \\TWCO\\ $J=6-5$ and \\CXVIIIO\\ $J=6-5$ data are used here to derive the gas excitation temperature and density of the OMC-1 core. In the submillimeter regime where the Rayleigh-Jeans approximation is often not valid, the observed radiation temperature in local thermodynamic equilibrium (LTE) can be expressed as \\begin{equation}\\label{cal-orion-1} {T^{*}_{\\rm R}}=\\frac{h\\nu}{k}\\biggl[\\frac{1}{e^{h\\nu/kT_{\\rm ex}}-1}-\\frac{1}{e^{h\\nu/kT_{\\rm bg}}-1}\\biggr](1-e^{-\\tau_{\\nu}}), \\end{equation} where $T_{\\rm ex}$ is the excitation temperature and $\\tau_{\\nu}$ is the optical depth at a specific molecular line transition. The background brightness temperature $T_{\\rm bg}$ includes the cosmic background radiation of 2.73 K and the radiation from warm dust, ranging from about 5 K for less dense gas to about 26 K for the Orion Hor Core \\citep{Goldsmith1997}. The second term inside the brackets can be neglected since it only contributes $\\lesssim2\\%$ of $T^{*}_{\\rm R}$. By making the assumption of the same excitation temperature for \\TWCO\\ and \\CXVIIIO\\ $J=6-5$, the optical depths of \\TWCO\\ and \\CXVIIIO\\ $J=6-5$ can be determined from the relation \\begin{equation}\\label{cal-orion-2} \\frac{T^{*}_{\\rm R}(\\rm{^{12}CO)}}{T^{*}_{\\rm R}(\\rm{C^{18}O)}}\\approx\\frac{1-e^{-\\tau(\\rm{^{12}CO})}}{1-e^{-\\tau(\\rm{C^{18}O})}}. \\end{equation} Since the optical depth of \\TWCO\\ $J=6-5 \\gg 1$, the optical depth of \\CXVIIIO\\ $J=6-5$ can be directly obtained ($\\sim 0.08$) by assuming that the optical depth ratio is approximated to the isotopologic abundance ratio, and here we adopted [\\TWCO]/[\\CXVIIIO] of 490 \\citep[][]{Boreiko1996,Wilson1992} and a beam filling factor of unity ($T^{*}_{\\rm R}=T_{\\rm MB}$). The excitation temperatures of \\TWCO\\ and \\CXVIIIO\\ $J=6-5$ can be estimated via \\begin{equation}\\label{cal-orion-3} T_{\\rm ex}=\\frac{h\\nu}{k}\\biggl[{\\rm ln}\\biggl(1+\\frac{h\\nu}{kT_{\\rm MB}\\rm(^{12}CO)}\\biggr)\\biggr]^{-1}. \\end{equation} Then the total \\CXVIIIO\\ column density can be derived from \\begin{equation} N({\\rm C^{18}O})=\\frac{3kQ_{\\rm rot}}{8\\pi^{3}\\nu S\\mu^{2}}e^{\\frac{E_{\\rm up}}{kT_{\\rm ex}}}\\int T_{\\rm MB}dV \\end{equation} \\begin{equation} \\simeq1.33\\times10^{12}\\ (T_{\\rm ex}+0.88)\\ e^{\\frac{E_{\\rm up}}{kT_{\\rm ex}}}\\int T_{\\rm MB}dV\\ \\rm{cm^{-2}}, \\end{equation} where $E_{\\rm up}$ is 110.6 K for the \\CXVIIIO\\ $J=6-5$ transition and $Q_{\\rm rot}$ is the rotational partition function. In the end, the \\HH\\ column density can be estimated by adopting a [\\TWCO]/[\\CXVIIIO] abundance ration of 490 and a \\TWCO\\ abundance of $8\\times10^{-5}$ \\citep{Wilson1992}. \\begin{figure*} \\centering \\includegraphics[angle=270,width=0.9\\textwidth]{Orion-A-12CO-C18O_6-5-den-ext-map.eps} \\caption{(a) The \\CXVIIIO\\ column density image of the OMC-1 core overlaid with black contours of the JCMT 850 \\micron\\ dust continuum image. The 850 \\micron\\ image is taken from the JCMT data archive. The contours represent 0.2$\\%$, 0.5$\\%$, 1$\\%$, 2$\\%$, 4$\\%$, 7$\\%$, 20$\\%$, 40$\\%$, 60$\\%$, and 80$\\%$ of the peak intensity (87 Jy beam$^{-1}$). (b) The excitation temperature image for \\TWCO\\ and \\CXVIIIO\\ overlaid with the 35 mm VLA-GBT continuum emission \\citep{Dicker2009} in black contours running from 15$\\%$ to 95$\\%$ in steps of 8$\\%$ of the peak intensity (0.94 Jy beam$^{-1}$). The stars mark the positions of the five Trapezium stars, and black triangles represent the position of the Orion Hot Core.\\label{den-temp}} \\end{figure*} The results of the \\HH\\ column density and excitation temperature distributions are shown in Figure \\ref{den-temp}, where the north-south dense ridge near the Trapezium cluster and the Orion Bar in the southeast are clearly seen. The average $\\tau(\\rm C^{18}O)$ at the $J=6-5$ transition over the whole OMC-1 core region is about 0.08 with an average excitation temperature of about 115 K, which are consistent with the \\TWCO\\ $J=9-8$ observations by \\citet{Kawamura2002}. Besides, the minimum excitation temperature in the OMC-1 core is $\\sim 30$ K, indicating a generally warm environment. The average \\CXVIIIO\\ column density in the OMC-1 core is about $8.7\\times10^{15}$ cm$^{-2}$, corresponding to an \\HH\\ column density of $5.5\\times10^{22}$ cm$^{-2}$ assuming a \\TWCO\\ abundance of $8\\times10^{-5}$ \\citep{Wilson1992}. Therefore, the dense gas mass is $\\sim 140$ \\Msol\\ in an area of 0.13 pc$^2$ in the OMC-1 core. The \\HH\\ column density agrees with the density of $5\\times10^{21}-7\\times10^{22}$ cm$^{-2}$ derived from the mid-$J$ CO isotopologue observations by \\citet{Wirstrom2006}. Our results agree with the recent measurement by \\citet{Wilson2011} using the \\THCO\\ $J=6-5$ line with a lower [\\TWCO]/[\\HH] ratio of $2\\times10^{-5}$. Additionally, \\citet{Habart2010} derived an \\HH\\ column density of about $9\\times10^{22}$ cm$^{-2}$ toward the Orion Bar from the {\\it Herschel}/SPIRE high-$J$ CO observations, assuming a lower excitation temperature of 85 K and a higher \\TWCO\\ to \\CXVIIIO\\ ratio of 560. Their result is consistent with ours, where we obtained an \\HH\\ column density of $\\approx10^{23}$ cm$^{-2}$ toward the Orion Bar with a higher excitation temperature ($\\gtrsim150$ K). \\begin{figure*} \\centering \\includegraphics[angle=270,width=0.9\\textwidth]{radex-co-13co-c18o-grid-dens-peak.eps} \\caption{The RADEX modeling results shown in the \\TWCO\\ $J=6-5$ peak temperatures and the \\THCO/\\CXVIIIO\\ ratios for three different \\TWCO\\ column densities. The column densities of \\THCO\\ and \\CXVIIIO\\ were fixed at the isotopologic abundance ratio of $1/60$ and $1/490$ of the \\TWCO\\ column density. The line widths of \\TWCO, \\THCO, and \\CXVIIIO\\ were fixed at 5, 5, and 4 \\kms, respectively. The red and blue contours denote the temperature and \\HH\\ number density, respectively, and the color points represent the data in the different regions of the OMC-1 core. Each point represents a peak \\TWCO\\ $J=6-5$ temperature and a temperature ratio between \\THCO\\ $J=8-7$ and \\CXVIIIO\\ $J=6-5$ from a single pixel with a resolution of 20$\\arcsec$. (a) The modeling input column density of \\TWCO\\ is $1\\times10^{18}$ cm$^{-2}$, and the data are plotted in a \\TWCO\\ column density density of $7.5\\times10^{17}-2.5\\times10^{18}$ cm$^{-2}$. (b) The modeling input column density of \\TWCO\\ is $5\\times10^{18}$ cm$^{-2}$, and the data are plotted in a density range of $2.5\\times10^{18}-7.5\\times10^{18}$ cm$^{-2}$. (c) The modeling input column density of \\TWCO\\ is $1\\times10^{19}$ cm$^{-2}$, and the data are plotted in the density of $\\geq7.5\\times 10^{18}$ cm$^{-2}$. Upper right panel shows the \\TWCO\\ column density map toward Orion BN/KL in contours of 5, 10, 20, 30, and $40\\times 10^{18}$ cm$^{-2}$. The stars mark the positions of the five Trapezium stars, and the Orion Hot Core (HC) position is also marked. \\label{radex}} \\end{figure*} The JCMT 870 $\\mu$m dust continuum emission image \\citep{Johnstone1999} shown in Figure \\ref{den-temp} agrees with our density map well, where the dust emission traces mostly the dense ridge (Orion BN/KL and South) and Bar regions. It is interesting to note that the dense ridge has an offset from the peak temperature positions, especially in the Orion South region ($\\sim30\\arcsec$) and the north of the Orion BN/KL region ($\\sim20\\arcsec$). This dense ridge seems to extend farther to the north of the OMC-1 core, and is probably related to the NH$_3$ filamentary structure seen by \\citet{Wiseman1998,Wiseman1996}. Additionally, as Figure \\ref{den-temp} (b) demonstrates, the interface between the ionized and the molecular warm dense gas is very pronounced, where Orion BN/KL and Orion South are located at the north-south dense ridge, and the Orion Bar and Orion East are part of the high temperature enclosed structure well illustrated by the free-free continuum emission. Furthermore, our large-scale images are consistent with the [\\CII] and [\\OI] emission maps obtained by \\citet{Herrmann1997}, where the rather uniform [\\CII] emission indicates that PDRs are present over the whole OMC-1 core region. \\subsection{RADEX modeling} The non-LTE radiative transfer program RADEX \\citep{vanderTak2007} was used to investigate the accuracy of the temperature and density calculation in LTE shown above. The optical depth effects are taken into account by RADEX using the escape probability approximation, where a uniform sphere geometry was chosen. The different transitions of \\TWCO, \\THCO, and \\CXVIIIO\\ were used in the modeling, and their line widths were fixed at 5, 5, and 4 \\kms, respectively, which are the average values from our spectra. The molecular data using in the modeling are taken from the Leiden Atomic and Molecular Database\\footnote{http://www.strw.leidenuniv.nl/$\\sim$moldata/} \\citep[LAMDA;][]{Schoeier2005}. In addition, only \\HH\\ was chosen as a collision partner, and 2.73 K was adopted as the background temperature. The input parameters are kinetic temperature and \\HH\\ number density. Each model was iterated with a kinetic temperature ranging from 50 to 450 K and an \\HH\\ number density ranging from $10^4$ to $10^8$ cm$^{-3}$. The input column densities of these three molecules were chosen, so that their ratios were fixed at their isotopologic abundance ratios, i.e., 490 for [\\TWCO]/[\\CXVIIIO] and 60 for [\\TWCO]/[\\THCO]. Three models were iterated with three different \\TWCO\\ column densities of $1\\times10^{18}$, $5\\times10^{18}$, and $1\\times10^{19}$ cm$^{-2}$ which were obtained from the LTE calculation. Since these models did not take outflows into account, the output molecular line radiation temperatures ($T_{\\rm R}$) were directly compared with the peak temperatures instead of the integrated intensities to avoid the strong line wing emission. The modeling results are shown in Figure \\ref{radex}. Figure \\ref{radex} reveals that most regions in the OMC-1 core have an \\HH\\ number density of $\\sim 10^{4}-10^{6}$ cm$^{-3}$. The Orion Bar and Orion East emission peaks (Fig. \\ref{radex} a), however, show higher densities. The distribution of the \\HH\\ number density in the Bar is consistent with $10^{4}-10^{5}$ cm$^{-3}$ for a homogeneous medium \\citep{Wyrowski1997}. The modeling results in the Orion BN/KL, South, and Bar regions all show similar trends of a diverse kinetic temperature (about 100--200 K), which are consistent with the LTE calculation results. In addition, the non-LTE modeling results indicate that the central region of Orion BN/KL close to the Hot Core (in 20\\arcsec\\ size) has a number density of about $10^5$ cm$^{-3}$ and a temperature of about 200 K, which is also close to the excitation temperature calculated in the LTE case. Noteworthily, the Orion BN/KL region and part of the Orion South region (Fig. \\ref{radex} a) show very high kinetic temperatures ($>350$ K) which may indicate an extra heating mechanism, e.g., outflow/shock heating. The suggestion of outflow/shock heating is supported by the larger \\TWCO\\ line widths (Fig. \\ref{map1}) and a great amount of known outflows in these regions \\citep[see][]{Henney2007}. However, the high kinetic temperatures ($\\geq250$ K) in the Orion Bar and Orion East regions (Fig. \\ref{radex} a) are hardly explained by the outflow/shock heating for the lack of outflow activities or broad \\TWCO\\ line widths. Instead, as PDRs, the Orion Bar and Orion East are expected to be heated by FUV photons from the Trapezium OB stars. Moreover, a recent study of the Orion Bar \\citep{Pellegrini2009} suggests an extra heating by the excess density of cosmic rays, which are trapped in the compressed magnetic field. \\subsection{Mass of the OMC-1 core} By adopting a radius of 40\\arcsec--60\\arcsec\\ and an \\HH\\ density of $5\\times10^{4}-2\\times10^{5}$ cm$^{-3}$ for Orion BN/KL, we can estimate a gas mass of 7--85 \\Msol\\ with the assumption of a spherical volume at a distance of 414 pc \\citep{Menten2007}. In a similar way, the gas mass in Orion South is 3--49 \\Msol\\ with a 30\\arcsec--50\\arcsec\\ radius and an \\HH\\ density of $5\\times10^{4}-2\\times10^{5}$ cm$^{-3}$. For Orion East, the \\HH\\ density is $\\sim2\\times10^5$ cm$^{-3}$ with a 15\\arcsec--25\\arcsec\\ radius, and the gas mass is 2--7 \\Msol. In contrast, it is very difficult to estimate the gas mass in the Orion Bar because of its geometry. We adopted a cylindrical geometry and an \\HH\\ density of $5\\times10^{4}-2\\times10^{5}$ cm$^{-3}$ as the lower limit since an edge-on plane may contain much more gas. With a length of 300\\arcsec\\ and a radius of 20\\arcsec--30\\arcsec, the gas mass of the Orion Bar is 9--79 \\Msol. The estimated total dense \\HH\\ gas mass in the OMC-1 core is 21--220 \\Msol\\ within a radius of 0.3 pc. This is also consistent with the mass estimated ($\\sim140$ \\Msol) from the LTE calculation above. If we include the warm diffuse gas ($10^{4}$ cm$^{-3}$) in the same volume, the total warm gas mass is 86--285 \\Msol\\ in the OMC-1 core. This mass estimate for the OMC-1 core is close to that derived from the Odin \\TWCO\\ and \\THCO\\ $J=5-4$ observations \\citep{Wirstrom2006}, i.e., 320 \\Msol\\ in the molecular ridge for a farther distance of 500 pc. The result of \\citet{Wilson2001} also suggests a similar estimate of a warm gas mass of 310--430 \\Msol\\ from the observations of \\TWCO\\ $J=7-6$ and $J=4-3$, again for a distance of 500 pc. Scaling these two literature values % to the distance of 414 pc, results in 213--295 \\Msol\\ for the total mass of the OMC-1 core. This is comparable to the value derived by us. Hence, we conclude that the higher-$J$ CO lines do trace most of the gas in the OMC-1 core which is heated mainly by the Trapezium stars." }, "1112/1112.3710_arXiv.txt": { "abstract": "A growing number of recent observations have revealed that the Galactic globular cluster (GC) $\\omega$ Cen is not the only GC that shows abundance spread in heavy elements (e.g., Fe). In order to understand the origin of the Galactic GCs with heavy element abundance spread (``HEAS''), we investigate the formation processes of massive GCs (MGCs) with masses larger than $10^{6} {\\rm M}_{\\odot}$ in gas-rich dwarf galaxies interacting and merging with the very young Galaxy. We find that massive and compact stellar clumps with masses larger than $10^{6} {\\rm M}_{\\odot}$, which can be regarded as progenitors of MGCs, can form from massive gas clumps that are developed through merging of gaseous regions initially at different radii and thus with different metallicities. Therefore it is inevitable that MGCs formed in dwarfs have HEAS. The abundance spread in each individual MGC depends on the radial metallicity gradient of the host dwarf such that it can be larger for the steeper metallicity gradient. For example, MGCs formed in a dwarf with a central metallicity of [Fe/H]$=-1.1$ and the radial gradient of $\\sim -0.2$ dex kpc$^{-1}$ can have the abundance spread of $\\Delta {\\rm [Fe/H]} \\sim 0.2$. The simulated MGCs appear to be significantly flattened owing to their dissipative formation from gas disks of their host dwarfs. Based on these results, we discuss possibly diverse formation mechanisms for the Galactic GCs such as M22, M54, NGC 2419, $\\omega$ Cen, and Terzan 5. ", "introduction": "One of remarkable recent developments in observational studies of the Galactic GCs is that most of the investigated GCs show varying degrees of chemical abundance spread: helium abundance spread in $\\omega$ Cen and NGC 2808 (e.g., Bedin et al. 2004; Piotto et al. 2007), a larger dispersion in $s$-process elements abundances in NGC 1851 (e.g., Yong \\& Grundahl 2008; Milone et al. 2011), abundance spread in light element in ``normal'' GCs (e.g., Carretta et al. 2010a), and HEAS in M22, Terzan 5, and NGC 2419 (e.g., Da Costa et al. 2009; Ferraro et al. 2009; Marino et al. 2009; Cohen et al. 2010). Lee et al. (2009) investigated color-magnitude diagrams of stars in $hk$-bands for the Galactic GCs and suggested that a significant fraction of the Galactic GCs have two different populations with HEAS (but see also Marino et al. 2009; Carretta et al. 2010b for similar and different claims). The origin of the HEAS in $\\omega$ Cen has been discussed by theoretical models in the context of formation and evolution of stellar galactic nuclei in dwarfs (``the stripped nucleus scenario''; e.g., Bekki \\& Freeman 2003; Romano et al. 2010). Since stellar populations of stellar galactic nuclei can have different metallicities and ages owing to their possible formation by merging of different stellar and gaseous clumps (e.g., Bekki 2007), it would not be so surprising that $\\omega$ Cen is observed to have HEAS owing to its possible origin from the stellar nucleus of its host dwarf. Also, it would be possible that the formation process of other GCs with HEAS (e.g., Terzan 5) can be different from that of $\\omega$ Cen (e.g., GC mergers; Br\\\"uns \\& Kroupa 2011; Bekki \\& Yong 2011, BY11). Ideally speaking, chemodynamical simulations of GC formation based on a theoretical model are essential for discussing both the formation processes of GCs from gas and the resultant internal chemical abundance spread within GCs in a fully self-consistent manner. However no chemodynamical simulations on the origin of GCs with HEAS (other than $\\omega$ Cen) have been done so far. The purpose of this Letter is to clearly show whether and how GCs with HEAS can be formed by using chemodynamical simulations of GC formation processes based on a GC formation scenario. We here consider that most of the Galactic GCs were formed in galactic building blocks (e.g., Searle \\& Zinn 1978), in particular, when they were interacting and merging with the very young Galaxy ($>10$ Gyr ago). We focus exclusively on [Fe/H] spread ($\\Delta {\\rm [Fe/H]}$) in MGCs with masses as large as or larger than $10^6 {\\rm M}_{\\odot}$ formed within their host dwarfs (i.e., galactic building blocks). \\begin{figure*} \\psfig{file=f1.eps,width=18.0cm} \\caption{ The time evolution of the mass distribution of the dwarf with respect to the Galactic center (left four) and to the dwarf's center (right four) projected onto the $x$-$y$ plane for the standard model (M1). Gas and new stars are shown by cyan and magenta, respectively, and $T$, which is the time that has elapsed since the simulation started, is shown in the upper left corner in units of Gyr for each panel. The green lines in the left four panels represent the orbit of the dwarf. The locations of MGC1, 2, and 3 are indicated by blue, red, and green circles at $T=1.7$ Gyr, respectively. } \\label{Figure. 1} \\end{figure*} ", "conclusions": "The present chemodynamical study has first demonstrated that MGCs with HEAS can be formed from massive gas clumps developed from merging of different gaseous regions with different metallicities. Therefore, MGCs with HEAS can be formed even without chemical enrichment by supernovae during their formation. The present study suggests that more massive GCs are more likely to have a larger degree of HEAS owing to their formation from gas clouds developed from merging between a larger number of different gaseous regions with different metallicities. The low-mass GCs can be formed from single gas clouds so that they are likely to have no/little HEAS. The present study thus suggests that the origin of HEAS in some Galactic GCs can be closely associated with HEAS of ISM in their host dwarfs. The large metallicity spread of various elements and possible age variation in $\\omega$ Cen (e.g., Sollima et al. 2005) can be consistent with the stripped nucleus scenario. However, this does not necessarily mean that all of the Galactic GCs with HEAS (e.g., NGC 2419, M22, and Terzan 5) were formed from stripped nuclei of dwarfs. A GC with distinct two peaks in the [Fe/H] distributions of the stars might well be consistent with GC merging with different [Fe/H] (e.g., Br\\\"uns \\& Kroupa 2011; BY11). A GC with a smaller degree of HEAS yet no clear bimodal [Fe/H] distribution might well form from massive gas clumps of the host dwarf and then be stripped from the dwarf without sinking into the center (i.e., without becoming the stellar nucleus) to finally become the Galactic halo GC. We lastly suggest that nucleation, GC merging, and merging of gas clouds with different [Fe/H] can be all promising mechanisms for the formation of the Galactic GCs with different degrees of HEAD." }, "1112/1112.2520_arXiv.txt": { "abstract": "{High-mass microquasars consist of a massive star and a compact object, the latter producing jets that will interact with the stellar wind. The evolution of the jets, and ultimately their radiative outcome, could depend strongly on the inhomogeneity of the wind, which calls for a detailed study.} {The hydrodynamics of the interaction between a jet and a clumpy wind is studied, focusing on the global wind- and single clump-jet interplay.} {We have performed, using the code \\textit{Ratpenat}, three-dimensional numerical simulations of a clumpy wind interacting with a mildly relativistic jet, and of individual clumps penetrating into a jet.} {For typical wind and jet velocities, filling factors of about $\\ga 0.1$ are already enough for the wind to be considered as clumpy. An inhomogeneous wind makes the jet more unstable when crossing the system. Kinetic luminosities $\\sim 10^{37}$~erg/s allow the jet to reach the borders of a compact binary with an O star, as in the smooth wind case, although with a substantially higher degree of disruption. When able to enter into the jet, clumps are compressed and heated during a time of about their size divided by the sound speed in the shocked clump. Then, clumps quickly disrupt, mass-loading and slowing down the jet.} {We conclude that moderate wind clumpiness makes already a strong difference with the homogeneous wind case, enhancing jet disruption, mass-loading, bending, and likely energy dissipation in the form of emission. All this can have observational consequences at high-energies and also in the large scale radio jets.} ", "introduction": "\\label{intro} Microquasars are binary systems hosting a star and an accreting black hole or neutron star. Matter from the star is transferred to the compact object, part of it being launched through magnetocentrifugal forces \\citep[e.g.,][]{bla77,bla82,bar11}. This triggers the formation of bi-polar jets, which generate non-thermal radio emission \\citep[e.g.,][]{mir99,rib05}, and are thought to be the location from where the gamma rays observed in some sources are emitted \\citep[e.g.,][]{alb07,tav09,abd09,sab10}. Jets could be magnetically dominated at their base, but magnetohydrodynamical processes occurring at higher jet height would accelerate the flow, efficiently converting magnetic energy into kinetic one \\citep[e.g.][]{kom07}. At the scales of the binary system ($\\sim 10^6\\,R_{\\rm Sch}$, where $R_{\\rm Sch}$ is the Schwarzschild radius), the jet is likely to be already a hydrodynamical (HD) flow. We focus here on the persistent jets thought to be present during the low-hard state of microquasars, although some considerations for transient ejecta, associated to low-hard to high-soft state transitions, are done below \\citep[see, e.g.,][for reviews on microquasar states]{fen04,fen09}. Part of the energy carried by the jet can be dissipated in the form of magnetic reconnection, recollimation and internal shocks, shear layers in the jet walls, and turbulence. Part of the dissipated energy can go to non-thermal particles, generating low- and high-energy emission via different mechanisms, synchrotron from radio to X-rays, and inverse Compton (IC) and hadron-related processes up to gamma rays \\citep[see, e.g.,][and references therein]{bos09}. As shown by \\cite{per08} (PB08 hereafter) and \\cite{per10} (PBK10 hereafter), in microquasars hosting an OB star (high-mass microquasars; hereafter HMMQ) the jet may be strongly influenced by the stellar wind. The one-side impact of the wind on the (presumably) already HD jet leads to strong and asymmetric recollimation shocks, bending and different types of instabilities: the recollimation shocks seem suitable candidates for particle acceleration and non-thermal emission; bending may be noticeable in radio at milliarcsecond scales; instabilities may destroy the jet flow even within the binary system. For typical wind and jet velocities, say $v_{\\rm w}\\sim 2\\times 10^8$ and $v_{\\rm j}\\sim 10^{10}$~cm/s, respectively, studies show that for compact binaries and jet-to-wind momentum flux ratios $\\la 0.1$ the jet can be already disrupted (PB08, PBK10). This number is rather constraining, since only a few HMMQ might be above this threshold. This could be the reason for the low number of HMMQ detected, as suggested in PBK10. In any case, even if not destroyed within the binary system, jets can suffer strong perturbations with both dynamical and radiative consequences. Previous work in HMMQ wind-jet interactions was done under the assumption that the wind is homogeneous, but in fact stellar winds are thought to be clumpy \\citep[e.g.,][]{owo06,mof08}. For this reason, it has been proposed that wind clumpiness should be taken into account when studying HMMQ (\\citealt{owo09,ara09} -ABR09 hereafter-; \\citealt{rom10,ara11}). An important parameter that determines the inhomogeneity of the wind is the wind filling factor $f$, which determines the wind volume fraction with higher density. For a significant departure from homogeneity, the intraclump medium mass and momentum fluxes will be negligible and only clumps will have a dynamical impact on the jet. Since the interaction between a HD jet and a clumpy wind has not been studied in detail, we have carried out 3-dimensional (3D) simulations of this scenario. Simulations have been done for two different jet powers and jet-to-wind momentum ratio, $L_{\\rm j}=3\\times 10^{36}-10^{37}$~erg/s and $\\approx 0.03-0.08$, respectively, to explore what could be the transition between jet destruction and long-term collimation. Another simulation has focused on the evolution of individual clumps injected in the jet at different heights. Unlike in PBK10, in which the simulation started with the jet being injected at its base, here the jet is conical and crossing the whole grid, and the clumpy wind is injected from one of the jet sides. The first goal of this work is the study of the hydrodynamical evolution of a jet when the wind interacting with it is clumpy. The second goal is to quantify for which values of the clump and jet parameters, clumpiness becomes a relevant factor. It is also interesting to study the evolution of a clump under the impact of a microquasar jet. The results can also be used to refine radiation models or interpret radio observations, although this will be treated qualitatively. The paper is organized as follows: in Section~\\ref{phys}, the scenario studied here is briefly introduced; in Sect.~\\ref{sim}, the simulations are described (Sect.~\\ref{sim1}); results are shown in Sect.~\\ref{sim2}; finally, in Sect.~\\ref{disc}, the results are discussed in the context of jet propagation in HMMQ (Sect.~\\ref{disc1}), individual clump-jet interactions (Sect.~\\ref{disc2}), and their implications for the non-thermal emission (Sect.~\\ref{disc3}). Throughout the paper, we will use cgs units. ", "conclusions": "\\label{disc} \\subsection{Jet propagation}\\label{disc1} The simulated jets are strongly deviated from their original direction and show disrupted morphologies at the end of the calculations. Initially, the first interactions at small values of $z$ trigger helical patterns that, in the absence of further perturbations, could couple to Kelvin-Helmholtz unstable modes. However, the reconfinement shock triggered by the ram pressure of the clumpy wind on the jet plus individual interactions with multiple clouds downstream of this shock, generate non-linear structures that are most important for the jet evolution. The position of the reconfinement shock is located within the binary region, as predicted by PB08. There, the jet is decelerated and heated, thus becoming even weaker relative to the wind thrust, as also shown in PB08 and PBK10. In the present case, the clumps from the up-wind region can penetrate in the jet already at $z\\sim 1-2\\times 10^{12}$~cm, generating additional bow-like shocks and decelerating even more the jet flow. The deviation of the jet favors the presence of a region, on the down-wind side of the jet, to which, by the end of simulations, the clump bow shocks and jet mixing have still not reached. In this region, the flow keeps a relatively large axial velocity in this direction. However, both shocks and clump-mixed jet material should eventually fill the whole jet at larger $z$, so the description just presented (see Figs~\\ref{fig:maps8}, \\ref{fig:maps9} and \\ref{fig:maps10}) can be extended to the whole jet cross section. As a result, farther downstream the jet becomes mass-loaded, slow and transonic. Although a similar result was already obtained in the homogeneous wind case (PB08, PBK10), wind clumping significantly potentiates jet destruction. When comparing jet~B in this work with jet~2 in PBK10, it is remarkable that the presence of a forward shock and the absence of clumps in the latter have a strong influence on the long-term evolution of the jet. The jet-driven forward shock in the wind and the corresponding cocoon (see PBK10) keep the wind at some distance from the jet. If a homogeneous wind were already in contact with jet~2 in PBK10, a smooth shear layer would form. Otherwise, the presence of clumps in jet~B enhances the thrust locally in the wind, increasing mass, momentum and energy exchange. Jet~2 in PBK10 thus shows a larger degree of collimation and is just slightly deviated from its original direction of propagation, contrary to what we observe in the case of jet~B in this paper. From this work we can conclude that jet luminosities $L_{\\rm j}\\gtrsim 10^{37}\\,\\dot{M}_{-6}\\mathrm{erg/s}$ ($\\dot{M}_{-6}=\\dot{M}/10^{-6}\\,M_\\odot$/yr) are needed if the jet is not to be destroyed when crossing the binary system. We notice that these dynamical arguments favor $L_{\\rm j}\\gtrsim 10^{37}$ and $10^{38}$~erg/s in the HMMQ Cygnus~X-1 and Cygnus~X-3, respectively. The cocoon/clumpy wind case deserves few words. This situation takes place when the forward shock is well within the binary system, and the cocoon pressure is high, preventing clumps from entering the cocoon. When the forward shock has reached the outskirts of the binary, the cocoon pressure drops quickly (PBK10), allowing wind clumps to penetrate into the cocoon and reach the jet, and eventually dissipate the cocoon away. \\subsection{Clump evolution}\\label{disc2} In jets~A and B, those clumps reaching the jet relatively close to its base are destroyed just by jet expansion or erosion. These interactions trigger a shock wave that propagates inside the jet, and when interactions are frequent such a wave forces the strongly disruptive asymmetric recollimation shock. This phenomenon is illustrated in Fig.~\\ref{fig:maps12}), in which the first clump is completely destroyed by jet expansion, whereas the second one, even when it does not fully penetrate into the jet, triggers a shock strong that propagates all through the latter lasting for the whole simulation (several $t_{\\rm d}$; see Sect.~\\ref{phys}). The clump at the highest $z$ in jet~A' ($1.4\\times 10^{12}$~cm) is shocked but still not significantly disrupted by instabilities after few $t_{\\rm d}$. Later, this clump could be destroyed or may eventually escape the jet, although jet bending in the down-wind direction makes the latter unlikely. When clumps are disrupted inside the jet, all the clump mass is entrained by the flow. The level of mass loading can be easily estimated from the amount of clumps entering into the jet per time unit: $\\dot{N}_{\\rm cj}\\sim (\\eta/4\\pi)(3\\dot{M}/4\\pi R_{\\rm c}^3\\rho_{\\rm c})\\approx 0.02$~clump/s or $\\approx 5\\times 10^{17}$~g/s. This is $\\sim 3$ times more mass flux than in the jet. Therefore, jet deceleration due to mass loading can be very efficient, as is most clearly seen in the simulation results for jet~B. The implications also apply to faster and lighter, but equally powerful jets. Even if a strongly relativistic flow is injected at the jet base, for similar jet (relativistic) ram pressures clump penetration into, and mixing with, the jet will also occur. This will mass-load and brake the jet very quickly. The qualitative dynamical scenario presented in ABR09 is validated by our numerical work in a semi-quantitative way. Remarkably, as hinted in ABR09, the clump interaction with many clumps will reduce the jet kinetic luminosity and also its ram pressure. In this way, clumps can actually keep their integrity longer, and penetrate farther inside the jet, even for moderate $f$-values. \\subsection{Radiation considerations}\\label{disc3} In ABR09, the radiation produced by a clump inside a HMMQ jet was calculated. Given the compactness of the considered region, i.e. the bow shock formed around the clump, radio emission was negligible. The luminosities of synchrotron X-ray, and (stellar photon) IC gamma-rays, dominated the non-thermal output, with their values anticorrelating depending on the magnetic field. Based on that work, and extending the study to the multi-clump/jet interaction case simulated here, we qualitatively discuss in what follows the expected emission. The typical number of clumps inside the jet, at the binary scales, can be estimated as $N_{\\rm cj}\\sim$ few times $t_{\\rm d}\\,N_{\\rm cj}$, i.e. $\\gtrsim 10$ for the wind and jet properties adopted in this work (recall that the real clump lifetime is longer than $t_{\\rm d}$). Since $t_{\\rm d}\\sim 100\\,{\\rm s}10^{10}$~cm are completely missing. Concerning (powerful) transient ejections, usually associated to X-ray state transitions, we note that such an ejections will require some time to form. If this takes hours, the wind will have time to surround the transient jet. Then, the clump impact will be as described here unless the jet is too powerful. If powerful blobs would appear as discrete even at the scales of the binary, they may have too much inertia to be significantly affected by the wind. The dynamical impact of the stellar wind on the jet, enhanced by clumpiness, should not be neglected when interpreting radio emission from HMMQ. Even if jets escape from disruption, the enhanced jet entropy and bending, even by small angles ($\\sim 10^\\circ$), could have observational consequences. The reason is that bending pushes jet material farther from the orbital plane axis, increasing the strength of the Coriolis force exerted by the stellar wind and the orbital motion. The farther away from this axis the jet reaches, the stronger this force gets, enhancing jet heating, turbulence and bending. Although quantitative predictions call for a detailed study, this region may be observationally probed using VLBI techniques. Far enough from the binary, once the jet has become too wide to be affected by the orbital motion, a collimated supersonic flow may form again. If the energy and momentum fluxes kept enough anisotropy after crossing the system and suffer from orbital motion, given a negligible external pressure a jet-like structure could form again and propagate unstopped up to pc scales, terminating in the ISM \\citep[e.g.][]{bor09,bos11,yoo11}. Despite the different outflow geometry, similar phenomena are also expected in high-mass binaries hosting non-accreting pulsars \\citep[see][]{bos11b}." }, "1112/1112.2158_arXiv.txt": { "abstract": "Pulsar timing now has a rich history in placing limits on the stochastic background of gravitational waves, and we plan soon to reach the sensitivity where we can detect, not just place limits on, the stochastic background. However, the capability of pulsar timing goes beyond the detection of a background. Herein I review efforts that include single source detection, localization, waveform recovery, a clever use of a ``time-machine\" effect, alternate theories of gravity, and finally studies of the noise in our ``detector\" that will allow us to tune and optimize the experiment. Pulsar timing arrays are no longer ``blunt\" instruments for gravitational-wave detection limited to only detecting an amplitude of the background. Rather they are shrewd and tunable detectors, capable of a rich and dynamic variety of astrophysical measurements. ", "introduction": "Pulsars are basically celestial clocks, and as such, can be used to construct a Galactic-scale gravitational wave detector using the same concept as ground-based interferometric detectors, i.e. one looks for phase changes in the arrival of the signal at the vertex station, in this case, earth. The length scales of our detector `arms' (1000 light years) as compared to the length of ground-based arms (4km) allow us to probe a different gravitational-wave frequency regime (nHz), a complement to the ground-based kHz regime \\citep{Yardley10}. For almost 30 years pulsar timers have been putting limits on the energy density of the stochastic background using pulsar timing \\citep{Romani83, Stinebring90, Kaspi94, Lommenthesis, Jenet06, vanHaasteren11, Yardley11}. They point out that at some moment in the future, we will detect rather than limit the stochastic background. This moment is predicted to be sometime within this decade \\citep{Demorest09, Verbiest09}. In the last 10 years the field of gravitational-wave detection using pulsars has matured, and we are now considering much more than just the background of gravitational waves. We are demonstrating that very precise work on specific sources can be done, and that we need to `tune' this detector in order to maximize our sensitivity to these sources. This manuscript briefly reviews these efforts, and is organized as follows. In \\S \\ref{sec:overview} I give some more details about the concept and current thought behind using pulsar timing arrays (PTAs) to detect gravitational waves. In the subsequent sections I review the work that shows that PTAs can be (\\S \\ref{sec:directional}) directional detectors, (\\S \\ref{sec:waveform}) used to recover the gravitational waveform, (\\S \\ref{sec:time}) used to recover information about the source at some time in past, (\\S \\ref{sec:distance}) used to measure luminosity distance to gravitational-wave sources, (\\S \\ref{sec:altgravity}) used to test alternate theories of gravity, and (\\S \\ref{sec:noise}) characterized as a formal `detector' using measurements of their noise. Finally, in \\S \\ref{sec:summary} I summarize the ways in which a PTA is no longer a `blunt' instrument for gravitational-wave detection, but rather a tunable, pointable, and adjustable detector that can be used to gain very specific astrophysical information about the gravitational-wave source being detected. ", "conclusions": "" }, "1112/1112.2534_arXiv.txt": { "abstract": "We report the observation in the GeV band of the blazar 1ES~0229+200, which over recent years has become one of the primary sources used to put constraints on the Extragalactic Background Light (EBL) and Extragalactic Magnetic Field (EGMF). We derive constraints on both the EBL and EGMF from the combined Fermi-HESS data set taking into account the direct and cascade components of the source spectrum. We show that the limit on the EBL depends on the EGMF strength and vice versa. In particular, an EBL density twice as high as that derived by Franceschini et al. (2008) is allowed if the EGMF is strong enough. On the other hand, an EGMF strength as low as $6 \\times 10^{-18}$~G is allowed if the EBL density is at the level of the lower bound from the direct source counts. We present the combined EBL and EGMF limits as an exclusion plot in two dimensional parameter space: EGMF strength vs. EBL density. ", "introduction": "Very-high-energy (VHE) \\gr\\ flux from distant blazars is absorbed on the way from the source to the Earth through its interaction with the Extragalactic Background Light (EBL) photons (\\cite{gould}). The measurement of the induced distortions of the VHE \\gr\\ flux from distant hard-spectrum blazars by the effect of absorption on the EBL was used to derive constraints on the EBL density (\\cite{AhaEBL,Aha1ES0229,Orr_EBL_1ES0229}). The conventional derivation of the upper bound on the EBL from \\gr\\ observations adopts the assumption that the intrinsic powerlaw-type spectrum of the primary source (a distant blazar) is characterized by the slope $dN_\\gamma/dE\\sim E^{-\\Gamma}$ with $\\Gamma \\ge 1.5$. This assumption appears reasonable in the framework of the most simple synchrotron-self-Compton (SSC) models for the broad band spectra of blazars. However, particular blazars considered for the derivations of EBL limits (the blazars with hardest intrinsic \\gr\\ spectra) may well not fit into this simplest SSC model framework, so that the assumption of $\\Gamma\\ge 1.5$ might not be applicable (\\cite{aharonian08,bottcher08,katarzinski06,Lefa_hard1,Lefa_hard2,neronov11}). If the intrinsic spectra of the blazars used for the derivation of the upper bound on the EBL density are harder than $\\Gamma=1.5$, this upper limit is relaxed (\\cite{mazin}). Constraints on the intrinsic slope of the spectra of blazars can be obtained from the observations by the Fermi Large Area Telescope (LAT) (\\cite{atwood09}) in the energy band below $\\sim 100$~GeV, where the effect of absorption on the EBL becomes negligible. However, the blazars used for the derivation of constraints on the EBL are characterized by hard spectra, which makes it difficult to observe their flux below 100~GeV. In fact, the blazar 1ES~0229+200, which provides the tightest constraints on the EBL (\\cite{Aha1ES0229}) is not listed in the catalogue of sources detected by LAT in two-year exposure (\\cite{fermi_catalog}), with only upper limits on the source flux derived from the LAT data (\\cite{NeronovEGMF,Tavecchio:2010mk,TaylorEGMF}) and a weak detection reported by \\cite{Orr_EBL_1ES0229}. An additional difficulty for such constraints is that the spectrum of hard blazars might be composed of two contributions. Apart from the direct \\gr\\ emission from the primary source, an additional contribution is expected from the \\gr\\ cascade initiated in the intergalactic medium (IGM) by the absorbed VHE $\\gamma$ rays (\\cite{aharonian94,plaga95}). The overall flux and the spectral shape of the cascade contribution are determined by the strength of the extragalactic magnetic field (EGMF) (\\cite{Aharonian_cascade,plaga95,neronov07,NerSem_prediction}). Uncertainty of the EGMF strength introduces an uncertainty of the importance of the cascade contribution and prevents the measurement of the slope of the intrinsic \\gr\\ spectrum of the source. In fact, the limits on the EBL derived up to now are based on an underlying assumption about the EGMF strength (the EGMF should be strong enough to suppress the cascade contribution up to the $\\sim$TeV energy band), which is not justified a-priori. If the assumption on the EGMF strength is relaxed, the \\gr\\ data can be used to measure the EGMF strength. The electron-positron pairs, created as a result of the absorption of multi-TeV photons, up-scatter the Cosmic microwave background (CMB) as they cool, creating secondary emission in the GeV domain (\\cite{NerSem_prediction}). With the mean-free-path of TeV $\\gamma$ rays being $D_\\gamma \\simeq 100~(E_\\gamma/10~ {\\rm TeV})^{-1}$~Mpc, the production of these multi-TeV energy $e^+$-$e^-$ pairs occurs predominantly in extragalactic space. Furthermore, if non-negligible magnetic fields are present in the region these pairs are born into, they deflect significantly from their initial directions, resulting in the secondary GeV photons having a reduced probability of reaching the observer. Thus, the apparent flux is suppressed at low energies. Observational evidence for the presence of such a suppression places constraints on the EGMF strength to be $\\gtrsim 10^{-17}$~G (\\cite{Dermer,Dolag:2010ni,NeronovEGMF,Tavecchio:2010mk,TaylorEGMF}). This limit on the EGMF is derived assuming a certain level of EBL density (a low EBL density derived by \\cite{Franceschini_EBL} is used in most of the publications on the EGMF in order to derive conservative bounds). However, a variety of EBL models do exist (\\cite{gilmore09, kneiske04, stecker_ebl}), which spans the full range of the present uncertainty in the EBL density. This uncertainty therefore introduces an uncertainty into the bounds on the EGMF derived from the \\gr\\ data. Higher(lower) EBL density leads to stronger(weaker) absorption of the primary \\gr\\ flux along with a stronger(weaker) cascade contribution to the observed flux in the GeV range. The suppression of such stronger(weaker) cascade contribution would require a stronger(weaker) EGMF. Thus, the limits on the EBL derived from the \\gr\\ data depend on the assumptions made about the EGMF strength and vice versa. This implies that the correct procedure for the derivation of limits on the EBL from the \\gr\\ data should include marginalization over the possible EGMF values. Conversely, the correct procedure for the derivation of the limits on the EGMF should include marginalization over the possible EBL densities and spectra. Practical implementation of this correct procedure for the derivation of the EBL and EGMF bounds, however, present a challenging task since the quality of the \\gr\\ data is usually insufficient for the exploration of the entire EBL-EGMF parameter space. Exploration of this parameter space requires the measurement of the source spectra both in the TeV and GeV energy ranges. In this Letter we report the observations of the blazar 1ES~0229+200 in the 1-300~GeV energy range using LAT's data with three years of exposure. The detection of the source below 100~GeV provides the information necessary for the correct analysis of the EGMF-dependent upper bound on the EBL density and of the EBL-dependent lower bound on the EGMF strength. We present such bounds in the form of a two-dimensional exclusion plot in the ``EBL density'' vs. ``EGMF strength'' parameter space. ", "conclusions": "Following our investigations into the newly detected blazar 1ES~0229+200 with Fermi/LAT we find that, if the \\cite{Franceschini_EBL} EBL level is adopted, an EGMF strength of at least $10^{-17}$~G is required in order for the inevitable GeV cascade spectral component to be consistent with observations. However, we find that, more generally, the EGMF lower bound is dependent on the EBL level adopted, as demonstrated in Fig.~\\ref{EBL_exclusion}. One should keep in mind that this bound is derived for large ($\\lambda_B > 1$~Mpc) correlation lengths of the EGMF, and scales approximately as $\\lambda_B^{-1/2}$ for $\\lambda_B \\lesssim 1$~Mpc. Our result lies in agreement with previous findings (\\cite{TaylorEGMF,Huan:2011kp}). Under the assumption that the source is stable on a time-scale much longer than 3 years, a larger bound on the EGMF is obtained (\\cite{NeronovEGMF,TaylorEGMF,Dolag:2010ni,Tavecchio:2010mk}). We also find that an intrinsic blazar spectral index of $\\Gamma=1.5$ is able to sit in agreement with observations of 1ES~0229+200 by Fermi/LAT ($\\Gamma=1.36\\pm 0.25$). Thus our results can find agreement with the underlying assumptions behind previous EBL constraint calculations by \\cite{AhaEBL}. However, the remaining uncertainty in the origin of the blazar's spectral shape in the GeV domain leaves open the possibility for a harder intrinsic spectrum ($\\Gamma<1.5$). We note, that in previous works of \\cite{AhaEBL} and \\cite{Orr_EBL_1ES0229} the authors made the implicit assumption for the absence of the cascade contribution in the GeV-TeV domain, which is equivalent to the presence of the strong EGMF with $B \\gtrsim 10^{-15}$~G. As can be seen from Fig. \\ref{EBL_exclusion}, for such a strong field our result is also compatible with the level of EBL found by \\cite{Orr_EBL_1ES0229}. For weaker fields, $B \\sim 10^{-17}$~G, we are also compatible with the findings of \\cite{dominguez11}. However, for such weak fields a proper account should be given of the GeV-TeV cascade component contribution. Future observations with ground-base Cherenkov telescopes, such as MAGIC Stereo, HESS 2 and CTA, will dramatically improve the measurements of blazar spectra below $\\sim 100$~GeV, allowing much better constraints on the EGMF and EBL. In the particular case of 1ES~0229+200, if the intrinsic spectrum is hard, the presented EGMF lower bound may be transformed to a \\textit{measurement}, helping to better understand the nature of the extragalactic magnetic field." }, "1112/1112.2228_arXiv.txt": { "abstract": "We report on the \\xray\\ and multiwavelength properties of 11 radio-quiet quasars with weak or no emission lines identified by the Sloan Digital Sky Survey (SDSS) with redshift $z=0.4$--$2.5$. Our sample was selected from the Plotkin et~al. catalog of radio-quiet, weak-featured AGNs. The distribution of relative \\xray\\ brightness for our \\hbox{low-redshift} weak-line quasar (WLQ) candidates is significantly different from that of typical radio-quiet quasars, having an excess of \\xray\\ weak sources, but it is consistent with that of high-redshift WLQs. Over half of the \\hbox{low-redshift} WLQ candidates are \\xray\\ weak by a factor of $\\gtrsim 5$, compared to a typical SDSS quasar with similar UV/optical luminosity. These \\xray\\ weak sources generally show similar UV emission-line properties to those of the \\xray\\ weak quasar \\phl\\ (weak and blueshifted high-ionization lines, weak semi-forbidden lines, and strong UV Fe emission); they may belong to the notable class of \\phl\\ analogs. The average \\xray\\ spectrum of these sources is somewhat harder than that of typical radio-quiet quasars. Several other \\hbox{low-redshift} WLQ candidates have normal ratios of \\xray-to-optical/UV flux, and their average \\xray\\ spectral properties are also similar to those of typical radio-quiet quasars. The \\xray\\ weak and \\xray\\ normal WLQ candidates may belong to the same subset of quasars having high-ionization ``shielding gas'' covering most of the wind-dominated broad emission-line region, but be viewed at different inclinations. The mid-infrared-to-X-ray spectral energy distributions (SEDs) of these sources are generally consistent with those of typical SDSS quasars, showing that they are not likely to be BL~Lac objects with relativistically boosted continua and diluted emission lines. The mid-infrared-to-UV SEDs of most radio-quiet weak-featured AGNs without sensitive X-ray coverage (34 objects) are also consistent with those of typical SDSS quasars. However, one source in our \\xray\\ observed sample is remarkably strong in \\hbox{X-rays}, indicating that a small fraction of \\hbox{low-redshift} WLQ candidates may actually be BL~Lacs residing in the radio-faint tail of the BL~Lac population. We also investigate universal selection criteria for WLQs over a wide range of redshift, finding that it is not possible to select WLQ candidates in a fully consistent way using different prominent emission lines (e.g., Ly$\\alpha$, \\civ, \\mgii, and H$\\beta$) as a function of redshift. ", "introduction": "Strong and broad line emission is a common feature of quasar spectra in the optical and UV bands. However, since multi-color quasar selection at high redshift in the Sloan Digital Sky Survey (SDSS; York et~al. 2000) is mostly based upon the presence of the Ly$\\alpha$ forest and Lyman break (e.g., Richards et~al.\\ 2002), the SDSS can also effectively select high-redshift quasars with weak or no emission lines. About 90 such weak-line quasars (WLQs) at high redshift have been found with \\lyanv\\ rest-frame equivalent widths of \\hbox{REW~$<15$~\\AA} (e.g., Fan et~al.\\ 1999, 2006; Anderson et~al.\\ 2001; Collinge et~al.\\ 2005; Diamond-Stanic et~al.\\ 2009, hereafter DS09). Some of these objects show a hint of weak Ly$\\alpha$ emission but no other lines; others are completely bereft of detectable emission lines even in high-quality spectra. High-redshift SDSS quasars show an approximately log-normal distribution of \\lyanv\\ REW with a mean of $\\approx 62$~\\AA\\ (DS09). The WLQs constitute $\\simgt 3\\sigma$ negative deviations from the mean, and there is no corresponding population with $\\simgt 3\\sigma$ positive deviations. The majority of these high-redshift WLQs are radio quiet ($\\alpha_{\\rm ro}>-0.21$; \\aro\\ is the slope of a nominal power law between 5~GHz and 2500~\\AA\\ in the rest frame; see \\S\\ref{xray} for a full definition). WLQs have mainly been studied at high redshifts due to the fact that the Ly$\\alpha$ forest enters into the SDSS spectroscopic coverage for quasars at $z>2.2$. However, there is no apparent reason to believe that these objects should not also exist at lower redshifts. Indeed, a few apparent analogs of WLQs at lower redshifts have been found serendipitously over the past $\\approx 15$ years; e.g., PG~1407+265 (McDowell et~al.\\ 1995; \\hbox{$z=0.94$}), 2QZJ2154--3056 (Londish et~al.\\ 2004; \\hbox{$z=0.49$}), and PHL~1811 (Leighly et~al.\\ 2007ab; \\hbox{$z=0.19$}). As a byproduct of a systematic survey for optically selected BL~Lacertae objects (hereafter BL Lacs) in SDSS Data Release~7 (DR7; Abazajian et~al. 2009), Plotkin et~al. (2010a) discovered about 60 additional radio-quiet WLQ candidates at $z<2.2$ for which all emission features have \\hbox{REW~$<5$~\\AA}. These objects are perhaps the first \\hbox{low-redshift} SDSS counterparts of the previously identified high-redshift SDSS WLQs. Following the nomenclature that has been established by previous work on WLQs (e.g., Shemmer et~al. 2009), we define ``high redshift'' as $z>2.2$ and ``low-redshift'' as $z\\leqslant 2.2$, because WLQs are selected with different approaches for these redshift ranges (see above). Although WLQs are rare, their exceptional characteristics constitute a challenge to our overall understanding of quasar geometry and physics, especially the quasar broad emission-line region (BELR). Analogously, physical insights have been gained by investigating other minority populations with exceptional emission-line or absorption-line properties, such as Narrow-Line Seyfert~1 (NLS1) galaxies and Broad Absorption Line (BAL) quasars. Therefore, extensive studies of the multi-band properties of WLQs should have scientific value. There are several candidate explanations for the physical nature of WLQs. Their UV emission lines may be weak due to an ``anemic'' BELR with a significant deficit of line-emitting gas (e.g., Shemmer et al. 2010). It has also been speculated that WLQs may represent an early stage of quasar evolution in which an accretion disk has formed and emits a typical continuum, but BELR formation is still in progress (e.g., Hryniewicz et~al. 2010; Liu \\& Zhang 2011). The weak UV emission lines may also be a consequence of a spectral energy distribution (SED) which lacks high-energy ionizing photons. This soft SED may be a result of unusual accretion rate. For example, an extremely high accretion rate might produce a UV-peaked SED (e.g., Leighly et~al. 2007). In this scenario, high-ionization lines, like \\civ, should be suppressed relative to low-ionization lines like H$\\beta$. However, Shemmer et~al. (2010) estimated the normalized accretion rates, $L/L_{\\rm Edd}$, of two high-redshift WLQs via near-infrared spectroscopy and found their accretion rates were within the range for typical quasars with similar luminosities and redshifts. Alternatively, a combination of low accretion rate and large black hole mass may lead to a relatively cold accretion disk that emits few ionizing photons. Laor \\& Davis (2011) predicted a steeply falling SED at $\\lambda\\;<\\;1000$~\\AA\\ for quasars with cold accretion disks, and such an SED was observed in the WLQ SDSS~J0945+1009 by Hryniewicz et~al. (2010). High-energy ionizing photons (including \\hbox{X-rays}) may be heavily absorbed before they reach the BELR. Wu et~al. (2011) studied a population of \\xray\\ weak quasars with unusual UV emission-line properties like those of \\phl\\ (weak and highly blueshifted high-ionization lines, weak semi-forbidden lines, and strong UV Fe emission). All of their radio-quiet \\phl\\ analogs were found to be \\xray\\ weak by a factor of $\\approx13$ on average. These objects also show a harder average \\xray\\ spectrum than those for typical quasars which suggests the presence of \\xray\\ absorption. \\phl\\ analogs appear observationally to be a significant subset ($\\approx 30\\%$) of WLQs. The existence of a class of quasars with high-ionization ``shielding gas'' covering most of the BELR, but little more than the BELR, could potentially unify the \\phl\\ analogs and WLQs via orientation effects (see \\S4.6 of Wu et~al. 2011). The shielding gas would absorb high-energy ionizing photons before they reach the BELR, resulting in weak high-ionization emission lines. When such a quasar is observed through the BELR and the shielding gas, a \\phl\\ analog would be seen; when it is observed along other directions, an \\xray\\ normal WLQ would be observed. Another possibility is that instead of being intrinsically weak, the UV emission lines of WLQs could in principle be diluted by a relativistically boosted UV/optical continuum as for BL~Lac objects. However, this scenario is not likely for most WLQs. Shemmer et~al. (2009) found that the X-ray properties of high-redshift WLQs are inconsistent with those of BL~Lac objects. Furthermore, there is no evidence of strong optical variability or polarization for these WLQs (see DS09; Meusinger et~al. 2011). The UV-to-infrared SEDs of high-redshift WLQs are also similar to those of typical quasars, while the SEDs of BL~Lac objects are much different (DS09; Lane et~al. 2011). Nevertheless, it is possible that the population of BL~Lac objects has a small radio-quiet tail (e.g., Plotkin et~al. 2010b) and that a small fraction ($\\lesssim5\\%$; see Lane et~al. 2011) of the general WLQ population may be BL~Lac objects. Most previous studies of WLQs were based on high-redshift objects. To investigate the nature of the overall WLQ population, we obtained new \\xray\\ observations of \\hbox{low-redshift} WLQs selected mainly from the catalog of radio-quiet BL~Lac candidates in Plotkin et~al. (2010a). We also utilized sensitive archival \\xray\\ coverage of the sources in their catalog. Our closely related science goals are the following: (1) enable comparison of the broad-band SEDs of \\hbox{low-redshift} WLQs to those of high-redshift WLQs, typical radio-quiet quasars, and BL~Lac objects; (2) provide basic constraints upon \\hbox{X-ray} spectral properties via band-ratio analysis and joint spectral fitting; (3) clarify if there is broad-band SED diversity among \\hbox{low-redshift} WLQs; and (4) allow reliable planning of future long, spectroscopic \\hbox{X-ray} observations. In \\S\\ref{sample} we describe the selection of our sample of \\hbox{low-redshift}, radio-quiet WLQ candidates. In \\S\\ref{uvo} we detail their UV/optical observations and the measurement of their rest-frame UV spectral properties. In \\S\\ref{xray} we describe the relevant \\hbox{X-ray} data analyses. Overall results and associated discussion are presented in \\S\\ref{discuss}. Throughout this paper, we adopt a cosmology with $H_0=70.5$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_{\\rm M}=0.274$, and $\\Omega_{\\Lambda}=0.726$ (e.g., Komatsu et~al. 2009). ", "conclusions": "\\label{discuss} \\subsection{Relative X-ray Brightness}\\label{discuss:daox} The \\daox\\ parameter (see Eqn.~\\ref{daoxeqn} for definition) is utilized to assess the \\xray\\ brightness of a quasar relative to typical radio-quiet quasars with similar UV luminosity. We compare the \\daox\\ distribution of our \\hbox{low-redshift}, radio-quiet WLQ candidates (see Fig.~\\ref{daox_fig}) to that of the 132 radio-quiet, non-BAL quasars in Sample B of Gibson et~al. (2008),\\footnote{We used an improved version of the Sample~B quasars in Gibson et~al. (2008) from which we further removed seven BAL quasars (see Footnote~16 in Wu et~al. 2011).} which represent typical radio-quiet SDSS quasars. All of the 132 Sample~B quasars are \\xray\\ detected. The Peto-Prentice test (e.g., Latta 1981), implemented in the Astronomy Survival Analysis package (ASURV; e.g., Lavalley et~al. 1992), is used to assess whether our \\hbox{low-redshift} WLQ candidates follow the same \\daox\\ distribution as that for typical quasars (see results in Table~\\ref{twost_table}). We prefer the Peto-Prentice test to other possible similar tests because it is the least affected by the factors of different censoring patterns or unequal sizes of the two samples which exist in our case. We also compare the \\daox\\ distribution of high-redshift, radio-quiet WLQs in Shemmer et~al. (2006, 2009) to that of our \\hbox{low-redshift}, radio-quiet WLQ candidates and that of typical SDSS quasars (also see Fig.~\\ref{daox_fig} and Table~\\ref{twost_table}). The \\daox\\ distribution of our \\hbox{low-redshift}, radio-quiet WLQ candidates is significantly different from that of typical SDSS quasars. The probability of null-hypothesis (two samples following the same distribution) is only $6.3\\times 10^{-7}$. This result is mainly due to the presence of a skew tail of \\xray\\ weak WLQs (see Fig.~\\ref{daox_fig}). Seven out of the 11 objects in our sample of \\hbox{low-redshift}, radio-quiet WLQ candidates have \\daox$\\;<-0.2$, giving a fraction of \\xray\\ weak objects of ($64^{+34}_{-24}$)\\% (68\\% confidence level). The mean \\daox\\ value for the \\hbox{low-redshift}, radio-quiet WLQ candidates is $-0.214\\pm0.078$, calculated with the Kaplan-Meier estimator also implemented in the ASURV package, while that for the Sample B quasars is $-0.001\\pm0.011$. The \\daox\\ distribution of the nine high-redshift, radio-quiet WLQs in Shemmer et~al. (2006, 2009) is also different from that of typical radio-quiet SDSS quasars, but less significantly (the probability of null-hypothesis is $7.1\\times 10^{-3}$).\\footnote{This result is somewhat inconsistent with the finding by Shemmer et~al. (2009) that their high-redshift WLQs have a similar \\daox\\ distribution to that of typical SDSS quasars. Shemmer et~al. (2009) used the Kolmogorov-Smirnov test and ignored the two high-redshift, radio-quiet WLQs with \\daox\\ upper limits. We include all nine high-redshift, radio-quiet WLQs since the Peto-Prentice test can properly treat censored data. The utilization of an improved Sample~B (see Footnote~17) does not substantially contribute to the inconsistency here. The Peto-Prentice test using all nine high-redshift, radio-quiet WLQs and the original Sample~B (as used by Shemmer et~al. 2009) provides a null-hypothesis probability of $1.53\\times 10^{-2}$.} Five of the nine objects in the \\hbox{high-redshift}, radio-quiet WLQ sample have \\daox$\\;<\\;-0.2$, giving a fraction of \\xray\\ weak objects of ($56^{+37}_{-24}$)\\% (68\\% confidence level); we note that the fraction could be somewhat higher (6/9) owing to the weak \\xray\\ upper limit for J1237+6301. The mean \\daox\\ value for the \\hbox{high-redshift}, radio-quiet WLQs is $-0.144\\pm0.075$. As expected the combined \\hbox{low-redshift} and high-redshift, radio-quiet WLQ sample (mean \\daox\\ value of $-0.187\\pm0.056$) also follows a different \\daox\\ distribution from that of typical SDSS quasars (null-hypothesis probability of $4.5\\times10^{-6}$). The \\daox\\ distribution of \\hbox{low-redshift}, radio-quiet WLQ candidates is consistent with that of high-redshift, radio-quiet WLQs (null-hypothesis probability of 0.35), though the sample sizes being compared are limited. \\subsection{Classifying Radio-Quiet WLQs}\\label{discuss:class} \\begin{figure*}[t] \\centering \\includegraphics[width=6.3in]{fig07color.ps} \\vspace{0.1cm} \\caption{\\footnotesize{$\\alpha_{\\rm ro}$--$\\alpha_{\\rm ox}$ diagram for WLQ candidates (red filled circles for \\hbox{low-redshift} objects in our sample; blue filled squares for high-redshift objects in Shemmer et al. 2006, 2009), BL~Lac objects (black open circles; Shemmer et al. 2009), and typical radio-quiet SDSS quasars (small green dots; Sample~B quasars of Gibson et~al. 2008). The red asterisk represents PHL~1811. Rightward (downward) pointing arrows represent \\aox\\ (\\aro) upper limits. The two dashed lines mark the criteria for radio-quiet (\\aro $> -0.21$), radio-intermediate ($-0.39 <$\\aro$<-0.21$), and radio-loud (\\aro$<-0.39$) objects. The boxes bordered by dotted lines show the three suggested groups of WLQ candidates based on their multi-band properties. Note that the WLQ sample has an excess of objects with large negative \\aox\\ values, compared to both typical radio-quiet quasars and BL~Lac objects.} \\label{aroaox_fig}} \\end{figure*} \\begin{figure*}[t] \\centering \\includegraphics[width=6.0in]{fig08color.ps} \\caption{\\footnotesize{REW(\\civ) plotted against the \\civ\\ blueshift for our radio-quiet WLQ candidates (filled squares for \\xray\\ detected sources, filled upside-down triangles for \\xray\\ undetected sources), \\phl\\ (asterisk), radio-quiet \\phl\\ analogs in Wu et~al. (2011) (stars), and \\hbox{radio-quiet}, \\hbox{non-BAL} quasars in Sample~B of Gibson et~al. (2008a) (circles). These sources are color-coded according to their \\daox\\ values (three color bins are used, corresponding to the \\xray\\ weak, \\xray\\ normal and \\xray\\ strong sources described in \\S\\ref{discuss:class}, respectively). The color bar shows the \\daox\\ range for each color. Source names for WLQs are labeled in the format of 'J$hhmm$' for brevity. The grey dots show the 13,582 \\hbox{radio-quiet} quasars in Sample A of Richards et~al. (2011; see their Fig.~7).} \\label{c4bew_fig}} \\end{figure*} To investigate the multi-band properties of \\hbox{low-redshift}, radio-quiet WLQ candidates, we plotted the sources of our sample in an \\aro-\\aox\\ diagram (Fig.~\\ref{aroaox_fig}) along with the high-redshift, radio-quiet WLQs in Shemmer et~al. (2006, 2009), the BL~Lac sample in Shemmer et~al. (2009), and the Sample~B quasar in Gibson et~al. (2008). The \\hbox{low-redshift}, radio-quiet WLQ candidates have similar multi-band properties to those of high-redshift, radio-quiet WLQs. They are generally much fainter in radio and \\hbox{X-rays} than most of the BL~Lac objects. The weak emission lines of \\hbox{low-redshift}, radio-quiet WLQ candidates are therefore not likely due to the dilution by relativistically boosted continua as for BL~Lac objects (see discussion in \\S4.1 of Shemmer et~al. 2009). However, it is possible that a small percentage of the WLQ candidates actually belong to the radio-faint tail of the BL~Lac population (see below). The population of radio-quiet WLQ candidates (both at low redshift and high redshift) has a wide dispersion of relative \\xray\\ brightness and UV emission-line properties. Motivated by their \\daox\\ distribution, their emission-line properties discussed below, and observations of related objects (e.g., Wu et~al. 2011), we will discuss them in three groups. The majority of WLQ candidates are not only \\xray\\ weaker than BL~Lac objects, but also weaker than typical radio-quiet SDSS quasars. These WLQ candidates may belong to the notable class of \\xray\\ weak quasars termed ``\\phl\\ analogs'' which were recently studied in detail by Wu et~al. (2011). The \\phl\\ analogs generally have weak and highly blueshifted high-ionization lines (e.g., \\civ, Si~{\\sc iv}), weak semi-forbidden lines (e.g., C~{\\sc iii}]), and strong UV Fe~{\\sc ii} and/or Fe~{\\sc iii} emission. Some of our \\hbox{low-redshift}, radio-quiet WLQs have similar UV emission-line properties to those of \\phl, as listed below: \\begin{enumerate} \\item J0812+5225 (\\daox$=-0.42$) has weak C~{\\sc iii}] and strong Fe~{\\sc ii} emission. \\item J0945+1009 (\\daox$\\;<\\;-0.34$) has weak \\civ\\ and C~{\\sc iii}] emission lines. Its \\civ\\ line is highly blueshifted ($\\approx -7000$~km~s$^{-1}$). \\item J1252+2640 (\\daox$=-0.39$) has weak C~{\\sc iii}] and strong Fe~{\\sc ii} emission. \\item J1139$-$0201 (\\daox$=-0.38$) has weak and highly blueshifted ($\\approx -2950$~km~s$^{-1}$) \\civ\\ emission, weak C~{\\sc iii}] emission, and strong Fe~{\\sc iii} emission. \\end{enumerate} A high-redshift, radio-quiet WLQ J1302+0030 (\\daox$=-0.38$) also has a weak and highly blueshifted \\civ\\ emission line (DS09; Wu et~al. 2011). All of the above mentioned sources are \\xray\\ weak by a factor of $>\\;7$ (see Table~\\ref{aox_table}). Fig.~\\ref{c4bew_fig} shows the distribution of our WLQ candidates in the \\daox$-$\\ion{C}{4} blueshift$-$ REW(\\ion{C}{4}) parameter space. J0945+1009, J1139$-$0201 and J1302+0030 are similar to \\phl\\ analogs in this diagram. Based on the model in \\S4.6 of Wu et~al. (2011), these \\phl\\ analogs may have high-ionization shielding gas with large column density and a large covering factor of the BELR which blocks most of the ionizing photons, resulting in weak high-ionization emission lines. If a quasar of this kind is viewed through the BELR and shielding gas, it would be an \\xray\\ weak WLQ with weak and highly blueshifted high-ionization lines (e.g., \\civ). Based on the estimate in \\S4.6 of Wu et~al. (2011), \\phl\\ analogs should make up $\\approx 30\\%$ of the total WLQ population. However, our sample of low-redshift, radio-quiet WLQ candidates appears to have a higher fraction ($\\gtrsim 50\\%$) of \\phl\\ analogs, which may indicate our sample has some selection bias toward \\xray\\ weak WLQ candidates. This bias could perhaps be the result of a more strict criterion upon the strengths of emission lines for most sources in our sample (REW~$\\lesssim$~5~\\AA). Quasars with weaker emission lines (e.g., \\civ) are perhaps more likely to be weak in \\hbox{X-rays} (e.g., see \\S4.5 of Wu et~al. 2011). The apparently higher fraction of \\phl\\ analogs in our sample than that in Wu et~al. (2011) may perhaps also be caused simply by small-sample statistics. A Fisher's exact test (Fisher 1922) gives an $11.1\\%$ probability for the different fractions of \\phl\\ analogs among these two samples under the null hypothesis (i.e., the two samples have the same fraction of \\phl\\ analogs). Some of our WLQ candidates have similar \\xray\\ brightness to that of typical radio-quiet quasars ($-0.2\\lesssim$~\\daox~$\\lesssim0.2$). Their high-ionization lines are also weak, but perhaps not highly blueshifted (e.g., see Fig.~\\ref{c4bew_fig} for J1408+0205). Some of them (e.g., J1612+5118) have very weak UV Fe~{\\sc ii} and/or Fe~{\\sc iii} emission. J1612+5118 does seem to have a highly blueshifted \\civ\\ line, for which the reason is unclear. However, the \\civ\\ line of this source is close to the blue border of its SDSS spectral coverage. Further UV spectroscopy with better \\civ\\ coverage is needed to confirm its \\civ\\ blueshift. Based on the model in Wu et~al. (2011), these sources are similar to \\phl\\ analogs physically, but they are viewed at different orientations. These sources are observed along lines of sight that avoid the shielding gas and the BELR. Therefore they appear normal in \\hbox{X-rays}. Their high-ionization lines are generally not highly blueshifted. In our WLQ candidate sample, one source (J1109+3736) is remarkably strong in \\hbox{X-rays}. It also shows similar UV/optical spectral properties to those of BL~Lac objects. This source may belong to the radio-faint tail of the BL~Lac population; we will discuss it further in \\S\\ref{discuss:j1109}. It is worth noting that the division of our radio-quiet WLQ candidates into the three groups discussed above (as shown in Fig.~\\ref{aroaox_fig}) is somewhat arbitrary. We do have some ``border-line'' sources with \\daox$\\;\\approx\\;\\pm0.2$ (e.g., J1212+5341). It is difficult to classify these sources clearly based on current information. \\subsection{The Infrared-to-X-ray SEDs of the Radio-Quiet WLQ Candidates}\\label{discuss:seds} \\begin{figure*}[t] \\centering \\includegraphics[width=6.2in]{fig09.ps} \\vspace{0.2cm} \\caption{\\footnotesize{Rest-frame spectral energy distributions (SEDs) of our \\hbox{low-redshift} WLQs with the best multiwavelength coverage, and of the radio-quiet BL~Lac candidate J1109+3736, ordered by \\daox. The photometric data points are from {\\it WISE} (open diamonds), 2MASS (open triangles), SDSS (filled circles), {\\it GALEX} (open circles), and X-ray observations (asterisks). The average SED of all SDSS quasars from the sample of Richards et al. (2006) is also shown (solid curve), scaled to the flux at rest-frame $10^{15}$~Hz. A parabolic SED for typical BL~Lac objects is shown by the dotted line in the top-left panel. The `+' signs in the bottom right-panel show the POSS photometry for J1109+3736.} \\label{sed1_fig}} \\end{figure*} For the purpose of investigating further the multi-wavelength SEDs of our \\hbox{low-redshift}, radio-quiet WLQs, we gathered photometry for our sample from the following bands: (1) near- and mid-infrared from {\\it WISE} (The Wide-field Infrared Survey Explorer; Wright et~al. 2010); (2) near-infrared from 2MASS (The Two Micron All Sky Survey; Skrutskie et~al. 2006); (3) optical from the SDSS; (4) UV from {\\it GALEX} (The Galaxy Evolution Explorer; Martin et~al. 2005); and (5) \\hbox{X-rays} from this work. Fig.~\\ref{sed1_fig} shows the SEDs of the five \\hbox{low-redshift} WLQs in our sample which have the best multi-band coverage. We also include the SED of the radio-quiet BL~Lac candidate J1109+3736 (see more discussion in \\S\\ref{discuss:j1109}). A key point to keep in mind is that these multi-band observations are non-simultaneous. The SEDs are therefore subject to potential distortions due to variability. We examined the {\\it WISE} and {\\it GALEX} image tiles by eye to identify potential cases of source blending, confusion, or incorrect matching caused by the low angular resolution of {\\it WISE} and {\\it GALEX}. None of the sources in our sample is subject to these kinds of problems. The 2MASS magnitudes in the SDSS DR7 quasar catalog were utilized; this catalog provided aperture photometry for additional sources detected down to $2\\sigma$ (see \\S5 of Schneider et~al. 2010). For the sources without detections at $\\geq2\\sigma$, we adopted flux upper limits obtained following the same photometry procedure (C.~M.~Krawczyk \\& G.~T.~Richards 2011, private communication). The first five sources in Fig.~\\ref{sed1_fig} are from the groups of \\xray\\ weak and \\xray\\ normal WLQ candidates discussed above. Their mid-infrared-to-UV SEDs are generally consistent with the composite SEDs of typical SDSS quasars in Richards et~al. (2006), and they are significantly different from the SEDs of BL~Lac objects (see the dotted parabolic line in the top-left panel of Fig.~\\ref{sed1_fig}; e.g., Nieppola et~al. 2006). We also investigated the SEDs for the {\\it WISE}-covered radio-quiet objects cataloged in Plotkin et~al. (2010a) which do not have sensitive \\xray\\ coverage (see the Appendix). The majority of them also have mid-infrared-to-UV SEDs consistent with those of typical radio-quiet quasars. Lane et~al. (2011) obtained similar results for their high-redshift WLQs; the composite SED of their high-redshift WLQs is inconsistent with SEDs of BL~Lac objects. For one source in our sample, J0945+1009, the flux in the UV band is lower than for typical SDSS quasars (see the {\\it GALEX} data points in the top-right panel of Fig.~\\ref{sed1_fig}). The UV deficiency of this source may be caused by Ly$\\alpha$-forest intervening absorption. However, Laor \\& Davis (2011) argued that such intervening absorption is not significant ($\\sim11\\%$ at most) for this source with $z=1.66$. The near-infrared-to-UV SED of J0945+1009 can be well fitted with their local black-body model for a cold accretion disk. \\subsection{X-ray Spectral Properties of \\hbox{Low-Redshift} WLQ Candidates}\\label{discuss:stack} Most of the \\hbox{low-redshift}, radio-quiet WLQ candidates do not have sufficient \\xray\\ counts for an individual \\xray\\ spectral analysis. We therefore investigate the average \\xray\\ spectral properties for these sources via stacking analyses and joint fitting. A stacked spectral analysis was performed for the six \\hbox{low-redshift}, radio-quiet WLQ candidates with \\daox$\\;<\\;-0.3$ (J0812+5225, J0945+1009, J1139$-$0201, J1252+2640, J2115+0001, and J2324+1443). These sources are the weakest in \\hbox{X-rays} among the full sample of \\hbox{low-redshift} WLQ candidates. The detected sources have similar numbers of \\xray\\ counts, so that any one of them will not dominate the stacking analysis. These six sources span a relatively wide range of redshift ($z=1.15$--$2.50$), and thus the observed-frame bands of each source correspond to different energy ranges in the rest frame. However, under the assumption of a simple power-law spectral model, one can stack the \\xray\\ counts to obtain the average effective power-law photon index. We added the \\xray\\ counts of these sources in the observed-frame soft band and hard band, respectively. The numbers of total net counts are $14.4^{+4.9}_{-3.7}$ in the soft band and $5.2^{+3.4}_{-2.2}$ in the hard band ($68\\%$ confidence level), and the resulting band ratio is $0.36^{+0.27}_{-0.18}$. With the average Galactic neutral hydrogen column density of these sources ($N_{\\rm H}=3.50\\times10^{20}$~cm$^{-2}$), the band ratio was converted to an effective power-law photon index $\\Gamma=1.66^{+0.63}_{-0.51}$. The average \\xray\\ spectrum of the \\xray\\ weak \\hbox{low-redshift} WLQs is perhaps somewhat harder than that for typical radio-quiet quasars ($\\Gamma\\approx2$), but it is consistent within the error bars. This average \\xray\\ spectrum is likely softer than that of the \\phl\\ analogs at $z=2.19$--$2.38$ ($\\Gamma=1.10^{+0.45}_{-0.40}$) in Wu et~al. (2011) which was also obtained via a stacking analysis, but consistent within 2$\\sigma$. Both stacking analyses suffer from large uncertainty due to limited \\xray\\ counts. For a sample combining the six \\xray\\ weak WLQs analyzed here (which are likely to be \\phl\\ analogs) and the radio-quiet \\phl\\ analogs in Wu et~al. (2011), the average \\xray\\ spectrum has a flat effective power-law photon index of $\\Gamma=1.35^{+0.33}_{-0.31}$. Deeper \\xray\\ observations are necessary to give tighter constraints on the \\xray\\ spectral properties of these \\xray\\ weak WLQ candidates. Two sources (J0945+1009 and J2115+0001) are undetected by \\chandra. Adding the \\xray\\ counts of these two sources cannot generate a stacked source that would be detected by \\chandra, because both of these two sources have zero \\xray\\ counts. However, we are able to obtain a tighter average constraint on their \\xray\\ brightness via stacking analysis. The upper limit upon the soft-band count rate of the stacked source is $2.51\\times10^{-4}$~s$^{-1}$. Average values of redshift, Galactic $N_{\\rm H}$, and $f_{2500\\mbox{\\rm~\\scriptsize\\AA}}$ are adopted in the following calculation. The upper limit upon the average flux density at rest-frame 2~keV is $5.16\\times10^{-33}$~erg~cm$^{-2}$~s$^{-1}$~Hz$^{-1}$ under the assumption of the Galactic-absorbed power law with $\\Gamma=2$. The upper limits upon \\aox\\ and \\daox\\ are calculated to be \\aox~$<\\;-2.21$ and \\daox~$<\\;-0.50$. Therefore, these two sources are \\xray\\ weak by a factor of $>\\;20$ on average. For the two \\xray\\ normal, \\hbox{low-redshift} WLQs (J1604+4326 and J1612+5118), we performed joint fitting to study the average \\xray\\ spectral properties of these sources. The \\xray\\ spectra were extracted from apertures of $3''$ radius centered on the \\xray\\ positions of these sources via the standard CIAO routine {\\sc psextract}. The background spectra were extracted from annular regions with inner radii of $6''$ and outer radii of $9''$, which are free of \\xray\\ sources. Two spectra were extracted individually for J1604+4326 from its two \\chandra\\ observations. Spectral fitting was performed with XSPEC v12.6.0 (Arnaud 1996). The $C$-statistic (Cash 1979) was used in the spectral fitting instead of the standard $\\chi^2$ statistic because the $C$-statistic is well suited to the limited \\xray\\ counts in our analysis (e.g., Nousek \\& Shue 1989). We fit the spectra jointly using a power-law model with a Galactic absorption component represented by the \\verb+wabs+ model (Morrison \\& McCammon 1983). We also used another model similar to the first, but adding an intrinsic (redshifted) neutral absorption component, represented by the \\verb+zwabs+ model. Both sources were assigned their own values of redshift and Galactic neutral hydrogen column density; the Galactic column density was fixed to the values calculated with COLDEN (Column~4 of Table~\\ref{aox_table}). The joint fitting results are shown in Table~\\ref{xspec_table}. The quoted errors or upper limits are at the 90\\% confidence level for one parameter of interest ($\\Delta\\;C=2.71$; Avni 1976; Cash 1979). The average \\xray\\ spectral properties of these two sources are similar to those of typical radio-quiet quasars. They have an average photon index ($\\Gamma=2.07^{+0.31}_{-0.30}$), consistent with that of typical radio-quiet quasars ($\\Gamma\\approx2$). The average photon index is also consistent with those from their band-ratio analyses. We did not find evidence of strong intrinsic neutral absorption for J1604+4326 and J1612+5118 ($N_{\\rm H}\\;\\lesssim\\;1.58\\times10^{22}$~cm$^{-2}$); the spectral fitting quality was not improved after adding the intrinsic neutral absorption component. \\begin{figure}[t] \\centering \\includegraphics[width=3.0in]{fig10a.eps} \\vspace{0.10cm} \\includegraphics[width=3.0in]{fig10b.eps} \\caption{\\footnotesize{\\hbox{X-ray} spectra of SDSS~J1530+2310 (top panel; binned to a minimum of 10 counts per bin) and SDSS~J1109+3736 (bottom panel; binned to a minimum of 20 counts per bin) fitted with a power-law model with Galactic absorption. The residuals are shown in units of $\\sigma$. The inset of each panel shows contours of the photon index vs. intrinsic neutral hydrogen column density parameter space, at confidence levels of 68\\%, 90\\%, and 99\\%, respectively.} \\label{xspec_fig}} \\end{figure} Two sources (J1109+3736 and J1530+2310) have sufficient \\xray\\ counts for individual spectral analysis. We will discuss the \\xray\\ spectral properties of J1530+2310 here and leave the radio-quiet BL~Lac candidate J1109+3736 for the next subsection. The \\xray\\ spectrum of J1530+2310 was extracted following the same procedure as for J1604+4326 and J1612+5118 (see above). The spectrum of J1530+2310 was first grouped to have at least 10 counts per bin (see Fig.~\\ref{xspec_fig}). We used the same spectral models as those for the joint fitting above. The standard $\\chi^2$ statistic was utilized. The fitting results are also shown in Table~\\ref{xspec_table}. Fig.~\\ref{xspec_fig} shows the \\xray\\ spectrum, the best-fit power-law model with Galactic absorption, and the contour plot of the $\\Gamma-N_{\\rm H}$ parameter space for the spectral model with intrinsic neutral absorption for J1530+2310. The \\xray\\ spectral properties of J1530+2310 are similar to those of J1604+4326 and J1612+5118. The photon index of J1530+2310 from the spectral fitting ($\\Gamma=2.11^{+0.37}_{-0.34}$) is consistent with that from the band-ratio analysis (see Table~\\ref{cts_table}). J1530+2310 also shows no evidence of strong intrinsic neutral absorption ($N_{\\rm H}\\;\\lesssim\\;2.67\\times10^{22}$~cm$^{-2}$). In summary, the \\xray\\ spectral properties of \\xray\\ normal \\hbox{low-redshift} WLQs are consistent with those of typical radio-quiet quasars. Shemmer et~al. (2009) found similar results for their high-redshift, radio-quiet WLQs. Their sources have a somewhat harder average \\xray\\ spectrum ($\\Gamma=1.86^{+0.72}_{-0.48}$), perhaps because they fit both \\xray\\ weak and \\xray\\ normal WLQs jointly. Shemmer et~al. (2009, 2010) also performed individual \\xray\\ spectral analysis on two high-redshift, radio-intermediate WLQs (J1141+0219 and J1231+0138). \\subsection{J1109+3736: The Radio-quiet BL Lac Candidate}\\label{discuss:j1109} J1109+3736 is a radio-quiet BL~Lac candidate based on its multi-band properties. It is strong in \\hbox{X-rays} by a factor of $\\approx6.3$ (\\aox$\\;=\\;-1.10$, \\daox$\\;=\\;0.31$). Its \\aox\\ value is similar to that of the majority of the BL~Lac population (see Fig.~\\ref{aroaox_fig}). \\xray\\ spectral analysis was carried out for J1109+3736 following similar procedures as those in \\S\\ref{discuss:stack}. The \\chandra\\ spectrum of J1109+3736 was grouped to have at least 20 counts per bin (see Fig.~\\ref{xspec_fig}). % The best-fit parameters are listed in Table~\\ref{xspec_table}. The \\xray\\ spectrum, the best-fit power-law model with Galactic absorption, and $\\Gamma-N_{\\rm H}$ contours for J1109+3736 are also shown in Fig.~\\ref{xspec_fig}. The best-fit photon index of J1109+3736 ($\\Gamma=1.77\\pm0.14$) is consistent with that from the band-ratio analysis. This photon index value indicates a harder \\xray\\ spectrum than those for the majority of the high-energy peaked BL~Lac objects (HBL),\\footnote{J1109+3736 would be classified as an HBL if it were indeed a BL~Lac object, based on the criterion $\\alpha_{\\rm rx}\\;<\\;0.75$ of Padovani \\& Giommi (1995), where $\\alpha_{\\rm rx}$ is the radio-to-X-ray spectral index, $\\alpha_{\\rm rx}=-0.130\\;\\log(f_{\\rm 1~keV}/f_{\\rm 5~GHz})$. The $\\alpha_{\\rm rx}$ value of J1109+3736 is estimated to be 0.46.} which have a mean photon index $\\Gamma\\approx2.2$, but it is still consistent with the broad distribution of HBL photon-index values (e.g., see the bottom panel of Fig.~1 in Donato et~al. 2005). Although this source has a high \\xray\\ count rate, we do not expect strong photon pile-up effects because a $1/2$ subarray mode was used for its \\chandra\\ observation. There is no evidence for significant intrinsic absorption in the J1109+3736 spectrum ($N_{\\rm H}\\;\\lesssim\\;1.4\\times10^{21}$~cm$^{-2}$). The SDSS spectrum of J1109+3736 (see the bottom-right panel of Fig.~\\ref{spec_fig}) shows a strong power-law continuum without any detectable emission lines, in particular no Balmer lines. Note that some of the \\xray\\ weak, \\hbox{low-redshift} WLQ candidates (e.g., \\phl; see Leighly et~al. 2007b) have fairly strong Balmer lines. The strength of its Ca~{\\sc ii} H/K break\\footnote{The strength of the Ca~{\\sc ii} H/K break is defined as the fractional change of the average flux densities in continuum regions blueward (3750--3950~\\AA) and redward (4050--4250~\\AA) of the H/K break (see Equation 1 of Plotkin et~al. 2010a).} ($C=0.185$) indicates a relatively small contribution to the SDSS spectrum from the host galaxy. J1109+3736 shows moderate \\xray\\ variability between its \\chandra\\ observation and the epoch of the \\rosat\\ All Sky Survey in 1990 November (RASS; Voges et~al. 1999). It was not detected by RASS. The upper limit for \\aox\\ is estimated to be \\aox$\\;<\\;-1.29$, showing a factor of $>3$ variation compared to its \\chandra\\ observation. The quasar should have been detected in the RASS if it had the same \\xray\\ brightness as that in the \\chandra\\ epoch (expecting $\\approx20$ counts in a 250~s RASS observation). There is also evidence for optical variability of J1109+3736. This source was optically identified during the Cambridge APM (Automated Plate Measuring Machine) scans of POSS (Palomar Observatory Sky Survey) plates (e.g., McMahon et~al. 2002). We converted the APM magnitudes in the $O$ and $E$ bands ($O=20.76,\\;E=19.12$) into a $B$-band magnitude ($B=20.57$) using Equations \\hbox{(1--3)} in McMahon et~al. (2002). We also converted the SDSS magnitudes to a $B$-band magnitude ($B=18.88$) using the equations in Table~1 of Jester et~al. (2005). The $B$-band magnitude difference is 1.69, indicating a factor of $\\sim 5$ variability over a time span of $\\sim50$ years. This source shows greater variability than most quasars during this time span (e.g., see Fig.~24 of Sesar et~al. 2006). Therefore, J1109+3736 is a BL~Lac candidate on the radio-faint tail of the full BL~Lac population. The broad-band SED of J1109+3736 is shown in Fig.~\\ref{sed1_fig}. It is relatively weak in the near infrared compared to its SDSS brightness. However, the SED profile of this source could be strongly affected by its variability (see the comparison between its SDSS and POSS fluxes in Fig.~\\ref{sed1_fig}) given its possible BL~Lac nature. Infrared photometry from {\\it WISE}, which will be available in the {\\it WISE} full data release,\\footnote{J1109+3736 is not covered by the currently available {\\it WISE} preliminary data release. See coverage map at http://wise2.ipac.caltech.edu/docs/release/prelim/.} will be helpful because the infrared-to-UV SEDs of BL~Lac objects are very different from those of typical quasars (e.g., Lane et~al. 2011). Future polarization measurements of J1109+3736 will also be useful in understanding its nature, since BL~Lac objects are usually highly polarized in the optical/UV band. However, polarimetry surveys of optically selected BL~Lac samples (e.g., Smith et~al. 2007; Heidt \\& Nilsson 2011) did not find any highly polarized radio-quiet BL~Lac candidates, indicating this kind of source, if it indeed exists, should make up only a small fraction of the total population of radio-quiet WLQ candidates." }, "1112/1112.3704_arXiv.txt": { "abstract": "We present near-infrared spectroscopic observations of star-forming galaxies at $z\\sim1.4$ with FMOS on the Subaru Telescope. We observed K-band selected galaxies in the SXDS/UDS fields with $K\\leq23.9$ mag, $1.2\\leq z_{ph} \\leq 1.6$, $M_{*}\\geq 10^{9.5} M_{\\odot}$, and expected F(H$\\alpha$) $\\geq$ 10$^{-16}$ erg s$^{-1}$ cm$^{-2}$. 71 objects in the sample have significant detections of H$\\alpha$. For these objects, excluding possible AGNs identified from the BPT diagram, gas-phase metallicities are obtained from [N\\emissiontype{II}]/H$\\alpha$ line ratio. The sample is split into three stellar mass bins, and the spectra are stacked in each stellar mass bin. The mass-metallicity relation obtained at $z\\sim1.4$ is located between those at $z\\sim0.8$ and $z\\sim2.2$. We constrain an intrinsic scatter to be $\\sim0.1$ dex or larger in the mass-metallicity relation at $z\\sim1.4$; the scatter may be larger at higher redshifts. We found trends that the deviation from the mass-metallicity relation depends on the SFR and the half light radius: Galaxies with higher SFR and larger half light radii show lower metallicities at a given stellar mass. One possible scenario for the trends is the infall of pristine gas accreted from IGM or through merger events. Our data points show larger scatter than the fundamental metallicity relation (FMR) at $z\\sim0.1$ and the average metallicities slightly deviate from the FMR. The compilation of the mass-metallicity relations at $z\\sim3$ to $z\\sim0.1$ shows that they evolve smoothly from $z\\sim3$ to $z\\sim0$ without changing the shape so much except for the massive part at $z\\sim0$. ", "introduction": "The gas-phase metallicity (hereafter, metallicity) is a fundamental parameter for the understanding of formation and evolution of galaxies, because it traces past star formation activity; metals in gas are produced in stars formed and returned into the inter-stellar medium (ISM) of galaxies. The metallicity is also affected by inflow and outflow of gas. Thus, by investigating the metallicity and its cosmological evolution, the star formation history of galaxies together with the gas inflow and outflow can be constrained. The presence of a correlation between stellar mass and metallicity (hereafter, mass-metallicity relation) in the local universe is well known. \\citet{Tremonti:2004p4119} established the mass-metallicity relation at $z\\sim0.1$ with a large sample ($\\sim53,000$) of Sloan Digital Sky Survey (SDSS) galaxies. Tracing the cosmological evolution of the mass-metallicity relation is indispensable to reveal how galaxies have been evolving. At $z\\sim1$, the mass-metallicity relations and the downsizing-like evolution to $z\\sim0.1$ were found; the less evolution can be seen in the massive part (e.g., \\cite{Savaglio:2005p3325,Zahid:2011p11939}). However, the anti-downsizing-like evolution of the mass-metallicity relations were also presented at similar redshifts \\citep{Lamareille:2009p5295, PerezMontero:2009p5308}. The mass-metallicity relations were obtained at $z\\sim2$ by using $\\sim9 0$ galaxies \\citep{Erb:2006p4143} and at $z\\sim3$ by $\\sim20$ galaxies \\citep{Maiolino:2008p5212, Mannucci:2009p8028} and the smooth evolution of the mass-metallicity relation from $z\\sim3$ to $z\\sim0$ is suggested. However, at $z\\sim2$, higher metallicities at a given stellar mass have been reported \\citep{Hayashi:2009p4235,Yoshikawa:2010p4286,Onodera:2010p4273} than those found by \\citet{Erb:2006p4143}. The discrepancy at $z\\sim2$ may be partly due to the small size of their samples ($10-20$). Since the redshift of $z=1-2$ is close to the peak epoch in the cosmic star-formation history, establishing the mass-metallicity relation at this redshift is very important. Although the metallicity correlates well with stellar mass, there is a scatter in the relation. Its origin may provide a clue to understand the process of the chemical enrichment and may be related to how the mass-metallicity relation evolves. At $z\\sim0.1$, \\citet{Tremonti:2004p4119} found that the mass-metallicity relation has an intrinsic scatter of $\\sim0.1$ and the scatter is larger at the lower stellar mass. At a given stellar mass, galaxies with higher surface stellar mass density tends to show higher metallicity, suggesting that they transformed more gas into stars raising the metallicity. \\citet{Ellison:2008p7997} showed that galaxies with larger specific star-formation rate (SFR) and size show lower metallicity at a given stellar mass at $z\\sim0.1$. More recently, \\citet{Mannucci:2010p8026} and \\citet{Yates:2011p16030} found that galaxies with larger SFRs tend to show lower metallicities at $z\\sim0.1$ and the scatter around the mass-metallicity relation is reduced very much (i.e., the fundamental metallicity relation) by introducing SFR as the second parameter affecting the metallicity. The scatter of the mass-metallicity relation, however, is still unclear at higher redshifts because of the limited sample size. In order to investigate these trends further, including the dependency of other parameters, at higher redshift, a large near-infrared spectroscopic survey is necessary, which is very time consuming and hard to achieve. The Fiber Multi Object Spectrograph (FMOS) on the Subaru Telescope enables us to survey a large spectroscopic sample at high redshifts. FMOS is a near infrared (NIR) fiber multi-spectrograph \\citep{Kimura:2010p11396}. The fiber positioner ``Echidna'' at the prime focus feeds 400 fibers in a FoV of \\timeform{30'} diameter. Two NIR spectrographs (IRS1 and IRS2) with an OH airglow suppression system are capable of obtaining both low resolution ($R\\sim650$) and high resolution ($R\\sim3000$) spectra in the wavelength range of $0.9-1.8$$\\mu$m. Taking advantage of the multiplicity available with FMOS, we are conducting a large spectroscopic survey for star-forming galaxies at $z=1-2$. This redshift range is of much interest, because it is near or shortly after the peak epoch in the cosmic star formation history (e.g., \\cite{Hopkins:2006p4539}) and thus important for the history of the chemical evolution of galaxies. It is also suitable for FMOS, because in the redshift range [N\\emissiontype{II}]$\\lambda\\lambda$6548,6584, H$\\alpha$, [O\\emissiontype{III}]$\\lambda\\lambda$4959,5007, and H$\\beta$ lines are located in the wavelength range observable with FMOS. In this paper we present the first results from this survey as to metallicity measurements of galaxies at $z\\sim1.4$. Throughout this paper, we adopt the concordance cosmology, ($\\Omega_{M}$ , $\\Omega_{\\Lambda}$ , $h$) = (0.3, 0.7, 0.7). All magnitudes are in the AB system \\citep{Oke:1983p15127}. ", "conclusions": "We present the first results obtained from near-infrared spectroscopic observations of star-forming galaxies at $z\\sim1.4$ with FMOS on the Subaru Telescope. We observed K-band selected galaxies in the SXDS/UDS fields with $K\\leq23.9$ mag, $1.2\\leq z_{ph} \\leq 1.6$, $M_{*}\\geq 10^{9.5} M_{\\odot}$, and expected F(H$\\alpha$)$\\geq10^{-16}$ erg s$^{-1}$ cm$^{-2}$. H$\\alpha$ emission lines of 71 objects are detected significantly. For these objects, excluding possible AGNs identified from the BPT diagram, gas-phase metallicities and upper limits are obtained from [N\\emissiontype{II}]/H$\\alpha$ line ratio by carefully considering the effects of the OH-masks. We separate the sample into three stellar mass bins and stack the spectra. We obtain a mass-metallicity relation at $z\\sim1.4$ from the stacking analysis. The mass-metallicity relation is located between those at $z\\sim0.8$ and $z\\sim2.2$. We tried to constrain an intrinsic scatter of the mass-metallicity relation and found that the scatter is $\\sim0.1$ dex and is comparable to that at $z\\sim0.1$ \\citep{Tremonti:2004p4119}. The scatter increases as stellar mass decreases if we take the intrinsic scatters at face-value. The scatters of the mass-metallicity relation at $z\\sim1.4$, however, should be lower limits; this implies that the scatter may be larger at higher redshifts. We found that the deviation from the mass-metallicity relation depends on the SFR and the half light radius: Galaxies with higher SFR and larger half light radii show lower metallicities at a given stellar mass. One possible scenario for the trends is the infall of pristine gas accreted from IGM or through merger events. Our results do not show clear FMR as proposed by \\citet{Mannucci:2010p8026}. Trends of the dependence of the SFR and the size on the mass-metallicity relation are, however, still not so clear. A larger sample from the further survey with FMOS in the future may be able to reveal clearer trends, not only the dependence of the SFR and the size but also that of other parameters. The compilation of the mass-metallicity relations at $z\\sim3$ to $z\\sim0.1$ shows that they evolve smoothly from $z\\sim3$ to $z\\sim0$ without changing its shape so much (except for the massive part at $z\\sim0.1$). The metallicity at $M_{*}=10^{10}M_{\\odot}$ on the mass-metallicity relation increases with decreasing redshift and can be described as $12+\\textrm{log(O/H)}=8.69 - 0.086 (1+z)^{1.3}$. The metallicity evolution rate was the highest at the cosmic age of $\\lesssim$ 3 Gyr which was before redshift of 2. However, the result may be influenced by sample selection and further studies with large samples at high redshifts are desirable. \\bigskip We would like to thank an anonymous referee for useful comments. We are grateful to the FMOS support astronomer Kentaro Aoki for his support during the observations. We also appreciate Soh Ikarashi, Kotaro Kohno, Kenta Matsuoka, and Tohru Nagao sharing fibers in their FMOS observations. KY is financially supported by a Research Fellowship of the Japan Society for the Promotion of Science for Young Scientists. KO's activity is supported by the grant-in-aid for Scientific Research on Priority Areas (19047003). We acknowledge support for the FMOS instrument development from the UK Science and Technology Facilities Council (STFC). DB and ECL acknowledge support from STFC studentships. We would like to express our acknowledgment to the indigenous Hawaiian people for their understanding of the significant role of the summit of Mauna Kea in astronomical research." }, "1112/1112.1701_arXiv.txt": { "abstract": "We present three new eclipsing white-dwarf / M-dwarf binary systems discovered during a search for transiting planets around M-dwarfs. Unlike most known eclipsing systems of this type, the optical and infrared emission is dominated by the M-dwarf components, and the systems have optical colors and discovery light curves consistent with being Jupiter-radius transiting planets around early M-dwarfs. We detail the PTF/M-dwarf transiting planet survey, part of the Palomar Transient Factory (PTF). We present a Graphics Processing Unit (GPU)-based box-least-squares search for transits that runs approximately 8$\\times$ faster than similar algorithms implemented on general purpose systems. For the discovered systems, we decompose low-resolution spectra of the systems into white-dwarf and M-dwarf components, and use radial velocity measurements and cooling models to estimate masses and radii for the white dwarfs. The systems are compact, with periods between 0.35 and 0.45 days and semimajor axes of approximately $\\rm2 R_{\\odot}$ (0.01 AU). The M-dwarfs have masses of approximately 0.35$\\rm{M_{\\odot}}$, and the white dwarfs are all hydrogen-atmosphere with temperatures of around 8000K, and have masses of approximately 0.5$\\rm{M_{\\odot}}$. We use the Robo-AO laser guide star adaptive optics system to tentatively identify one of the objects as a triple system. We also use high-cadence photometry to put an upper limit on the white dwarf radius of 0.025$\\rm{R_{\\odot}}$ (95\\% confidence) in one of the systems. Accounting for our detection efficiency and geometric factors, we estimate that $\\rm 0.08\\%^{+0.10\\%}_{-0.05\\%}$ (90\\% confidence) of M-dwarfs are in these short-period, post-common-envelope white-dwarf / M-dwarf binaries where the optical light is dominated by the M-dwarf. Similar eclipsing binary systems can have arbitrarily small eclipse depths in red bands and generate plausible small-planet-transit light curves. As such, these systems are a source of false positives for M-dwarf transiting planet searches. We present several ways to rapidly distinguish these binaries from transiting planet systems. ", "introduction": "Large numbers of non-eclipsing white-dwarf / main-sequence binaries have been discovered in the Sloan Digital Sky Survey and other surveys (e.g. \\citealt{Rebassa2011}; \\citealt{Bianchi2007} and references therein). For low-mass stars in particular there is a bridge in color between white dwarfs and M-dwarfs. The bridge is interpreted as being due to rare white-dwarf / M-dwarf binaries, at a ratio with respect to single stars of $\\sim$1:2300 \\citep{Smolcic2004}. White-dwarf / M-dwarf \\textit{eclipsing} systems are much rarer, and almost all have been discovered by searching for white dwarfs displaying very deep eclipses of up to several magnitudes (e.g. \\citealt{Drake2010}). These searches find systems containing relatively hot ($>12000$K) white dwarfs and mid-to-late M-dwarfs. The discovery rate of these systems (e.g. \\citealt{Drake2010, Parsons2011, Parsons2011b}) is increasing with the advent of large sky surveys. These binaries survived the common-envelope phase of their evolution and many will become cataclysmic variables (e.g. \\citealt{Nebot2011}), and so the properties and number statistics of these systems can provide windows into two important areas of stellar evolution. Precision measurements of the systems allow the determination of the masses and radii of two types of stars for which there are relatively few measurements \\citep{Nebot2009, Kraus2011, Pyrzas2011, Pyrzas2009}. In this paper we present three eclipsing white-dwarf/M-dwarf systems discovered during the PTF/M-dwarfs search for transiting planets around M-dwarfs. In contrast to most known eclipsing systems of this type, the systems detected in this survey have optical and infrared emission dominated by the M-dwarf component and contain relatively low-temperature (8000K) white dwarfs and relatively early M-dwarfs. The shape of the light curves of the detected systems is similar to that expected for transiting giant planets around M-dwarfs, in particular in having a flat-bottomed eclipse with a depth of 1-20\\% in red optical bands. The PTF/M-dwarfs survey \\citep{Law2011} is a search for transiting planets around 100,000 M-dwarfs. The survey is performed with the Palomar Transient Factory (PTF) camera \\citep{Rahmer2008, Law2009, Law2010} on the 48-inch Samuel Oschin telescope at Palomar Observatory, and is a Key Project of the Palomar Transient Factory \\citep{Law2009, Rau2009}. The PTF/M-dwarfs survey is designed to complement other M-dwarf transiting planet surveys such as MEarth (e.g. \\citealt{Charbonneau2009, Irwin2010}), the WFCam transit survey \\citep{Sipocz2011} and the \\mbox{M-dwarfs} in the Kepler mission target list \\citep{Borucki2011}, by covering a much larger number of M-dwarfs at somewhat lower sensitivity. The survey achieves photometric precisions of a few percent for $\\approx$100,000 targets, and few-millimag precision around a subset of $\\approx$10,000 M-dwarfs. These systems offer much larger transit depths compared to solar-type stars, while their very red colors compared to most other stars in the field greatly reduce the probability of a blended eclipse producing a difficult-to-detect transit false positive. \\begin{figure*} \\centering \\resizebox{1.0\\textwidth}{!} { \\includegraphics{field.eps} } \\caption{A PTF Camera image of a survey field centered at 17:28, +57:22 and covering $3.50^{\\circ}\\times2.31^{\\circ}$. The highlighted points show the 2,851 stars with photometrically-estimated spectral types later than K4 and photometric stability better than 5\\%. The colors of the points correspond to the stellar temperatures, with K4 as yellow and late-M-dwarfs as red. The pipeline has removed stars with possible photometric precision problems such as proximity to a bright star or a bright ghost image. North is up and East is to the right.} \\label{fig:field} \\end{figure*} The three eclipsing systems presented here were originally detected as Jupiter-sized planet candidates during the first year of operations of the survey. Follow up of the candidates showed large color changes during eclipse and very large radial velocity signals, suggesting a hidden hot companion. In this paper we detail the properties of these eclipsing M-dwarf - white-dwarf systems and explore ways to distinguish them from true planetary transits. The paper is organized as follows: in section \\ref{sec:survey} we describe the PTF/M-dwarfs survey, its precision photometry methods, and its target detection strategies, including a new method of performing a box-least-squares transit search in parallel on GPU hardware. In section \\ref{sec:disc_and_followup} we detail the three new eclipsing white-dwarf - M-dwarf systems and describe follow up photometric, low-resolution spectroscopic, and radial velocity observations, which are used to determine masses and radii for the system components in section \\ref{sec:models}. Section \\ref{sec:discussion} determines the frequency of eclipsing binaries such as these and discusses ways to distinguish them from transiting planets. ", "conclusions": "\\label{sec:discussion} The discovery of only three white-dwarf / M-dwarf binaries in the PTF/M-dwarfs survey suggests a low incidence for such systems. This phase of the survey has covered $\\approx$45,000 M-dwarfs with sufficient precision to detect short-period systems with similar eclipse depths to those shown here (figure \\ref{fig:n_targets}). The survey's detection efficiency for systems with fractional-day periods is near 100\\% (figure \\ref{fig:detec_effic}). The geometric probability of eclipse is 5-15\\% for an early-M-dwarf primary in a roughly equal-mass half-day-period binary with a white dwarf. Taken together, the detection of three systems implies that $\\rm 0.08\\%^{+0.10\\%}_{-0.05\\%}$ (90\\% confidence) of M-dwarfs are in short-period post-common-envelope white-dwarf / M-dwarf binaries where the optical light is dominated by the M-dwarf. \\begin{figure} \\centering \\resizebox{1.0\\columnwidth}{!} { \\includegraphics{n_targets.eps} } \\caption{The cumulative number of M-dwarf targets searched by the PTF/M-dwarfs survey in the first year of operations, as a function of achieved photometric stability. M-dwarfs in the PTF/M-dwarf fields are selected and confirmed on the basis of 2MASS, USNO-B1 and (where available) SDSS colors, along with proper motions determined from 2MASS and USNO-B1 positions. We require a high-confidence photometric color fit, along with a 2$\\sigma$ proper motion detection to rule out giant interlopers.} \\label{fig:n_targets} \\end{figure} \\begin{figure} \\centering \\resizebox{1.0\\columnwidth}{!} { \\includegraphics{period_detect.eps} } \\caption{The eclipse detection efficiency for a typical PTF/M-dwarfs observing scenario, based on simulations including the actual observing cadences of the survey. The detection efficiency is the probability that at least two eclipses will be observed from an eclipsing target with a specified period, after observing window and weather effects have been taken into account. The low efficiencies around integer numbers of days at short periods are due to the likelihood of an eclipse always occurring in daytime; the same effect allows some long period binaries to be picked up if their eclipses always occur at night.} \\label{fig:detec_effic} \\end{figure} The periods of the detected systems are all longer than most known eclipsing M-dwarf / white-dwarf systems (figure \\ref{fig:period_dist}). Our survey is designed to detect longer-period transiting exoplanets, so this is not in itself surprising. However, our lack of detections at shorter periods, despite our near-100\\% detection efficiency for such systems, suggests that binaries including these relatively low-temperature white dwarfs are preferentially found at relatively large orbital radii. \\begin{figure} \\centering \\resizebox{1.0\\columnwidth}{!} { \\includegraphics{period_hist.eps} } \\caption{The periods of our three detected systems compared to the distribution of periods of the white-dwarf dominated eclipsing systems found by \\citet{Drake2010}, which were all found in a single survey and so offer a useful comparison sample.} \\label{fig:period_dist} \\end{figure} These eclipsing binaries appear as M-dwarfs when selected by optical and infrared colors, and they display transit-like light curves which can have arbitrarily small depths. They appear to be at least three times more common than transiting giant planets in the PTF/M-dwarfs survey. These properties make these systems an important false-positive for current and future transiting planet searches around M-dwarfs. Fortunately, they can be distinguished from true transiting planets with small amounts of extra data, using the methods detailed below. The methods are ordered by difficulty, starting with those which require only the discovery light curves. \\begin{enumerate} \\item{Eclipse duration. The increased mass of the system compared to a M-dwarf/planet system leads to a shorter eclipse; high-SNR and high-cadence light curves are however required to distinguish this case from a lower-inclination system.} \\item{Eclipse shape. The eclipse of the white dwarf by the M-dwarf provides an exactly-flat-bottomed eclipse outside the short ingress and egress periods. With high-precision photometry this can be distinguished from a true transiting planet which will show the effects of limb darkening as it passes in front of the M-dwarf.} \\item{Out-of-eclipse variability. All our systems show out-of-eclipse variability at levels which are orders of magnitude greater than that expected for planetary systems (see \\citet{Drake2003} for details of a similar selection method for planetary transits around solar-type stars).} \\item{UV emission. Two of our new systems showed UV emission and the third has possible emission. Where data is available, the presence of UV (or even u-band) emission suggests the existence of something other than an M-dwarf in the system.} \\item{Low-resolution spectroscopy. Like the three targets presented in this paper, a low-resolution spectrum could be decomposed into white dwarf and M-dwarf components. Faint, low-temperature white dwarfs may be, however, hard to detect with this technique.} \\item{Multi-color photometry. These systems show a strong variation in eclipse depth with wavelength (depending on the relative temperatures of the white dwarf and M-dwarf and their ratio of radii).} \\item{Radial velocities. These systems have radial velocity amplitudes at least 500$\\times$ larger than expected for a planet with the same period. Just two RV few-km/sec-precision datapoints are sufficient to discriminate these systems from transiting planet systems.} \\end{enumerate} Of these methods, mulitcolor photometry through eclipse is probably the most time-efficient method of detecting systems like this. As large scale transiting planet surveys of M-dwarfs start up and continue, many more systems in this interesting parameter space are likely to be found. Follow-up precision photometry and radial velocities will allow direct measurements of the masses and radii of all components of these systems. As the white dwarf transits across the disk of the M-dwarf the transit depth is expected to be around 1 millimag, although lensing by the white dwarf (e.g., \\citealt{Marsh2001}) will make the transit shallower than expected from only geometric considerations (e.g., \\citealt{Steinfadt2010}). High-cadence and high-precision photometry of the white dwarf eclipse (occultation) ingress and egress may be the best approach to directly measure the white dwarf radii. Furthermore, small asymmetries in the ingress and egress light curve due to the photometric Rossiter-McLaughlin effect \\citep{Shporer2012, Groot2011} can allow a measurement of the white dwarf spin-orbit alignment and rotation velocity. With these methods, this new group of systems will fill a poorly-covered range of the white-dwarf and M-dwarf mass/radius relations." }, "1112/1112.6121_arXiv.txt": { "abstract": "The application of high end computing to astrophysical problems, mainly in the galactic environment, is developing for many years at the Dep. of Physics of Sapienza Univ. of Roma. The main scientific topic is the physics of self gravitating systems, whose specific subtopics are: i) celestial mechanics and interplanetary probe transfers in the solar system; ii) dynamics of globular clusters and of globular cluster systems in their parent galaxies; iii) nuclear clusters formation and evolution; iv) massive black hole formation and evolution; v) young star cluster early evolution. In this poster we describe the software and hardware computational resources available in our group and how we are developing both software and hardware to reach the scientific aims above itemized. ", "introduction": "Celestial mechanics is one of the most classic examples of chaos in physics: the mutual gravitating systems show a chaotic behaviour, being extremely sensitive to differences in initial conditions. This problem can be only partially controlled using high-order integration algorithms. The intrinsic difficulty of the problem is summarized by the so called double divergence of the pair interaction potential $U_{ij} \\propto 1/r_{ij}$, where $r_{ij}$ is the distance between particle $i$ and particle $j$. The ``ultraviolet'' divergence corresponds to gravitational encounters at vanishing distance, while the ``infrared'' divergence means that the force never vanishes. The computational problems arising from these divergences make the classic gravitational \\emph{N}-body problem unique. ", "conclusions": "" }, "1112/1112.2064_arXiv.txt": { "abstract": "Current observations suggest that the universe was reionized sometime before $z\\sim 6$. One way to observe this epoch of the universe is through the Near Infrared Background (NIRB), which contains information about galaxies which may be too faint to be observed individually. We calculate the angular power spectrum ($C_l$) of the NIRB fluctuations caused by the distribution of these galaxies. Assuming a complete subtraction of any post-reionization component, $C_l$ will be dominated by galaxies responsible for completing reionization (e.g., $z\\sim 6$). The shape of $C_l$ at high $l$ is sensitive to the amount of {\\it{non-linear}} bias of dark matter halos hosting galaxies. As the non-linear bias depends on the mass of these halos, we can use the shape of $C_l$ to infer typical masses of dark matter halos responsible for completing reionization. We extend our previous study by using a higher-resolution $N$-body simulation, which can resolve halos down to $10^8~M_\\sun$. We also include improved radiative transfer, which allows for the suppression of star formation in small-mass halos due to photo-ionization heating. As the non-linear bias enhances the dark-matter-halo power spectrum on small scales, we find that $C_l$ is steeper for the case with a complete suppression of small sources or partial suppression of star formation in small halos (the minimum galaxy mass is $M_{\\rm min}=10^9~M_\\sun$ in ionized regions and $M_{\\rm min}=10^8~M_\\sun$ in neutral regions) than the case in which these small halos were unsuppressed. In all cases, we do not see a turn-over toward high $l$ in the shape of $l^2 C_l$. ", "introduction": "\\label{sec:introduction} Probing the beginnings of star and galaxy formation in the early universe is one of the goals of modern cosmology. These high redshift stars provided a wealth of ionizing photons, which caused all of the hydrogen in the universe to be ionized, a process called reionization. Therefore, understanding high-redshift star formation is closely coupled with our understanding of reionization history. Recent measurements of the polarization of the Cosmic Microwave Background have shown that reionization started early and was extended in time, with an equivalent instantaneous reionization redshift of $z \\sim 11$ \\citep{kogut/etal:2003,spergel/etal:2003,spergel/etal:2007, page/etal:2007,dunkley/etal:2008,komatsu/etal:2008}. Therefore, we expect there to be significant star formation before this. With improving observations, there are many ways to be able to observe these first few generations of stars and galaxies. We can now observe high-redshift ($z>6$) star formation via galactic surveys, which provide statistical properties of high-redshift sources. These surveys provide a wealth of information on early star formation; however, they only probe the ``tip of the iceberg,'' e.g., a small fraction of the whole population, mainly those galaxies that are bright enough to be identified in surveys. {\\footnote{Those exceptionally bright galaxies also might be missed in surveys, because they are rare enough that they will not be found within the survey's limited area \\citep{trenti/stiavelli:2008}}}. Another way to observe these high-redshift galaxies is to look for the {\\it diffuse} background light originating from them. Ultra-violet photons produced by these early galaxies at $z\\gtrsim 6$ would be redshifted into the near infrared band ($\\lambda\\gtrsim 1~\\mu$m). Therefore, any background above and beyond that from the low-redshift galaxies could be attributable to these early galaxies. Observing the Near Infrared Background (NIRB) has the benefit of probing a population of galaxies {\\it{as a whole}}, not just the unusually bright objects which can be detected individually. For this purpose, one can observe the mean NIRB \\citep{santos/bromm/kamionkowski:2002, magliocchetti/salvaterra/ferrara:2003,salvaterra/ferrara:2003, cooray/yoshida:2004,madau/silk:2005, fernandez/komatsu:2006} as well as fluctuations \\citep{kashlinsky/odenwald:2000,kashlinsky/etal:2002, kashlinsky/etal:2004,kashlinsky/etal:2005,kash/etal:2007,kashlinsky/etal:2007, kashlinskyb/etal:2007c,kashlinsky:2005,magliocchetti/salvaterra/ferrara:2003, odenwald/etal:2003,cooray/etal:2004,matsumoto/etal:2005,thompson/etal:2007a, thompson/etal:2007b, fernandez/etal:2010}. In our previous work \\citep{fernandez/etal:2010}, we calculated the angular power spectrum of the NIRB fluctuations using an $N$-body simulation to trace the large-scale structure and the formation of galactic halos, % coupled with radiative transfer simulations of reionization. Combining this numerical data with the latest analytic modeling of the internal, unresolved properties of radiation processes such as stellar emission, Lyman-$\\alpha$ line, two-photon emission, free-free and free-bound emission from the formed galaxies \\citep{fernandez/komatsu:2006}, we analyzed the effects of changing the star formation efficiency $f_*$, the escape fraction $f_{\\rm esc}$\\footnote{Here, by the ``escape fraction,'' we refer to a fraction of hydrogen-ionizing photons escaping from the galaxies into the intergalactic medium. We ignore dust extinction.}, and the mass and metallicity of the stars on the NIRB angular power spectrum. Our previous simulation was able to resolve dark matter halos down to the minimum mass of $M_{\\rm min}=2.2\\times 10^9~M_\\sun$. In the current paper, we extend this analysis by improving the mass resolution by more than an order of magnitude to $M_{\\rm min}=10^8~M_\\sun$. We also include a new physical effect: the suppression of star formation in small-mass halos due to photo-ionization heating (the so-called ``Jeans-mass filtering'', which prevents baryons from collapsing into small dark matter halos \\citep[e.g.,][]{shapiro/etal:1994,2007MNRAS.376..534I}). The structure of this paper is as follows. The simulations used for this analysis are described in \\S~\\ref{sec:simulation}. The fiducial stellar model used is discussed in \\S~\\ref{sec:stellarpop}, and our results are presented in \\S~\\ref{sec:results}. This model is then expanded upon in \\S~\\ref{sec:pops}. We compare our results to observations in \\S~\\ref{sec:obs}, and conclude in \\S~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} The NIRB mean and fluctuation signals can be a very powerful probe of the high-redshift star formation ($z\\gtrsim 6$). Building upon our previous work on this subject \\citep{fernandez/komatsu:2006,fernandez/etal:2010}, we have extended the calculation of the angular power spectrum of the NIRB fluctuation by improving the mass resolution of $N$-body simulations by an order of magnitude. With the simulation now resolving the dark matter halo mass down to $M=10^8~M_\\sun$, we have confirmed our previous finding: due to the non-linear halo biasing, the shape of $l^2C_l$ of the NIRB fluctuation does not exhibit a turn over. However, we do observe a flattening of the shape as we lower the minimum halo mass (because the non-linear bias decreases as masses go down), and thus the shape of $l^2C_l$ can be used to infer typical masses of dark matter halos hosting galaxies in a high redshift universe. We have gone beyond simply increasing the mass resolution of the simulation. For the first time, we consider cases with a radiative feedback suppressing the star formation in low-mass galaxies with $M<10^9~M_\\sun$, due to the Jeans-mass filtering in the ionized and heated IGM. We find that, when low-mass sources are partially suppressed by photo-ionization heating, the predicted angular power spectrum becomes similar to the one with a complete suppression of halos below $M_{\\rm min}=10^9~M_\\sun$, yielding a steep power spectrum at high $l$. Therefore, the shape of the angular power spectrum at high $l$ can directly provide information on the typical mass of sources responsible for completing reionization. The amplitude of the angular power spectrum is less robust, as it depends on a number of parameters such as the stellar initial mass spectrum, metallicity of stars, star formation efficiency, and escape fraction. However, one robust feature is that it is largely determined by the properties of the halo populations at late stages of reionization. Therefore, the angular power spectrum is higher if the stars produce more ionizing photons that do not escape from the halo. (Therefore, maximizing $f_*$ and $N_i$ while minimizing $f_{\\rm esc}$ within the limitations of reionization.) Finally, we find that our predictions, all of which are tuned to satisfy the reionization constraints, are below the current measurements. Given that these measurements should probably be taken as upper limits on the contributions from galaxies in $z\\gtrsim 6$, we conclude that our calculations are consistent with the current measurements." }, "1112/1112.5191_arXiv.txt": { "abstract": "We present \\textsc{NBursts+phot}, a novel technique for the parametric inversion of spectrophotometric data for unresolved stellar populations where high-resolution spectra and broadband SEDs are fitted simultaneously helping to break the degeneracies between parameters of multi-component stellar population models. ", "introduction": "Optical spectra of galaxies contain important information about their internal kinematics and stellar content that can be extracted by fitting them against stellar population models. There is a number of approaches for the full spectrum fitting that were proved to be efficient in studying different classes of galaxies and star clusters. On the other hand, the broad-band photometry in different spectral domains that became available for large samples of galaxies thanks to modern wide-field surveys allows one to study certain properties that cannot be derived from the spectra (e.g. internal extinction) and to brake degeneracies between parameters in case of complex star formation histories (SFH). Usually, photometric measurements and spectra are used independently. Here we propose a new approach \\textsc{NBursts+phot} that fits in a single minimization loop spectral and photometric information and recovers both, parametric SFH, and internal kinematics of galaxies. ", "conclusions": "" }, "1112/1112.5158_arXiv.txt": { "abstract": "We consider the possibility that the cosmological dark matter consists of particles very close in mass to new colored particles below the TeV scale. While such a scenario is inherently difficult to directly confirm at colliders, we find that indirect dark matter searches may be a powerful alternative. In particular, we show that in this case dark matter annihilation to $\\bar q q g$ final states can give rise to significant antiproton (but also gamma-ray) fluxes, and compare the resulting constraints to bounds from direct searches at LEP, the Tevatron and the LHC. For supersymmetric neutralinos degenerate with squarks, \\fex, antiprotons can give rise to more stringent constraints for masses below around 45--80~GeV. ", "introduction": "Collider experiments have pushed the scale for possible new physics beyond the standard model (BSM) to ever higher energies in recent years. The CERN LHC, in particular, is now running at a center of mass energy of $7\\,$TeV and the non-observation of any clear BSM signal, so far, has allowed to place impressive limits~\\cite{Aad:2011qa,Chatrchyan:2011zy}; after an integrated luminosity of $1.04$ fb$^{-1}$, scalar quarks (squarks) in supersymmetric extensions of the standard model, \\fex, have already been excluded for masses below $\\sim 900$ GeV~\\cite{Aad:2011ib,cmssimp}. One has to keep in mind, however, that these results rest on the assumption that the new colored states quickly decay into the lightest neutralino, which is assumed to be massless for the sake of the analysis, and emit high-energy QCD jets in this process. These limits thus do not apply if the only accessible lighter states are very close in mass~\\cite{ATLAS:limit,Chatrchyan:2011ek,LeCompte:2011cn,LeCompte:2011fh}, in which case the resulting jets and missing transverse energy would be too soft to pass the signal selection criteria. Scenarios with degenerate particle spectra are thus generically very difficult to probe at hadron colliders; the situation is considerably better for electron-positron colliders like LEP, but even in this case the limits from direct searches for colored states do not apply to highly degenerate spectra~\\cite{Heister:2002hp, Abbiendi:2002mp, Achard:2003ge, Abdallah:2003xe, LEP_SUSYWG}. Here, we investigate whether such a situation could be probed by indirect dark matter (DM) searches, assuming that the lightest BSM state is a weakly interacting massive particle (WIMP) that makes up the cosmological DM. The self-annihilation of WIMP pairs could then leave an imprint in the spectrum of cosmic rays (see, \\fex, Ref.~\\cite{Bertone:2004pz} for a review on particle DM and indirect searches). As we will demonstrate, WIMP annihilation into $\\bar qq g$ final states, a channel so far hardly explored in indirect DM searches, may give rise to the most stringent constraints in degenerate scenarios (mostly through antiproton, but also through gamma-ray production). In fact, this channel may help to fill remaining loop-holes for the existence of new colored states below masses of around 100 GeV that are left from direct collider searches for such particles. The rest of this Letter is organized as follows: In Section 2 we discuss the importance and main characteristics of DM annihilation into $\\bar{q}qg$ final states, focussing on the case of neutralino DM; in Section 3 we present the energy spectra of antiprotons that result from $\\bar{q}qg$ hadronization and discuss antiproton propagation through our Galaxy; Section 4 contains the limits on the annihilation cross-section that can be derived from cosmic-ray observations and Section 5 confronts these limits with direct collider searches for charged colored particles. Finally, in Section 6 and 7, respectively, we present a discussion of our results and conclude. ", "conclusions": "In this Letter, we have described a DM annihilation channel, and discussed its phenomenological implications in some detail, that so far has hardly been explored in the context of indirect dark matter searches. If DM is a Majorana fermion $\\chi$, or a scalar, direct annihilation via $\\chi\\chi\\to\\bar{q}q$ is strongly helicity suppressed. In this case, the overall annihilation rate \\emph{today} can be dominated by internal gluon Bremsstrahlung, $\\chi\\chi\\to\\bar{q}qg$, and can be boosted by several orders of magnitude with respect to the tree-level result; while this enhancement is much smaller in the early universe, during \\emph{freeze-out}, it can also have a non-negligible effect on the relic density~\\cite{Flores:1989ru,hep-ph/9306325,hep-ph/0608215}. The precise branching ratios depend on the difference between the DM particle's mass $m_\\chi$ and the squark mass $m_{\\tilde{q}}$; the process $\\chi\\chi\\to\\bar{q}qg$ becomes strongest in the mass degenerate case, i.e.~for $m_{\\tilde{q}}/m_\\chi\\simeq1$. We studied the spectra of cosmic rays generated by this channel, and found that the energy spectrum of antiprotons coming from $\\bar{q}qg$ is generically enhanced with respect to the spectrum from $\\bar{q}q$ final states (Fig.~\\ref{fig:dNdx}). The corresponding limits on the annihilation cross-section for $\\chi\\chi\\to\\bar{q}qg$ are hence somewhat stronger than limits on the traditional $\\bar{b}b$ or $W^+W^-$ final states (Fig.~\\ref{fig:svlim}). From this figure we can see, as noted before~\\cite{Evoli:2011id,Bottino:2005xy,Bringmann:2009ca,Lavalle:2010yw}, that antiprotons are most constraining for light DM masses, which in particular seems to disfavor the simplest DM interpretations of recent results from direct DM detection experiments~\\cite{Bernabei:2010mq,Aalseth:2011wp,Angloher:2011uu}. Even more interestingly, the region of small mass splittings between squarks and the LSP (where the annihilation fluxes are largest) is precisely the region that is generically difficult to probe directly with collider searches for squarks; this is because the energy stored in hadronic jets coming from the decay $\\tilde{q}\\to\\chi q$ becomes too small to pass the trigger selection in this limit. In this sense, indirect DM searches are complementary to collider searches and constrain a part of the parameter space that is not directly accessible by either LEP, Tevatron or LHC (Fig.~\\ref{fig:muconstraints}).\\footnote{ Note that also in the situation where the scalar tops are heavy, which generally helps to lift the lightest neutral Higgs boson mass to the high value of $\\sim125\\,$GeV consistent with recent indications found by both ATLAS and CMS, our results from indirect DM searches would not change and constitute important independent limits on light first or second generation squarks.} Depending on the cosmic-ray propagation model, these limits can exclude Bino masses up to 45--80 GeV in the degenerate case. Ongoing AMS-02 measurements are expected to strengthen the limits on the annihilation cross-section by up to an order of magnitude (see \\fex~Ref.~\\cite{Evoli:2011id}), which will make it possible to probe Bino masses of 100 GeV and beyond. \\medskip \\paragraph{Note added} During the final stages of this work, we became aware of another project analyzing the annihilation of DM into $\\bar q q g$ final states~\\cite{TAS}. \\paragraph" }, "1112/1112.1955.txt": { "abstract": "We present new radial velocity measurements from the Bulge Radial Velocity Assay ({\\sl BRAVA}), a large scale spectroscopic survey of M-type giants in the Galactic bulge/bar region. The sample of $\\sim$4500 new radial velocities, mostly in the region $-10^\\circ < l < +10^\\circ$ and $b\\approx -6^{\\circ}$ more than doubles the existent published data set. Our new data extend our rotation curve and velocity dispersion profile to $+20^\\circ$, which is $\\sim$2.8 kpc from the Galactic Center. The new data confirm the cylindrical rotation observed at $-6^\\circ$ and $-8^\\circ$, and are an excellent fit to the Shen et al. (2010) N-body bar model. We measure the strength of the TiO$\\varepsilon$ molecular band as a first step towards a metallicity ranking of the stellar sample, from which we confirm the presence of a vertical abundance gradient. Our survey finds no strong evidence of previously unknown kinematic streams. We also publish our complete catalog of radial velocities, photometry, TiO band strengths, and spectra, which is available at the IRSA archive: {\\tt http://irsa.ipac.caltech.edu/} as well as at UCLA: {\\tt http://brava.astro.ucla.edu/}. ", "introduction": "Only for stars in the Milky Way is it currently possible to examine both the three dimensional kinematics and composition of a central bulge/bar population, offering a unique laboratory for the study of galaxy formation and evolution. Up until now, there have been very few optical spectra and radial velocities of bulge stars, published in a catalog form, with the spectra and measurements made publicly available. Here we present a catalog of the {\\it Bulge Radial Velocity Assay} (BRAVA) that has a sample of $\\sim 10,000$ M giant stars selected from the red giant branch (RGB) of the Two micron All-sky Survey \\citep[2MASS;][]{skrutskie06}. This catalog gives the 2MASS magnitudes and positions, our measured radial velocities, and our spectra. This publication presents our complete sample of low resolution spectra for the {\\sl BRAVA} survey, which covers approximately $-10^\\circ < l <+10^\\circ $ and $-4^\\circ $4 M$_\\odot$, based on the lithium excess (Tatarnikova et al., 2003a,b; Paper I; Munari et al. 2011). Only T CrB and RS Oph, both SyRNe, show significantly enhanced $^7$Li (Wallerstein et al. 2008), reinforcing the identification of V407 Cyg as a member of this class notwithstanding that the 2010 outburst is the first {\\it known} such event in this system. It is the only SyRN system to show maser emission (Deguchi et al. 2011), the individual components of which varied during the outburst. But perhaps most unusual was the event that led to its being labeled ``nova-like'' in the literature, a symbiotic nova-like event in 1936 that lasted for about five years during which time the visual brightness of the system increased by more than a factor of 10, a behavior that is {\\it not} seen historically among the other SyRNe (in addition to T CrB and RS Oph, neither V745 Sco nor V3890 Sgr appear to have such variability). There were no contemporaneous spectroscopic observations during the event of the 1930s but, if it were like other systems, there would have been continuous mass loss through a wind from the white dwarf induced by the onset of a thermonuclear envelope process following a low level of accretion.\\footnote{The historical record is quite poorly sampled before 1990 and we know very little of what V407 Cyg was doing between the outburst of the 1930s and the beginning of photometric and spectroscopic monitoring in the 1990s. The unanticipated explosion of V407 Cyg means it was not included in the photographic survey by Schaefer et al. (2009) of Galactic recurrent novae.} The nova explosion produced a rich variety of phenomena, as we have detailed, that distinguish this outburst from other SyRNe. The first indication of [O III] 4959, 5007\\AA\\ was on JD 55272 (Day 6) ; [N II] 6548, 6583\\AA\\ only appeared after JD 55289 (Day 24). The narrow [O I] 6300, 6363\\AA\\ and Mg I 4671\\AA\\ emission lines were, in contrast, clearly visible from Day 4. In the Day 4 spectrum, narrow emission lines from permitted Fe-peak ions were seen without the underlying broad wings; those appeared by Day 6 and had become asymmetric in opposite senses (as we discussed above) by Day 13. The first appearance of the double peak emission on [O III] 4959\\AA\\ can be estimated from the comparison of the lower resolution Ond\\v{r}ejov spectrum on JD 55292 and the first NOT spectrum on JD 55288. This gives a limit to the radius of the ionized region of $\\leq$2$\\times$10$^{16}$ cm. Assuming that this is matter ejected during the outburst in the 1930s, the width is $\\Delta R/R \\leq 0.15$ based on the duration of the event (about 5 yr) and using the electron density from the [N II] analysis, $n_e \\approx 3\\times$ 10$^4$ cm$^{-3}$, the ionized mass sampled by the nova line profile is $\\leq$10$^{-4}$ M$_\\odot$. If this is due to a wind during a symbiotic nova-like event, the inferred mass loss rate was $\\leq 2\\times$ 10$^{-5}$ M$_\\odot$yr$^{-1}$ independent of the distance. The density is lower than, but roughly compatible with, the electron density inferred from the recombination analysis of the Na I D lines at this velocity discussed in Paper I, about 10$^5$ cm$^{-3}$. The disappearance of the blue component can be interpreted as recombination in a zone located on the side of the giant toward the observer and shadowed from the WD. The redward extension would be, in this picture, the low density periphery that is still illuminated by the WD and also collisionally excited by the shock. The similar maximum positive radial velocity on all the forbidden transitions and the absence of the permitted lines, point to the emission arising in a very low density region. The distinct taxonomic separation of the line profile asymmetry between permitted and forbidden transitions supports the picture that the blue wing of the E2 and M1 lines is suppressed by higher densities being encountered in the approaching part of the shock. and the disappearance of this component on almost all of the strong lines in the last spectrum being due to recombination. The gap between the 2010 and 2011 observing sessions is, however, too long to constrain the densities. \\begin{figure} \\centering \\includegraphics[width=8cm]{p2f34.ps} %%%\\includegraphics{empty.eps} %%%\\includegraphics{empty.eps} \\caption{An example of the separation of the various contributors to the observed line profiles in the 2011 Aug. 21 NOT spectrum. Top: [O I] 6300\\AA. Bottom: [O III] 5007\\AA\\ (solid), Mg I 4571\\AA\\ (dot), [S II] 6716\\AA\\ (dash), [N II] 6548\\AA\\ (dot-dash). The Mira continuum has been removed in each but the scaling is for display (arbitrary flux units). The spike at +60 km s$^{-1}$ is a cosmic ray hit near Mg I. See text for description.} \\end{figure} The late-epoch spectrum from JD 55794 further clarifies the various contributions to the profiles. As an example, we used the [O I] 6300\\AA\\ line (which was identical in profile to the related 6363\\AA\\ line) for the decomposition, shown in the upper panel of Fig. 33. In the lower panel of Fig. 33 we show an empirical decomposition using a set of representative profiles from the data itself with the continuum of the Mira subtracted. The relative intensities are chosen for display, they are not results from a fitting procedure. The Mg II line was narrow with a FWHM of 25 km s$^{-1}$ centered at -61$\\pm$1 km s$^{-1}$, the same radial velocity previously derived for the Mira wind absorption lines. The [N II] 6548\\AA\\ showed an extended red wing to +200 km s$^{-1}$ and the persistent red peak. This latter emission peak coincided with one on the [S II] 6716 and 6730\\AA\\ lines that also showed the reduced blue peak, like [Ar III] 7135 (Fig. 32). The [O III] 4959\\AA\\ and 5007\\AA\\ lines no longer displayed the low velocity emission peaks, instead having a pure shock profile similar to [Ca V] 5309\\AA. The comparison indicates,however, that a part of the [O I] profile is still being contributed by the part of the wind that has not been completely ionized and, as we proposed in Paper 1, the wing may be due to charge exchange behind the shock.\u00caThe [S II] and [N II] lines still showed contributions from the extended environment. Deguchi et al. (2011) present the variations of the SiO maser during the first month of outburst. Taking $V_{Mira,LSR} = -26$ km s$^{-1}$, the main SiO (J=1-0, v=2) component is at -30 km s$^{-1}$ (LSR), corresponding in heliocentric $v_{rad}$ to -58 km s$^{-1}$, at slightly higher velocity ($\\approx$ 4 km s$^{-1}$) than the peak of the [O I] 6300\\AA\\ line on JD 55286 and the mean absorption line velocity but compatible with the wind line velocities. \\footnote{There is an ambiguity here. Deguchi et al. give the system heliocentric radial velocity of -41 km s$^{-1}$, and discuss orbital motion with a 43 year period citing the radial velocity from Tatarnikova et al. (2003b) using the Ca I 6573\\AA\\ line. We detected this line in the last NOT spectra from 2010 and in the 2011 spectrum at $v_{\\rm rad}$=-43 km s$^{-1}$. There was, as we mentioned above, a weak absorption feature on Na I D at a similar velocity and a weak absorption component on Mg I 5167\\AA. The velocity disagreed, however, with all other absorption line values and we have no explanation for this other than either a blend or emission filling in the blue wing. We adopt the velocity determined from the Li I 6707\\AA\\ and [Ca II] 7291, 7323\\AA\\ lines in our comparison.} The (J=1-0, v=1) transition also shows a weaker line at -47 km s$^{-1}$ that does not corresponding to any absorption on the Na I D lines. The other SiO maser line (J=1-0, v=2) shows only the -58 km s$^{-1}$ line. The velocity is consistent with that of the Mira wind as determined from the absorption features discussed in the previous sections. The rapid turn-off of the maser after the explosion, if due to the irradiation by the initial shock and the advance of the ionized region as a precursor, indicates that the cold material is located at a distance of about 1 to 10 AU from the Mira, consistent with the distance in Deguchi et al (2011). %We should note that the timescale for the disappearance of the wind absorption lines, $\\leq$45 days, is longer than the timescale of about 20 days on which the maser lines turned on again. The orientation proposed by Deguchi et al. (2011) from the maser emission places the WD near quadrature at the time of the nova explosion. The chromospheric lines, and the ``emission hole'' in the forbidden lines, are more consistent with a phase closer to superior conjunction for the WD -- that is, being seen through a substantial portion of the Mira wind. Two results point to this. The absorption line spectrum during the early outburst was strong, both on P Cyg components and in isolated absorption lines, and are more consistent with a long pathlength. The other is the asymmetry of the high velocity component of the lines. Were the WD seen nearer quadrature, it would not be possible to produce so asymmetric a shock line profile. The persistence of the neutral lines also support the picture that the inner portion of the wind was screened from the UV of the explosion and shock while the bulk of the wind was ionized within the first 60 days. If the fluorescent emission was excited by the UV of the WD, the disappearance of those transitions could indicate either that the ionization had removed the ions (mainly singly ionized Fe-peak transitions) or that the WD had turned off. Monitoring observations of V407 Cyg at centimeter wavelengths with the EVLA have also shown the appearance, after Day 160, of a resolved nonthermal component at a few times $10^{15}$ cm, depending on the distance adopted for the Mira.\\footnote{The EVLA spectra are available at the URL:\\\\ https://safe.nrao.edu/evla/nova/\\#v407cyg.} This is compatible with the location of the gas that we suggest was ejected during the 1930s event. The shock, if expanding as inferred in Paper I with a simple power law $v \\sim t^{-1/3}$, should have reached a distance of a few times $10^{15}$ cm in about one year. The 2011 Apr. 23 Ond\\v{r}ejov spectrum shows that the [N II] lines had developed extended redward emission with a maximum velocity of about +300 km s$^{-1}$, which is the velocity we expect from the shock at that time.\\footnote{As an aside we note that from the [O III] and [N II] diagnostics, the density at this velocity is compatible with a compression ratio of about an order of magnitude, indicating that the shock is likely now isothermal, and for this compression ratio we would expect a steep spectrum with an exponent of about -3.3 (Blandford \\& Ostriker 1978) and a strong shock of order $10^3$ in pressure ratio. } % correction for LSR to Hel = 12.5 km/s" }, "1112/1112.4532_arXiv.txt": { "abstract": "We extend the two-dimensional Cartesian shapelet formalism to $d$-dimensions. Concentrating on the three-dimensional case, we derive shapelet-based equations for the mass, centroid, root-mean-square radius, and components of the quadrupole moment and moment of inertia tensors. Using cosmological $N$-body simulations as an application domain, we show that three-dimensional shapelets can be used to replicate the complex sub-structure of dark matter halos and demonstrate the basis of an automated classification scheme for halo shapes. We investigate the shapelet decomposition process from an algorithmic viewpoint, and consider opportunities for accelerating the computation of shapelet-based representations using graphics processing units (GPUs). ", "introduction": "Complex, three-dimensional structures abound in astronomy on all scales from ``fluffy'' dust aggregrates in molecular clouds (Ossenkopf 1993; Stepnik et al. 2003), to cosmological large-scale structure that has been described as ``sponge-like'' (Gott, Dickinson \\& Melott 1986), or a ``skeleton'' (Sousbie et al. 2008) of clusters, filaments and voids (Barrow, Bhavsar \\& Sonoda 1985; White et al. 1987). While aspects of these structures can be expressed in terms of simple, geometrically-motivated properties such as their triaxiality or quadrupole moment, these quantities are not able to capture the higher order complexity of the true shape. The challenge, therefore, is to provide an accurate description of an arbitrary three-dimensional (3-d) shape, possibly over many physical length scales, in the hope that this can lead to improved theoretical or analytical insight into the structure in question. The human visual system is more than capable of identifying structures and sub-structures for an individual 3-d object, but such qualitative interpretations only have limited use -- it is not practical to attempt a classification of shapes by eye when there are many thousands of objects to inspect.\\footnote{Although, if there are enough individual eyes available to assist, then this approach is feasible, as the Galaxy Zoo project ({\\tt http://www.galaxyzoo.org}) has demonstrated.} The preferred alternative is an automated approach including: \\begin{itemize} \\item decomposition via an appropriate basis set (e.g. Fourier analysis, wavelet transformations); \\item partitioning [e.g. Voronoi tesselation - see Icke \\& van de Weygaert (1987) for an early cosmological application]; \\item the use of minimal spanning trees to identify connected structures (Barrow et al. 1985; Pearson \\& Coles 1995); \\item Minkowski functionals [which return global geometric properties such as volume, surface area and edge density -- Mecke, Buchert \\& Wagner (1994); Sahni, Sathyaprakash \\& Shandarin (1998)]; and \\item segementation [e.g. ``dendrograms'' used by Goodman et al. (2009) to identify self-gravitating structures in molecular clouds]. \\end{itemize} The approach we present in this paper is the extension of the two-dimensional (2-d) shapelet method (Refregier 2003) to three dimensions. Shapelets are sets of orthonormal basis functions based on the Hermite polynomial solutions of the quantum harmonic oscillator (QHO). Simple analytic forms can be derived for the physical properties of 3-d structures (e.g. centre of mass, root-mean-square radius and the components of the quadrupole moment and moment of intertia tensors), which can be efficiently calculated in shapelet space. In astronomy, 2-d shapelets have been applied to image simulation (Massey et al. 2004; Ferry et al. 2008), the morphological classification of galaxies (Kelly \\& McKay 2004; Andrae, Jahnke \\& Melchior 2011) and sunspots (Young et al. 2005), and weak gravitational lensing (Refregier \\& Bacon 2003). The latter includes the measurement of shear (Kuijken 2006), flexion (Goldberg \\& Bacon 2005), point-spread function modelling and deconvolution (Melchior et al. 2009; Paulin-Henriksson, Referegier \\& Amara 2009), and weak lensing by large-scale structure from the FIRST radio survey (Chang, Refregier \\& Helfand 2004). Massey et al. (2007) investigated weak lensing with polar shapelets (Massey \\& Refregier 2005), a form more suitable for images with rotational symmetry. Further properties of shapelets, including integral relations and convolution sums are presented in Coffey (2006). The importance of the shapelet approach lies not so much in the basis functions, but in the simplifed computation of quantities relating to shape and structure that can be determined once a shapelet decomposition has been obtained. These analytic quantities are expressed as linear sums of weighted shapelet states, greatly reducing the calculation complexity compared to (numerically) solving the related integral formulations. Shapelet decomposition is not without its problems [see Berry, Hobson \\& Withington (2004) for an extensive discussion]. Melchior, Meneghetti \\& Bartelmann (2007) examined the limitations of shapelet image analysis in cases where the orthonormality condition [see equation (\\ref{eqn:ortho}) below] fails, and proposed a decomposition procedure that preserves physical properties of images. Melchior et al. (2010) and Bosch (2010) considered problems with using circular Gaussian basis functions to model galaxies with high ellipticity or a large S\\'{e}rsic index. Ngan et al. (2009) proposed an alternative orthonormal basis based on the S\\'{e}rsic profile (hence S\\'{e}rsiclets) for use in weak lensing analysis. While helping to avoid issues with poor shape recovery from overfitting low signal-to-noise galaxies, and fitting with too many degrees of freedom, S\\'{e}rsiclets do not possess the analytic properties of shapelets, and the basis functions must be generated numerically. Indeed, it is the existence of analytic functions that has motivated our choice of 3-d Cartesian shapelets as an appropriate tool for quantifying properties of three-dimensional structures. The remainder of this paper is set out as follows. In Section \\ref{sec:cartesian}, we present the mathematics of 3- and $d$-dimensional Cartesian shapelets. New analytic expressions are presented for several important physical properties of 3-d structures in Section \\ref{sct:analytic}. In Section \\ref{sct:implement}, we describe issues relating to implementing an efficient 3-d shapelet decomposition code. We highlight the inherent high-degree of paralellism in the shapelet decomposition algorithm, which makes it a promising target for graphics processing units. In Section \\ref{sct:cosmoapp}, we present first applications of 3-d shapelets to problems in cosmological simulations, with an emphasis on studying sub-structure in dark matter halos, demonstrating how an automated shape classifier can work in shapelet space. We end with a summary and outlook for 3-d shapelets in astronomy in Section \\ref{sct:conc}. \\begin{figure*} \\begin{center} \\includegraphics[width=7in]{figure1.pdf} \\end{center} \\caption{Examples of three-dimensional Cartesian shapelets ($\\beta = 1$). Top row: (left) ${\\bmath n} = (0,0,0)$; (right) ${\\bmath n} = (0,1,0)$. Bottom row: (left) ${\\bmath n} = (2,0,2)$; (right) ${\\bmath n} = (1,2,4)$. For each panel, we calculate the maximum data value, $f_{\\rm max}$, and generate 10 equally spaced iso-surfaces over the range $(-f_{\\rm max}, f_{\\rm max})$. Individual isosurfaces are coloured with a two-ended intensity colour map: blue $\\rightarrow$ black $\\rightarrow$ orange. \\label{fig:B000}} \\end{figure*} ", "conclusions": "\\label{sct:conc} We have extended the two-dimensional Cartesian shapelet formalism of Refregier (2003) to three dimensions, deriving analytic expressions for the zeroth moment, object centroid, root-mean-square radius, and the components of the quadrupole moment and moment of inertia tensors. We also presented generalisations to $d$-dimensions. Further work is necessary to develop a robust and systematic optimisation strategy for the decomposition parameters, and the development of specfic applications for the three-dimensional shapelet technique requires such a strategy. There are also opportunities to develop the formalism further, specifically extending it to include spherical shapelet functions [c.f. the alternative presentation of two-dimensional Cartesian shapelets as polar shapelets by Massey \\& Refregier (2005)]. The shapelet decomposition algorithm exhibits attributes that make it an ideal target for implementation on modern, massively-parallel GPUs. Our algorithm analysis demonstrates that the computation is entirely (or embarassingly) parallel; has minimal or no branching; maintains a high ratio of arithmetic operations to memory-access operations; and has a memory access pattern that will result in aligned or contiguous access to memory, required for achieving a high memory throughput. With our proposed scheme of precomputing shapelet voxel-integral terms, the computation reduces to a parallel series of multiply-add operations, which are almost ideal for GPUs -- we anticipate achieving close to peak processing performance. Significantly reducing the computation time for the shapelet decomposition, compared to CPU, means that more processing time is then available for optimisation. As an example application, we have demonstrated how three-dimensional shapelets can be used to study the complex sub-structures of dark matter haloes from cosmological $N$-body simulations, including providing an alternative approach to classifying the properties of haloes. Our preliminary investigation suggests that halo triaxiality measured purely from the moment of inertia tensor may be incorrect due to limitations of group finders that are not able to separate out what may be truly distinct sub-clumps. Improvements to our current `by eye' approach to classification could include development of a decision tree or neural network classifier, or the use of principle component analysis to significantly reduce the number of shapelet terms required for classification (Kelly \\& McKay 2004). The shapelet formalism is virtually unexplored in the three-dimensional domain, offering opportunities for the further development of a methodology that can be used to quantify and analyse complex three-dimensional structures. Future applications of the three-dimensional shapelet techinique may include classification and parameterisation of sources identified in H{\\sc i} spectral line data cubes; studying the shapes of voids in cosmological simulations (by considering an inverted density field); and the possibility to generate mock dark matter haloes through an extensive study of the distribution of shapelet amplitudes as a function of mass and triaxiality." }, "1112/1112.1910_arXiv.txt": { "abstract": "{It has been proposed that the envelopes of luminous stars {{\\changedA may}} be subject to substantial radius inflation. The peculiar structure of such inflated envelopes, with an almost void, radiatively dominated region beneath a thin, dense shell could mean that many in reality compact stars are hidden below inflated envelopes, displaying much lower effective temperatures. {{\\changedA The inflation effect has been discussed in relation to the radius problem of Wolf-Rayet (WR) stars, but has yet failed to explain the large observed radii of Galactic WR stars.}} } { We wish to obtain a physical {{\\changedA perspective of the inflation effect}}, and study the consequences for the radii of Wolf-Rayet (WR) stars, and luminous blue variables (LBVs). For WR stars the observed radii are up to an order of magnitude larger than predicted by theory, whilst S Doradus-type LBVs are subject to humongous radius variations, which remain as yet ill-explained.} {{{\\changedA We use a dual approach to investigate the envelope inflation, based on numerical models for stars near the Eddington limit, and a new analytic formalism to describe the effect. An additional new aspect is that we take the effect of density inhomogeneities (clumping) within the outer stellar envelopes into account.}}} {{{\\changedA Due to the effect of clumping we are able to bring the observed WR radii in agreement with theory. Based on our new formalism, we find that the radial inflation is a function of a dimensionless parameter $W$, which largely depends on the topology of the Fe-opacity peak, i.e., on material properties.}} For $W>1$, we {{\\changedA discover}} an instability limit, for which the stellar envelope becomes gravitationally unbound, i.e.\\ there no longer exists a static solution. Within this framework we {{\\changedA are also able to explain the S\\,Doradus-type instabilities}} for LBVs like AG\\,Car, with a possible triggering due to changes in stellar rotation. } {The stellar effective temperatures in the upper Hertzsprung-Russell (HR) diagram are potentially strongly affected by the inflation effect. This may have particularly strong effects on the evolved massive LBV and WR stars just prior to their final collapse, as the progenitors of supernovae (SNe)\\,Ibc, SNe\\,II, and long-duration gamma-ray bursts (long GRBs).} ", "introduction": "\\label{sec:intro} In the standard picture for the evolution of the most massive stars (with $M \\ga 30\\,\\msun$), O stars have been proposed to evolve through successive luminous blue variable (LBV) and Wolf-Rayet (WR) phases before exploding as hydrogen (H) free supernovae (SNe) Ibc \\citep[e.g.,][]{lan1:94,mey1:03,heg1:03}. This evolutionary path is however still open as LBVs have more recently been suggested to be the direct progenitors of some H-rich type II SNe \\citep{kot1:06,smi2:07,gal1:09}. Additional renewed interest in WR stars arises from their connection to long-duration gamma ray bursts \\citep[long GRBs, cf.][]{yoo1:05,woo1:06}. What LBVs and WR stars have in common is their proximity to the Eddington limit. Not only is this thought to be instrumental for their mass-loss behavior \\citep{vin1:02,gra1:08,vin1:11,gra1:11}, but it may also have key consequences for their stellar structures. Using modern OPAL opacities, \\citet{ish1:99}, \\citet{pet1:06}, and \\citet{yun1:08} studied the interior structure of massive stars, revealing a ``core-halo'' configuration that involves a relatively small convective core and an extended radiative envelope. This is referred to as the ``inflation'' of the outer envelope. It has been known for many decades that the observed radii of WR stars are almost an order of magnitude larger than canonical WR stellar structure models indicate. This is oftentimes attributed to their so-called pseudo-photospheres, which concern an ``effective'' photosphere several times larger than the hydrostatic radius, as a result of a dense stellar outflow \\citep[e.g.,][]{cro1:07}. Pseudo-photospheres have also been discussed in the context of LBVs, although the issue is still under debate \\citep[e.g.,][]{smi1:04}. An alternative explanation for the large radii of WR stars may involve envelope inflation. The issue is particularly relevant in the context of WR stars as GRB progenitors, as there have been several suggestions that the required radii of GRB progenitors are up to an order of magnitude ($\\sim$\\,$1\\,R_\\odot$ vs. $\\sim$\\,$10\\,R_\\odot$) smaller than the radii of observed WR stars \\citep{mod1:09,cui1:10}. The WR radius issue is thus highly relevant for GRB modeling. For LBVs the issue of their radii is equally interesting. The defining property of LBVs concerns their ``S Doradus'' cycles. On timescales of years to decades LBVs are seen to vary between effective temperatures of $\\sim$30\\,kK (early B spectral type) to $\\sim$8\\,kK (early F). Due to the lack of convincing counter-evidence, these S Dor excursions in the upper Hertzsprung-Russell diagram have generally been assumed to occur at constant bolometric luminosity \\citep{hum1:94}, but recent studies have challenged this behavior for a couple of objects \\citep{gro1:09,cla1:09}. Either way, the issue of the increased LBV radii during S Dor cycles remains un-questioned, but a satisfying explanation for it has yet to be provided \\citep{vin1:09}. Outer envelope inflation may turn out to be an interesting explanation for S Dor variations, as will be detailed in the following. The paper is organized as follows. In Sect.\\,\\ref{sec:modelcomp} we present numerical stellar structure models that show the envelope inflation effect. In Sect.\\,\\ref{sec:anal} we describe the effect analytically, {\\changedA including a new instability limit. We also provide a recipe that can be used to estimate the inflation effect for arbitrary stars with given (observed) parameters}. The implications of our results are discussed in Sect.\\,\\ref{sec:implications}, with focus on WR\\,stars and LBVs. The conclusions are summarized in Sect.\\,\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} In the present work, we have discussed the possible formation of radially inflated stellar envelopes for stars approaching the Eddington limit, as originally proposed by \\citet{ish1:99}, and \\citet{pet1:06}. {\\changedA In addition to numerical models, we could provide an analytic description of this process, leading to the discovery of a new instability limit, and a clumping dependence. Both effects turn out to be of profound importance. While the former may be connected to S Doradus-type instabilities in LBVs, the latter can account for the large observed radii of Galactic WR stars, and thus resolve the long standing WR radius problem.} {\\changedA Within our analytical approach we could describe the inflation effect in terms of a dimensionless parameter $W$} (Eqs.\\,\\ref{eq:Qrecipe}, \\ref{eq:W}) that characterizes the ratio between internal and gravitational energy at the base of the envelope. $W$ can be estimated on the basis of opacity tables and a combination of stellar parameters. The envelope inflation occurs when $W$ approaches one. For $W\\ge 1$, we find that the envelope becomes gravitationally unbound, i.e., there exists no static solution. A prerequisite for envelope inflation is the proximity of the star to the Eddington limit, with Eddington factors of $\\Gamma_{\\rm e}>0.3$--0.35, at solar metallicity. For a given chemical composition and $\\Gamma_{\\rm e}$, the resulting stellar radius depends on the ratio $M/R$, and on the degree of (in)homogeneity of the material within the envelope (characterized by a clumping factor $D$). Envelope inflation has a strong impact on the effective temperatures of stars in the upper HR diagram. For Wolf-Rayet stars it can account for the 'radius problem', i.e., that the observed stellar temperatures of H-free WR stars are much lower than predicted by canonical stellar structure models. To reproduce the observed HRD positions of these objects, it is necessary to assume that the material within the envelope is clumped. The resulting clumping factors lie in the range of $D=1$--16, in notable agreement with typical clumping factors detected in the winds of WR stars \\citep[e.g.,][]{ham1:98}. As the envelope inflation is expected to depend on metallicity \\citep{ish1:99,pet1:06}, it may also account for the observed $Z$-dependence of the spectral subtype distribution of WR stars, i.e., that WR stars in high-$Z$ environments show lower effective temperatures \\citep[e.g.,][]{cro1:07}. For the luminous blue variable AG\\,Car we suggest that the observed HRD position of a $\\sim$\\,$70\\,M_\\odot$ star, with a relatively compact core, might be subject to substantial envelope inflation. During its S\\,Dor cycle, AG\\,Car is found to increase its radius by a factor of two, while its luminosity decreases by a factor of 1.5. The luminosity decrease might be explained as a result of the formation of a dense outer shell with a mass of $\\sim 2$$\\cdot$$10^{-3}\\,M_\\odot$ as predicted by our models. The energy needed to lift this material from the core to the outer radius is of the same order of magnitude as the 'missing energy' due to the decrease in luminosity. Similar to the mechanism proposed by \\citet{gro2:09,gro1:11}, the S\\,Dor variability of AG\\,Car may be due to the effect of stellar rotation, which reduces the effective stellar mass, and thus leads to an instability with $W>1$, that forces the star to expand. We conclude that the stellar effective temperatures in the upper HR diagram are potentially strongly affected by the inflation effect. The peculiar structure of the inflated envelopes, with an almost void, radiatively dominated region, beneath a thin and dense shell could mean that many, rather very compact stars, are hidden below inflated envelopes, and thus display much lower effective temperatures to the observer. Also compact, fast rotating stars may be hidden this way. This may particularly affect massive stars just before the final collapse, and the question of whether they form a SN\\,Ibc, or a GRB \\citep[for the latter smaller core radii are inferred than the observed WR radii, cf.][]{cui1:10}. The complex envelope structure may also affect the early X-ray afterglow of GRBs \\citep[e.g.][]{li1:07}." }, "1112/1112.0322.txt": { "abstract": "%(Note: I will write abstract and conclusions once we have decided the very final contents of the paper) We discuss chemical enrichments of $\\sim$4000 SDSS early-type galaxies using as tracers a large variety of element abundance ratios, namely [C/Fe], [N/Fe], [O/Fe], [Mg/Fe], [Ca/Fe] and [Ti/Fe]. We utilise the stellar population models of absorption line indices from \\citet*{TMJ10} which are based on the MILES stellar library. We confirm previous results of increasing age, [Z/H] and [O/Fe] ratios (most often represented by [$\\alpha$/Fe] in the literature) with velocity dispersion. We further derive identical correlations with velocity dispersion for the abundance ratios [O/Fe], [Mg/Fe] and [C/Fe], implying that C/Mg and C/O are close to solar values. This sets a lower limit on the formation time-scales and star-burst components of early-type galaxies to $\\sim$0.4 Gyr, which is the lifetime of a 3M$\\odot$ star, since the full C enrichment must be reached. [N/Fe] correlates with velocity dispersion, but offset to lower values and with a steeper slope compared to the other element ratios. We do not find any environmental dependencies for the abundances of C and N, contrary to previous reports in the literature. [Fe/H] does not correlate with velocity dispersion over the entire parameter range covered, but for fixed age we find a steep trend for the [Fe/H]-$\\sigma$ relation. This trend is weaker than the analogous for total metallicity (which also shows steeper trends at fixed age) owing to the lower Fe contribution from SN Ia for more massive early-type galaxies. We find [Ca/Fe] ratios that are close to solar values over the entire velocity dispersion range covered. Tentative, due to large scatter, the results for [Ti/Fe] indicate that Ti follows the trends of Ca. This implies a significant contribution from SN Ia to the enrichment of heavy $\\alpha$-elements and puts strong constraints on supernova nucleosynthesis and models of galactic chemical evolution. %Interestingly, we find no correlations with environment for any of the element abundance ratios. This is in disagreement with previous studies that found indications for over-abundances of N and C in low density environments and poses tight constrains to the formation histories of massive elliptical galaxies. %We present a method for simultaneously deriving the stellar population parameters log(age), [Z/H] and the element abundance ratios [C/Fe], [N/Fe], [O/Fe], [Mg/Fe], [Ca/Fe] and (for the first time) [Ti/Fe] for unresolved stellar populations. We utilise the new flux-calibrated stellar population models of absorption line indices from \\citet{TMJ10}. %The new method is used for deriving full chemical enrichment histories for a sample of 3802 early-type galaxies from the Sloan Digital Sky Survey \\citep[SDSS,][]{york00}. We confirm previous results of increasing age, [Z/H], [O/Fe] (most often represented by [$\\alpha$/Fe] in the literature), [Mg/Fe], [C/Fe] and [N/Fe] with stellar %velocity dispersion. As expected we find very similar trends for [O/Fe] and [Mg/Fe], but also for [C/Fe] which means that C/Mg and C/O are close to solar values. This sets a lower limit on the formation time-scales for early-type galaxies to $\\sim$0.3 Gyr, which is the age of a 4M$\\odot$ star, since the full C enrichment must be reached. %the elements [C/Fe], [N/Fe] and [O/Fe] confirming the production of C and N in massive stars. %On the contrary we find an under-abundance of Ca compared to Mg and O. Although tentative due to large errors, a similar pattern is found for Ti as for Ca. %scale with Fe, hence there are no such correlations with galaxy mass for either [Ca/Fe] %or [Ti/Fe]. %These results indicate that the contribution from SNIa to the enrichment of $\\alpha$-elements is dependent on atomic number. %more to the enrichment of heavier alpha-elements than previously %thought. %This puts strong constraints on supernova nucleosynthesis and models of galactic chemical evolution. %Interestingly, we find no correlations with environment for any of the element abundance ratios. This is in %disagreement with previous studies that found indications for over-abundances of N and C in low density %environments and poses tight constrains to the formation histories of massive elliptical galaxies. ", "introduction": "\\label{intro} %- Search for different formation histories for galaxies, bensby, feltzing used for Milky-Way %- Search for different formation mechanisms for different elements %- unresolved SP, integrated light important to reveal parameters %- introduce models %- systematic differences in models, how they affect the result. in this case differences in % stellar libraries and in stellar evolution. %- introduce the production of elements, Ca thought to be similar to Mg but evidence of not being % (cenarro et al 2003, thomas et al. 2003, smith et al), % C and N produced in low mass star according to smith et al, but in massive stars according to % sanchez-blazquez et al. %- smith et al. find a slope for Ca/Fe with Fe/H and a dependence of C/Fe with age that when taken % into account give a similar behavior for Ca with sigma as Mg, indicating that %- C and N thought to be produced in low- and intermediate mass stars, however evidence % for production of C in massive stars (intro sanchez-blazquez). Could differences found % for N and C both here and for GC be evidence for different production mechanisms for % C and N. %- Kelson find differences between N and C %- sanchez-blazquez show Evidence for environmental differences for C and N. %- environmental study in clemenz et al 2006, only influenced, not alpha, C or metallicity %- arheological downsisizing get support from xxx downsizing %- Stellar light useful for formation histories and chemical evolution %- Search for different formation histories for galaxies, %- downsizing found - results. Support from xxx downsizing %- more ratios to reveal different histories - results for galaxies and milky-way %- for first time, full range of elments up to Fe, detailed differences %- Search for different formation mechanisms for different elements %- introduce the production of elements, Ca thought to be similar to Mg but evidence of not being % (cenarro et al 2003, thomas et al. 2003, smith et al), % C and N produced in low mass star according to smith et al, but in massive stars according to % sanchez-blazquez et al. %- C and N thought to be produced in low- and intermediate mass stars, however evidence % for production of C in massive stars (intro sanchez-blazquez). Could differences found % for N and C both here and for GC be evidence for different production mechanisms for % C and N. %- Ca underabundance possible evidence for SNIa contributing to alpha-elements close to Fe-peak elements % the inclusion of Ti strengthen this, Ti included here for first time. %- unresolved SP, integrated light important to reveal parameters %- Lick indices and models %- systematic differences in models, how they affect the result. in this case differences in % stellar libraries and in stellar evolution. %Stars carry useful information on the conditions present at the time of their formation. The chemical compositions of stellar atmospheres are tracers of the element abundances of the parent gas clouds forming the stars throughout the formation history of a galaxy. Some elements are also affected by dredge-up during stellar evolution \\citep*[e.g.][]{sweigart89}. %up to several billion of years ago. Stellar populations are therefore a powerful tool to extract information on chemical evolution in the Universe. Element abundances can be directly determined for individual stars of resolved stellar populations in the Milky Way or in nearby dwarf galaxies, using absorption lines measured in high resolution stellar spectra \\citep*[e.g.][]{edvardsson93,fuhrmann98,bensby04,feltzing09,bensby10}. Light averaged spectra must instead be used for determining element abundances of distant unresolved stellar populations. The absorption features of such spectra are sensitive to multiple elements due to velocity dispersion broadening. The Lick system of absorption line indices \\citep*[e.g.][]{worthey94,trager98} have been frequently used for measuring 25 prominent absorption features in galaxy spectra. Elements are produced in stellar nucleosynthesis besides the primordial nucleosynthesis of H and He. The chemical enrichment of stellar populations depends on the star formation history, initial mass function, fraction of exploding supernovae etc. The chemical pattern of the parent gas clouds will be carried on to new stellar generations. Thus the chemical enrichment of stellar populations is also affected by mechanisms affecting the interstellar medium (ISM) such as the efficiency of stellar winds to mix newly synthesised elements with the ISM, efficiency of outflow from galactic winds to remove enriched gas, inflow of less enriched gas from gas reservoirs etc. \\citep[e.g.][]{matteucci89,matteucci94}. Chemical enrichment sets stringent constraints on galaxy formation and evolution. %, which can be done by studying element abundance ratios of stellar populations and the ISM. %The chemical pattern of a stellar population will be affected by several different mechanisms such as the time-scale of star formation, the efficiency to mix newly synthesised elements with the interstellar medium (ISM), the initial mass function, efficiency of galactic winds to remove the ISM, fraction of Supernovaes etc. Tracing the chemical enrichment of stellar populations is therefore useful for constraining galaxy formation which can be done by studying element abundance ratios of the stellar populations and the ISM. Studies beginning in the late 1970's have revealed non-solar abundance ratios for the stellar populations of early-type galaxies \\citep*[e.g.][]{oconnell76,peterson76,burstein84,worthey92,davies93,surma95}, indicating different chemical enrichment histories. %This information can be used for setting constraints on star formation histories and the evolution %of the Universe. %on the production of various elements. This triggered more detailed investigations showing that the ratio between $\\alpha$-elements and Fe-peak elements increases with increasing galaxy mass for early-type galaxies \\citep[e.g.][]{trager00b,thomas05,bernardi06,clemens06,thomas10}. %Containing most of the stellar mass in the Universe, early-type galaxies are %good tracers of mass assembly and have been used extensively to constrain galaxy formation. Most interestingly, the [$\\alpha$/Fe] ratio participates in the E-E dichotomy, i.e. elliptical galaxies with low [$\\alpha$/Fe] ratios have core-less central profiles, while [$\\alpha$/Fe]-enhanced galaxies with short formation time-scales have cores \\citep{Kormendy09}. %While the $\\alpha$-elements are believed to mainly be ejected into the interstellar %medium through Type II supernovae (SNII), Fe-peak elements are also contributed by Type Ia supernovae (SNIa). %The delay of SNIa relative star formation results in weaker $\\alpha$/Fe for more extended periods of star formation. %Hence the observed $\\alpha$/Fe-mass trends have been interpreted as shorter formation time-scales for more massive early-type galaxies%(mass most often representented %by velocity dispersion) %, referred to as %astro-archeological downsizing. %a contribution of $\\alpha$-elements to the interstellar medium (ISM) %from Type II supernovea (SNII) only, while SNIa contribute Fe-peak elements as well under delayed time-scales relative to star formation. %The consequence is that a higher $\\alpha$/Fe ratio indicate shorter formation time-scales for more massive %early-type galaxies, referred to as astro-archeological downsizing. %Other explanations to the observed $\\alpha$/Fe-$\\sigma$ relationship have been proposed, in addition to astro-archeological downsizing, %such as mass dependent initial mass function (IMF) and more effective SNII yields at higher velocity dispersion \\citep{trager00b}. %In the debate, astro-archeological downsizing gets support %from the cosmological downsizing found by \\citet{cimatti06} and \\citet{pozzetti09}, where the populations of more massive galaxies %are in place at higher redshifts. The light-averaged ages of early-type galaxies have also been found to increase with increasing velocity dispersion \\citep[e.g.][]{thomas10}, giving %further support to the downsizing scenario. %The downsizing scenario is in contradiction to todays models of hierarchical galaxy formation that instead predict %the most massive systems to have the longest formation time-scales \\citep[e.g][]{lucia07}. Hierarchical models are unable to %reproduce the observed $\\alpha$/Fe-$\\sigma$ relationship \\citep{thomas99}, but are on the other hand %able to very well reproduce the observed large scale clustering in the universe \\citep[e.g.][]{springel05}. %The picture of galaxy evolution is clearly far from fully understood. Individual element abundance ratios, in addition to $\\alpha$/Fe, can further disentangle the formation of different stellar populations. Since the individual elements are produced in different stellar evolutionary phases they trace varying formation histories. \\citet{edvardsson93} and \\cite{bensby10} derive as many as 12 different element abundance ratios to distinguish between the different formation histories of the stellar populations in the Milky~Way. %Several studies have derived individual element abundances for early-type galaxies, in most cases for fairly small samples, %with varying results. \\citet{sanchez03} predict over-abundances of carbon and nitrogen %in environments of lower density for 98 early-type galaxies, while \\citet{clemens06} study a large sample of SDSS %\\citep[Sloan Digital Sky Survey,][]{york00} early-type galaxies and find environment not to influence the enhancement of carbon. %\\citet{sanchez06} study 98 early-type galaxies and find evidence for $\\alpha$/Fe and C/Fe to increase with increasing $\\sigma$ with a similar slope, while %N/Fe show a steeper slope and Mg/Fe an even steeper slope. %\\citet{kelson06} on the other hand find no significant trends for neither $\\alpha$/Fe nor C/Fe with $\\sigma$, but a strong anticorrelation of %$\\alpha$/N with $\\sigma$. %\\citet{graves08} introduced a method of simultaneously derviing individual element abundance ratios for unresolved stellar populations. %This method was adopted in \\citet{graves07} and \\citet{smith09} for a large of SDSS galaxies and 147 cluster galaxies, respectively. \\citet{smith09} %find similar trends of increasing abundancs for Mg, C and N with increasing $\\sigma$ for a sample of 147 red-sequence galaxies. %The $\\alpha$-element Ca is a puzzle as it is has been found to trace Fe instead of other $\\alpha$-elements for early-type galaxies %\\citep{cenarro03,thomas03b,smith09}. A number of studies have derived individual element abundance ratios for early-type galaxies \\citep{sanchez03,sanchez06,clemens06,kelson06,graves07,graves08,smith09,price11}, in several cases for fairly small samples. %Results have been \\citet{sanchez03}, \\citet{sanchez06}, \\citet{clemens06}, \\citet{kelson06}, \\citet{graves07}, \\citet{graves08} and \\citet{smith09}, leaving the picture of chemical evolution in early-type galaxies incomplete. Different methods are applied in these studies, but they are all based on absorption line indices. The results are dependent on the method applied and the sample used. The aim of this work is to simultaneously derive all element abundance ratios allowed by the sensitivity of the adopted absorption line indices. The maximum amount of information is extracted from the indices to reliably derive the abundance ratios and state of the art models of stellar populations of absorption indices are utilised to obtain as accurate results as possible. To fully interpret observed element abundance ratio trends, % derived for unresolved stellar populations stellar nucleosynthesis needs to be understood. \\citet{pipino10} find up-to-date models of chemical evolution to struggle in simultaneously reproducing observed abundance ratios for Carbon and Nitrogen from the unresolved stellar populations of early-type galaxies. The $\\alpha$-element Ca is a puzzle as it is has been found to trace Fe instead of other $\\alpha$-elements for early-type galaxies \\citep{cenarro03,thomas03b,smith09,Saglia02}. This has been interpreted as Ca being contributed by SNIa as well as SNII \\citep*{TJM10}. Thus element abundance ratios of unresolved stellar populations are also useful for constraining stellar nucleosynthesis. % \\citet{sanchez03}, predict over-abundances of carbon and nitrogen %in environments of lower density for 98 early-type galaxies, while \\citet{clemens06} study a large sample of SDSS %\\citep[Sloan Digital Sky Survey,][]{york00} early-type galaxies and find environment not to influence the enhancement of carbon. %\\citet{sanchez06} study 98 early-type galaxies and find evidence for $\\alpha$/Fe and C/Fe to increase with increasing $\\sigma$ with a similar slope, while %N/Fe show a steeper slope and Mg/Fe an even steeper slope. %\\citet{kelson06} on the other hand find no significant trends for neither $\\alpha$/Fe nor C/Fe with $\\sigma$, but a strong anticorrelation of %$\\alpha$/N with $\\sigma$. %\\citet{graves08} introduced a method of simultaneously derving individual element abundance ratios for unresolved stellar populations. %This method was adopted in \\citet{graves07} and \\citet{smith09} for a large sample of SDSS galaxies and 147 cluster galaxies, respectively. \\citet{smith09} %find similar trends of increasing abundancs for Mg, C and N with increasing $\\sigma$ for a sample of 147 red-sequence galaxies. %The $\\alpha$-element Ca is a puzzle as it is has been found to trace Fe instead of other $\\alpha$-elements for early-type galaxies %\\citep{cenarro03,thomas03b,smith09}. %The picture is again far from fully understood and studies of the full range of element abundances for large samples of galaxies %are needed. %More precisely, this work presents We present a technique for deriving a wide range of element abundance ratios %the full chemical enrichment histories for unresolved stellar populations, including %the abundance ratios [O/Fe] (representing [$\\alpha$/Fe]), [C/Fe], [N/Fe], [Mg/Fe], [Ca/Fe] and [Ti/Fe]. The method is based on new flux-calibrated stellar population models of absorption line indices presented in \\citet{TMJ10}. %The sensitivities to individual %element abundances of the absorption line indices are used for simultaneously derive all element abundance ratios. In \\citet{TJM10} we applied the method to galactic globular clusters and found it to be well calibrated. We analyse a sample of 3802 SDSS early-type galaxies for which we investigate element ratio scaling relations with velocity dispersion.% and environmental density. The paper is organised as follows. The SDSS early-type galaxy sample used is presented in Section~\\ref{data} and the technique for deriving the element abundance ratios is described in Section~\\ref{spp}. The results of derived element abundance ratios for the data sample are presented in Section~\\ref{results} and further discussed and compared with the literature in Section~\\ref{disc}. Conclusions are given in Section~\\ref{conc}. %The use of individual element abundances can not only be used for %The varying results of individual element abundances questions the %Individual element abundance ratios are not only useful for ratios obviously can be used for SFH as well as constraining processes producing elements, IMF variations etc. %For the first time include all elements. %The picture of galaxy formation and evolution is clearly far from fully understood %and the star, also chemical evolution not well understood. Studying several ratios can disentangle galaxy evolution and %chemical evolution. For the first time include all elements %To make the picture even more complex, several studies have found that not all $\\alpha$-like elements are equally enhanced %in early-type galaxies \\citep*[e.g.][]{vazdekis97,worthey98,trager98,henry99,TMB03,thomas03b,smith09}. %This implies that %The picture of galaxy evolution is clearly not fully understood and further analyses are needed to break down %the details of galaxy formation. To do so %All these studies indicated that the formation and evolution of galaxies are %more complex than previously believed. Stellar population models of absorption %indices are a key tool for unveiling these complexities surrounding galaxy %evolution, since absorption indices are sensitive %to the fundamental stellar population parameters age, metallicity and %element abundance ratios. %Integrated light for unresolved stellar populations. %It is believed that in the explosion of SNII the full range of elements up to Fe is produced, %including the $\\alpha$-elements. The explosion of SNIa is instead believed mainly produce %the F-epeak elements. The ratio %Stellar population models of absorption line indices are a key tool for %the analysis of galaxy absorption spectra and the derivation of fundamental %galaxy properties such as formation age and element abundance %ratios (e.g. [$\\alpha$/Fe]. The absorption %Several studies have later found that not all $\\alpha$-like elements are equally enhanced %in early-type galaxies \\citep*[e.g.][]{vazdekis97,worthey98,trager98,henry99,TMB03,thomas03b,smith09}. %Important cosmological results have also been made from studying element abundance ratio effects using stellar %population models of absorption indices, with the indication of %down-sizing %effects for early-type galaxies (early-type galaxies), where the most massive early-type galaxies seem to %have formed under the shortest time-scales \\citep{trager00b,thomas05}. %For such studies the %[$\\alpha$/Fe]-ratio is useful as it traces formation time-scales, with %the $\\alpha$-elements being produced during short time-scales when massive %stars end their life as SNII and the Fe-elements being produced under more %extensive %time-scales when binaries end up as SNIa. %The indications of down-sizing effects %were in contradiction to hierarchical models of galaxy formation %that could not reproduce the enhancement of $\\alpha$-elements in luminous %ellipticals \\citep{thomas99} %uitable for tracing mass assembly %- element abundance ratios %- Enivronment studies for early-type galaxies %- Clemens et al. (2006): no gradient of Ca4227 with velocity dispersion % Velocity dispersion correlate with environment, higher sigma in denser areas %- Introduce the use of absorption line indices %- Motivate the study of early-type galaxies %----------------------------- % CHAPTER 2 %----------------------------- ", "conclusions": "\\label{conc} We present light-averaged ages, metallicities and element abundance ratios for 3802 SDSS early-type galaxies drawn from the MOSES catalogue \\citep{schawinski07} with visual morphology classifications. %The 25 Lick indices measured on the observed galaxies are compared to the flux-calibrated models of stellar populations based on absorption line indices from \\citet*{TMJ10}. The sensitivity of the Lick indices to the variation of different element abundances allow us to determine element abundance ratios for C, N, O, Mg, Ca, Ti and~Fe. %In this work we use updated stellar population models and a new method that on top of age and metallicity computes [C/Fe], [N/Fe], [O/Fe], [Mg/Fe], [Ca/Fe] and [Ti/Fe]. This method is of an iterative nature due to the sensitivity to multiple elements for several Lick indices. Using the flux-calibrated TMJ models of absorption line indices, which are based on the MILES stellar library, we have developed a method for simultaneously deriving the element abundance ratios [C/Fe], [O/Fe] (inferred from [$\\alpha$/Fe]), [N/Fe], [Mg/Fe], [Ca/Fe] and [Ti/Fe]. The models are well calibrated with galactic globular clusters with independent measurements of stellar population parameters and element ratios \\citep[TMJ,][]{TJM10}. %Light-average ages, total metallicities and [$\\alpha$/Fe] ratios were derived for the same data sample in \\citet{thomas10}. In this work we use updated stellar population models and a new method that on top of age and metallicity computes [C/Fe], [N/Fe], [O/Fe], [Mg/Fe], [Ca/Fe] and [Ti/Fe]. This method is of an iterative nature due to the sensitivity to multiple elements for several Lick indices. %The element abundance ratios [C/Fe], [N/Fe], [O/Fe], [Mg/Fe], [Ca/Fe] and [Ti/Fe] are discussed along with light-average age, total metallicity and iron abundance. %Despite new models and method the results of this work agree with the results of \\citet{thomas10}, i.e. strong correlations for age, metallicity, [O/Fe] and [Mg/Fe] with velocity dispersion. Hence the the conclusions from \\citet{thomas10} remain valid. We study the relationships between the stellar population parameters and galaxy stellar velocity dispersion. In agreement with the literature stellar population age and total metallicity correlate with velocity dispersion. [Fe/H] instead does not show such a correlation over the entire parameter range covered, but for a fixed age a steep trend is found for the [Fe/H]-$\\sigma$ relation. This trend is shallower than the analogous for [Z/H] due to suppressed Fe enrichment in more massive galaxies because of time-scale dependent contribution from SN Ia Similar trends are found for [O/Fe], [Mg/Fe] and [C/Fe], i.e. strong correlations with velocity dispersion in agreement with the literature. The first two are expected to be similar, since both O and Mg belong to the group of $\\alpha$-elements produced in massive stars through type II Supernovae. This is also in favour of the down-sizing scenario of early-type galaxies that set an upper limit on the star formation time-scales and where more massive systems experience shorter time-scales \\citep[e.g.][]{thomas10}. The C/Mg ratios are close to solar values, which instead sets a lower limit for the formation time-scales of early-type galaxies. Stars with masses down to $\\sim$3 M$_{\\odot}$ contribute significantly to the production of C. To reach solar C/Mg ratios formation time-scales need to be long enough for such stars to eject C into the ISM. The inferred lower formation time-scale limit is then $\\sim$0.4 Gyr, which is the life-time of a 3 M$_{\\odot}$ star. %contribute significant amounts of C and have a life-times of $\\sim$0.4 Gyr, which then sets the lower formation time-scale limit. %A slightly shallower and steeper slope compared to the [Mg/Fe]-$\\sigma$ relationship is found for [O/Fe]-$\\sigma$ and [C/Fe]-$\\sigma$, respectively. This can be explained by metallicity dependent mass-loss rates in massive stars, ejecting more C before it can be converted to O. The [N/Fe] ratios are overall lower by $\\sim$0.2 dex compared to [O/Fe] and [Mg/Fe] and the trend with velocity dispersion is very steep, i.e. more massive galaxies have significantly higher [N/Fe] ratios. The observed [N/Fe]-$\\sigma$ trends are difficult to interpret due to uncertainties in the origin of N. %Systematic uncertainties in the modelling of N abundances arise from the choice of stellar evolutionary tracks and further complicates the interpretation. These systematic uncertainties affect the zero-point of the [N/Fe]-$\\sigma$ relationship, while the slope remains steep. The zero-point and slope of this relationship can not be simultaneously matched by up-to-date models of chemical evolution \\citep{pipino10}. Either the theoretical stellar yields have to be increased by a significant factor or other prescriptions have to be incorporated into the models that affect the N yields. Such prescriptions could be: 1. N yields with a stronger dependence on metallicity, since more massive early-type galaxies are more metal-rich. 2. A dependence on galaxy mass for the ratio between the time-scale of star formation and the time-scale of primordial gas inflow, which affects the N/O, N/Mg and N/Fe ratios due to the secondary nature of N. %3. Star formation fraction in globular clusters, which show high N abundances, that is dependent on galaxy mass. %At least three different scenaThis slope is stronger than for [Mg/Fe]-$\\sigma$ and [O/Fe]-$\\sigma$, indicatingThe N abundances are under-abundant compared to O and Mg with $\\sim$0.2 dex. We do not find any dependence on environmental density for the element ratios studied. This is in contradiction to previous studies that have reported environmental dependencies for C and N abundances. Hence difference formation scenarios for field and cluster early-type galaxies can not be inferred from the element ratios studied in this work. The [Ca/Fe] ratios do not correlate significantly with velocity dispersion and are close to solar values over the entire velocity dispersion range covered. Although tentative, due to large errors, Ti shows a behaviour similar to Ca. %abundances are also overall under-abundant compared to O and Mg. This indicates an atomic number dependent contribution from type Ia Supernovae to the production of $\\alpha$-elements, i.e. the yields from type Ia Supernovae are higher for heavier $\\alpha$-elements. This is now universally found since similar patterns have been found in the stellar populations of the Milky Way \\citep[][and references therein]{TJM10} and puts strong constraints on supernova nucleosynthesis. %The ? elements follow a pattern such that the elements with higher atomic number, namely Ca and Ti, are less enhanced. More specifically, [Ca/Fe] ratios are lower than [O/Fe] and [Mg/Fe] by about 0.2 dex. Ti continues this trend. We compare this result with recent determinations of element abundances in globular cluster and field stars of the Milky Way. We come to the conclusion that this pattern is now universally found. It suggests that Type Ia su- pernovae contribute significantly to the enrichment of the heavier ? elements as predicted in supernova explosion calculations and galactic chemical evolution models. This explains the presence of a Ca under-abundance (close to solar [Ca/Fe] ratios) in massive early-type galaxies and predicts similarly low [Ti/Fe] ratios in pop- ulations with short formation time-scales. %[Ti/Fe] is for the first time derived for extragalactic unresolved stellar populations in this work." }, "1112/1112.1067_arXiv.txt": { "abstract": "We present a resolved-star spectroscopic survey of \\ndwarfs{} dwarf spheroidal (dSph) satellites of the Andromeda Galaxy (M31). We filter foreground contamination from Milky Way (MW) stars, noting that MW substructure is evident in this contaminant sample. We also filter M31 halo field giant stars, and identify the remainder as probable dSph members. We then use these members to determine the kinematical properties of the dSphs. For the first time, we confirm that And XVIII, XXI, and XXII show kinematics consistent with bound, dark matter-dominated galaxies. From the velocity dispersions for the full sample of dSphs we determine masses, which we combine with the size and luminosity of the galaxies to produce mass-size-luminosity scaling relations. With these scalings we determine that the M31 dSphs are fully consistent with the MW dSphs, suggesting that the well-studied MW satellite population provides a fair sample for broader conclusions. We also estimate dark matter halo masses of the satellites, and find that there is no sign that the luminosity of these galaxies depends on their dark halo mass, a result consistent with what is seen for MW dwarfs. Two of the M31 dSphs (And XV, XVI) have estimated maximum circular velocities smaller than $12$ km/s (to 1$\\sigma$), which likely places them within the lowest mass dark matter halos known to host stars (along with Bo\\\"{o}tes I of the MW). Finally, we use the systemic velocities of the M31 satellites to estimate the mass of the M31 halo, obtaining a virial mass consistent with previous results. ", "introduction": "\\label{sec:intro} Dwarf spheroidal (dSph) galaxies are among the most extreme objects in the pantheon of galaxies. Their low luminosities ($10^3 < L/L_\\odot < 10^8$), lack of significant gas \\citep{grcevich09}, and low numbers compared to \\LCDM{} expectations \\citep{kl99ms,moo99ms} are all puzzles that remain to be solved. The difficulty in understanding the count of dSph galaxies around the Milky Way (MW) and M31 is known as the missing satellites problem, an issue that has prompted a flurry of activity modeling these galaxies \\citep[recently,][and references therein]{krav10satrev,bullock10msp,kaz11stirr,font11}. Most models rely heavily on feedback scenarios that are tied directly to the masses of the dark matter halos that presumably host dSph galaxies. In this sense, mass determinations for dwarfs are among the most important diagnostic measurements for testing theoretical predictions at the frontier of galaxy formation. { \\let\\thefootnote\\relax\\footnotetext{* 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} } \\addtocounter{footnote}{-1} While the brightest dSphs can be detected at the distance of nearby clusters \\citep[e.g.,][]{hilker03,durrell07}, faint, diffuse dSphs can only be detected via resolved star counts, which limits detection to the Local Group (LG). Kinematics of these galaxies thus requires resolved star spectroscopy at extragalactic distances. Thus, despite the motivations outlined above, detailed study of a large population of dSphs has been limited to the MW satellites \\citep[e.g.,][]{mateo98,sandg,walker09,simon11seg1}. Studies of the MW dSph population have resulted in puzzles that have provided interesting challenges to \\LCDM{} and galaxy formation models. The missing satellite problem noted above is the classic example, for which a variety of solutions have surfaced \\citep[e.g,][] {bullock00,stri07msp,tollerud08,bovill09,kop09,krav10satrev}. However, there remain other questions such as the cause of their low gas fractions \\citep {grcevich09,nichols11}, their morphologies \\citep{kaz11stirr}, or the curiously small number with the high densities expected by \\LCDM{} \\citep{BKBK11}. These studies are based entirely on the MW dSph population, as this has been the only available data set. Yet there is evidence that the MW has had an unusual merger history relative to similarly bright galaxies such as M31 \\citep{guhathakurta06,hammer07}. Furthermore, there are hints that the dSph populations of M31 and the MW exhibit different scaling relations \\citep{mcc05andbig,kalirai10}. Hence, expanding the sample of satellite systems is crucial to generalizing the information dSphs provide about galaxy formation. Fortunately, the past few years have seen much growth in the known satellite population of M31. This is primarily due to the advent of deep surveys of the region surrounding M31 specifically designed to search for substructure like dSphs or their remnants \\citep[e.g.,][]{ibata07,pandas09nat}. While the distance to M31 means that detection limits do not reach those of the MW's ultra-faint dSph system, smaller angular coverage is needed to survey M31's environs. Thus, the full population of known M31 satellites (shown in Figure \\ref{fig:overview}) now includes 27 dSphs \\citep{zucker04,martin06,zucker07,maj07and14,irwin08,mc08,martin09,richardson11,Slater11And28,Bell11And29}. These are further supplemented by 4 dwarf Elliptical (dE) satellites, which are similar in morphology to the dSph, but have somewhat higher luminosities. The imaging surveys that detected these dSphs provide data sets that allow characterization of the photometric properties of M31 satellites, but do not provide the kinematics of these dSphs necessary for characterizing their masses and dark matter content. These kinematical data sets are also crucial for confirming the candidates' status as self-bound galaxies. This is illustrated by two M31 dSphs (And IV and VIII) that were originally identified as dSph satellites, but were later shown by spectroscopic follow-up to be non-satellites \\citep{ferg00no4,ibata04and8,merrett06and8}. Beyond this, understanding the kinematics of M31 satellite system as a whole also provides insight into M31, its dark matter halo, and its accretion history \\citep[e.g.,][]{evans00,vdm08}. While kinematics exist for some of the M31 dSphs \\citep{cote99and2,guhathakurta00,chapman05and9,Chapman07,chapman08,Letarte09,collins10,kalirai10}, the large number of recent discoveries leave many yet to be spectroscopically confirmed, and a homogeneously observed and reduced sample of a significant fraction of these satellites is necessary to properly determine characteristics of the satellites system \\emph{as a whole}. With these ends in mind, we report here on kinematics of \\ndwarfs{} dSphs from the Spectroscopic and Photometric Landscape of Andromeda's Stellar Halo (SPLASH) Survey. This ongoing survey of the environs of M31 aims to characterize the stellar halo of M31 and its satellite population via resolved star spectroscopy. A companion paper on structural parameters and photometric properties is forthcoming \\comppapp. This paper is organized as follows. In \\S \\ref{sec:obs}, we describe the observations performed for this data set. In \\S \\ref{sec:analysis}, we describe the reduction and membership analysis performed homogeneously across the spectroscopic data set, as well as our method for estimating total velocity dispersions for each satellite in the sample. In \\S \\ref{sec:data}, we present the results of our full spectroscopic sample, and consider each satellite in turn, describing the results of our analysis and unique aspects of each galaxy. In \\S \\ref{sec:scaling}, we consider the scaling relations of M31 dSphs and compare them to the MW. In \\S \\ref{sec:M31mass}, we use the galaxies' $\\vsys$ to estimate the mass of M31. Finally, we present our conclusions in \\S \\ref{sec:conc}. ", "conclusions": "\\label{sec:conc} In this paper, we have described spectroscopy of M31 dSph satellites as part of the SPLASH Survey. We filter out MW foreground and M31 halo field contamination to identify M31 dSph member stars, and use these data to determine $\\vsys$ and $\\slos$ for the satellites. Based on these kinematics, we determine for each dSph the implied mass within the half light radius, and (under the assumption that these objects are dark matter-dominated) we estimate their dark halo properties. This paper can be summarized as follows: \\begin{enumerate} \\item We provide a homogenous spectroscopic survey of \\ndwarfs{} M31 dSphs and provide radial velocities of resolved stars in these galaxies. \\item We confirm that And XVIII, XXI, and XXII are kinematically cold and hence likely true satellite galaxies. \\item We find that And XXII has a $\\vsys$ close to with M33, suggesting it is associated with M33 rather than M31. If so, this is likely the first large mass ratio sub-subhalo (or satellite of a satellite) known. \\item We find that the M31 dSphs obey very similar mass-size-luminosity scalings to those of MW satellites. This suggests that the MW satellite population is not particularly unique and may be typical of a starforming $L_*$ galaxy. \\item We use the scalings of the M31 dSphs to infer properties of their dark matter halos. The masses of these halos show no sign of scaling with luminosity, similar to the MW dSphs \\citep{stri08commonmass,walker09}. \\item The density of And XIV, as well as perhaps And XV and And XVI, is consistent with dark matter halos with $\\vmax<10$ \\kps{} (although consistent with higher masses at $\\sim 1\\sigma$). If the most-likely masses for these systems are correct, these (along with the MW satellite Bo\\\"{o}tes I) are the lowest-mass dark matter halos hosting stars, with potential well depths indicative of field halos that are below the atomic hydrogen cooling limit. \\item Using the systemic velocities of M31 dSphs as tracer particles and adopting an empirical mass estimator suggested by n-body simulations, we estimate the mass of M31 within \\satsrad{} kpc to be $\\andmass$. This corresponds to a virial mass for M31's dark matter halo of $\\andmvir$. \\end{enumerate} This analysis of M31 dSphs thus represent a major step forward in understanding the faintest known class of galaxies. The M31 satellites present an opportunity to understand these galaxies in a new way, as a system, along with their host halo, providing a rich set of opportunities for examining galaxy formation and \\LCDM. Their similarity to MW dSphs also confirm of the Copernican principle, affirming that the MW may be a typical galaxy with typical satellites, albeit in an extraordinary universe." }, "1112/1112.4752_arXiv.txt": { "abstract": "We consider the effect of galaxy intrinsic alignments (IAs) on dark energy constraints from weak gravitational lensing. We summarise the latest version of the linear alignment model of IAs, following the brief note of Hirata \\& Seljak (2010) and further interpretation in Laszlo et al. (2011). We show the cosmological bias on the dark energy equation of state parameters $w_0$ and $w_a$ that would occur if IAs were ignored. We find that $w_0$ and $w_a$ are both catastrophically biased, by an absolute value of just greater than unity under the Fisher matrix approximation. This contrasts with a bias several times larger for the earlier IA implementation. % Therefore there is no doubt that IAs must be taken into account for future Stage III experiments and beyond. We use a flexible grid of IA and galaxy bias parameters as used in previous work, and investigate what would happen if the universe used the latest IA model, but we assumed the earlier version. We find that despite the large difference between the two IA models, the grid flexibility is sufficient to remove cosmological bias and recover the correct dark energy equation of state. In an appendix, we compare observed shear power spectra to those from a popular previous implementation and explain the differences. ", "introduction": "The gravitational lensing of distant galaxy images has the potential to be a powerful cosmological tool. The lensing effect directly probes the matter distribution as a function of redshift, and thus tells us about the expansion history and growth of structure in the Universe. In this way it allows us to constrain the dark energy equation of state, or whatever is causing the apparent accelerated expansion. Weak gravitational lensing poses a number of tough technical challenges if its true potential is to be exploited. The typical cosmic shear induced on galaxies of interest % is of order 1\\%, which is significantly smaller than the intrinsic ellipticity of the galaxies themselves. Of course we, as observers, have no access to the \\textit{unlensed} galaxy images so we must treat a population of galaxies statistically to recover the cosmological information contained in the cosmic shear signal. The correlation function of galaxy shapes as a measure of gravitational lensing was first proposed by~\\citet{kaiser92} and first observed by \\citet{Bacon:2000yp,Kaiser:2000if,Wittman:2000tc} and \\citet{van_Waerbeke:2000rm}. For reviews see \\citet{Bartelmann:1999yn,Munshi:2006fn,Refregier:2003ct}, and \\citet{hoekstra_jain_2008}. A naive approach to cosmic shear assumes that the intrinsic distribution of galaxy ellipticities is random across the sky. If this was the case observed ellipticities on a certain patch of sky could be averaged to recover the cosmic shear, because the intrinsic ellipticity would average to zero. However it was soon pointed out that this assumption of random intrinsic ellipticity distribution is unjustified \\citep{HRH,catelankb01,crittendennpt01,croftm00}. In fact galaxies may be expected to align with the large scale gravitational potentials in which they form so we expect physically close galaxies to be preferentially aligned with each other (known as the Intrinsic-Intrinsic (II) correlation). \\citet{hiratas04} noted an additional negative correlation between foreground galaxies shaped by a particular gravitational potential and background galaxies which are lensed by the same potential (known as the Gravitational-Intrinsic (GI) correlation) which can be of greater magnitude than the II term. After the alignment of galaxy ellipticities from linear response to the gravitational potential % was proposed as an effect % by \\citet{catelankb01} the study was put on a firm analytic footing by the introduction of the Linear Alignment (LA) model by \\citet{hiratas04}, hereafter HS04. This approach, in which the orientation of galaxies responds linearly to the large-scale gravitational potential in which they form, has become the standard model when including the effects of Instrinsic Alignments (IA). Correlated shears of galaxies have been observed at low redshift, where they can be attributed to intrinsic alignments, by Brown et al. 2002, and have been measured in galaxies selected to be physically close in \\citet{mandelbaumea06,hirataea07,Okumura:2008bm,brainerdea_2009} and \\citet{mandelbaumea06}. % Number density - shear correlations are easier to observe since the random galaxy ellipticity only enters the calculation once, and this has been constrained in \\citet{mandelbaumea06,hirataea07,Okumura:2008,mandelbaumea09} and \\citet{joachimiea_megazlrg}. \\cite{hirataea07} noted that the scale dependence of the signal they measured was better matched to theory if the non-linear matter power spectrum was used in the LA instead of the linear power spectrum implied in HS04. \\cite{bridleandking} then used the non-linear matter power spectrum in their cosmological forecast calculations. This approach has been called the Non-Linear Alignment (NLA) ansatz. \\citet{schneiderb09} attempted a more motivated solution for assigning power to the IA model at small scales through the use of the halo model of galaxy clustering. By design their halo model reproduced the LA model at large scales. The II contribution to intrinsic alignments can be taken into account in cosmological analyses by removing galaxies with small physical separation from the analysis e.g. \\citet{kings02,kings03}; \\citet{heymansh03}; \\citet{takadaw04}. This was attempted in a real analysis of COSMOS data in \\citet{schrabbackea09} by removing the autocorrelation tomographic bin. An extension of this method for GI was suggested in \\citet{king05} and developed to a sophisticated level in \\citet{joachimi_schneider_2008,joachimi_schneider_2009}. Alternatively a model may be assumed for all the intrinsic alignment contributions, and the free parameters can be marginalised over, as demonstrated at the Fisher matrix level in \\citet{bridleandking,bernstein_2008,joachimi_bridle_2009,MGpaper1,MGpaper2} and demonstrated on real data in \\citet{kirk_bs_2010}. The redshift evolution of the IA contributions in the LA model was found to be incorrect due to a mistake in HS04 in the conversion between the primordial potential and the matter power spectrum. This was corrected in a new version of HS04, issued as \\citet{hiratas10_posterratum}, hereafter HS10. In addition there is an ambiguity in HS04 as to which cosmological epoch is responsible for the ``imprinting'' of galaxy intrinsic alignments. Most starkly the question is, are IAs frozen in at some redshift of formation or do they evolve with the growth of structure, particularly nonlinear clustering on small scales? \\citet{blazek2011} note this as an issue in the HS04 approach and estimate its effect on the GI amplitude as of order 20\\%. \\citet{MGpaper1}, \\citet{MGpaper2} implement one physically motivated solution to this question (as well as including the redshift evolution correction of \\citealt{hiratas10_posterratum}) which we adopt in this work. The result of this evolution in the treatment of IAs is that much useful theoretical and observational work has been conducted using an incorrect implementation of the LA model, using often unjustified treatments of the small scale seeding of galaxy IAs. In this work our aim is to present basic results for the most up to date implementation of the LA model for IAs. We present angular power spectra for components of the shear-shear ($\\epsilon\\epsilon$), position-position ($nn$) and position-shear ($n\\epsilon$) observables as well as the reduction in constraining power in measuring dark energy caused by a robust treatment of IAs and the biasing of cosmological parameters that results from ignoring IAs or treating them using an old model. Many of these results reproduce previous work in the literature which was conducted using an old implementation of the LA model. Specifically we reproduce Fig. 1 from \\citet{joachimi_bridle_2009} for the new implementation of the LA model which we hope will act as a reference for those wishing to apply it in the future. Furthermore we investigate the ability of a flexible parameterisation of intrinsic alignments to compensate for using the old LA model compared to the latest version we summarise here. This paper is organised as follows: In section \\ref{sec:LA_now} we describe the most up to date implementation of the LA model of IAs, detailing our formalism for shear-shear ($\\epsilon\\epsilon$), position-position ($nn$) and position-shear ($n\\epsilon$) correlations. Section \\ref{sec:bias} presents the bias on cosmological parameter estimation caused by mistreatment of IAs for the latest model, the old model and an intermediate model; we then conclude in section \\ref{sec:conclusions}. In an appendix we summarise the history of the LA model and present the major differences between the latest implementation and the previous widely used version. ", "conclusions": "\\label{sec:conclusions} Galaxy IAs are the most important physical systematic in the study of cosmic shear. As increasing volumes of data become available from WGL surveys more interest is being paid to their correct treatment. Much of this work has been dominated by the LA model, originally introduced in HS04 (and the NLA extension which we refer to as HS04NL). This original implementation was subsequently corrected in \\citet{hiratas10_posterratum} and detailed attention paid to the treatment of non-linear clustering in \\citet{MGpaper1,MGpaper2}. However, much of the existing literature has been produced using the uncorrected HS04NL model. In this paper we have provided a brief explanation of the evolution of, and context surrounding, the LA model for IAs, highlighting the most important differences between HS04NL and the latest implementation. We have calculated the angular power spectra of the cosmic shear observables, correcting the implicit mistake of \\citet{joachimi_bridle_2009}, which was based on HS04, which we hope will act as a new reference for those interested in applying IAs to their cosmic shear analysis. The main motivation for the study of IAs is the measurement of unbiased constraints of key cosmological parameters. We show that the new LA implementation significantly reduces the impact of IAs, and hence the bias, but that the effect is still very significant, producing a bias at the tens of $\\sigma$ level when we assume perfect knowledge of IAs. If we did know the IA signal perfectly then we could produce unbiased measurements by simply subtracting the IA signal from our measured cosmic shear. In practice it is useful to parameterise our ignorance of the true IA signal through a set of nuisance parameters which are marginalised over to produce weaker but hopefully unbiased cosmological estimates. We show that a robust grid of 130 nuisance parameters for IAs and magnification uncertainties, % allowed to vary in scale and redshift, effectively removes the bias due to assuming an incorrect IA model. The same effect holds when our observables are extended to include shear-shear, position-position and shear-position correlations. The extra observables increase constraining power so that we are able to produce unbiased constraints on $\\epsilon\\epsilon+n\\epsilon+nn$ which recover 40\\% of the constraining power of the highly biased constraint from $\\epsilon\\epsilon$ alone when the wrong IA model is assumed. The LA model has allowed a firm foothold on the study of IAs to develop over the last decade. We have detailed an updated implementation of the most used IA model and its physical motivation, showing a reduction in biasing of cosmological constraints. In addition we reiterate that a robust nuisance parameter model can control the biasing due to IAs for either the old or new implementations. Together, these results should give us confidence that the IA effect is under control as we begin to analyse the first data from large cosmic shear surveys. However our somewhat pessimistic approach carries the cost of reduced constraining power. Future work, focused on accurate simulation and measurement of IAs is sure to provide more detail on the physical mechanisms responsible for the initial intrinsic ellipticity distribution and its evolution. This better knowledge of IAs will improve our ability to model them and reduce our dependence on brute-force marginalisation over nuisance parameters. As such the marginalised constraints which we present in Fig.~\\ref{fig:bias_oldnew} may be a worst-case scenario. With better knowledge of IAs, better constraints on cosmology will be possible." }, "1112/1112.3943_arXiv.txt": { "abstract": "We report the petrology, O isotopic composition, and Al-Mg isotope systematics of a chondrule fragment from the Jupiter-family comet Wild 2, returned to Earth by NASA's Stardust mission. This object shows characteristics of a type II chondrule that formed from an evolved oxygen isotopic reservoir. No evidence for extinct $^{26}$Al was found, with ($^{26}$Al/ $^{27}$Al)$_0$ $<$ 3.0$\\times$10$^{-6}$. Assuming homogenous distribution of $^{26}$Al in the solar nebula, this particle crystallized at least 3 Myr after the earliest solar system objects---relatively late compared to most chondrules in meteorites. We interpret the presence of this object in a Kuiper Belt body as evidence of late, large-scale transport of small objects between the inner and outer solar nebula. Our observations constrain the formation of Jupiter (a barrier to outward transport if it formed further from the Sun than this cometary chondrule) to be more than 3 Myr after calcium-aluminum-rich inclusions. ", "introduction": "High-temperature objects---fragments of chondrules and calcium-aluminum-rich inclusions (CAIs)---have been discovered in the collection of samples returned by NASA's Stardust mission from comet Wild 2 (e.g. \\citet{nak08}, \\citet{mat10}). The diversity of the Wild 2 samples shows that this comet does not consist entirely of unaltered nebular material rich in presolar grains and interstellar amorphous silicates \\citep{bro06}, nor is comet Wild 2 close in composition to a known class of meteoritic material or chondritic-porous interplanetary dust particles (e.g. \\citet{ogl10}). The presence of chondrules and CAIs in a Jupiter-family comet requires a transport mechanism between the hot inner solar system, where these objects likely crystallized, and the scattered disk, where they were eventually accreted into the comet \\citep{dun04}. Large-scale transport in the disk must have occurred before the formation of Jupiter because the growing embryo of Jupiter strongly suppressed outward transport across its orbit \\citep{cie09}. A nascent Jupiter efficiently accreted material that diffused into its orbital gap, effectively creating a barrier to outward transport of high-temperature objects if its orbit was farther from the Sun than where these objects formed. Measurements of the decay products of short-lived radionuclides in high-temperature objects in the Stardust collection will yield an earliest time when this large-scale transport occurred. It is then possible to constrain the timing of the formation of Jupiter, which because of its large mass controlled much of the dynamics of planetesimals forming in the young solar system (e.g \\citet{wal11}). ", "conclusions": "Since CAIs are believed to have formed when the Sun was a class 0 or class I protostar, Iris formed late in the evolution of the solar nebula, at a time when $\\sim$95\\% of the scattered disk is thought to have cleared \\citep{her01}. The residual disk material may have coated Jupiter-family comets with a ``late veneer'' enriched in inner solar system material \\citep{ogl10}, and this may explain significant differences between the Stardust sample, which sampled Wild 2 coma material ejected by jets that entrained near-surface dust \\citep{bel10}, and chondritic porous interplanetary dust particles, which probably sample the bulk of Kuiper Belt comets \\citep{nes10}. These analyses are for a single object in the Stardust collection. However, Iris is not unique or even particularly unusual in the suite of particles returned from comet Wild 2. \\citet{nak08} identified four chondrule-like objects in the Stardust samples which are more $^{16}$O-enriched than Iris (Figure \\ref{oiso}) and olivines that are less Fe-rich (Fo$_{79\\text{--}80}$, Fo$_{95}$, Fo$_{91}$). At least one chondrule fragment was also identified during the Stardust preliminary examination \\citep{zol06}. All of these objects show igneous textures similar to Iris, though Iris appears to have formed in a more oxidizing environment. Recent measurements by \\citet{jos11} show two fragments of Fe-rich olivine (Fo$_{62\\text{--}67}$, Fo$_{58\\text{--}61}$), that also have O isotopic composition close to Iris. Chondrule-like objects from Stardust show a broad range of isotopic and mineralogical compositions; a subset of these objects could be genetically related to Iris. The formation time of Iris is most consistent with the late-forming chondrules in CR chondrites. Chondrules from CB chondrites also formed relatively late (as determined by $^{207}$Pb--$^{206}$Pb measurements), likely by a giant impact in the early solar system \\citep{kro05}. The parent bodies of both CR and CB chondrites were scarcely heated \\citep{sco06}, similar to Wild 2 \\citep{bro06}. However, type II chondrules like Iris (and similar Stardust fragments) are very rare in CR and CB meteorites: $<$1\\% of all chondrules \\citep{wei93,kro02}. The Iris mesostasis is Fe-poor and Na-rich compared to mesostasis in type II chondrules in unequilibrated CR chondrules, and the Iris olivine is Ca- and Al-rich relative to olivine in CR type II chondrules \\citep{ber11}. Although Iris is similar to chondrules from CR and CB meteorites in that it formed relatively late, Iris and similar Stardust fragments are unlikely to have originated from the CR- or CB-chondrite-forming region. The Fe-rich olivines in Iris must have formed in an environment with oxygen fugacity higher than the typical redox conditions of the steady-state solar nebula \\citep{kro00}. Shocks in the outer solar nebula, beyond the water-snow line at $\\sim$5 AU \\citep{cyr98}, were rich in water vapor \\citep{cie03} which could have formed a chondrule like Iris near the current orbit of Jupiter. However, the predicted composition of chondrule olivine in such a shock is strongly peaked between Fo$_{76}$ and Fo$_{89}$, Fo$_{70}$ olivine and lower is predicted to be $\\sim$3\\% \\citep{fed08} of all chondrule olivine generated in the shock. Therefore we conclude that Iris was probably not created in an outer-nebula shock, but likely formed in the inner solar nebula, from material with relatively high Fe/Mg or in a region of high oxygen fugacity \\citep{jon90}, more than 3 Myr after CAIs. Recent measurements of oxygen isotope variations in the rim of a CAI from the CV3 chondrite Allende \\citep{sim11} indicate that these early objects experienced circulation in the solar nebula. A Type C CAI found in the Stardust samples was constrained to have crystallized at least 1.7 Myr after the onset of CAI formation (assuming a homogeneous nebular reservoir of canonical ($^{26}$Al/$^{27}$Al)$_0$), though it likely experienced a complex history \\citep{mat10}. Additionally, high-temperature components (CAIs and chondrules) do not appear to be scarce in the Stardust samples. These observations provide strong evidence of a dynamic early solar system transporting material between the inner and outer nebula. The formation of Iris in the inner nebula requires it to be transported to the scattered disk at $\\sim$35 AU \\citep{dun04,tir09} where it was incorporated into comet Wild 2. This transport could have occurred outside the plane of the disk (e.g. \\citet{cie07}), by diffusion (e.g. \\citet{cuz03}), aerodynamic lofting and radial drift (e.g. \\citet{cie08}), or outward advective flows (e.g. \\citet{hug11}). Jupiter's growing embryo would open a gap in the disk \\citep{bat03} which would prohibit outward transport \\citep{cie09} if Iris was created in an inner-nebula shock (unless Jupiter had migrated inside of 2 AU and was closer to the Sun than Iris when Iris formed \\citep{wal11}). Iris likely formed in an event prior to Jupiter's formation, such as in an inner-nebula spiral shock that existed before Jupiter opened a gap in the disk \\citep{bol05}. Therefore, our measurements set a constraint on the formation time of Jupiter of at least 3 Myr after CAI formation making it unlikely that Jupiter formed early (e.g., by disk instability \\citep{bos01}). Our constraint on the formation time of Jupiter is consistent with arguments based on the requirement to accrete asteroids (with constituent radiometric-dated chondrules and CAIs) before Jupiter grows large enough to inhibit accretion ($>$3--5 Myr, \\citet{sco06}). Jupiter was estimated to form $\\sim$3.3 Myr after the onset of planetesimal fragmentation in the main belt by \\citet{bot05}, also consistent with the outward transport of Iris. The formation of chondrules in planetesimal bow shocks caused by Jovian resonances \\citep{wei98} requires Jupiter to form $\\sim$1 Myr after CAIs, a scenario which is disallowed by our measurements." }, "1112/1112.4614_arXiv.txt": { "abstract": "The Fermi Large Area Telescope (Fermi LAT) provides long term systematic monitoring observations of the gamma-ray emission from blazars. The variability properties and the correlation with other wavelength bands are important clues for the evaluation of blazar models. We present results from timing and multiwavelength correlation analysis and discuss differences between blazar classes. ", "introduction": "More than 60 AGNs, almost all blazars, were detected by EGRET on Compton GRO, which established these sources as a powerful class of gamma-ray emitters [1]. In its first year of operation Fermi LAT has increased the number of known gamma-ray blazars by a factor of 10. Even more importantly the instrument is mapping the full sky every three hours, which allow regular monitoring of these sources on time scales from hours to years. These monitoring observations now form an important part of many ongoing efforts to study the variability and multiwavelength properties of blazars. Such observations contains information about the relative origin of different spectral components and about dynamical and radiation processes in blazar jets. ", "conclusions": "" }, "1112/1112.1031_arXiv.txt": { "abstract": "The $\\lya$ emission has been observed from galaxies over a redshift span $z \\sim 0 - 8.6$. However, the evolution of high-redshift $\\lya$ emitters (LAEs), and the link between these populations and local galaxies, remain poorly understood. Here, we investigate the $\\lya$ properties of progenitors of a local $L^{*}$ galaxy by combining cosmological hydrodynamic simulations with three-dimensional radiative transfer calculations using the new $\\art$ code. We find that the main progenitor (the most massive one) of a Milky Way-like galaxy has a number of $\\lya$ properties close to those of observed LAEs at $z \\sim 2 - 6$, but most of the fainter ones appear to fall below the detection limits of current surveys. The $\\lya$ photon escape fraction depends sensitively on a number of physical properties of the galaxy, such as mass, star formation rate, and metallicity, as well as galaxy morphology and orientation. Moreover, we find that high-redshift LAEs show blue-shifted $\\lya$ line profiles characteristic of gas inflow, and that the $\\lya$ emission by excitation cooling increases with redshift, and becomes dominant at $z \\gtrsim 6$. Our results suggest that some observed LAEs at $z \\sim 2-6$ with luminosity of $\\La \\sim 10^{42-43}~\\ergs$ may be similar to the main progenitor of the Milky Way at high redshift, and that they may evolve into present-day $L^{*}$ galaxies. ", "introduction": "The $\\lya$ emission from young galaxies can be a powerful probe of the early universe \\citep{Partridge67, Charlot93}. Recent narrow-band deep imaging surveys using large-aperture telescopes have detected a large number of $\\lya$ emitting galaxies, or $\\lya$ emitters (LAEs), at redshifts $z \\gtrsim 3$ \\citep[e.g.,][]{Hu96, Cowie98, Steidel00, Malhotra04, Taniguchi05, Kashikawa06, Shimasaku06, Iye06, Hu2006, Gronwall07, Ouchi08, Hu2010, Ouchi10, Lehnert10}. By combining $\\lya$ emission with broad-band continuum, multi-wavelength observations are beginning to address the physical properties of these high-redshift LAEs \\citep[e.g.,][]{Gawiser06, Gronwall07, Lai07, Nilsson07, Pirzkal07, Lai08, Ouchi08, Pentericci09, Ono10A, Ono10B, Hayes10, Finkelstein11, Nilsson11, Acquaviva11}. It has been suggested that these objects are mostly compact, young galaxies with low metallicity. In addition, \\citet{Ouchi08} studied the evolution of equivalent widths (EWs) and the characteristic $L^{*}_{\\lya}$ with redshift from $z \\sim 3$ to $z\\sim 6$, and found that the mean EW increased with redshift, while the $L^{*}_{\\lya}$ did not change significantly. More recently, \\citet{Ciardullo11} studied the evolution of luminosity function (LF) and EW from $z=2.1$ to $3.1$, and found that $L_{*}$ increases from $10^{42.3}~\\rm \\ergs$ at $z=2.1$ to $10^{43}~\\rm \\ergs$ at $z=3.1$. \\citet{Blanc11} studied the $\\lya$ properties of LAEs in the redshift range $z=1.9-3.8$ from the Hobby Eberly Telescope Dark Energy Experiment (HETDEX) Pilot Survey, and showed that the median $\\lya$ escape fraction ($\\fesc$) was $\\sim 29 \\%$, and it does not evolve significantly with redshift. On the other hand, \\citet{Hayes11} suggested that $\\fesc$ monotonically increases between redshift 0 and 6, which implies that high-z galaxies tend to be LAEs. While high-redshift LAEs have been studied with large samples in the redshift range of $z \\sim 2.2-6.6$, there is only a limited number of observations on LAEs at $z \\lesssim 1$. Some local star-forming galaxies have been studied by various wavelengths and show a complex structure of $\\lya$ and UV continuum \\citep{Hayes07, Ostlin09}. \\citet{Atek09} showed that the $\\fesc$ of local LAEs have a large dispersion, ranging from $\\sim 3$ to $100~\\%$. In addition, \\citet{Deharveng08} studied a sample of 96 local LAEs at $z=0.2-0.35$ from UV space telescope {\\it GALEX}, and found that these LAEs have similar EW distribution as those at $z=3.1$. Recently, \\citet{Cowie10} have studied $z \\sim 0.3$ LAEs with a larger sample, and showed that these LAEs are more compact, and have lower metallicity than UV-continuum selected galaxies at the same redshift. In addition, \\citet{Finkelstein09} suggested, from fitting of spectral energy distributions (SEDs), that low-z LAEs are significantly more massive and older galaxies than their high-z counterparts. One of the important issues in galaxy evolution is how high-redshift LAEs evolve into galaxies in the local universe. \\citet{Gawiser07} suggested, from clustering analysis, that most $z=3.1$ LAEs evolve to present-day galaxies of $\\lesssim 2.5~ L^{*}$, unlike other populations which typically evolve into more massive galaxies. Moreover, \\citet{Guaita10} indicated that LAEs at $z=2.1$ were building blocks of present-day $L^{*}$ galaxies such as the Milky Way (MW). However, the link between high-redshift LAEs and local galaxies, and the probability of these LAEs evolving into present-day $L^{*}$ galaxies are not well constrained from observations. In order to address these questions, one may use the Milky Way as a local laboratory. Moreover, since $\\lya$ emission has been detected from the most distant galaxies, understanding of the $\\lya$ properties of the Milky Way progenitors will provide an important clue to the formation of early galaxies. To date, there are only a limited number of theoretical studies on this important issue \\citep[e.g.,][]{Salvadori10, Dayal11b}. Both \\citet{Salvadori10} and \\citet{Dayal11b} focused on MW progenitors at $z\\sim 6$ constructed from semi-analytical merger trees and a cosmological smoothed particle hydrodynamics (SPH) simulation, respectively. They both used the same analytical prescription of $\\lya$ emission in which the intrinsic $\\lya$ luminosity scales linearly with the star formation rate \\citep{Dayal08}. However, because the $\\lya$ properties depend sensitively on a number of factors, such as the scattering and propagation of the photons in the inhomogeneous medium, the dust content of the gas, the ionization structure, the UV continuum, and the photon escape fraction. Such a complicated process can only be probed by comprehensive $\\lya$ radiative transfer calculations combined with realistic simulation of galaxy formation. As we will show in this work, our detailed $\\lya$ modeling on a high-resolution cosmological simulation produce a number of $\\lya$ properties such as the luminosity functions at different redshifts in good agreement with observations. Moreover, in order to investigate the evolution of LAEs, we need to study the progenitors of the MW at different redshifts systematically, not just at a specific time. In this paper, we investigate the $\\lya$ properties of MW progenitors over a wide redshift range of $z \\sim 0 - 10$, by combing cosmological SPH simulation of a MW-like galaxy from Zhu et al. (in preparation) with 3D RT calculations using the newly improved $\\art$ code by \\cite{Yajima11A}. Our hydrodynamic simulation includes important physics of both dark and baryonic matter, and has high resolutions to track the formation history of the MW. Our RT calculations include both $\\lya$ resonant scattering and continuum emission, and are done on an adaptive-mesh refinement grid, which covers a large dynamical range and resolves the small-scale structures in high-density region. Interstellar dust is also taken into account to accurately estimate the $\\fesc$ of $\\lya$ photons and UV continuum, and the EWs. The paper is organized as follows. We describe our cosmological simulation in \\S2, and the RT calculations in \\S3. In \\S4, we present results of the $\\lya$ properties of MW progenitors from redshift 10 to 0, which include the $\\lya$ surface brightness, $\\lya$ luminosity, $\\fesc$, EW and line profile. In \\S5, we discuss the dependence of $\\fesc$ on physical properties, LAE fraction in our galaxy sample, $\\lya$ escaping angle and the contribution from excitation cooling to $\\lya$ emissivity, and we summarize in \\S6. ", "conclusions": "\\subsection{Dependence of $\\lya$ Properties on Galaxy Properties} \\label{sec:fesc} \\begin{figure*} \\begin{center} \\includegraphics[scale=0.9]{fesc_all.eps} \\caption{Dependence of $\\lya$ properties on various physical properties of a galaxy. {Upper~:} The relation between $\\lya$ escape fraction and halo mass, SFR and metallicity. The different color represents the different redshift, as indicated in the color bar. The dash lines are least-absolute-deviations fittings. {Lower~:} The relation between $\\lya$ luminosity and halo mass, SFR and metallicity. } \\label{fig:fesc_all} \\end{center} \\end{figure*} \\begin{figure} \\begin{center} \\includegraphics[scale=0.4]{frac.eps} \\caption{ Number fraction of LAEs in our galaxy sample. The green, blue and red filled circles represent the fraction at $\\lya$ luminosity threshold, $10^{40}, ~10^{41}$, and $10^{42}~\\ergs$, respectively. } \\label{fig:frac} \\end{center} \\end{figure} As shown in previous Sections, the $\\lya$ properties vary significantly in different galaxies. Here we explore the dependence on a number of physical properties of a galaxy. Figure~\\ref{fig:fesc_all} shows the dependence of escape fraction $\\fesc$ (top panels) and $\\lya$ luminosity $\\La$ (bottom panels) on the galaxy mass, SFR, and metallicity Z. We apply a least-absolute-deviations fitting to the data using a power-law function, ${\\rm log} Y = \\alpha {\\rm log}X + \\beta$. The mass dependence of $\\fesc$ has a large dispersion, but from our fitting, $\\alpha \\sim - 0.02$ and $\\beta \\sim -0.17$, which suggests that $\\fesc$ roughly decreases with the total mass, consistent with the results of \\citet{Laursen09b}. At $M \\sim 10^{10-11}~\\Msun$, the $\\fesc$ is mostly constant at $\\sim 10 - 30~\\%$. In contrast, $\\fesc$ is more tightly correlated with the SFR, with $\\alpha \\sim -0.08$ and $\\beta \\sim -0.41$. At high SFR, dust can be enriched quickly by type II supernovae, and can effectively absorb the $\\lya$ photons. In addition, galaxies with high SFR have more hydrogen gas. The gas decreases the mean free path of $\\lya$ photons, resulting in the increase of the dust optical depth which reduces the escape fraction. In addition, the $\\fesc$ decreases with metallicity, $\\alpha \\sim -0.35$ and $\\beta \\sim -0.65$. Since the dust content linearly increases with metallicity in our model, the $\\lya$ photons can be absorbed effectively by gas with high metallicity. This trend is consistent with observational indication by \\cite{Atek09} and \\cite{Hayes10}. On the other hand, the luminosity $\\La$ has different relationships with these properties from the $\\fesc$. The $\\La$ is also roughly correlated with the mass, $\\La \\simeq 10^{37.7}\\times\\Mtot^{0.38}$, with a large dispersion. Only massive galaxies with $\\Mtot \\gtrsim 10^{11}~\\Msun$ have the $\\lya$ luminosity of $\\La \\ge 10^{42}~\\ergs$. This is consistent with suggestions from clustering analysis of observed LAEs at $z=2-3$ \\citep[e.g.,][]{Gawiser07, Guaita10}. In our model, the massive galaxies at $z=2-3$ evolve into $L^{*}$ galaxies at $z=0$. Hence, our results support the suggestion by \\cite{Gawiser07} that the observed LAEs with $\\La \\gtrsim 10^{42}~\\ergs$ at $z=2-3$ are likely progenitors of local $L^{*}$ galaxies. The $\\La$ has the tightest correlation with SFR among the properties investigated here: $\\La \\simeq 10^{41.7}\\times \\rm SFR^{0.53}$ In the literature, a simple linear relation is commonly used, with $\\La \\;(\\ergs) = 1.1 \\times 10^{42}\\times \\rm{SFR} \\;(\\Msunyr)$, assuming that $\\La / L_{\\rm H\\alpha} = 8.7$ (case B). However, our result suggests that the relation between $\\La$ and SFR becomes somewhat shallower due to the dependence of $\\fesc$ on SFR. Finally, the emergent $\\La$ does not show a strong dependence on metallicity, $\\La \\simeq 10^{41.1}\\times (Z/Z_{\\odot})^{-0.27}$. This is due to the fact that, although the intrinsic $\\La$ increases with halo mass (so does SFR and metallicity), the $\\fesc$ decreases with metallicity, so $\\La$ of higher-metallicity galaxies is suppressed by dust absorption. We should point out that the large scatter in the correlations in Figure~\\ref{fig:fesc_all} may be due to the small volume of our simulation and the small number of our galaxy sample. In addition, as we discuss in Section~\\ref{sec:limit}, a number of limitations of our model, such as the simplified ISM model and insufficient resolutions, may contribute to uncertainty in these relations. Moreover, the luminosity scaling relations may change under some specific detection limits. We will study these relations in detail with improved model and simulations in future work. \\subsection{Redshift Dependence of LAE Fraction} The number fraction of LAEs ($\\fa$) in our sample is shown in figure~\\ref{fig:frac}. The detection limit of $\\lya$ varies in different surveys. At high redshifts ($z \\gtrsim 3$), the LAE detection in most of observations has been confined to $\\La \\gtrsim 10^{42}~\\ergs$. Here, we derive the $\\fa$ with three $\\La$ thresholds, $\\La \\gtrsim 10^{40}, ~10^{41}, ~10^{42}~\\ergs$ with EW of $\\gtrsim 20~\\A$. The $\\fa$ with $\\La \\gtrsim 10^{40}, 10^{41}~\\ergs$ rapidly increases from $z=0$ to $\\sim 5$, and then remains nearly constant with higher values $\\gtrsim 0.8$ at $z \\gtrsim 5$. The trend is roughly similar to the SFR history (Figure~\\ref{fig:sfr}). Since the $\\La$ is tightly correlated with the SFR (Figure~\\ref{fig:fesc_all}), the number of galaxies with $\\La \\gtrsim 10^{40}, ~10^{41}$ increases at $z \\sim 0 - 4$. On the other hand, although the SFR decreases at $z \\gtrsim 4$, the $\\fesc$ increases due to low metallicity. Hence, the $\\fa$ does not decrease at $z \\gtrsim 4$. Meanwhile, the $\\fa$ with $\\La \\gtrsim 10^{42}~\\ergs$ is nearly constant, and shows $\\sim 2 - 10~\\%$. Since the SFR tightly correlates with $\\La$, and it roughly increases with the galaxy mass, some massive galaxies can be LAEs with $\\La \\gtrsim 10^{42}~\\ergs$. In addition, the $\\fa$ of LAEs having intrinsic $\\La \\gtrsim 10^{42}~\\ergs$ change with cosmic star formation history. However, the $\\fesc$ decreases around the phase of SFR peak, and therefore suppresses the $\\fa$. On the other hand, at lower redshift, the observations indicate that number density of LAEs decreases by some factors \\citep[e.g.,][]{Cowie10}. The discrepancy may come from the difference in density field and the small box in our simulation. Our initial condition is a somewhat special one which is focused on a MW-size galaxy, and the zoom-in simulation region is $\\sim 5^3~ h^{-3} \\Mpc^{3}$. Therefore, our simulation cannot reproduce the global statistics in observations. Moreover, the LAEs fraction having $EW > 25~\\A$ in this work shows $\\sim 44\\; \\%$ at z=4 and $\\sim 100\\; \\%$ at $z = 6$, which is somewhat higher than the LAE fraction in LBG sample \\citep{Stark10, Stark11, Pentericci11, Schenker12, Ono12}. However, in observation, the LAE fraction increases with decreasing UV brightness. Most of our model galaxies at $z \\gtrsim 3$ are fainter than the detection threshold in the LBG observation. Since the number of galaxies brighter than the threshold of LBG observation is quite small (less than ten), we need a larger sample covering a wide mass range to verify the model of LAEs. In addition, although some LAEs have been observed with UV continuum, and hence categorized as LBGs, it is inadequate to study LAEs from LBG-only sample, because a large fraction of LAEs may have UV continuum under the detection limit of current observations. We will address the general properties such as luminosity function, EW distribution and clustering systematically by using a set of uniform simulations with mean density field in larger volumes in future work. \\subsection{The Viewing-angle Scatter of Escaping $\\lya$ Photons} \\label{sec:angle} \\begin{figure} \\begin{center} \\includegraphics[scale=0.45]{fescmap.eps} \\caption{The ``viewing-angle scatter'' of escaping $\\lya$ photons depends strongly on galaxy morphology and orientation. Shown here is the $\\lya$ escape probability in a irregular galaxy progenitor at z=3.1 (top panel) and the spiral MW galaxy at z=0 (bottom panel). The color bar indicates the probability per unit $\\rm d\\phi, dcos\\theta$. There is no clear direction in the irregular galaxy, but in the disk galaxy, the $\\lya$ photons escape in a preferred direction normal to the disk. } \\label{fig:fescmap} \\end{center} \\end{figure} Despite their high metallicity, a fraction of galaxies at low redshift $z \\lesssim 1$ show high escape fraction $\\fesc$ of $\\lya$ photons (Figure~\\ref{fig:fesc_all}). We find that the escaping angle of the $\\lya$ photons depends strongly on the galaxy morphology and orientation, a phenomenon we dub as the ``viewing-angle scatter''. % Disky objects seen edge on can be hundred times fainter than the same objects seen face on. In a galaxy which has a gas disk, the $\\lya$ photons escape in a preferred direction normal to the disk, but there is no clear escaping direction in compact or irregular galaxies without a gas disk. We demonstrate this effect in Figure~\\ref{fig:fescmap}. We first estimate the normal direction to the gas disk according to the total angular momentum of the gas, and set $\\theta = 0^{\\circ}$ along this direction. In a galaxy with irregular morphology such as the main progenitor at $z = 3.1$, there is no clear preferred escaping angle, as illustrated in the top panel of Figure~\\ref{fig:fescmap}. However, in a spiral galaxy with rotationally supported gas disk such as the MW galaxy at z=0 in our simulation, the escaping angle is strongly confined to $\\rm{cos\\theta \\simeq \\pm 1}$, corresponding to $\\theta \\simeq 0^{\\circ}$ or $180^{\\circ}$, as shown in the bottom panel of Figure~\\ref{fig:fescmap}. This is due to the fact that the $\\lya$ photons have the minimum optical depth along the normal direction to the gas disk. More than $60~\\%$ $\\lya$ photons escapes to the direction of $\\rm |cos~\\theta| \\lesssim 0.5$. Generally the $\\lya$ flux from our model galaxies can scatter around the mean value typically by a factor of ten just from different orientations. As illustrated in Figure~\\ref{fig:img}, most galaxies in our simulation have highly irregular shapes at high redshift due to accretion and gravitational interaction. At z=0, a number of them evolve into spiral disks. The ``viewing-angle scatter'' explains why we see high $\\lya$ escape fractions in a number of low-z galaxies, and the fact that $\\lya$ is detected in a large number of face-on spiral galaxies in the nearby universe \\citep[e.g.,][]{Cowie10}. \\subsection{Contribution of Excitation $\\lya$ Cooling} \\label{sec:exc} \\begin{figure} \\begin{center} \\includegraphics[scale=0.4]{exc.eps} \\caption{ The fraction of excitation cooling $\\lya$ to the total intrinsic $\\lya$ luminosity as a function of redshift. The red filled circles indicate the values of the main progenitor, while the blue filled circles represent the median value of the galaxy sample indicated with black open circles. The error bars show the quartiles. } \\label{fig:exc} \\end{center} \\end{figure} There are two major mechanisms to generate $\\lya$ emission, the recombination of ionizing photons and the collisional excitation of hydrogen gas. However, the relative contribution between the two mechanisms is not well understood. From our calculations, we find that the contributing fraction of excitation $\\lya$ emission to the total intrinsic $\\lya$ luminosity increases with redshift, as shown in figure~\\ref{fig:exc}. In our cosmological simulation, galaxy evolution is accompanied by cold, filamentary gas streams with temperature $T\\sim 10^{4-5}~\\rm K$, which penetrate deep inside dark matter halos (Zhu et al. in preparation, Yajima et al. in preparation), a phenomenon also reported by other groups \\citep{Katz03, Keres05, Keres09, Birnboim03, Dekel06, Ocvirk08, Brooks09, Dekel09}. Such cold gas can efficiently produce the excitation $\\lya$ cooling photons \\citep{Dijkstra09, Faucher09, Goerdt10}. At higher redshift, galaxies form through more efficient gas accretion and more frequent merging event. As a result, the contributing fraction increases with redshift, and becomes dominant at $z \\gtrsim 6$. This excitation mechanism does not depend on the stellar radiation, and can therefore produce high $\\lya$ EWs. We find that the EWs of LAEs increases significantly at $z \\gtrsim 6$, reaching $\\gtrsim 10^3~\\A$ at $z \\sim 10$. This is larger than the upper-limit of EW, $240~\\A$, which considers only stellar sources assuming a Salpeter IMF with solar abundance of metallicity \\citep{Charlot93}. Although the upper-limit increases with decreasing metallicity, it was suggested that top-heavy IMF like Pop III stars are needed for making $\\rm EW > 400~\\A$ \\citep{Schaerer03, Raiter10}. However, even though Salpeter-IMF is used in this work and the stellar metallicity is mostly $Z / Z_{\\rm \\odot} \\gtrsim 10^{-3}$, we find that the EW can be higher than the upper-limit by the efficient excitation $\\lya$ emission. On the other hand, the $\\lya$ line is strongly damped by IGM correction at $z \\gtrsim 6$ \\citep{Haiman02, Laursen11, Dayal11b}, which can result in a lower EW. The suppression by IGM highly depends on the inhomogeneous ionization structure around LAEs \\citep[e.g.,][]{McQuinn07, Mesinger08, Iliev08}. We will address the detectability of high-redshift LAEs and EW after IGM correction by running large-scale $\\lya$ RT in IGM in future work. \\subsection{$\\lya$ Luminosity Functions} \\label{sec:LF} The simulation box in this work is too small to study global statistics directly. As a rough estimate, we use the luminosity -- halo mass correlation we find above may be used to construct $\\lya$ luminosity functions (LFs) at different redshift when combined with halo mass functions from large-box, general cosmological simulations. For example, at $z = 3.1$, we divide all galaxies in the snap shot by the halo mass with 0.25 dex, and fit to the median value of each bin, this gives a correlation of $\\La ~({\\ergs})= 10^{32.94} \\times (M_{\\rm halo}^{0.79} / \\Msun)$. We then use this to convert the halo mass function of \\citet{Sheth99} to the $\\lya$ LF. Figure~\\ref{fig:LF} shows the resulting $\\lya$ LFs in comparison with observations at redshift $z = 3.1, 5.7$, respectively. The red solid curves are LFs above a detection threshold of $S_{\\rm Ly\\alpha} = 10^{-18}~\\rm ergs \\; s^{-1} \\; cm^{-2} \\; arcsec^{-2}$ \\citep{Ouchi08, Ouchi10}, while the red dashed lines represent LFs from total luminosity (counting all escaped photons without a flux cut). While the un-filtered LFs seem to agree with observations of \\cite{Gronwall07} and \\cite{Ouchi08}, the filtered ones are significantly off. The difference comes from the reduction of $\\La$ and $\\fesc$ due to the flux cut. Moreover, the dispersion in the luminosity -- halo mass relation at different redshift may cause a large scatter in the LFs. This plot suggests that the current simulation in this work is not suitable to study a large galaxy population and its statistical properties, because there are too few observable LAEs. Moreover, as discussed earlier, the predicted $\\lya$ properties may be affected by a number of numerical and physical limitations of our model. For example, the one-phase model currently used in the present work may underestimates the density of cold hydrogen gas, and hence underestimates the $\\lya$ flux. We will study the $\\lya$ LFs at different redshift in a forthcoming paper with the improved $\\art$ which incorporates a two-phase ISM model, and a general simulation with mean overdensity in a larger volume (Yajima et al, in preparation). \\begin{figure} \\begin{center} \\includegraphics[scale=0.4]{LF_wSB.eps} \\caption{ Derived $\\lya$ luminosity functions from our model in comparison with observations. The red, solid lines are filtered LFs above a detection threshold of $S_{\\rm Ly\\alpha} = 10^{-18}~\\rm ergs \\; s^{-1} \\; cm^{-2} \\; arcsec^{-2}$ \\citep{Ouchi08}, while the red, dashed lines represent unfiltered ones. The black dashed lines are the Schechter function at $z=3.1$ and $5.7$ derived from LAEs observation by \\citet{Ouchi08}, and the black dotted line is from observation at $z=3.1$ by \\citet{Gronwall07}. } \\label{fig:LF} \\end{center} \\end{figure} \\subsection{Limitations of Our Model} \\label{sec:limit} As demonstrated above, our model is able to explain a number of observed properties of LAEs at different redshift. However, we should point out that our current simulations suffer from a number of major limitations which may affect the predicted $\\lya$ properties. \\begin{itemize} \\item In the current work, we use a one-phase ISM model, which considers the average density and temperature of the gas. Such a model likely underestimates the density of cold hydrogen gas, which may lead to significant underestimate of the $\\lya$ emission coming from cold ($\\sim 10^4$~K), dense gas. On the other hand, such a model also underestimates the amount of dust associated with cold molecular gas, which likely results in underestimate of absorption of $\\lya$ photons by gas and dust. We will investigate the $\\lya$ RT and ionization structures in a two-phase ISM model in a forthcoming paper. \\item The absorption and transmission of IGM are not taken into account in the RT calculations. As discussed in the previous section, these two effects may suppress the $\\lya$ flux and change the line profiles. \\item The simulations do not have sufficient resolutions to resolve dense regions and outflow, which requires a high spatial resolution of $\\sim \\rm pc$ \\citep[e.g.,][]{Fujita09}. It is a challenge for cosmological simulations to resolve both the inflow gas from large scales of $\\sim \\rm Mpc$ and the outflow from pc-scale star forming regions. For a simulation with a box of 100 Mpc like the one we have, this requires a large dynamical range over eight orders of magnitude, which is beyond the scope of our current work. \\item The simulation box in this work is too small to study a large galaxy population, as well as effects of environment on galaxy properties and their evolution. One needs uniform simulations in large volumes in order to systematically investigate the formation and evolution of L* galaxies. \\end{itemize} Finally, we stress once again that caution should be taken when comparing directly the results from our calculations to data from a given survey, because, as discussed above, the observed $\\lya$ properties depend sensitively on a number of factors, including galaxy properties, viewing angle, model parameters, and observational threshold." }, "1112/1112.0037_arXiv.txt": { "abstract": "We present the results of numerical simulations of wave propagation and jet formation in solar atmosphere models with different magnetic field configurations. The presence in the chromosphere of waves with periods longer than the acoustic cutoff period has been ascribed to either strong inclined magnetic fields, or changes in the radiative relaxation time. Our simulations include a sophisticated treatment of radiative losses, as well as fields with different strengths and inclinations. Using Fourier and wavelet analysis techniques, we investigate the periodicity of the waves that travel through the chromosphere. We find that the velocity signal is dominated by waves with periods around 5~minutes in regions of strong, inclined field, including at the edges of strong flux tubes where the field expands, whereas 3-minute waves dominate in regions of weak or vertically oriented fields. Our results show that the field inclination is very important for long-period wave propagation, whereas variations in the radiative relaxation time have little effect. Furthermore, we find that atmospheric conditions can vary significantly on timescales of a few minutes, meaning that a Fourier analysis of wave propagation can be misleading. Wavelet techniques take variations with time into account and are more suitable analysis tools. Finally, we investigate the properties of jets formed by the propagating waves once they reach the transition region, and find systematic differences between the jets in inclined field regions and those in vertical field regions, in agreement with observations of dynamic fibrils. ", "introduction": "Chromospheric wave propagation has been an extensively studied, but poorly understood, subject in solar physics for some time. One puzzle has been the presence of propagating waves with periods on the order of 5 minutes or more. Such waves were observed by, e.g., \\citet{Orrall1966} and \\citet{Giovanelli+etal1978}; this was considered surprising since the acoustic cutoff period in the chromosphere, above which waves should not be able to propagate, is on the order of 200~s. Such long-period propagation was later found to be widespread, generally occurring wherever the local magnetic field is strong. \\citet{Lites+etal1993} show an example of observations of neighboring internetwork and network regions that have a marked difference in their Fourier spectra. It was realized \\citep{Michalitsanos1973,Bel+Leroy1977,Suematsu1990} that, since magnetoacoustic waves in a strongly magnetized medium are restricted to propagating along field lines, the effective gravity (i.e., the component of gravity along the magnetic field) would be reduced in magnetic regions. Since the cutoff frequency (in an isothermal atmosphere) is given by \\begin{equation} \\nu_{ac}=\\frac{\\gamma g_{eff}}{4\\pi c_s}, \\label{cutoffeq} \\end{equation} where $c_s$ is the sound speed and $\\gamma$ is the ratio of specific heats, the cutoff frequency would also be lower in regions of strong inclined field, potentially allowing 5-minute ($\\sim$3~mHz) waves to propagate. This hypothesis has since been tested in a number of increasingly advanced numerical simulations, e.g. \\citet{DePontieu+etal2004}, \\citet{Hansteen+etal2006}, and \\citet{Heggland+etal2007}. All have found that it is an effective mechanism for transmitting 5-minute power through the chromosphere; some models have suggested that the leakage of 5-minute waves into the chromosphere can also explain the presence of 5-minute oscillatory signal in the corona \\citep{DePontieu+etal2005}. One criticism of this explanation has been that not all observations of 5-minute propagation are in regions of obviously inclined field. An alternative explanation has therefore been suggested, originally by \\citet{Roberts1983} and later developed and tested by \\citet{Centeno+etal2006,Centeno+etal2009} and \\citet{Khomenko+etal2008}, in which changes in the radiative relaxation time associated with small scale magnetic structures are responsible for increasing the cutoff period. This mechanism has also been demonstrated to enable propagation of 5-minute oscillations, even in vertical magnetic structures. However, the underlying basis of this theory and the related simulations is a highly simplified energy equation in which radiative losses are approximated by Newton's law of cooling (\\citeauthor{Khomenko+etal2008} \\citeyear{Khomenko+etal2008} and references therein). A related subject that has been studied extensively is the generation and propagation of chromospheric jets such as spicules (type I and II), macrospicules, surges, fibrils, and mottles. A variety of models have been proposed over the years (see \\citeauthor{Sterling2000} \\citeyear{Sterling2000} for a review), and although it remains unclear whether the many types of jets are different manifestations of the same underlying physical phenomenon or not, there are at least strong indications that the jets known as dynamic fibrils are driven by shock waves traveling through the chromosphere \\citep{Hansteen+etal2006,DePontieu+etal2007,Heggland+etal2007,Martinez+etal2009a}. Many of these jets have lifetimes around 5~minutes, and so the waves driving them need to have been channeled into the upper chromosphere via one of the processes outlined above. In this paper, we present the results of two-dimensional simulations of wave propagation from the convection zone to the transition region and corona. The simulations include a sophisticated treatment of radiative losses and heat conduction, and study the effects of different magnetic field geometries and strengths on the propagation of waves through the chromosphere. We also look at the jets that these waves produce once they reach the transition region, and perform a statistical comparison of the jets produced in a model with vertical field and in one with inclined field. In \\textsection{}2 we describe the different simulations and the code. \\textsection{}3 contains an analysis of wave propagation and periodicities in the various models, and the results are discussed in \\textsection{}4. \\textsection{}5 looks at jet formation and properties, and a summary and conclusions follow in \\textsection{}6. ", "conclusions": "\\begin{figure*} \\plottwo{f24a.ps}{f24b.ps} \\caption{Fourier spectra of the vertical velocity taken from our simulations (left panels) and the Doppler velocity derived from synthesized spectral lines (right panels). The synthesized spectra have been smoothed over 2 arcseconds. In all panels, we show the spectrum of case B in the center (between $x=16.5$~Mm and $x=33$~Mm), flanked on both sides by the spectrum of the non-magnetic case A. This figure can be compared with Figure 4 in \\citet{Lites+etal1993}.} \\label{litesrad} \\end{figure*} The main result from our simulations is that 5-minute waves are able to propagate through the chromosphere in regions where the magnetic field is inclined and sufficiently strong. In regions where the field is vertical or weak, the velocity field is dominated by waves with periods of around 3 minutes. As mentioned in the introduction, a model which uses the radiative relaxation time, rather than the inclination of the magnetic field, as the mechanism for increasing the cutoff period, has been proposed, and has been demonstrated to work in numerical simulations \\citep{Khomenko+etal2008}. However, it relies on a simple Newtonian cooling model for approximating radiative losses. Our simulations use an advanced and realistic method for computing the radiative losses \\citep{Gudiksen+etal2011} that includes all the important mechanisms involved in radiative losses in the photosphere and chromosphere, and as such it is considerably more realistic than a simple Newtonian cooling model. While the radiative relaxation time model expects 5-minute propagation above all strong small-scale magnetic structures, regardless of the field inclination, our simulations show 3-minute propagation above the central region of flux concentrations, and 5-minute propagation in inclined-field regions to the sides. This is exactly the result predicted by the field inclination model, but is in conflict with the predictions of the radiative relaxation model. Our conclusion, based on the simulations presented here, is therefore that the radiative relaxation model is not effective when a more realistic energy equation is considered. While large-scale regions of homogeneous inclined field, like our cases D and E, represent idealizations of the conditions at the edge of plages, our case C represents realistic conditions for an isolated magnetic element or pore. In this model, the magnetic field is largely vertical in the photosphere and above the transition region, while the field expansion leads to significant inclination where it is needed the most, in the chromosphere. This can then account for the 5-minute propagation, but we expect 3-minute propagation in the center of the flux concentration. Is this supported by observations? The answer is in fact not very clear. \\citet{Lites+etal1993} show an example of a network region which has very little power at frequencies above 4~mHz over most of its area. There is, however, a pronounced spike of higher-frequency power located in the center of the network ($x=30$~Mm in their Figure 4). The authors attribute this to noise from seeing, which it may very well be; it is visible at frequencies up to 20~mHz, and a study of the coherence spectra also lends credence to this explanation. However, the signal at frequencies of 5-6~mHz in this spike is much stronger than that at higher frequencies, and indeed is of the same magnitude as the (real) signal at those frequencies in the neighboring quiet regions. It is notable that such a spike, with a spatial extent of 1-2~Mm, is in fact exactly what we find in our case C. The central spike observed by \\citeauthor{Lites+etal1993} could thus be real. In order to make more direct comparisons to their results, which were based on Doppler shifts in observations of the Ca~{\\sc ii}~H line at $3968.49$~\\AA{} and the Fe~{\\sc i} line at $3966.82$~\\AA, we have computed synthetic spectra in these lines from our simulations using the non-LTE radiative transfer code MULTI \\citep{Carlsson1986}. The Ca line is treated in full non-LTE, while the Fe line, which is formed in the wing of the Ca line, is included in the same computation using additional localized opacity and source function terms based on LTE. The calculations have been performed column by column, i.e., neglecting any radiative interaction in the $x$-direction. The computed spectra have been smoothed over $2''$, which appears to be close to the effective resolution of the observations in \\citet{Lites+etal1993}, and we have then calculated Doppler shifts from these smoothed spectra. It should be noted that Ca~{\\sc ii}~H has a very complex line profile, usually with several emission peaks within the deeper absorption line. In a dynamic atmosphere, particularly in the presence of strong shocks, these emission peaks can become very large and are often asymmetric (so called bright grains), and the line can undergo central reversal as well. Defining and identifying the line center of such a line is a non-trivial task. We have used a method that finds the center of the region where the intensity is below a certain threshold above the minimum intensity. This method generally gives acceptable results, but can not be expected to correspond directly to the velocity at any one given height in the simulations. The power spectra of the calculated Doppler velocities are shown in the two right-hand panels of Figure~\\ref{litesrad}; the two left-hand panels show the power spectra of the vertical velocity taken directly from our simulation data, at heights corresponding to the approximate formation heights of the lines. Since the observations of \\citet{Lites+etal1993} covered a larger area than our simulation boxes, we show a combination of the spectra from two of our simulations in the figure. In the center of all panels, between $x=16.5$~Mm and $x=33$~Mm, we show the spectrum of our case B, representing network conditions. On both sides, from $x=0$~Mm to $x=16.5$~Mm and from $x=33$~Mm to $x=49.5$~Mm, we show the spectrum of our case A, representing conditions in the weakly magnetized internetwork. This figure should be compared with Figure 4 in \\citet{Lites+etal1993}. While our synthesized Ca power spectrum (upper right) is not a perfect match to the observations of \\citeauthor{Lites+etal1993}, there are many similarities. Of particular note is the difference in the dominant frequencies between the non-magnetic regions on both sides and the network region in the center. The network (case B) is dominated by lower frequencies, particularly in the area between $x=22$~Mm and $x=29$~Mm, where the dominant frequencies are 3-$3.5$~mHz. The internetwork (case A) has a more scattered spectrum, but the dominant frequencies are mostly between 5 and 7~mHz. Furthermore, the network has significant power at very low frequencies, below 3~mHz. These effects are also found by \\citeauthor{Lites+etal1993}. This power at low frequencies is not found in the simulation velocity at $z=1$~Mm (upper left), and although the velocity spectrum and the Ca spectrum are similar in many ways, there are also several differences. These differences are partly due to the smoothing applied to the Ca data, partly due to the difficulty of defining a meaningful Doppler shift of the highly complex Ca line profile, and partly due to the fact that the line is formed over a range of heights rather than at one given height, and this height range can also vary with both horizontal position and time. Using the Ca Doppler velocity as a proxy for atmospheric velocity on the real Sun can therefore be misleading. The velocity spectrum of the Fe line (lower right) shows a very good correspondence with the velocity at $z=0.25$~Mm in the models (lower left). This line has a much simpler profile and is formed in a region with few strong shocks affecting local conditions. The dominant frequency is 3~mHz in both the network and the internetwork, though there is also some power at 5~mHz in most locations. \\citeauthor{Lites+etal1993} also find that the power in this line is in the 3-5~mHz band, although most of it is between 3 and 4~mHz. Like us, they find no significant difference between the spectra in the internetwork and in the network in this line. More recent observations related to the question of wave periodicity have been performed by \\citet{Centeno+etal2009}, \\citet{deWijn+etal2009}, and \\citet{Stangalini+etal2011} \\citep[see also][]{Jefferies+etal2006}. \\citeauthor{Centeno+etal2009} used the Tenerife Infrared Polarimeter of the German Vacuum Tower Telescope at the Observatorio del Teide, with a seeing-limited spatial resolution of $1''-1\\farcs{}5$. They then found propagating 5-minute waves in the chromosphere above a facular region. The photospheric magnetic field as determined from Stokes inversions was within 20$^{\\circ}$ of the vertical. \\citeauthor{deWijn+etal2009}, using the Solar Optical Telescope on {\\it Hinode} with a resolution of $0\\farcs{}16$, found 3-minute signal in the center of a plage region, but more 5-minute propagation towards the sides in the direction of the expanding field. \\citeauthor{Stangalini+etal2011} used a combination of data from {\\it Hinode} and the IBIS instrument at the Dunn Solar Telescope, with an estimated average resolution of $0\\farcs{}36$. They found propagating 5-minute waves along the inclined field on the edges of a pore, and some power in 3-minute oscillations at the center. They also found both 5-minute and 3-minute propagation, though with more power in the 5-minute band, in a nearby region with small magnetic elements where they estimate that the chromospheric magnetic field is close to vertical, based on a force-free field extrapolation. The results of our simulations are in agreement with \\citet{deWijn+etal2009}, and with the propagation patterns observed by \\citet{Stangalini+etal2011} around their pore. In the more vertical magnetic structures observed by \\citet{Stangalini+etal2011} and \\citet{Centeno+etal2009}, we would expect more power in the 3-minute band than in the 5-minute band based on the results of our simulations, if indeed the field is mainly vertical and the flux tubes do not move around very much. There are, however, several possible mechanisms that could explain the 5-minute dominance and resolve this apparent difference. For one, all flux tubes naturally expand with height, and this expansion creates a region of inclined field (as illustrated, on a large scale, by our case C). Thus, even if the field is close to vertical in the photosphere, there will be regions between the photosphere and the chromosphere where the field at the edges of the flux tubes is inclined, and the long-period waves can propagate there. \\citet{Centeno+etal2009} do study coherence spectra to look for signs of a possible horizontal shift in the signal as a result of propagation along inclined fields, and find good coherence between the photospheric and chromospheric signal, but our results show that the field does not need to be inclined throughout the photosphere and chromosphere in order to enable 5-minute wave propagation. A few hundred km along the edges of an expanding flux tube may be enough, and any horizontal shift could then be less than one resolution element ($1''-1\\farcs{}5$ in their data). In the observations of \\citet{Stangalini+etal2011}, the photospheric field is not uniformly vertical, and a force-free extrapolation is not a very good approximation in the chromosphere. Furthermore, although there is more power in the 5-minute signal, they also find significant signal at periods around 3 minutes in the region with smaller magnetic elements. We believe that our model, where field inclination is the dominant mechanism for allowing long-period wave propagation, is compatible with these findings. Flux tube movement and limited resolution may also be partially responsible for the relative dominance of 5-minute power in the observations of \\citet{Centeno+etal2009}. In case B, we found that the strongest 3-minute power appeared above flux tubes that undergo little horizontal motion, while 5-minute power was found in inclined field regions at the edges of flux tubes. If the flux tubes move around, both the 5-minute and 3-minute power will be spread out and one would not see a clear distinction between the (average) flux tube center and the sides in a Fourier analysis. The flux tube center also covers a rather small area at any given time, and this makes the related 3-minute waves difficult to observe in low-resolution data. In the higher-resolution data of \\citet{Stangalini+etal2011}, regions of 3-minute propagation are found, and this could possibly be because the flux tubes at these locations move around less. We would encourage observers to look for differences in the periodicity of oscillations at the center and edges of flux tubes in future high-resolution datasets. A third possibility is that heating may play a role. The temperature structure of the chromosphere is in general not well known. In particular, the real Sun may have more magnetic heating of the upper chromosphere than our 2D models. Higher temperature would reduce the cutoff frequency and allow 5-minute waves to propagate more easily. Yet another suggestion, as mentioned by \\citet{deWijn+etal2009}, is that the field may be twisted. The waves could then travel along field lines that are everywhere inclined with respect to the local vertical, but without significant horizontal displacement. Such field twist is a 3D effect and can not be tested in our 2D simulations, but should be considered in later work. A different point, that we have already mentioned in our analysis, is also worth making: although a Fourier analysis can be a powerful tool, it has some important limitations. In order to achieve sufficient spectral resolution, one usually needs time series on the order of one hour. The solar atmosphere, however, is dynamic on timescales of minutes. Atmospheric conditions can and do change, and the Fourier transform is not well suited for picking up such changes. At least in areas where the general signal is weak, non-recurring events can end up dominating the power spectrum (see Figure~\\ref{bampang456} for an example; there are several other examples in the dataset). In such cases, the Fourier analysis does not say all that much about the general conditions at that location. This can be particularly dangerous because the eye is naturally drawn to peaks in the power spectrum. Also, there can be times when local conditions are notably different from the time average, and these can be correlated with changes in the signal (e.g., Figure~\\ref{vertpwrudang185}). A wavelet analysis, which takes time variations into account, or at least a careful examination of the time series, is essential for identifying the actual processes going on in the atmosphere. We have presented several simulations investigating the periodicity of waves propagating through the chromosphere. We find that waves with periods of around 5 minutes can propagate in regions where the magnetic field is strong and inclined, including at the edges of flux tubes. In regions where the magnetic field is weak or vertical, we find primarily 3-minute waves; this also applies above vertical flux tubes and in the center of strong expanding flux tubes. These results indicate that field inclination is critical to the propagation of long-period waves. Where the flux tubes undergo significant horizontal motion, both the 5-minute and the 3-minute power is spread out and the distinction is not as clearly visible. Since we have included an advanced treatment of radiative losses in our simulations and find 3-minute propagation above vertical field regions, we conclude that variation in the radiative relaxation time is not an effective mechanism for increasing the cutoff period. Our simulations are in agreement with the results of recent high-resolution {\\it Hinode} observations. We have also studied the jets produced by these waves once they reach the transition region. We find systematic differences between the jets produced in a model with mostly vertical field, and in a model with mostly inclined field. The results are in agreement with observations of dynamic fibrils. It is also important to point out, for the purpose of future analyses of wave propagation and periodicity, whether from simulations or observations, that the Fourier analysis can be misleading and hide important information about the state of the medium. This is because the solar atmosphere is dynamic and changes on timescales much shorter than one hour, which is often the minimum timescale needed to achieve sufficient spectral resolution in a Fourier analysis. A wavelet analysis takes variations in time into account and can be an invaluable tool for figuring out what processes are important for the dynamics." }, "1112/1112.2497_arXiv.txt": { "abstract": "{} {We investigate the surface nitrogen content of the six magnetic O stars known to date as well as of the early B--type star $\\tau$ Sco. We compare these abundances to predictions of evolutionary models to isolate the effects of magnetic field on the transport of elements in stellar interiors.} {We conduct a quantitative spectroscopic analysis of the sample stars with state-of-the-art atmosphere models. We rely on high signal-to-noise ratio, high resolution optical spectra obtained with ESPADONS at CFHT and NARVAL at TBL. Atmosphere models and synthetic spectra are computed with the code CMFGEN. Values of $N/H$ together with their uncertainties are determined and compared to predictions of evolutionary models.} {We find that the magnetic stars can be divided into two groups: one with stars displaying no N enrichment (one object); and one with stars most likely showing extra N enrichment (5 objects). For one star ($\\Theta^{1}$ Ori C) no robust conclusion can be drawn due to its young age. The star with no N enrichment is the one with the weakest magnetic field, possibly of dynamo origin. It might be a star having experienced strong magnetic braking under the condition of solid body rotation, but its rotational velocity is still relatively large. The five stars with high N content were probably slow rotators on the zero age main sequence, but they have surface $N/H$ typical of normal O stars, indicating that the presence of a (probably fossil) magnetic field leads to extra enrichment. These stars may have a strong differential rotation inducing shear mixing. Our results should be viewed as a basis on which new theoretical simulations can rely to better understand the effect of magnetism on the evolution of massive stars.} {} ", "introduction": "The evolution of massive stars is governed by two main physical processes: their mass loss \\citep{cm86} and their rotation \\citep{mm00}. All additional processes directly affecting these properties are thus major contributors to the fate of massive stars. This is the case of magnetism. Since the pioneering work of \\citet{bm97}, it is known that the presence of a large scale surface magnetic field will modify massive stars winds. By deflecting material along the field lines, magnetism affects the wind geometry and structure. Simulations by \\citet{ud02,ud08,ud09} have shown that equatorial overdensities can be created. Material can either fall back onto the stellar surface or be ejected in the magnetic equatorial plane depending on the interplay with rotation. In the case of very strong magnetic fields, observations and simulations show that a rigid magnetosphere strongly affects the wind structure, creating caps of stellar material in the minima of the magneto-rotational potential \\citep{townsend07}. Since the wind flow is deflected by the presence of a magnetic field, the way the star loses its angular momentum is affected. In addition, the internal structure of the star is also affected, forcing the star to follow new evolutionary paths. Given the relevance of massive star evolution to many fields of astrophysics (galactic chemical evolution, interstellar dynamics, supernovae and gamma--ray bursts, nucleosynthesis...), it is crucial to understand the magnetic properties of these objects in order to constrain their influence on the evolution and fate of OB stars. The recent development of powerful spectropolarimeters has revolutionized our view of massive stars magnetism. The first detection of a magnetic field on the Trapezium O--type star $\\Theta^{1}$~Ori~C \\citep{donati02} dates back to less than a decade. This spectroscopically varying and strong X--ray emitting star has revealed a surface kilogauss field organized mostly as a dipole. Subsequent detections have been made on HD~191612 \\citep{donati06}, HD~148937 \\citep{hubrig08,wade11}, HD~57682 \\citep{grunhut09} and HD~108 \\citep{martins10}. Field strengths are of several hundred Gauss to a few kilogauss and the geometry, when constrained, is consistent with a dipole. The only exception is the O supergiant $\\zeta$~Ori~A \\citep{bouret08} for which a more complex field topology was tentatively derived, together with a remarkably weak field of 50--100 G. This raises the question of the origin of magnetism in massive stars. The debate is ongoing. A fossil origin, stable over Myrs and with a rather simple geometry, is supported by some simulations \\citep{brait04,duez10}. In that case, the star retains the original magnetic field of its parental molecular cloud or dynamo field of its convective progenitor. Alternatively, dynamo processes may be at work. In analogy with low mass stars where a convective envelope (or even a fully convective structure) exists, magnetic field might be generated in the convective core of high mass stars and subsequently transported to the surface \\citep{charb01,brun05}. A dynamo may also operate in the radiative zones of massive stars \\citep{spruit02,mcdo04}. The field would be produced by the shear in this region. This idea has been tested in numerical simulations, with contradictory results \\citep{brait06,zahn07}. However, assuming that this dynamo process was able to produce and maintain a magnetic field, evolutionary calculations have been conducted to test the effects of this type of magnetism on the evolution of massive stars \\citep{mm04}. They find that the internal structure is strongly affected, with a completely different rotation profile in the star compared to non-magnetic calculations including rotation. A consequence of this re-organization of the stellar interior is a modification of the transport of angular momentum and of chemical elements. In particular, \\citet{mm05} studied the effects on the surface chemical appearance of the star. They show that the presence of a dynamo generated magnetic field significantly increases the efficiency of mixing, leading to strong nitrogen and even helium surface enhancement. Further analysis of the feedback effects of magnetic braking balanced these results though, highlighting the importance of the type of rotation profile (differential versus solid-body) in the final chemical appearance of the star \\citep{meynet11}. Although still limited, the sample of magnetic O stars is now populous enough to test these predictions. An important side effect of the spectropolarimetric observations is the production of very high S/N ratio optical spectra at high resolution. These are perfectly suited for abundance studies. In this paper, we thus analyze the surface nitrogen content of the magnetic O stars known to date as well as of the B0.2V star $\\tau$ Sco. Our aim is to see if peculiar chemical patterns are observed and if so, to what extent they can be explained by the current generation of evolutionary models. The goal is to identify specific patterns that can only be attributed to the presence of a magnetic field. This is particularly important in the context of explaining the population of N--rich slowly rotating stars in the Magellanic Cloud reported by \\citet{hunter08,hunter09}. Such stars are not explained by rotating evolutionary tracks and have been suggested to be magnetic. The observations are presented in Sect.\\ \\ref{s_obs}. We describe the analysis with atmosphere models in Sect.\\ \\ref{s_mod}. The results are discussed in Sect.\\ \\ref{s_mag}. We show how the chemical properties of the magnetic O stars can be interpreted in terms of evolutionary models with rotation and with or without magnetic field. We finally summarize our conclusions in Sect.\\ \\ref{s_conc}. ", "conclusions": "\\label{s_conc} We have conducted a quantitative spectroscopic analysis of the six magnetic O stars known to date together with the early B star $\\tau$ Sco. Spectra collected with ESPADONS at CFHT and NARVAL at TBL have been used. Atmosphere models computed with the code CMFGEN have been computed. We have determined the stellar parameters and the surface nitrogen abundance. A careful investigation of the uncertainties associated with $N/H$ has been pursued. Our results can be divided into two parts: \\begin{itemize} \\item[$\\bullet$] The surface $N/H$ of $\\zeta$ Ori A -- the star with the weakest magnetic field -- is lower than the predictions of evolutionary models. For that star the presence of the magnetic field probably quenches the transport of chemical elements. The conditions under which such a process can happen are not clear. Evolutionary models accounting for magnetic braking predict no chemical enrichment in solid body rotating stars. But their rotational velocities are also rather low, whereas $\\zeta$~Ori~A appears to have a rather standard \\vsini. \\item[$\\bullet$] The other magnetic O stars, especially the three Of?p stars, display surface nitrogen content consistent with those of non-magnetic O stars with initial \\vsini\\ of 300 \\kms. However, these magnetic O stars were most likely slow rotator on the zero age main sequence (with values of \\vsini\\ of a few tens of \\kms\\ at most). Consequently, they feature larger surface $N/H$ than they should if they did not host magnetic fields as also indicated by their nitrogen excess compared to rotating evolutionary models. \\end{itemize} The presence of a magnetic field in O stars thus appears to produce two types of chemical peculiarities: either an extra enrichment, or the absence of enrichment. Confrontations with evolutionary predictions in the case of dynamo fields indicate that solid body rotation might explain the properties of the non--enriched star, but its relatively normal rotational velocity is a puzzle in that context. For the N--rich magnetic O stars, a fossil field is usually assumed. Our results should be a basis for comparison with future evolutionary calculations including fossil fields since such evolutionary models do not exist at present." }, "1112/1112.6388_arXiv.txt": { "abstract": "Earliest AUGER (the Pierre Auger Observatory) UHECR (Ultra High Energy Cosmic Rays) anisotropy correlated with AGN (active galactic nuclei) within a GZK (Greisen-Zatsepin-Kuz'min cut off) Universe almost fade away. Recent UHECR mass compositions did show a negligible nucleon composition and an UHECR nuclei (light or heavy) signature. Indeed last map miss the Super-Galactic Plane. The absence of UHECR events toward the Virgo cluster, an unique spread clustering of events around Cen-A, our nearest AGN, suggested a He-like nuclei as the main extragalactic UHECR component from Cen A, coexisting with Auger and HIRES (High-Resolution Fly's Eye) composition. Because the light nuclei fragility such He UHECR cannot arrive from Virgo (being too far). Multiplet at twenty EeV along Cen A recently discovered by Auger confirm this interpretation as being foreseen to be indebt to fragments ($D,He^{3}, p$) that had to reach us along the same UHECR. However remaining majority of UHECR clustering are partially correlated with a gamma noise at (1-3 MeV) in Comptel sky, linked to $Al^{26}$ galactic radioactive map as well as to a few TeV gamma (ICECUBE-ARGO) anisotropy maps; rare UHECR triplet are overlapping on Vela TeV anisotropy and other nearest galactic gamma sources (as partially Crab and a Galactic core corona). Therefore UHECR might be also (or mostly) heavy radioactive galactic nuclei as $Ni^{55}$, $Ni^{56}$, $Ni^{57}$ and $Co^{57}$ bent from the sources whose $\\beta$ and $\\gamma$ radioactivity and decay in flight is boosted (by Lorentz factor $\\Gamma_{Ni}\\simeq 10^{9}- 10^{8}$), leading to TeV correlated sky anisotropy. Galactic UHECR signals inside the inner center maybe suppressed by the largest spreading repulsive Lorentz bending forces. More clustering around external galactic plane is nevertheless expected in present and future data. Magellanic Cloud and Magellanic Stream may also rise more and more in UHECR maps (as well as in multiplet signals). Future UHECR clustering might be observed around Cas A and Cygnus by T.A. (Telescope Array). The UHECR spectra cut off may be not an extragalactic GZK feature but just the imprint of a galactic confinement and-or a spectroscopic heavy composition decrease step. ", "introduction": "Ultra High Energy Cosmic Rays, UHECR, Astronomy and nuclear composition are more and more in severe conflict. This contradiction was already inscribed in early apparent discover of a Super Galactic UHECR correlation \\cite{Auger-Nov07} with a first UHECR nuclei composition signature. It was note realistic for UHECR to arrive exactly from Super Galactic plane, as apparent map were suggesting, while being heavy nuclei (and not nucleon). A very heavy UHECR composition being greatly bent by galactic magnetic fields cannot open to any sharp astronomy nor correlate to Super Galactic Plane, nor it can explain easily the unique observed mild UHECR Cen A event clustering. Heavy nuclei as Fe, Ni, Co may however induce large scale anisotropy or spread inhomogeneity (as possibly from a nearby source, Vela). Moreover let us remind that UHECR (if nucleon and if extragalactic) are making photo-pion interacting on cosmic CMBR (Cosmic Microwave Background Radiation) and a consequent EeV gamma and neutrinos secondaries. These expected signals at EeV for gamma and tau neutrinos are more and more constrained being at the edge of AUGER detection, see \\cite{Auger08}. If UHECR are light nuclei they may produce fragment secondaries nuclei or nucleon, as well as gamma-ray and neutrino tails at tens of TeV-PeV because nuclei photo-dissociation; also heavy nuclei, if extragalactic, may lead by photodisintegration on extragalactic flight leading to such TeVs-PeV gamma and neutrino tail. If UHECR are very heavy they maybe bent and bound and confined within their local birth source group. For instance UHECR heavy nuclei maybe bent and confined within Virgo group ; in analogy heaviest UHECR may be bent, crowded and nearly contained within our galaxy, but they cannot produce much photodisintegration secondaries. However \\emph{UHECR radioactive nuclei} as $Ni^{56}$, $Ni^{57}$ and $Co^{57}$ $Co^{60}$, while flying and decaying, may trace their presence by boosted gamma (and positron) tails and beta decay neutrinos. If such UHECR are produced by Super-novae remanent, SNRs (SuperNova Remanents) , or better by their twin micro-quasars jets (in GRB-SN asymmetric jet explosions and late precessing jet model, see \\cite{Fargion1998}) these tails might be observable. The relic gamma radiation (in rest frame of the nuclei around hundred keV) will be boosted by UHECR $E \\geq 6 \\cdot 10^{19} eV$ energy, heavy nuclei huge Lorentz factor ($\\Gamma_{Ni^{55}} \\simeq 10^{9}$), shining at our laboratory system around tens-hundred TeV gamma region, observable at Milagro-ARGO-ICECUBE TeVs CR maps as their correlated anisotropy, see Fig \\ref{fig4-5}. A comparable trace somehow correlated, but at more non relativistic stages, is offered by heavy nuclei (as the $Al^{26}$ at MeV Comptel gamma maps see Fig. \\ref{1r}) in galactic plane \\cite{Fargion2011}. Inspired by this first radioactive gamma-UHECR connection possibility now we extend such understanding for the partial gamma TeV-UHECR connection anisotropy assumed born by boosted nuclei decay, see Fig.\\ref{fig4-5}. As a natural consequence present UHECR cut off may be related not to any real nucleon extragalactic GZK cut-off but to a heavy nuclei spectroscopy step, a confinement in our own galaxy, an nearby anisotropic distances dilution. The same heavy nuclei ejection by nearby Supernova maybe the source of local nuclei anisotropy made by asymmetric SN-jet explosions (see:\\cite{Hansen:SN-Uranium}). \\begin{figure}[htb] \\begin{center} \\includegraphics[width=3.1 in]{Disk-Ellipse1.eps} \\includegraphics[width=3.1 in]{Comptel-1-2-3.eps} \\caption{ Left: The AUGER UHECR event map and two of the three AUGER multiplet clustering toward Cen A; a thin narrow elliptical area and a small disk mark the place. A third Multiplet clustering points toward Large and Small Magellanic clouds and it overlaps on Magellanic stream. Right: last UHECR event map by AUGER where the clustering near Cen A overlap the MeV Comptel gamma (1-3 MeV) map; note connected map beside nearby AGN Cen A, along the nearest pulsar Vela, the Magellanic stream and the galactic core. }\\label{1r} \\end{center} \\end{figure} \\begin{figure}[!t] \\vspace{5mm} \\centering \\includegraphics[width=4.8 in]{ALL-TEV-ARGO-AUGER-HIRES.eps} \\caption{ The AUGER 2010 UHECR (red) and Hires (blue) event map in celestial coordinate on recent (2010) TeV diffused CR map (ARGO-Milagro-ICECUBE) and labels. See \\cite{ARGO},\\cite{Desiati}, and references. The triplet UHECR event clustering toward Vela and the TeV spread activity around is remarkable. The Cen A clustering at the fig. center is the main feature in the UHECR map. ARGO TeV anisotropy born around Crab connect and overlap the UHECR events in nearby Orion TeV region. Note doublet along the galactic plane and the TeV near Cygnus and Cas A regions, where Hires did and we foresee TA may detect UHECR events.} \\label{fig4-5} \\end{figure} Extragalactic UHECR around CenA formed (mostly) by lightest nuclei may explain a partial clustering of events, and the UHECR absence around Virgo. Light nuclei are fragile and fly few Mpc before being halted by photo-disruption \\cite{Fargion2008},\\cite{Fargion2009},\\cite{Fargion2010}. The fragments $He + \\gamma \\rightarrow D+D,D+\\gamma \\rightarrow p+n+ \\gamma , He + \\gamma \\rightarrow He^{3}+n, He + \\gamma \\rightarrow T +p $ may nevertheless trace the same UHECR maps by a secondary clustering at half or even fourth of the UHECR primary energy \\cite{Fargion09a}, \\cite{Fargion2011}. At lower energy (at ten EeV or below) the huge smeared cosmic ray isotropy and homogeneity may hide these tiny inhomogeneity traces. Nevertheless the very recent AUGER EeV map did not show the expected solar Compton Getting anisotropy but an anisotropy in a different region. This result partially disagree with a (widely accepted) extragalactic EeV UHECR component but it favors a more local (galactic) EeV UHECR presence whose smearing is related to the large Lorentz angle bending and whose non-Compton-Getting inhomogeneity is related to nearby sources, as we are also suggesting in present article. ", "conclusions": "$ , $Co^{60}$ radioactive nuclei? } The correlation of UHECR with Cen A, the absence of Virgo, the hint of correlation with Vela and with galactic TeV anisotropy, might be in part solved by an extragalactic lightest nuclei, mainly He, for Cen A , (see Fig.\\ref{1r}) and mostly by heavy radioactive nuclei for other sources mostly in our own galaxy (see Fig.\\ref{fig4-5}). A partial confirm is the predicted \\cite{Fargion09b} ,\\cite{Fargion2011} and observed \\cite{Auger11} multiplet clustering (as Deuterium or proton fragments) at half UHECR edge energy aligned (see Fig.\\ref{1r}): He like UHECR maybe bent by a characteristic angle as large as $\\delta_{rm-He} \\simeq 16^\\circ$; expected proton or D fragments multiplet along tails spread at $\\delta_{rm-p} \\simeq 32^\\circ$ \\cite{Fargion2011},\\cite{Auger11}. Moreover as shown here, UHECR Ni,Co and heavier nuclei too may be deflected by $\\delta_{rm-Ni} \\simeq 18,7^{o}$ for nearby Vela, $\\delta_{rm-Ni} \\simeq 128^{o}$ (or less) for Crab, Gum Nebulae at few kpc away, forcing UHECR toward nearby TeV inhomogeneities. Such TeV gamma anisotropy is made by boosted UHECR decay at hundred keV gamma and beta positrons shining to us at tens TeV. In conclusion therefore we foresee analogous UHECR traces around Cygnus and Cas A (as in TeV map) in growing future TA map. We foresee also crowding and clustering around ($\\mp 18^{o}$) the galactic plane for similar arguments, but not much in galactic center where the stronger magnetic field may spray away the UHECR signals around an external corona ($\\mp 20^{o}$). The UHECR radioactive beta decay in flight may spread as gamma TeV traces as well as $\\nu$ TeVs-PeVs anisotropy whose muon signal is sink into dominant atmospheric muon $\\nu_{\\mu}$ noise. A new spread $\\nu_{\\tau}$ neutrino astronomy, noise free, related to astronomical (parasite oscillated) tau neutrino and its boosted tau (\\emph{mini-double bang} within a 5-50 meter size) in Deep Core or Antares may reveal hundred TeV tau decay (seeing similar PeVs ones in ICECUBE \\cite{Learned}). Also Tau airshowers may rise in Cherenkov beamed air-showers. \\cite{Fargion1999}, \\cite{FarTau} or fluorescence telescopes at higher energies \\cite{FarTau},\\cite{Feng02},\\cite{Auger08}. The discover of such expected Neutrino astronomy may shed additional light on the UHECR nature, origination and mass composition, while opening our eyes to mysterious UHECR sources. Also future gamma TeV maps and UHECR multiplet correlation as foreseen may lead to a more conclusive fit of this unsolved, century old, cosmic ray puzzle. \\\\" }, "1112/1112.6341_arXiv.txt": { "abstract": "{The formation of spectroscopic binaries (SB) may be a natural byproduct of star formation. The early dynamical evolution of multiple stellar systems after the initial fragmentation of molecular clouds leaves characteristic imprints on the properties of young, multiple stars. The discovery and the characterization of the youngest SB will allow us to infer the mechanisms and timescales involved in their formation. Our work aims to find spectroscopic companions around young stellar objects (YSO). We present a near-IR high-resolution (R~$\\sim$~60000) multi-epoch radial velocity survey of 7 YSO in the star forming region (SFR) $\\rho$ Ophiuchus. The radial velocities of each source were derived using a two-dimensional cross-correlation function, using the zero-point established by the Earth's atmosphere as reference. More than 14 spectral lines in the CO $\\Delta\\nu$ = (0-2) bandhead window were used in the cross-correlation against LTE atmospheric models to compute the final results. We found that the spectra of the protostars in our sample agree well with the predicted stellar photospheric profiles, indicating that the radial velocities derived are indeed of stellar nature. Three of the targets analyzed exhibit large radial velocity variations during the three observation epochs. These objects - pending further confirmation and orbital characteristics - may become the first evidence for proto-spectroscopic binaries, and will provide important constraints on their formation. Our preliminary binary fraction (BF) of $\\sim$71\\% (when merging our results with those of previous studies) is in line with the notion that multiplicity is very high at young ages and therefore a byproduct of star formation.} ", "introduction": "\\indent Since the early twenties that there is growing evidence that multiple stars may be the rule and not the exception. This conception is more recentlt supported by observations of \\cite{duq91} who found that the stellar multiplicity among main sequence (MS) stars in the solar neighborhood may be as high as $\\sim$ 60\\%. Since then, more refined surveys probed the dependencies of the properties of stellar systems on physical parameters like age, mass, separation, and environment. These observations are highly relevant, because there is still no comprehensive theory that explains the properties of multiple stars and their relation to star formation. Recently, \\cite{raga10} confirmed that stellar multiplicity decreases with stellar mass (or later spectral type) in the vicinity of the Sun, a finding that was already suggested earlier \\cite[\\textit{e.g.}][]{sieg05,lada06}. Because the stellar population in the solar neighborhood is dominated by late-type M-stars, this implies that about half of the solar-like main sequence stars in the solar neighborhood occur in multiple systems. \\\\ \\indent It is likely that the multiplicity rate of stars is established in the early phases of star formation or during the dynamical evolution that occurs afterwards \\citep{good04, dedo04, bate09}. And it is plausible that both initial and environmental conditions constrain multiple star formation, and play a vital role in explaining their evolution over time, and their relation to the observed MS stellar multiplicity \\citep{mat00,ster03,cone08,ko08,kacz11}. It is therefore essential to probe the stellar multiplicity as early as possible, to gain a full understanding of the formation and the dynamical evolution of multiple systems. \\\\ \\indent \\cite{gh93,lein93,simon95,pat02,bec03} and \\cite{duc04,duc07} all noted differences in the BF of pre-main-sequence (PMS) stars in different SFRs. In particular, the multiplicity of T Tauri stars with ages $\\lesssim$ 10$^6$ yrs, appears to be higher by a factor of 2 in less dense regions such as Taurus compared to denser SFR such as Ophiuchus \\citep[\\textit{e.g.}][]{simon95,bec03,ratz05} or the field \\citep{pat02, duc04}. Toward even younger ages, radio continuum \\citep[\\textit{e.g.}][]{rei04} and near- and mid- Infrared (IR) \\citep{bars05, hai02, hai04, hai06} investigations found additional evidence for an even higher multiplicity among Class I and Flat-spectrum ($\\sim$ 2 - 5 x 10$^5$ yrs) protostars. \\\\ On the other hand, \\cite{mau10} failed to detect companions in a small sample of 5 protostars at separations 50 \\textless $a$ \\textless 5000 AU ($a$ being the semimajor axis) in a millimeter study of self-embedded, young (Class 0) objects ($\\sim$ 10$^4$ - 10$^5$ yrs). Taken together with the observations of \\cite{loon00}, the authors argue that multiplicity in the separation ranges from 100 to 600 AU would only be defined in a later stage of star formation (namely after the Class 0 phase) and that early multiplicity in pristine systems may not be as ubiquitous and primeval. \\\\ \\indent However, all multiplicity studies in the earliest phases of stellar evolution are plagued with relatively small number statistics. In particular, very little is known about companions of embedded protostars at the sub-AU separation scales, a regime impossible to address with conventional imaging techniques. Spectroscopic Binaries (SB) may provide additional important constraints on the star-formation mechanism itself. \\cite{toko02} and \\cite{toko06} estimated the SB frequency in MS stars and inferred that 65\\% of their sample of 165 spectroscopic binaries were members of higher order multiple systems, often found in a hierarchical configuration. Remarkably, the frequency of SB in multiple systems strongly depends on the orbital period of the SB: 96 of the SBs with orbital periods shorther than 3 days are in multiple systems, while this rate drops to 30\\% for SBs with orbital periods longer than 13 days. The higher frequency of spectroscopic system within triple or higher order systems suggests an imprint of the formation mechanism itself. Spectroscopic pairs may have lost their orbital angular momentum owing to the presence of a third body that interacts with the inner binary via Kozai cycles and tidal friction \\citep{kos72, ste05, fabr07}. This interaction has the effect of tightening the spectroscopic binary orbit, while the tertiary would eventually be evacuated to an outer region of the system. The SB may therefore hold an important fossil record of (prevailing) initial conditions during their formation. However, this scenario is not conclusive by itself, and it is not clear at which evolutionary stage this process operates efficiently. Observations are required to better constrain the detailed mechanisms of how tight binaries are formed at very young ages. \\\\ \\indent The first successful PMS SB detections were performed by \\cite{herbig77,hart86,mat89} and \\cite{mat94} in the Taurus and Ophiuchus SFRs with a radial velocity precision limited to $\\sim$1kms$^{-1}$ by optical high-resolution spectroscopy. Since then, improving instrumental capabilities and better radial velocity precision allowed to address PMS sources and the lower mass regimes \\citep[\\textit{e.g.}][]{mel03, cov06, ku06, joe06, joe08, prat07}. Only recently, infrared (IR) high-resolution spectroscopy allowed the search for SB among the youngest, embedded PMS and the very low mass stars \\citep[\\textit{e.g.}][]{prat07,bla07,bla10,fi10b,rice10,cro11}. In particular \\citep[][]{fi10b,bla10,cro11} reached a precision of less than 50 ms$^{-1}$ in PMS stars and very low mass stars. Even the planetary mass regime may be in reach using high-stability, high-precision IR spectroscopy. \\\\ \\indent Our own work aims to find SBs in a sample of 38 Class I/ II protostars in the SFR $\\rho$ Ophiuchus. We performed a high-resolution IR spectroscopic survey to derive precision radial velocities (RV) using the NIR spectrograph CRIRES at the VLT \\citep{Kaeufl04}. We describe our sample, the data analysis methodology and results in Sect. 2. Sect. 3 discusses the RV derived in the context of the $\\rho$ Ophiuchus SFR and its implications on the formation of SB at very young ages in general. In Sect. 4. we draw our conclusions. \\begin{figure}[ht] \\centering \\includegraphics[height=8.0cm]{tel_star.eps} \\caption{Process of telluric signature removal in DET4. The extracted nodded spectra of GSS26 (A) is divided by the spectrum of the telluric standard stars with the same exposure time and observed closest in time (same airmass) (B). Lower panel exhibits 'cleaned' nodded spectrum (C).} \\label{allspec8} \\end{figure} ", "conclusions": "" }, "1112/1112.4491_arXiv.txt": { "abstract": "We explore a class of dark matter models with two dark matter candidates, only one interacts with the standard model sector. One of the dark matter is thermalized with the assistance of the other stable particle. While both stable particles contribute to the total relic density only one can elastically scatter with nuclei, thus effectively reducing the direct detection rate. ", "introduction": "The most natural explanation for the astrophysical and cosmological indications of a large component of dark matter in the Universe is a new weakly interacting massive particle that behaves as cold dark matter. Such dark matter candidates are found in different extensions of the standard model (SM)~\\cite{Bertone:2004pz}. However, dark matter (DM) could be composed of more than one particle, and several multi component DM models have been suggested recently~\\cite{Zurek:2008qg, Profumo:2009tb, Feldman:2010wy, Winslow:2010nk, Batell:2010bp, Liu:2011aa, Adulpravitchai:2011ei}. Interest for multi component DM arose, in particular, from hints for signals in cosmic rays as well as in direct detection experiments corresponding to two completely different DM mass scales. On the one hand, the PAMELA experiment~\\cite{PAMELA} has reported an cosmic ray excess of positrons which would be consistent with a TeV scale DM candidate annihilating mainly into leptons. On the other hand, DAMA~\\cite{DAMA}, CoGeNT~\\cite{Cogent} and CRESST~\\cite{CRESST} have all reported excesses in the direct detection rate that would be compatible with DM around 10 GeV. Even though there is no conclusive evidence that these signals are due to DM, it is interesting to investigate multi component DM models to expand the range of DM models at a time when DM searches in indirect, direct and collider experiments are increasing their sensitivities. Furthermore, replacing the usual R-parity symmetry that guarantees the stability of the lightest R-parity odd particle by an enlarged symmetry group allows not only multiple DM candidates but also new freeze-out mechanisms. In particular, the semi-annihilation mechanism where two DM particles annihilate into another DM particle and a SM particle was proposed in Ref.~\\cite{semiannihilation}. In this paper, we propose a new type of freeze-out mechanism, assisted freeze-out, within the framework of multi component DM models. In this new freeze-out mechanism, one DM candidate can be thermalized only through the assistance of the other stable particle. Thus, the decoupling of one DM particle from the thermal bath is influenced by the other DM particle. Consequently, the relic density of DM is solved by using two coupled Boltzmann equations for two stable fields. In the analysis of the right-handed sneutrino DM~\\cite{arXiv:1105.1652}, a similar situation has been already considered even though this model has only one stable particle. After setting the Boltzmann equations, we construct a simple model with two hidden DM sectors corresponding to two new $U(1)$ gauge symmetries, only one of which interacts with the SM sector. This is achieved through kinetic mixing of the new and standard gauge bosons. The DM particles are assumed to be Dirac fermions. We then show how both particles can contribute to the relic density of DM while only one can scatter elastically on nucleons. Although the direct detection rate tends to be rather high in this model, we show examples, where all constraints can be satisfied, including a case with a DM candidate around 10 GeV. This paper is organized as follows. The basic set-up is presented in section 2. An explicit model is constructed in section 3 and the implications for the relic density of DM as well as for the direct detection rates on nucleons are studied. Section 4 contains our conclusions. ", "conclusions": "We have illustrated with a simple toy model containing two stable dark matter particles, how the assisted freeze-out mechanism worked and could reproduce the measured value for the relic density of dark matter. The main feature of this type of model is that only one of the DM particles is involved in direct detection searches while both contribute to the relic density. In particular, when the DM particle that interacts with SM particles is the subdominant DM component, it is possible to reconcile models with large elastic scattering rates on nuclei with the exclusion bounds of XENON100. Moreover, the lighter particle can be a light DM candidate around 10 GeV as indicated by the DAMA, CoGeNT and CRESST results." }, "1112/1112.3564_arXiv.txt": { "abstract": "Global frequentist fits to the CMSSM and NUHM1 using the {\\tt MasterCode} framework predicted $\\Mh \\simeq 119 \\gev$ in fits incorporating the \\gmt\\ constraint and $\\simeq 126 \\gev$ without it. Recent results by ATLAS and CMS could be compatible with a Standard Model-like Higgs boson around $\\Mh \\simeq 125 \\gev$. We use the previous {\\tt MasterCode} analysis to calculate the likelihood for a measurement of any nominal Higgs mass within the range of 115 to 130~GeV. Assuming a Higgs mass measurement at $\\Mh \\simeq 125 \\gev$, we display updated global likelihood contours in the $(m_0, m_{1/2})$ and other parameter planes of the CMSSM and NUHM1, and present updated likelihood functions for $\\mgl, \\msqr$, \\bmm\\ and the spin-independent dark matter cross section \\ssi. The implications of dropping \\gmt\\ from the fits are also discussed. We furthermore comment on a hypothetical measurement of $\\Mh \\simeq 119 \\gev$. \\bigskip \\begin{center} {\\tt KCL-PH-TH/2011-40, LCTS/2011-21, CERN-PH-TH/2011-305, \\\\ DCPT/11/168, DESY 11-242, IPPP/11/84, FTPI-MINN-11/31, UMN-TH-3023/11, {arXiv:1112.3564 [hep-ph]}} \\end{center} \\vspace{2.0cm} ", "introduction": "\\label{sec:intro} Taking into account the relevant experimental constraints, the CMSSM and NUHM1 predict that the lightest Higgs boson should have couplings similar to those of the Standard Model (SM) Higgs boson~\\cite{Ellis:2001qv,Ambrosanio:2001xb,mc7}, and that it should weigh no more than $\\sim 130 \\gev$~\\cite{mh,FeynHiggs,asbstev}. We recently reported the results of global frequentist fits within the CMSSM and NUHM1 to the first $\\sim 1$/fb of LHC data, also including precision electroweak and flavour measurements and the XENON100 upper limit on elastic spin-independent dark matter scattering~\\cite{mc7}, updating the results of previous global fits by ourselves~\\cite{mc1,mc2,mc3,mc35,mc4,mc5,mc6,mcweb} and others~\\cite{pre-LHC,post-LHC} (see also \\cite{LR}). The results reported in~\\cite{mc7} included likelihood contours in the $(m_0, m_{1/2})$, $(\\tan \\beta, m_{1/2})$ and $(\\MA, \\tan \\beta)$ planes of the CMSSM and NUHM1, as well as $\\Delta \\chi^2$ functions for $\\mgl$, \\bmm, $\\Mh, \\MA$ and sparticle production thresholds in $e^+ e^-$ annihilation. Notable predictions of these global fits included $\\Mh = 119.1^{+3.4}_{-2.9} \\gev$ in the CMSSM and $\\Mh = 118.8^{+2.7}_{-1.1} \\gev$ in the NUHM1 (which should be combined with an estimated theory error $\\Delta \\Mh = \\pm 1.5 \\gev$). These two fits are based solely on the Higgs-{\\it independent} searches including the \\gmt\\ constraint, i.e., they do not rely on the existing limits from LEP~\\cite{Barate:2003sz,Schael:2006cr}, the Tevatron~\\cite{TevHiggs}, or the LHC~\\cite{ATLASHA,CMSHA}. These predictions increase to $\\Mh = 124.8^{+3.4}_{-10.5} \\gev$ in the CMSSM and $126.6^{+0.7}_{-1.9} \\gev$ in the NUHM1 if the \\gmt\\ constraint is dropped. Subsequently, the ATLAS and CMS Collaborations have released their official combination of the searches for a SM Higgs boson with the first $\\sim 1$/fb of LHC luminosity at $E_{\\rm cm} = 7$~TeV~\\cite{ATLAS+CMS}. Impressively, the combination excludes a SM Higgs boson with a mass between 141 and 476~GeV. Most recently, the ATLAS and CMS Collaborations have presented preliminary updates of their results with $\\sim 5$/fb of data~\\cite{Dec13}. These results may be compatible with a SM-like Higgs boson around $\\Mh \\simeq 125 \\gev$, though CMS also report an excess at $\\Mh \\simeq 119 \\gev$ in the $ZZ^*$ channel. We recall that, for low values of $\\Mh$, the SM electroweak vacuum would be unstable~\\cite{unstable}, decaying into a state with Higgs vev $> 10^8 (10^{10}) \\gev$ if $\\Mh = 119 (125) \\gev$, and that a very plausible mechanism for stabilizing the vacuum is supersymmetry (SUSY)~\\cite{ER}. In this paper, we first report the likelihood function for an LHC measurement of $\\Mh$ with a nominal value $\\in (115, 130) \\gev$, incorporating the theoretical error $\\pm 1.5 \\gev$ and an estimate $\\pm 1 \\gev$ of the possible experimental error. In both the CMSSM and NUHM1, this likelihood function is minimized for $\\Mh \\simeq 119 \\gev$ if \\gmt\\ is included, and is contained within the theoretical uncertainty range shown previously as a `red band'~\\cite{mc7}. We then discuss the consequences of combining a measurement of $\\Mh \\simeq 125 \\gev$ (assuming that the current excess will be confirmed with more integrated luminosity) with our previous analysis~\\cite{mc7} of constraints on the CMSSM and NUHM1 including \\gmt. We find that the best-fit values of $m_0$ and $m_{1/2}$ in the CMSSM and NUHM1 are moved to substantially higher values, especially in the case of $m_{1/2}$. We also update our results on the best-fit regions in the $(m_{1/2}, \\tb)$ and $(\\MA, \\tb)$ planes, where we find again the substantial increase in $m_{1/2}$, as compared with our pre-LHC $\\Mh$ results. We present the corresponding one-dimensional likelihood functions for the gluino mass $\\mgl$, an average right-handed squark mass $\\msqr$, the lighter scalar tau mass, $\\mstaue$, as well as in the $(\\mneu{1}, \\ssi)$ plane, where $\\mneu{1}$ is the mass of the lightest neutralino and \\ssi\\ is the spin-independent dark matter scattering cross section. As could be expected, we find larger values of $\\mgl, \\msqr, \\mneu{1}$ and $\\mstaue$ than in our pre-LHC $\\Mh$ fit, and smaller values of \\ssi, though \\bmm\\ is little affected. Since $\\Mh \\simeq 125 \\gev$ is the value that was favoured in the CMSSM/NUHM1 fits omitting the \\gmt\\ constraint~\\cite{mc7}, we also show some results for fits where \\gmt\\ is dropped. In this case, we find that preferred regions of the $(m_0, m_{1/2})$ planes are localized at relatively high values, corresponding to relatively large sparticle masses. Correspondingly, the spin-independent dark matter scattering cross section \\ssi\\ would be relatively small in this case, though again there would be relatively little effect on \\bmm. Finally, we show selected results for a hypothetical measurement of $\\Mh \\simeq 119 \\gev$. ", "conclusions": "The ATLAS and CMS searches for the Higgs boson have already excluded a very large range of masses, with the only remaining windows for a SM-like Higgs boson being in the ranges $\\Mh \\in (115.5, 127) \\gev$ or $> 600 \\gev$~\\cite{ATLAS+CMS,Dec13}. The latter range is disfavoured by precision electroweak data, so attention naturally focuses on the low-mass range. It may or not be a coincidence that this range includes the range $\\Mh \\lsim 130 \\gev$ accessible in simple supersymmetric models such as the CMSSM and NUHM1. Within this range, our previous global fits of these models including \\gmt\\ predicted $\\Mh \\sim 119 \\gev$ if the \\gmt\\ constraint was included in the fit, and $\\Mh \\sim 126 \\gev$ if \\gmt\\ was omitted~\\cite{mc7}. The latest ATLAS and CMS results display an interesting fluctuation at $\\Mh \\sim 125 \\gev$ (i.e.\\ close to the latter result from \\cite{mc7}) and we have combined a hypothetical measurement of $\\Mh = 125 \\gev$ with the global likelihood functions obtained in our previous fits~\\cite{mc7}. As we have shown in this paper, this combination refines our previous predictions for the CMSSM and NUHM1 model parameters within global fits incorporating \\gmt. In particular, the combination prefers a range of larger values of $m_{1/2}$, resulting in larger values of $\\mgl$ and other sparticle masses being preferred, restricting the prospects for discovering supersymmetry at the LHC within these models. The predictions for \\ssi\\ are pushed to higher masses and lower cross sections, particularly in the CMSSM. There are also smaller changes in the predictions for other observables such as \\bmm\\ . We have also shown the analogous CMSSM and NUHM1 fit results for a hypothetical measurement of $\\Mh \\simeq 125 \\gev$ if the \\gmt\\ constraint is omitted. In this case we find a stronger preference for larger values of $(m_0, m_{1/2})$, and correspondingly larger values of $\\tb$ and $\\MA$, as well as larger values of $\\mgl, \\msqr$, potentially lying beyond the reach of the LHC. We have also commented on the potential implications of a hypothetical Higgs discovery at $\\Mh \\simeq 119 \\gev$. Time will soon tell where the LHC experiments are indeed discovering the Higgs boson. However, we have shown that $\\Mh = 125 \\gev$ is a possibility within the CMSSM and NUHM1, although it lies at the upper range of what is possible within the CMSSM or NUHM1, and might suggest reduced prospects for discovering these particular models of supersymmetry at the LHC. Alternatively, it could well be that one should look beyond the frameworks of the models discussed here. \\subsubsection*{Note Added} After acceptance of this paper for publication, we became aware of issues in the implementation of the {\\tt FeynHiggs} code and in the cold dark matter density calculation, which required extra sampling and reprocessing of the NUHM1 parameter space. We are grateful to Nazila~Mahmoudi and Azar Mustafayev for discussions on dark matter density calculations. \\subsubsection*{Acknowledgements} The work of O.B., K.J.D., J.E., J.M. and K.A.O. is supported in part by the London Centre for Terauniverse Studies (LCTS), using funding from the European Research Council via the Advanced Investigator Grant 267352. The work of S.H. is supported in part by CICYT (grant FPA 2010--22163-C02-01) and by the Spanish MICINN's Consolider-Ingenio 2010 Program under grant MultiDark CSD2009-00064. The work of K.A.O. is supported in part by DOE grant DE-FG02-94ER-40823 at the University of Minnesota." }, "1112/1112.1561_arXiv.txt": { "abstract": " ", "introduction": "The Wide-field ASKAP Legacy L-band All-sky Blind surveY (WALLABY)\\footnote{Principal Investigators: Baerbel Koribalski and Lister Staveley-Smith. See www.atnf.csiro.au/research/WALLABY for more details about the survey.} (\\citet{Koribalski_2009}; Koribalski, B., Staveley-Smith, L. et~al., in preparation) is an ambitious project that aims to detect neutral hydrogen to a redshift of $z \\sim 0.26$, across $\\sim 70\\%$ of the sky. It is one of the two top ranked projects that will be carried out using the Australian SKA Pathfinder (ASKAP). WALLABY is possible because of ASKAP's unprecedented $\\sim 30$ sq. degree field-of-view, which is achieved using Phased Array Feeds (PAFs). WALLABY will use all 36 of ASKAP's antennae, but due to limitations on computing resources will only process the inner 30 antennae (with a maximum baseline of 2km) to image the sky with a 30$^{\\prime\\prime}$ synthesised beam and produce datacubes with voxels\\footnote{Voxels are often referred to as pixels when discussing a single channel of a datacube. Technically though these `pixels' are still voxels. For this reason the term voxel is used throughout instead of pixel to aid consistency.} that project to $\\sim$10$^{\\prime\\prime}$ on the sky. The high spatial resolution is complemented by an anticipated spectral resolution of 3.86 km s$^{-1}$. ASKAP spectral datacubes will therefore cover a large area of the sky to high resolution, which results in very large datacubes containing at least 2048 x 2048 x 16 384 voxels. WALLABY will consist of $\\sim$ 1200 of these large datacubes. The size and number of these datacubes renders manual source finding unfeasible. The performance of the automatic source finder used by WALLABY will determine how many (\\HI) galaxies are found by WALLABY. The majority of source finders in existence use intensity thresholding to find sources. {\\sc SExtractor} \\citep{1996A&AS..117..393B}, {\\sc SFind} \\citep{2002AJ....123.1086H} and {\\sc Duchamp} \\citep{2008glv..book..343W,Duchamp2} are good examples of source finders based on intensity thresholding. Conceptually, intensity threshold source finders check every pixel (voxel) in an image (datacube) to see if the pixel (voxel) value is sufficiently extreme that it's unlikely to be noise. Once all of the source pixels (voxels) have been identified, they are combined into objects. The various intensity threshold source finders differ in how they estimate the noise, set a threshold for identifying source pixels (voxels), pre-process the image (datacube) to improve the source finder results and the manner in which they create objects from source pixels (voxels). All intensity threshold source finders share an inherent limitation though. Consider an arbitrary source in a spectral datacube. Improved spatial and spectral resolutions result in the source occupying more voxels in the datacube. Dispersing the source's signal over more voxels means that it contributes less to the flux value of each voxel that it occupies. This makes it harder for an intensity threshold method to detect the source. Using a simple model this effect is illustrated in Figure \\ref{fig:VoxelSN}, where the maximum voxel $S/N$ of an object with an integrated $S/N$ of 5 is plotted for various asymmetries. The maximum voxel $S/N$ is calculated to be $S/N_{integrated} \\times \\beta / \\sqrt{n}$, where $\\beta$ describes the asymmetry of the object's flux distribution and $n$ is the number of voxels. By overlaying the minimum expected size (in voxels) of WALLABY sources on Figure \\ref{fig:VoxelSN}, we can assess the impact of this inherent limitation on WALLABY. The neutral hydrogen detected in emission is warm, which gives it an intrinsic amount of dispersion. We will assume that any real WALLABY source extends over at least 3 channels. We will also assume that in every channel, a source occupies at least 3x3 voxels for a 30$^{\\prime\\prime}$ synthesised beam and 10$^{\\prime\\prime}$ voxels). Galaxies rarely lie at the middle of a voxel though, so a more realistic minimum is a grid of 4x4 or 5x5 voxels in every channel. It is expected that most galaxies detected by WALLABY will be unresolved or at most marginally resolved (Duffy, A., Meyer, M. \\& Staveley-Smith, L. 2011, in preparation), so we also consider a 7x7 grid of voxels in every channel to account for marginally resolved, off-centre galaxies. After multiplying by the minimum number of channels to obtain the expected minimum size of WALLABY galaxies in voxels, the minimum size for these different grids is overlaid in Figure \\ref{fig:VoxelSN}. This demonstrates that off-centre and/or marginally resolved galaxies will be difficult to detect with a basic implementation of an intensity thresholding source finder, unless the flux is asymmetrically distributed. Figure \\ref{fig:VoxelSN} also illustrates that this effect is amplified in 3-D datasets such as future WALLABY datacubes. For example, in a 2-D image the 7x7 and 5x5 vertical lines would approximately lie at the position of the 4x4 and 3x3 vertical lines. \\begin{figure}[ht] \\begin{center} \\includegraphics[width=0.98\\linewidth]{VoxelSN_5.ps} \\caption{The maximum voxel $S/N$ of an object (with an integrated $S/N = 5$) plotted against the number of voxels comprising the object, n. The various lines correspond to different asymmetries in the distribution of the voxel's flux over the n voxels. The vertical lines (labelled) denote the minimum size of a point source extending over three channels and occupying 3x3, 4x4, 5x5 or 7x7 voxels in every channel.} \\label{fig:VoxelSN} \\end{center} \\end{figure} This inherent limitation is compounded further by using a size-based rejection criterion to weed out false detections. If you only detect a few, unconnected voxels the source will be flagged as a false positive by a size-based rejection criterion. Off-centre and/or marginally resolved galaxies that are detected because of asymmetric flux distributions are most likely to be detected in the form of a few, unconnected voxels. This effect will also show up as an enhanced fracture rate of extended, well resolved sources. The inherent limitation of intensity thresholding based source finders can be offset by using more aggressive intensity thresholds (i.e., lower intensity thresholds), but it often results in many false (source) detections. The solution to this problem is to run the source finder on the datacube multiple times with the datacube smoothed to a different scale each time. This is however a very inefficient solution to the problem. {\\sc Duchamp} provides multiple options for dealing with this inherent limitation: a secondary `growth' threshold, smoothing and a 3-dimensional wavelet reconstruction of the dataset. As part of its design study, WALLABY has investigated novel methods of source detection as an alternative to multiple passes of an intensity threshold source finder. The goal of this investigation was to develop source detection methods that are optimised for large datacubes with high spatial and spectral resolution. The Characterised Noise \\HI ({\\sc CNHI}) source finder that I present here is one of the novel source detection methods that have been developed. The rest of the paper is structured as follows. The conceptual framework for the {\\sc CNHI} source finder is presented in Section \\ref{section:concept}. The inherent limitations of the {\\sc CNHI} source finder are then discussed in Section \\ref{section:limits}. Next, the current implementation of the {\\sc CNHI} source finder is presented in Section \\ref{section:implement}. Finally, some example results are discussed in Section \\ref{section:examples} before finishing with a summary. ", "conclusions": "\\label{section:summary} WALLABY and other projects that will be carried out on next-generation radio telescopes ASKAP and MeerKAT herald the start of the data deluge era in radio astronomy. The sheer size of WALLABY datacubes necessitates automation of many tasks in the data reduction pipeline, that previously would have been carried out with some level of manual input by an astronomer. Complete automation of finding \\HI galaxies in spectral datacubes is one of the challenges that is actively being investigated by WALLABY. The resolution and size of WALLABY observations poses a challenge for many existing automated source finders. This challenge arises from the underlying conceptual framework and algorithm that these source finders are based on, and not a flaw in the implementation. The {\\sc CNHI} source finder has been developed using a conceptual framework that can handle the large size of WALLABY datacubes and takes advantage of the resolution. Treating a datacube as a set of spectra (akin to an IFU observation), it attempts to find sources by looking for regions in each spectrum that do not look like noise. This is achieved using a novel implementation of matched filtering. Instead of using multiple filters that describe various types of sources, a single filter describing the noise is used. Sources are detected using this noise filter by identifying regions that do not look like noise. The performance of the {\\sc CNHI} source finder was tested using the PS and ES datasets in \\citet{WP_2011}. Analysis of the {\\sc CNHI} source finder output demonstrated that a reasonable combination of completeness and refined reliability can be achieved. A refined completeness of $\\sim 80\\%$ and $\\sim 50\\%$ was achieved for the PS and ES datasets, respectively, with a refined reliability of $\\sim 95\\%$. The PS dataset is better than the $80\\%$ completeness would suggest though, because the {\\sc CNHI} source finder found $\\sim 95\\%$ of all PS objects with a maximum voxel flux $\\geq 5\\sigma$, with a refined reliability of $\\sim 95\\%$. This analysis also demonstrated that the {\\sc CNHI} source finder recovers a significant fraction of the source flux. The recovery fraction asymptotes towards $100\\%$ as the total flux increases. More aggressive thresholds (larger) result in a recovery fraction that asymptotes faster, and starts higher. This suggests an alternative use of the {\\sc CNHI} source finder as a source parametrisation tool, that is used in tandem with another source finder. Finally, the performance analysis demonstrated that {\\sc CNHI} detections of a source are an unbiassed representation of the source and its flux distribution. These results are very promising, and warrant further testing and refinement of the {\\sc CNHI} source finder. There are three development goals for the {\\sc CNHI} source finder. Further development of the {\\sc CNHI} source finder will initially focus on incorporating multi-scale bundling. This will effectively achieve independently-scaled matched filtering in both the frequency and spatial dimensions. Additionally, the {\\sc CNHI} source finder will have a simple intensity thresholding test added to it. Incorporating an intensity thresholding test will make the {\\sc CNHI} source finder sensitive to sources occupying 3 or fewer channels. The final development goal is to incorporate fourier analysis, polynomial fitting and existing baseline structure removal techniques. Upon completing this next development cycle, the {\\sc CNHI} source finder will be tested and tweaked using the next round of ASKAP simulations and the \\HI Parkes All Sky Survey (HIPASS) \\citep{2000ASPC..217...50S} datacubes. The HIPASS datacubes have been selected because they have a well defined source catalogue, contain a mixture of resolved and unresolved sources, have known artifacts and calibration issues and there is a potential for the {\\sc CNHI} source finder to detect new sources." }, "1112/1112.4805_arXiv.txt": { "abstract": "DAV stars, also called ZZ Ceti variables, are pulsating white dwarfs with atmospheres rich in H. Asteroseismology of DAV stars can provide valuable clues about the origin, structure and evolution of DA white dwarfs. Recently, a new DAV star, WD J191643.83$+$393849.7, has been discovered in the field of the \\emph{Kepler} spacecraft. It is expected that further monitoring of this star in the next years will enable astronomers to obtain the best lightcurve of a pulsating DA white dwarf ever recorded, and thus to know with unprecedented precision the hidden details of the internal structure of this star. In this paper, we perform a first asteroseismological analysis of WD J191643.83$+$393849.7 on the basis of fully evolutionary DA white-dwarf models. Specifically, we employ a complete set of evolutionary DA white-dwarf structures covering a wide range of effective temperatures, stellar masses, and H envelope thicknesses. These models have been obtained on the basis of a complete treatment of the evolutionary history of progenitors stars. We compute $g$-mode adiabatic pulsation periods for this set of models and compare them with the pulsation periods exhibited by WD J191643.83$+$393849.7. Based on a tentative estimation of the mean period spacing of the star, we find that the stellar mass should be substantially large ($\\gtrsim 0.80 M_{\\odot}$), in agreement with the spectroscopically derived stellar mass. Also, from period-to-period fits we find an asteroseismological model characterised by a low effective temperature, rather high stellar mass and a thin H envelope. The possibility that this rather massive pulsating white dwarf can be further monitored with \\emph{Kepler} with a high degree of detail turns the star WD J191643.83$+$393849.7 into a promising and unique object to study the physics of crystallization and carbon/oxygen phase diagrams at high densities. ", "introduction": "Pulsating DA (H-rich atmospheres) white dwarfs, also known as ZZ Ceti or DAV variable stars, comprise the most numerous and studied class of degenerate pulsators. Their instability strip, probably a pure one, is centered at an effective temperature of around $12\\,000$ K (Winget \\& Kepler 2008; Fontaine \\& Brassard 2008; Althaus et al. 2010a). The luminosity variations (of up to $0.30$ mag) are caused by nonradial $g$-mode pulsations of low degree ($\\ell \\leq 2$) and periods between $\\approx 70$ and $\\approx 1500$ s, excited through a combination of the $\\kappa-\\gamma$ mechanism acting in the hydrogen partial ionization zone (Dolez \\& Vauclair 1981; Winget et al. 1982) and the ``convective driving'' mechanism --- once a sufficiently thick outer convective zone has been developed (Brickhill 1991; Goldreich \\& Wu 1999). ZZ Ceti asteroseismology --- the comparison between the observed periods and the periods computed for a suite of representative stellar models --- has the potential of disentangle the internal structure of DA white dwarfs, allowing to place constraints on the stellar mass, the thickness of the outer envelopes, the core chemical composition, weak magnetic fields and slow rotation rates from the observed period patterns (Winget \\& Kepler 2008; Fontaine \\& Brassard 2008; Althaus et al. 2010a). A total of 148 ZZ Ceti stars are currently known (Castanheira et al. 2010). To this list we must add the recently discovered DAV star WD J191643.83$+$393849.7 (\\emph{Kepler} ID 4552982 hereinafter WD J1916$+$3938), a ZZ Ceti located in the \\emph{Kepler} mission field and identified through ground-based times series photometry (Hermes et al. 2011; hereinafter HEA11). This star ($T_{\\rm eff}= 11\\, 129 \\pm 115$ K ; $\\log g= 8.34 \\pm 0.06$) exhibits low-amplitude luminosity variations with periods between $\\approx 800$ and $\\approx 1450$ s (see column 2 of Table \\ref{table1}). It is expected that an extended monitoring of this object with \\emph{Kepler} could bring the best lightcurve of a pulsating white dwarf ever recorded (HEA11), even with a better quality than the lightcurves of the brighter stars observed through uninterrupted ground-based campaigns of the Whole Earth Telescope (WET; Nather et al. 1990). In order for white dwarf asteroseismology to provide realistic constraints on the internal structure of white dwarfs, it is crucial the use of stellar models characterised by consistent and detailed chemical profiles to accurately assess the adiabatic pulsation periods. This requirement has been tackled successfully in the case of the hot DOVs or GW Vir stars (see C\\'orsico et al. 2007a, 2007b, 2008, 2009). Very recently, Romero et al. (2011) have performed the first asteroseismological analysis of a set of 44 ZZ Ceti stars based on the new generation of fully evolutionary DA white dwarf models presented in Althaus et al. (2010b). These models are characterised by realistic chemical profiles throughout the star and cover a wide range of stellar masses, thicknesses of the H envelope and effective temperatures. The analysis of a large number of ZZ Ceti stars has the potential to characterise the global properties of the class, in particular the thicknesses of the H envelope and the stellar masses. The study of Romero et al. (2011) revealed that DA white dwarfs in the solar neighbourhood could harbor a broad range of H-layer thickness. On the other hand, this analysis was successful in finding, for the first time, an univocal asteroseismological model for the archetypal ZZ Ceti star G117$-$B15A, one of the targets of that analysis. Motivated by the exciting discovery of the first pulsating DA white dwarf in the field of the \\emph{Kepler} mission, and encouraged by the availability of the most complete set of detailed evolutionary and pulsation models for DAV stars up to date (Romero et al 2011), we present in this paper a first asteroseismological analysis of WD J1916$+$3938. We are well aware that the asteroseismological potential of this pulsating star will probably be fully exploited with further observations by \\emph{Kepler}, which will allow to have available a period spectrum substantially richer than that we have at hand at present just from ground-based photometry. However, we think that a first asteroseismological analysis of this star based on the seven periods --- a representative number of periods for ZZ Ceti standards --- currently available is worth being attempted. By considering the mean period spacing exhibited by the star, we found that WD J1916$+$3938 should be substantially massive, in excellent agreement with the spectroscopic inference. We also found that period-to-period fits favour an asteroseismological model characterised by a high stellar mass, a low effective temperature, and a rather thin H envelope. The paper is organized as follows. In the next Section we briefly describe the computer codes and stellar models employed. The short Section \\ref{smd} is devoted to the derivation of the spectroscopic mass of WD J1916$+$3938. We make inferences about the stellar mass from the period spacing data in Section \\ref{period-spacing}, taking into account its dependence not only with the effective temperature and mass, but instead also with the thickness of the H envelope. In Section \\ref{period-fit} we extract information of the star through asteroseismological period fits. Section \\ref{rotation} is devoted to infer a rotation period of the star. Finally, we enumerate our results in the Conclusions (Section \\ref{conclu}). \\begin{table} \\centering \\caption{The observed frequencies, periods, and amplitudes of WD J1916$+$3938 (columns 1 to 3), according to HEA11, and the periods of a phenomenological model (see Section \\ref{period-spacing}) with a fixed ($\\ell= 1$) period spacing of $38.54$ s (columns 4 to 6).} \\begin{tabular}{cccccc} \\hline \\hline $\\nu$ & $\\Pi_i^{\\rm obs}$ & $A_i$ & $\\Delta k$ & $\\Pi_{\\rm mod}$ & $\\Pi_i^{\\rm obs}-\\Pi_{\\rm mod}$ \\\\ $[\\mu$Hz$]$ & $[$ s $]$ & $[\\%]$ & & $[$ s $]$ & $[$ s $]$ \\\\ \\hline 1213.7 & 823.9 & 0.38 & $2$ & $819.53$ & $ 4.37$ \\\\ 1198.9 & 834.1 & 0.32 & $5$ & $935.60$ & $-1.01$ \\\\ 1070.1 & 934.5 & 0.36 & --- & --- & --- \\\\ 1032.1 & 968.9 & 0.44 & $6$ & $974.14$ & $-5.24$ \\\\ 918.3 & 1089.0 & 0.25 & $9$ & $1089.75$& $-0.75$ \\\\ 854.8 & 1169.9 & 0.23 & $11$ & $1166.82$& $ 3.08$ \\\\ 696.0 & 1436.7 & 0.24 & $18$ & $1436.58$& $ 0.12$ \\\\ \\hline \\hline \\end{tabular} \\label{table1} \\end{table} ", "conclusions": "\\label{conclu} Tempted by the exciting discovery of the first pulsating DA white dwarf in the \\emph{Kepler} mission field by HEA11, and encouraged by the availability of the most complete set of detailed evolutionary and pulsation models for DAV stars up to date (Romero et al. 2011), in this paper we have presented a first asteroseismological analysis of WD J1916$+$3938. This, even with the certainty that the full asteroseismological potential of this DAV star will be exploited with further observations by \\emph{Kepler} in the next years. We summarize our main findings below: \\begin{itemize} \\item[-] We derive a spectroscopic mass of $M_*= 0.797-0.805\\ M_{\\odot}$ for WD J1916$+$3938 on the basis of the new DA white dwarf evolutionary tracks of Romero et al. (2011). This value is smaller than the estimate of HEA11, of $M_*= 0.82\\ M_{\\odot}$, based on the older models of Wood (1990). In our analysis, we have taken into account the dependence of the tracks in the $T_{\\rm eff}-\\log g$ diagram with the thickness of the H envelope (see Fig. \\ref{gteff}). \\item[-] By assuming that all the periods exhibited by WD J1916$+$3938 are normal modes of the star, and supposing that all the periods except one (at 834.1 s) are associated to $\\ell= 1$ modes, we derive an average period spacing of $\\overline{\\Delta \\Pi}_{\\rm obs}= 38.54 \\pm 0.29$ s. On the basis of the asymptotic theory of stellar pulsations, we estimate a stellar mass of $M_*= 0.851 \\pm 0.031 M_{\\odot}$, corresponding to a thick H envelope. If we take into account the strong dependence of the asymptotic period spacing with the thickness of the H envelope, we found an evident degeneracy of $\\Delta \\Pi_{\\ell}^{\\rm a}$ with $M_*$ and $M_{\\rm H}$. In virtue of this, we can conclude that $M_* \\gtrsim 0.837 M_{\\odot}$, but we cannot say anything about the thickness of the H envelope. Finally, if we ignore the constraint imposed by the effective temperature as derived by spectroscopy, we found that WD J1916$+$3938 should have a stellar mass $M_* \\gtrsim 0.77 M_{\\odot}$ in order for the star to be within the ZZ Ceti instability strip. \\item[-] By using the individual pulsation periods exhibited by WD J1916$+$3938, and taking into account several assumptions for their $\\ell$- and $k$-identification, we found an asteroseismological model with $T_{\\rm eff}= 11\\,380$ K, $\\log g= 8.39$, $M_*= 0.837 M_{\\odot}$, $\\log(M_{\\rm H}/M_*)= -7.36$, and $\\log(M_{\\rm He}/M_*)= -2.50$. The stellar mass of the asteroseismological model is in good agreement with the spectroscopic estimate ($M_*\\approx 0.8 M_{\\odot}$) and with the derivation from the period spacing ($M_*\\gtrsim 0.837 M_{\\odot}$). This is quite encouraging, because the three estimates have been derived from very different approaches and assumptions. On the other hand, the effective temperature of the asteroseismological model is somewhat higher than the spectroscopic one ($T_{\\rm eff}= 11\\,129 \\pm 115$ K). \\item[-] In passing, we have estimated a rotation period of 18.77 hs for WD J1916$+$3938. The reliability of this result rests on the assumption that the modes with periods at 823.9 s and 834.1 are the $m= +1$ and $m= -1$ components, respectively, of a $\\ell= 1$ frequency triplet due to rotation. \\end{itemize} In closing, we think that the conclusion that WD J1916$+$3938 should be a rather massive ZZ Ceti star, with $M_* \\gtrsim 0.8 M_{\\odot}$, appears to be robust. The possibility that this rather massive pulsating white dwarf can be further monitored with \\emph{Kepler} with a high degree of detail turns this star into a promising and unique target to study the physics of crystallization and carbon/oxygen phase diagrams at high densities. In this regard, we note that, for a stellar mass of $M_* \\sim 0.88 M_{\\odot}$, crystallization of matter should occur at the centre of a white dwarf star at $T_{\\rm eff} \\approx 10\\,500$ K according to the phase diagram of Segretain \\& Chabrier (1993), and at $T_{\\rm eff} \\approx 9\\,600$ K if the new phase diagram recently proposed by Horowitz et al. (2010) is considered (Althaus et al. 2011). Indeed, based on direct molecular dynamics simulations for dense carbon-oxygen mixtures, Horowitz et al. (2010) find substantially lower crystallization temperatures and that the shape of the carbon-oxygen phase diagram is of the azeotropic form --- and not of the spindle type --- as previously believed. Our analysis from the period spacing of WD J1916$+$3938 suggests that this star could well have a stellar mass much higher than $0.8 M_{\\odot}$ and necessarily a rather thin H envelope. If this were the case, this star could have a substantial fraction of its interior crystallized depending on the adopted carbon-oxygen phase diagram, an issue that could be explored with future asteroseismological analysis that exploit the new high-quality photometric data coming from the ongoing monitoring by the \\emph{Kepler} mission." }, "1112/1112.3108_arXiv.txt": { "abstract": "We have catalogued and analysed cosmological parameter determinations and their error bars published between the years 1990 and 2010. Our study focuses on the popularity of measurements, their precision and their accuracy. The accuracy of past measurements is gauged by comparison with the most recent WMAP results of Komatsu et al. (2011). The 637 measurements in our study are of 12 different parameters and we place the techniques used to carry them out into 12 different categories. We find that the popularity of parameter measurements (published measurements per year) in all 12 cases except for the dark energy equation of state parameter $w_0$ peaked between 1995 and 2004. Of the individual techniques, only Baryon Oscillation measurements were still rising in popularity at the end of the studied time period. The quoted precision (fractional error) of most measurements has been declining relatively slowly, with several parameters, such as the amplitude of mass fluctutations $\\sigma_{8}$ and the Hubble constant $H_{0}$ remaining close to the $10\\%$ precision level for a 10-15 year period. The accuracy of recent parameter measurements is generally what would be expected given the quoted error bars, although before the year 2000, the accuracy was significantly worse, consistent with an average underestimate of the error bars by a factor of $\\sim 2$. When used as complement to traditional forecasting techniques, our results suggest that future measurements of parameters such as fNL, and $w_{a}$ will have been informed by the gradual improvment in understanding and treatment of systematic errors and are likely to be accurate. However, care must be taken to avoid the effects of confirmation bias, which may be affecting recent measurements of dark energy parameters. For example, of the 28 measurements of $\\Omega_{\\Lambda}$ in our sample published since 2003, only 2 are more than 1 $\\sigma$ from the WMAP results. Wider use of blind analyses in cosmology could help to avoid this. ", "introduction": "Modern cosmological parameters have been measured since Hubble's (1929) discovery of the expansion of the Universe. The number of model parameters increased during the late 1980s with the introduction of what is often referred to as the ``Standard cosmological model'' (e.g. Dodelson 2005). The idea of ``Precision cosmology'' emerged more recently, and by the present time, many of the parameters in this model are well known (see e.g., Komatsu \\etal 2011, hereafter WMAP7). This presents us with an interesting opportunity: by comparing the past measurements of parameters and their error bars with the currently known values, we can evaluate how well the measurements were carried out in the past, how realistic the quoted uncertainties were, and which methods gave the most statistically reliable results. We can also study how both their precision and accuracy has varied with time. Such research will help us in our quest to make critical evaluations of what will be possible in the future, and by working with past data serves as a complement to more conventional future extrapolations of technology and techniques (e.g., the report of the Dark Energy Task Force, hereafter DETF, Albrecht \\etal, 2006). In the present paper we make a first attempt at such a study, by compiling published parameter values taken from the NASA Astrophysics Data System over the years 1990-2010. Previous studies of cosmological parameter determinations have tended to focus on the Hubble Constant, $H_{0}$, for which there is a longer than 80 year baseline for analysis. Several papers have used the comprehensive database compiled by John Huchra\\footnote{https://www.cfa.harvard.edu/~dfabricant/huchra/} to generate their dataset, such as the study of the non-Gaussian error distribution in those measurements (Chen \\etal 2003). Gott \\etal (2001) used median statistics in a metanalysis of these $H_{0}$ measurements to find the most probable value (and also analysed early measurements of $\\Omega_{\\Lambda}$). This median statistics approach has also been used to combine individual estimates of $\\Omega_{m0}$, the present mean mass density in non-relativistic matter by Chen and Ratra (2003). In the present paper we do not seek to combine the measurements from various works into best determinations of parameters. Instead we start from the assumption that the parameters we look at have been well measured (and their correct values are close to the WMAP7 values) and see what this implies about past measurements. We therefore will be starting with the assumption that $\\Lambda$CDM is the correct cosmological model. This should be borne in mind when interpreting our results. Even if the true cosmology turns out in the future to be something else, we expect that the effective values of the $\\Lambda$CDM parameters are not likely to be very different (given the good fits to current data), so that our approach will have some value even then. Parameters which at the moment are unknown, or very poorly constrained, such as the non-Gaussianity parameter fNL (e.g., Slosar \\etal 2008), or the time derivatives of the dark energy equation of state parameter $w$ (e.g., Chevallier \\& Polarski 2001) can obviously not be studied at present with our approach. Instead we hope that the general lessons from the past about the reliability of error bars, methods and achievable precision and accuracy can usefully to inform future efforts to measure those parameters. The DETF report explains how four different techniques are being used and will be used in the future to constrain dark energy parameters. These techniques, gravitational lensing, baryon oscillations, galaxy cluster surveys and supernova surveys all have a history and have been involved in a large number of previous measurements of different parameters. It is interesting to see how they have performed in the past, and evaluate them based on this data. By looking over the published record, we can also show how measurement precision has changed, in terms of the quoted fractional error bars, and see how this compares with predicted future trends. One can ask whether for example the earlier error bars were unrealistically small, so that the quoted precision of measurements has not changed much. This should have consquences for the accuracy of measurements, which we will define and measure. In general, our motivation for this study can be summarized by the idea that once cosmological parameter measurements are published, for the most part they are ignored when future work arrives. The dataset left behind can instead become a valuable resource to inform future work. Our plan for the paper is as follows. In Section 2., we detail the source for the cosmogical parameter estimates and how the data was collated. We explain the different categorizations of measurements and methods and standardization that was carried out. In Section 3 we outline the steps involved in our analysis of the data, and present results including historical trends in some individual parameters and the precision and accuracy of measurements. In Section 4 we summarize our findings and discuss our results. ", "conclusions": "\\subsection{Summary} We have compiled cosmological parameter measurements published between 1990 and 2010 and the techniques used to measure them. Using this data we have carried out an analysis of historical trends in popularity, precision and accuracy. The accuracy of past measurements has been estimated by assuming that WMAP7 parameter values of Komatsu \\etal (2011) (combined with $\\Lambda$CDM standard values for e.g. $w_{0}$) are the correct ones. Our findings can be summarised as follows: (1) The number of published measurements for different parameters peaks between 1995 and 2004 for all cases, except for $w_{0}$ for which the number was still rising in 2010. (2) Of all techniques used to measure the parameters, only baryon oscillation and ``combined'' measurements were still rising in terms of publications per year by 2010. (3) The quoted precision of measurements has been declining relatively slowly for most parameters, with several (e.g. $\\sigma_{8}$, $H_{0}$ remaining flat for 10-15 years. (4) The accuracy of recent parameter measurements is generally what should be expected based on the quoted error bars i.e. the error bars overall are neither understimated nor overestimated (an accuracy, $N_{\\sigma}=1.0$, within the Poisson uncertainty on the measurement). Before 2000, the accuracy $N_{\\sigma}$ as closer to 2, indicating underestimation of the error bars by a factor of 2. Overall, there is a small non-Gaussian tail to the error distributions (we find that $20\\%$ of measurements are more that $2\\sigma$ away from the true values. (5) The accuracy of most methods has become consistent with $N_{\\sigma}=1.0$, with the historically most innaccurate parameter measurement technique being the use of galaxy peculiar velocities. Measurements of $\\Omega_{M}$ and particularly $\\Omega_{\\Lambda}$ made since 2000 tend to have accuracy $N_{\\sigma}$ significantly less than 1.0, indicating ``confirmation bias'' and/or an overestimation of error bar sizes. \\subsection{Discussion} Over the 20 year period covered in this study, it is apparent that many of the parameters in what is now the concordance CDM cosmological model went from the status of no information or only limits to being known at the $10\\%$ level or better. It is also apparent from Figure 2 that there was a ``golden age'' of parameter measurements between $\\sim 1995$ and $\\sim 2005$ during which the number of published measurements peaked sharply and then declined. This seems to indicate that for many purposes (such as the use of a background cosmology in galaxy formation models), the precision to which the $\\Lambda$CDM parameters were known by the time of the first WMAP results is sufficient, and many of the reasons for pinning down the model better had diminished after that. This said, however, the exception to this rule, measurements of $w_{0}$ (which are still rising in terms of number per year at the end of our study) seems to point to a coming new era in parameter measurement. Certainly, the motivation for the large number of ongoing and future large-scale structure, lensing and other surveys is to hunt for the signatures of dynamical dark energy and modified gravity, and given the number of researchers carrying out these studies it is likely that measurements will continue to rise. Many parameters which we have not catalogued are now within reach of quantitiative study. These include the modified gravity parameter, $E_{G}$ (Reyes \\etal 2010) and the time derivative of the equation of state parameter, $w_{a}$. Measurements of such parameters involve searching for deviations from the concordance CDM model and fall into a different category from most of the parameters we have studied in this paper. Inflationary parameters such as the non-Gaussianity fNL, or tensor to scalar ratio $r$ will pinned down with higher precision in the future, and these should also represent a growth area. The motivation for most future measurements being largely framed in terms of a quest for fundamental physics, it would be logical to assume that they will continue until the cause for the Universe's acceleration are better understood. Likewise, parameters describing the dark matter particle should be added to this category. Possible behaviours for the precision of future parameter measurements can be predicted by looking at the past results (Figure \\ref{prepa}). There is a very wide range, but most parameters improve slowly, with a factor of 10 improvement in precision over the 20 years representing the extreme (2 out of 12 parameters). The precision of some parameters has remained relatively flat for the whole period, so this is a possibility for future so far unconstrained parameters. An argument against this slow progress however is the fact that many new surveys (such as of Baryon Oscillations) are targeted primarily at measures of specific parameters, and this aggressive approach (for example including specific precisions to be obtained at a given time in survey proposal documents) could lead to faster progress. Our investigation of the accuracy of results could potentially lead to some of the most interesting findings. We have seen that in the earlier half of our studied time period there is evidence that the error bars were significantly underestimated, but that this has changed over time. When discussing the accuracy, we are should be aware that it was not possible in our analysis to take into account several factors which have the potential to affect our conclusions. For example, we do not keep track of the priors that people have assumed in their measurements, and in many of the later cases, this may include the WMAP results as priors. That this is happening is likely to be responsible for much of the post WMAP1 tightening of constraints seen in Figures \\ref{h0}- \\ref{w0}. When computing the error bars on the mean accuracies of measurements (Figs \\ref{accp} and \\ref{accm}) we have used Poisson errors based on the number of measurements in each bin. This will tend to underestimate the uncertainty on the accuracy because some of those measurements could be using the same underlying data, or be using similar priors, or a combination of the two. There will therefore be correlations between the error bars so that our estimates of the accuracy will be affected. Equivalently the chi-square of the fiducial result compared to the data points will be incorrectly determined to be low because of the correlations are not included. Bearing the above points in mind, we return to the panels in Figures \\ref{accp} and \\ref{accm} where the accuracy seems to be significantly below $N_{\\sigma}=1$. This is most obvious in the second panel ($\\Omega_{\\Lambda}$) of Figure \\ref{accp}. Such as result could be a sign that either the error bars have been significantly overestimated, or else that researchers have been influenced by prior results (``confirmation bias''), or a combination of the two. If we return to the data points which led to the last two bins of panel two of Figure \\ref{accp}, we find something especially striking. Of those 28 measurements, only 2 are more than $1 \\sigma$ from the fiducial results of Table 1. These 28 measurements were carried out by approximately 11 separate groups (as determined by authorship lists) using several different techniques. This closeness of published results to the ``correct'' ones is somewhat worrying for future measurements. One can interpret this as coming partly from error bars being overerestimated by cautious cosmologists, for example by including possible systematic errors in the error bars which are not actually present to such a large degree, or in a related point authors marginalizing over parameters which are actually better known than was assumed. We note that including or excluding the actually quoted systematic error bars (Section 3.4) has little effect on this result. An additional question is why some parameters have $N_{\\sigma} <1$ and others do not (e.g., $\\sigma_{8}$). The relatively low number statistics of our whole dataset preclude us from making any strong statements about this issue. If it does partly result from confirmation bias, one can also wonder how observers knew which value of $\\Omega_{\\Lambda}$ (for example) would be the ``correct'' one, given that our fiducial (mostly WMAP7) results from Table 1 were published in 2011. If this bias is present, it is probably related to the mean level for $\\Omega_{\\Lambda}$ resulting from several prior measurements. For example in Figure \\ref{lambda} and others, the value of the parameter seems to be pretty well determined at least by 2003. If we look at the techniques which are often associated with dark energy measurements, SN and BAO, we can see in Figure \\ref{accm} that these two have low $N_{\\sigma}$ for recent measurements. Of the 23 measurements which where included in the last bins of the SN panel of Figure \\ref{accm}, only 2 are more than 1 $\\sigma$ from the fiducial result. We note that this fiducial result from Table 1 does include BAO measurements, but not SN estimates of dark energy. In the case of SN, however, only 4 measurements of $\\Omega_{\\Lambda}$ are included in these bins, and only 2 separate groups of researchers, so that for that subset of data, statistical fluctuations may well be responsible for the low $N_{\\sigma}$ seen. If confirmation bias is present, on the other hand, one could argue about who is confirming who- certainly the first SN results on dark energy predate those from BAO and from most other techniques. These sorts of questions might be addressed by a more detailed look at the published measurements, including details of priors, jointly used datasets and analysis techniques. Then again small number statistics probably would not allow firm conclusions to be drawn. These hints should instead serve as a warning that care and perhaps concrete steps be taken to avoid any confirmation bias in the future. In conclusion, we have seen that huge progress has been made in the 20 year period covered by our study. Important questions have been resolved (e.g., is the Universe open?, do massive neutrinos contribute substantially to the dark matter density?), a model has been found which agrees with essentially all observational data so far ($\\Lambda$CDM), and the parameters of that model have been pinned down at the $1-10\\%$ level. The first WMAP results (e.g. as presented in Spergel \\etal (2003)) form a watershed which is easy to pick up in most plots of parameters with time, and serves as a reminder that statistically measurable progress is not always gradual. Perhaps the most interesting aspect of our study, of the accuracy of past results compared to our most recent knowledge has found that understanding of systematic errors and uncertainties in cosmological measurements has demonstrably improved since the early 1990s. On average, results in the last 10 years are consistent with expectations, given their error bars, something which should instill confidence in future measurements. There are some signs that recent measurements of dark energy parameters are closer to the ``expected'' values for $\\Lambda$CDM than statistically likely. These may be explainable by correlations between measurements which we have not included. On the other hand this may serve as a sign that as cosmology collaboration sizes increase carrying out more blind analyses (as in particle physics) may be a good idea." }, "1112/1112.2612_arXiv.txt": { "abstract": "We have used the Robert C.\\ Byrd Green Bank Telescope to time nine previously known pulsars without published timing solutions in the globular clusters M62, NGC~6544, and NGC~6624. We have full timing solutions that measure the spin, astrometric, and (where applicable) binary parameters for six of these pulsars. The remaining three pulsars (reported here for the first time) were not detected enough to establish solutions. We also report our timing solutions for five pulsars with previously published solutions, and find good agreement with past authors, except for PSR J1701$-$3006B in M62. Gas in this system is probably responsible for the discrepancy in orbital parameters, and we have been able to measure a change in the orbital period over the course of our observations. Among the pulsars with new solutions we find several binary pulsars with very low mass companions (members of the so-called ``black widow'' class) and we are able to place constraints on the mass-to-light ratio in two clusters. We confirm that one of the pulsars in NGC~6624 is indeed a member of the rare class of non-recycled pulsars found in globular clusters. We also have measured the orbital precession and Shapiro delay for a relativistic binary in NGC~6544. If we assume that the orbital precession can be described entirely by general relativity, which is likely, we are able to measure the total system mass ($2.57190(73)\\; \\Msun$) and companion mass ($1.2064(20)\\; \\Msun$), from which we derive the orbital inclination ($\\sin{i} = 0.9956(14)$) and the pulsar mass ($1.3655(21)\\; \\Msun$), the most precise such measurement ever obtained for a millisecond pulsar. The companion is the most massive known around a fully recycled pulsar. ", "introduction": "} Millisecond pulsars (MSPs) form by ``recycling'' a dormant neutron star through the accretion of mass and angular momentum from a binary companion \\citep{acr+82}. This leads to a very rapidly rotating and stable pulsar with a relatively low magnetic field ($\\sim 10^9\\; \\gauss$) and long lifetime ($\\gtrsim 10^9\\; \\yr$). MSPs form naturally in the dense environments of globular clusters (GCs), thanks to frequent exchange interactions that may lead to the formation of mass transferring binaries \\citep{cr05}. Sensitive searches of GCs have uncovered 144 pulsars in 28 clusters\\footnote{See \\url{http://www.naic.edu/~pfreire/GCpsr.html} for an up-to-date list.}, and the vast majority of these are recycled MSPs. Indeed, nearly half of all MSPs have been discovered in GCs\\footnote{There are 160 field pulsars with period $P < 20\\; \\ms$ in the ATNF catalog, and 131 GC pulsars that meet the same criteria (see \\url{http://www.atnf.csiro.au/people/pulsar/psrcat/}).}. The same interactions that form MSPs so efficiently in clusters also lead to many exotic systems that are rarely seen in the disk of the Galaxy. These include the fastest spinning MSP \\citep{hrs+06}, highly eccentric binaries \\citep{rhs+05,frg07}, massive neutron stars \\citep{frb+08}, pulsar-main sequence binaries \\citep{dpm+01}, and many ``black widow'' systems \\citep{kbr+05}. These discoveries demonstrate the huge scientific payoff that can come from the discovery of unique pulsars thanks to two factors---the extreme nature of neutron stars, which opens windows on otherwise inaccessible realms of physics, and the extraordinary clock-like nature of MSPs. The bedrock of pulsar astronomy is \\emph{timing}, the process of creating a model that unambiguously accounts for every rotation of the pulsar, thus probing the pulsar and its environment. For MSPs, the arrival time of a pulse can typically be measured to within a few microseconds or better \\citep[e.g.][]{vbv+08}, which enables very precise timing models. Timing GC pulsars leads to unique challenges, though. Because GCs are usually at a distance of several kiloparsecs, the flux density of their constituent MSPs is usually very low, necessitating very long integration times. Another consequence of cluster distances is very high dispersion measures, which necessitate moving to higher observing frequencies (e.g. $2\\; \\GHz$) where pulsars tend to be weaker. Luckily, since many clusters contain several MSPs that can each be observed during a single observation, the required observing time is well spent. In addition to recycled MSPs, a small population of slow, non-recycled pulsars have also been observed in a handful of clusters \\citep{lmd96,blt+11}. Their presence is something of a mystery, because they resemble in every way the ``normal'' pulsars that are so numerous in the Galactic disk, but which have lifetimes $\\sim 10^7$--$10^8\\; \\yr$ and are thought to form through core collapse supernovae of massive stars. However, GCs are old stellar systems with typical ages $10^{10}\\; \\yr$ \\citep{cgcf00}, so all stars massive enough to form pulsars should have died some $10^{10}\\; \\yr$ ago. As such, the core collapse of massive stars cannot be the avenue through which non-recycled pulsars in GCs have formed. The most popular alternative formation scenario involves the collapse of a massive white dwarf via electron capture \\citep{nom84,nom87}, though the details of these so-called electron capture supernovae are not well understood. In this paper, we present new timing solutions for nine pulsars spread across three GCs---M62, NGC~6544, and NGC~6624. Each of these clusters contains multiple MSPs and NGC~6624 contains two non-recycled pulsars. All of these pulsars were discovered elsewhere, but full, phase-connected timing solutions have only been published for five of them (see Table \\ref{table:psrs} for a summary and list of references). We also present our solutions for these five. We began a timing campaign using the Robert C.\\ Byrd Green Bank Telescope (GBT) for the nine pulsars without full solutions, starting in 2009 February. We have managed to obtain full solutions that include measurement of the first period derivative for six of these pulsars. A combination of low flux densities and an irreversible data processing error conspired to prevent us from obtaining full solutions for the two remaining MSPs, but we are in a position to comment on them in further detail. Our most exciting result is the discovery of a massive companion orbiting a fully recycled MSP. In \\S\\ref{sec:obs} we describe our observations and in \\S\\ref{sec:timing} outline our method for timing the pulsars and further data analysis. A discussion of individual systems can be found in \\S\\ref{sec:psrs}, and we provide a summary in \\S\\ref{sec:conc}. ", "conclusions": "} We set out to obtain full timing solutions for nine previously known GC pulsars in M62, NGC~6544, and NGC~6624 (three of which have never been published). Our efforts were successful for six of these pulsars, and a seventh now has a partial solution. We have confirmed the nature of three black widow pulsars (M62E and F, and NGC~6544A). NGC~6544A has $\\dot{P} < 0$, providing unambiguous evidence that $\\dot{P}$ is dominated by acceleration in the cluster potential, and providing us with a probe of $\\Upsilon\\rmsub{V}$ in the cluster core. We find $\\Upsilon\\rmsub{V} \\gtrsim 0.072$ which, while not very constraining, is consistent with small amounts of low-luminosity matter such as massive black holes or other stellar remnants. We are also able to confirm that NGC~6624C belongs to a rare class of non-recycled GC pulsars similar to the slow pulsars found in the Galactic disk. The highlight of our work is unquestionably NGC~6544B. This binary MSP is in a highly eccentric orbit and exhibits clear orbital precession and Shapiro delay. Under the well justified assumption that this is due almost entirely to GR, we are able to obtain a most-probable total system mass of $2.57190(73)\\; \\Msun$, companion mass of $1.2064(20)\\; \\Msun$, and pulsar mass of $1.3655(21)\\; \\Msun$. This is the highest mass companion known to orbit a fully recycled MSP, and raises the possibility that NGC~6544B is part of either double neutron star system, which could form via exchange interactions in the GC. We are deeply grateful to Andrea Possenti and Alessandro Corongiu for providing us with data collected by the Parkes Telescope that was used to check for phase connection in our solutions for M62D, E, and F. We would also like to thank an anonymous referee for helpful comments. R.\\ Lynch acknowledges support from the GBT Student Support program and the National Science Foundation grant AST-0907967 during the course of this work. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc." }, "1112/1112.0567_arXiv.txt": { "abstract": "Weak-gravitational-lensing distortions to the intensity pattern of 21-cm radiation from the dark ages can be decomposed geometrically into curl and curl-free components. Lensing by primordial gravitational waves induces a curl component, while the contribution from lensing by density fluctuations is strongly suppressed. Angular fluctuations in the 21-cm background extend to very small angular scales, and measurements at different frequencies probe different shells in redshift space. There is thus a huge trove of information with which to reconstruct the curl component of the lensing field, allowing tensor-to-scalar ratios conceivably as small as $r\\sim 10^{-9}$---far smaller than those currently accessible---to be probed. ", "introduction": " ", "conclusions": "" }, "1112/1112.0084_arXiv.txt": { "abstract": "We discuss the effects of cosmic phase transition on the spectrum of primordial gravitational waves generated during inflation. The energy density of the scalar condensation responsible for the phase transition may become sizable at the epoch of phase transition, which significantly affects the evolution of the universe. As a result, the amplitudes of the gravitational waves at high frequency modes are suppressed. Thus the gravitational wave spectrum can be a probe of phase transition in the early universe. ", "introduction": " ", "conclusions": "" }, "1112/1112.0745_arXiv.txt": { "abstract": "The cross-correlation of the Ly-$\\alpha$ forest and redshifted 21-cm emission has recently been proposed as an observational tool for mapping out the large-scale structures in the post-reionization era $z \\le 6$. This has a significant advantage as the problems of continuum subtraction and foreground removal are expected to be considerably less severe in comparison to the respective auto-correlation signals. Further, the effect of discrete quasar sampling is considerably less severe for the cross-correlation in comparison to the Ly-$\\alpha$ forest auto-correlation signal. In this paper we explore the possibility of using the cross-correlation signal to detect the baryon acoustic oscillation (BAO). To this end, we have developed a theoretical formalism to calculate the expected cross-correlation signal and its variance. We have used this to predict the expected signal, and estimate the range of observational parameters where a detection is possible. For the Ly-$\\alpha$ forest, we have considered BOSS and BIGBOSS which are expected have a quasar density of $16 \\ {\\rm deg}^{-2}$ and $64 \\ {\\rm deg}^{-2}$ respectively. A radio interferometric array that covers the redshift range $z=2$ to $3$ using antennas of size $2 \\, {\\rm m} \\times 2 \\, {\\rm m}$, which have a $20^{\\circ} \\times 20^{\\circ}$ field of view, is well suited for the redshifted 21-cm observations. It is required to observe $25$ independent fields of view, which corresponds to the entire angular extent of BOSS. We find that it is necessary to achieve a noise level of $1.1 \\times 10^{-5} \\ {\\rm mK}^2$ and $6.25 \\times 10^{-6} \\ {\\rm mK}^2$ per field of view in the redshifted 21-cm observations to detect the angular and radial BAO respectively with BOSS. The corresponding figures are $3.3 \\times 10^{-5} \\ {\\rm mK}^2$ and $1.7 \\times 10^{-5} \\ {\\rm mK}^2$ for BIGBOSS. We also discuss possible observational strategies for detecting the BAO signal. Four to five independent radio interferometric arrays, each containing $400$ antennas uniformly sampling all the baselines within $50 \\, {\\rm m}$ will be able to carry out these observations in the span of a few years. ", "introduction": "Neutral hydrogen (HI) in the post-reionization epoch ($z < 6$) is known to be an important cosmological probe seen in both emission and absorption. Here, the redshifted 21-cm emission \\citep{hirev1, hirev3} and the transmitted QSO flux through the Ly-$\\alpha$ forest \\citep{wien, Mandel} are both of utmost observational interest. In a recent paper \\citet{tgs5} have proposed the cross-correlation of the 21-cm signal with the Lyman-$\\alpha$ forest as a new probe of the post-reionization era. While it is true that the emission and the absorption signals both originate from neutral hydrogen (HI) at the same redshift (or epoch), these two signals, however, do not originate from the same set of astrophysical sources. The 21-cm emission originates from the HI housed in the Damped Lyman-$\\alpha$ systems (DLAs) which are known to contain the bulk of the HI at low redshifts \\citep{proch05}. The collective emission from the individual clouds appears as a diffuse background in low frequency radio observations \\citep{poreion2}. On the contrary, the Ly-$\\alpha$ forest consists of a large number of Ly-$\\alpha$ absorption lines seen in the spectra of distant background quasars. These absorption features arises due to small density fluctuations in the predominantly ionized diffuse IGM. On large scales, however, the fluctuations in the 21-cm signal and the the Ly-$\\alpha$ forest transmitted flux are both believed to be excellent tracers of the underlying dark matter distribution. \\citet{tgs5} have proposed that the cross-correlation between these two signals can be used to probe the power spectrum during the post-reionization era. The HI power spectrum can be determined separately from observations of the Ly-$\\alpha$ forest \\citep{pspec} and the redshifted 21-cm emission \\citep{poreion1}. The detection of the individual signals, however, face severe observational challenges. The Ly-$\\alpha$ auto-correlation power spectrum is expected to be dominated by the Poisson noise arising from the discrete sampling of QSO lines of sight. The cross-correlation signal has the advantage that it is not affected by the Poisson noise which affects only its variance. Uncertainties in fitting the QSO continuum \\citep{pspec4, kim04, bolton} pose another challenge for using Ly-$\\alpha$ observations to determine the power spectrum. Astronomical sources like the galactic synchrotron emission, extra-galactic point sources, etc. appear as foregrounds for the redshifted 21-cm signal. These foregrounds, which are several orders of magnitude larger than the signal, pose a great challenge for detecting the post-reionization 21-cm signal \\citep{fg10,fg11}. The foregrounds and systematics in the Ly-$\\alpha$ forest and the redshifted 21-cm emission are, however expected to be uncorrelated and we therefore expect the problem to be much less severe for the cross-correlation signal. A detection of a cross-correlation signal shall, hence, conclusively ascertain its cosmological origin. Apart from being an independent probe of the large scale matter distribution, the cross-correlation signal can potentially unveil the same astrophysical and cosmological information as the individual auto-correlations. Cosmological density perturbations drive acoustic waves in the primordial baryon-photon plasma which are frozen once recombination takes place at $ z \\sim 1000$, leaving a distinct oscillatory signature on the CMBR anisotropy power spectrum \\citep{peeb70}. The sound horizon at recombination sets a standard ruler that maybe used to calibrate cosmological distances. Baryons contribute to $15 \\% $ of the total matter density, and the baryon acoustic oscillations are imprinted in the late time clustering of non-relativistic matter. The signal, here, is however suppressed by a factor $ \\sim \\Omega_b/\\Omega_m \\sim 0.1$, unlike the CMBR temperature anisotropies where it is an order unity effect \\citep{komatsu}. The baryon acoustic oscillation (BAO) is a powerful probe of cosmological parameters \\citep{seoeisen,white}. This is particularly useful since the effect occurs on large scales ($\\sim 150 \\, \\rm Mpc$), where the fluctuations are still in the linear regime. It is possible to measure the angular diameter distance and the Hubble parameter as functions of redshift using the the transverse and the longitudinal oscillations respectively. These provide means for estimating cosmological parameters and placing stringent constraints on dark energy models. Nonlinear effect of gravitational clustering tend to wipe out the BAO signal, and it is preferable to avoid very low redshifts where this is a potential problem. However, very high redshifts too are not very useful for constraining dark energy models. Several authors have reported a $2$ to $3 \\sigma$ detection of the BAO in low redshift galaxy surveys \\citep{baoeisen05, percival07, baosdss}. The possibility of detecting the BAO signal in the Ly-$\\alpha$ forest has been extensively studied by \\citet{cosparam1}. Several groups have also considered the possibility of detecting the BAO signal using redshift 21-cm emission \\citep{pen2008, maowu, masui10}. Several QSO surveys are now being considered with the intent of measuring the BAO using the Ly-$\\alpha$ forest (eg. BOSS \\citep{mcd1} and BIGBOSS \\citep{bigboss}). The possibility of a wide field redshifted 21-cm survey to detect the BAO is also under serious consideration. In this paper, we consider the possibility of studying the BAO using the cross-correlation signal. In Section 2. of this paper we quantify the cross-correlation between the Ly-$\\alpha$ forest and the 21-cm emission, and present theoretical predictions of the expected signal. We have used the multi-frequency angular power spectrum $C_{\\ell}(\\Delta z)$ (MAPS, \\citealt{datta1}) in preference to the more commonly used three dimensional power spectrum $P(k)$ to quantify the cross-correlation signal. This has several advantages which we briefly discuss here. MAPS refers to the angular multipole $\\ell$ (or equivalently angle) and redshift interval $\\Delta z$ which are the directly relevant observational quantities for both Ly-$\\alpha$ surveys and redshifted 21-cm observations. This is particularly important if we wish to determine the angular scales and redshift intervals which need to be covered in order to detect a particular feature in the signal. The foregrounds in the redshifted 21-cm observations and the continuum in the Ly-$\\alpha$ forest are both expected to have a smooth, slow variation along the frequency axis, and this plays a crucial role in removing these from the respective data. It is therefore advantageous to use MAPS which retains the distinction between the frequency (redshift) and angular information, unlike $P(k)$ which mixes these up. \\citet{fg3} have analyzed GMRT observations using MAPS to jointly characterize the angular and frequency dependence of the foregrounds at $150 \\, {\\rm MHz}$. \\citet{fg10} and \\citet{fg11} have applied MAPS to Analyze $610 \\, {\\rm MHZ}$ GMRT observations, and have used this to characterize the foregrounds for the post-reionization 21-cm signal. In fact, they show that it is possible to completely remove the foregrounds from the measured $C_{\\ell}(\\Delta \\nu)$ by subtracting out polynomials in $\\Delta \\nu$. Finally, the signal could have a significant contribution from the light cone effect, particularly if the observations span a large redshift interval. It is, in principle, possible to account for this in the MAPS, though we have not done this here. It is, however, not possible to account for this effect in the three dimensional power spectrum which mixes up the information from different epochs through a Fourier transform along the radial direction. In Section 3 of this paper we quantify the imprint of the BAO feature on the cross-correlation signal. Here MAPS has the advantage that it allows us to separately study the radial and the transverse oscillations through the $\\Delta \\nu$ and the $\\ell$ dependence respectively. In Section 4 we introduce an estimator for the MAPS of the cross-correlation signal and derive its statistical properties. In particular, we present a detailed analysis of the noise for the cross-correlation estimator. In Section 5 we present several observational considerations which are relevant for the cross-correlation signal, some pertaining to the Ly-$\\alpha$ forest and others to the redshifted 21-cm signal or both. Finally, in Section 6 we discuss the detectability of the cross-correlation signal and the BAO features. We have used the cosmological parameters $(\\Omega_m h^2, \\Omega_b h^2, \\Omega_{\\Lambda}, h, n_s, \\sigma_{8})= (0.136, 0.023, 0.726, 0.71, 0.97, 0.83)$ from \\citet{komatsu} throughout this paper. ", "conclusions": "In this paper we have developed a theoretical formalism for estimating the cross-correlation signal between the fluctuations in the Ly-$\\alpha$ forest and the fluctuations in the redshifted 21-cm emission from neutral hydrogen. Both of these quantities are measured as functions of frequency (redshift) and angular scale. Consequently, we have used the Multi-frequency Angular Power Spectrum (MAPS) to quantify the statistical properties of the cross-correlation signal. This deals directly with the observed quantities, and retains the distinction between the angular and the frequency information. This, as we shall elaborate shortly, is very important in the light of foreground removal and continuum subtraction. Continuum fitting and subtraction is a very critical step in calculating $\\delta_{\\F}$ for the Ly-$\\alpha$ forest and several different methods have been proposed for handling this \\citep{pspec4, mcd06}. Errors in continuum subtraction can be a serious problem for the Ly-$\\alpha$ forest auto-correlation signal \\citep{kim04}. The redshifted 21-cm signal is buried deep under astrophysical foregrounds which are several orders of magnitude larger \\citep{shaver99,fg4,fg1,fg6, fg3, pen09, fg10}. Several different techniques have been proposed for separating the cosmological 21-cm signal from the foregrounds \\citep{ fg3, jelic, fg7, fg8, fg10}. Foreground removal is a rather severe problem for observations of the redshifted 21-cm auto-correlation signal. In addition to the 21-cm signal, the foregrounds make a very large contribution to the expectation value of the auto-correlation estimator. However, the foregrounds are believed to have a slowly varying, smooth frequency (or $\\Delta z$) dependence which is quite distinct from the signal which decorrelates rapidly with increasing $\\Delta z$ (Figure \\ref{fig:kappa1}). It is therefore possible to remove the foregrounds from the measured MAPS ($P_{\\F T o}(\\ell,\\Delta z)$ ) by subtracting out any component that varies slowly with $\\Delta z$. In fact, \\citet{fg11} have recently used MAPS to analyze 610-MHZ GMRT observations and show that it is possible to remove the foregrounds from the m auto-correlation by fitting and subtracting out slowly varying polynomials in $\\Delta \\nu$. We do not expect the continuum in the Ly-$\\alpha$ forest to have any correlation with the foregrounds in the redshifted 21-cm observations, and consequently they will not contribute to the expected cross-correlation signal. The continuum and the foregrounds will, however, appear as extra contributions to the variance of the cross-correlation estimator. Since these contributions appear in the variance, it is possible to reduce these by combining different independent estimates of the cross-correlation. We can reduce the continuum and foreground contributions by combining estimates of the cross-correlation at different baselines $\\U$ and different fields of view. The problem, therefore, is much less severe in comparison to the auto-correlation. Further, the continuum and the foregrounds are both expected to have a slowly varying frequency (or $\\Delta z$) dependence and it should be possible to remove these from the measured MAPS by subtracting out the component that varies slowly with $\\Delta z$. We plan to perform a detailed analysis of these issues in future. Our study shows that it is possible to have a $5-\\sigma$ detection of the imprint of the first BAO peak in the cross-correlation signal using BOSS, an upcoming QSO survey. For this, we have considered a radio interferometric array that covers the $z$ range $z=2$ to $3$ using antennas of size $2 \\, {\\rm m} \\times 2 \\, {\\rm m}$ which have a $20^{\\circ} \\times 20^{\\circ}$ field of view. We find that we need to observe $25$ fields of view, approximately the full angular coverage of BOSS. with a noise level of $N_T=1.1 \\times 10^{-5} \\, {\\rm mK}^2$ and $N_T= 6.25 \\times 10^{-6} \\, {\\rm mK}^2$ in order to achieve a $5-\\sigma$ detection of the angular and radial oscillations respectively. The corresponding noise levels are $N_T=3.3 \\times 10^{-5} \\, {\\rm mK}^2$ and $N_T= 1.7 \\times 10^{-5} \\, {\\rm mK}^2$ for BIGBOSS whose quasar density is expected to be four times larger than that of BOSS. We now briefly discuss how it may be possible to carry out such observations. In our analysis we have made the simplifying assumption that the antennas are distributed such that all the baselines within $50 \\, {\\rm m}$ are uniformly sampled. We consider an interferometric array with $N_{ann}=400$ antennas which roughly corresponds to the maximum number of $2 \\, {\\rm m} \\times 2 \\, {\\rm m}$ antennas that can fit in a $50 \\, {\\rm m} \\times 50 \\, {\\rm m}$ region. We see that we need $\\Delta t = 5700 \\, {\\rm hr}$ and $10,000 \\, {\\rm hr}$ of observation per field of view (\\ref{eq:oc1}) to reach the noise levels required to detect the angular and radial oscillations respectively with BOSS. The corresponding figures are $1900 \\, {\\rm hr}$ and $3700 \\, {\\rm hr}$ for the BIGBOSS. Note that we it is required to observe a single field of view for $8 \\, {\\rm hr}$ a day for a whole year in order to achieve $3000 \\, {hrs}$ of observing time. It is quite evident that we require to observe $25$ fields of view, with $2$ to $3$ years of dedicated observations for each field, in in order to detect the BAO. Operating sequentially, considering one field after the next, the required observations would possibly run over a period of $50 \\, {\\rm yr}$ to a century, which raises the need to consider alternative observational strategies. It is a viable possibility to have antennas that can simultaneously observe several independent fields of view. However, it is unlikely (if not impossible) to have antennas that can simultaneously observe $16$ or $25$ such $20^{\\circ} \\times 20^{\\circ}$ fields of view. For the purpose of this discussion, we assume that we have antennas that can simultaneously observe $4$ fields of view. We then see that it would approximately require observations over a decade (or more) in order to detect the BAO. Another possibility is to have several radio interferometric arrays, each located at a different location and observing a different parts of the sky. Four to five separate arrays, each with $400$ antennas, would be required to carry out these observations in the span of a few years. It is important to note that it may be possible to reduce the observational requirements to some extent by optimally distributing the baselines instead of considering them to be uniformly distributed. We propose to investigate these issues in a future study. Observations of the BAO can be used to constrain the values of various cosmological parameters. The equation of state of the Dark Energy is particularly important in this context. In this paper we have mainly estimated the range of observational parameters for which it will be possible to detect the BAO using the cross-correlation signal. We plan to study a variety of issues including the optimal array configuration and cosmological parameter estimation in future." }, "1112/1112.2789_arXiv.txt": { "abstract": "{ The majority of VO-compatible spectra handling applications operates only with a few spectra entirely downloaded from single or several SSAP servers. We try to identify the scientific cases which could immediately benefit from future SSAP applications designed for GRID deployment. Their key feature is the sophisticated spectra pre-selection and preprocessing done on distributed servers using the intelligent agent summarising the results and performing final high-level processing or analysis. ", "introduction": "There is a wealth of calibrated astronomical spectra accessible in current data archives, however most of them are not suitable for direct physical analysis. Being pre-reduced by some automatic pipeline or individually by manual reduction, they are mostly stored in archives in form of FITS file or ASCII table as the relation of intensity in arbitrary numbers (instrumental counts, digital numbers etc.) and wavelength (or frequency). The scientific analysis of such spectra requires further processing by the variety of different methods. In certain studies a huge number of spectra has to be collected from different servers (e.g. in different spectral regions ) and transformed into common units. The clear definition of metadata description and easy unit conversion as well as transparent lookup and download of spectra in Virtual Observatory has a great potential to become a new advanced way of astronomical research. The connection of powerful distributed computing available by GRID, Web Services and VO protocols can establish an innovative research environment allowing the data mining of huge datasets. ", "conclusions": "Astronomical spectroscopy uses a wide range of techniques with different level of complexity to achieve its final goal --- to estimate the most precise and reliable information about celestial objects. The large part of spectroscopic analysis today has been accomplished by several independent non VO-compatible legacy packages, where each works with different local files in its own data format. Analysis of large number of spectra is thus very tedious work requiring good data bookkeeping. Accomplishing the analysis in VO infrastructure may benefit from automatic aggregation of distributed archive resources (e.g. the multispectral research), seamless on-the-fly data conversion, common interoperability of all tools (using PLASTIC protocol) and powerful graphical visualisation of measured and derived quantities. The currently available VO clients supporting the SSA protocol can provide with only basic functions in a interactive environment. The more advanced work with spectra (e.g. automatic preprocessing or advanced analysis) is not supported in VO at all. The most of the legacy applications should be turned into the VO server side services and the conversion of legacy scripts and recipes into Workflows will allow their easy deployment on GRIDs. Modern techniques of astronomical analysis require the considerable amount of computing power and very offten the iterative comparison with theoretical models is indispensable. The properly designed VO service getting both observed and synthetic spectra transparently from VO resources and processing the number of objects in parallel on GRID might become a killer application turning the attention of wide astronomical community to the Virtual observatory as a viable and innovative way of modern astronomical research. By introduction of modern VO-aware tools into the astronomical spectral analysis a remarkable increase of effectiveness of astronomical research can be achieved." }, "1112/1112.2740_arXiv.txt": { "abstract": "{ Studies have suggested that there is a strong correlation between the masses of nuclear star clusters (NSCs) and their host galaxies, a correlation which said to be an extension of the well-known correlations between supermassive black holes (SMBHs) and their host galaxies. But careful analysis of disk galaxies -- including 2D bulge/disk/bar decompositions -- shows that while SMBHs correlate with the stellar mass of the \\textit{bulge} component of galaxies, the masses of NSCs correlate much better with the \\textit{total} galaxy stellar mass. In addition, the mass ratio $\\mnsc/\\mtot$ for NSCs in spirals (at least those with Hubble types Sc and later) is typically an order of magnitude smaller than the mass ratio $\\mbh/\\mbulge$ of SMBHs. The absence of a universal ``central massive object'' correlation argues against common formation and growth mechanisms for both SMBHs and NSCs. We also discuss evidence for a break in the NSC--host galaxy correlation: galaxies with Hubble types earlier than Sbc appear to host systematically more massive NSCs than do types Sc and later. ", "introduction": "As far as we can tell, all massive galaxies in the local universe harbor supermassive black holes (SMBHs, with masses $\\mbh \\sim 10^{6}$--$10^{9} \\Msun$). The masses of these SMBHs correlate strongly with several global properties of the host galaxies, particularly with the central velocity dispersion \\sigzero{} \\citep{ferrarese00,gebhardt00} and with the bulge luminosity or mass \\citep[e.g.,][]{marconi03,haring04}. These correlations imply that the processes which drive galaxy growth and the processes which drive black hole growth are intimately linked -- perhaps even the \\textit{same} processes. It is now also clear that many galaxies, particularly later-type spirals, host luminous nuclear star clusters \\citep[NSCs; e.g.,][]{carollo97,boker02}, with masses in the range $10^{5}$--$10^{8} \\Msun$; see the review by \\citet{boker08} for more details. Recently, several authors have argued that NSCs and central SMBHs have the \\textit{same} host-galaxy correlations: in particular, that SMBHs and NSCs have the same correlation with bulge luminosity and mass \\citep{wehner06,ferrarese06,cote06,rossa06} (but see \\citet{balcells07}). The suggestion, then, is that NSCs and SMBHs are in a sense members of the same family of ``Central Massive Objects'' (CMOs), and thus that they may have grown via the same mechanisms \\citep[e.g.,][]{mclaughlin06,li07,nayakshin09,devecchi10}. We argue, however, that one should be cautious about assuming that NSCs and SMBHs are really part of the same family, with the same host-galaxy relationships. To begin with, the samples of \\citet{wehner06} and \\citet{ferrarese06}, which were used to make the CMO argument, were almost entirely early-type galaxies -- mostly ellipticals and dwarf ellipticals. These are galaxies which are, in essence, ``pure bulge'' systems, so one could just as easily argue for a correlation with \\textit{total} galaxy mass. But we know that SMBHs in \\textit{disk} galaxies correlate better with just the bulge, and \\textit{not} with the total galaxy mass or light \\citep[e.g.,][]{kormendy-gebhardt01,kormendy11}. Given that there have been previous claims that NSCs in spiral galaxies correlate with the \\textit{total} galaxy light \\citep[e.g.,][]{carollo98}, we are prompted the ask the question: do nuclear clusters \\textit{in disk galaxies} correlate with the bulge (like SMBHs), or with the whole galaxy? ", "conclusions": "" }, "1112/1112.5356_arXiv.txt": { "abstract": "The most extended, closed magnetic loops inferred on T Tauri stars confine hot, X-ray emitting plasma at distances from the stellar surface beyond the the X-ray bright corona and closed large-scale field, distances comparable to the corotation radius. Mechanical equilibrium models have shown that dense condensations, or ``slingshot prominences'', can rise to great heights due to their density and temperatures cooler than their environs. On T Tauri stars, however, we detect plasma at temperatures hotter than the ambient coronal temperature. By previous model results, these loops should not reach the inferred heights of tens of stellar radii where they likely no longer have the support of the external field against magnetic tension. In this work, we consider the effects of a stellar wind and show that indeed, hot loops that are negatively buoyant can attain a mechanical equilibrium at heights above the typical extent of the closed corona and the corotation radius. ", "introduction": "Observations of T Tauri Stars (TTS) have revealed the presence of magnetic structures much larger than those seen on our Sun. Indeed, analysis of these structures has invoked solar flare-based models: to infer the length scales of confining loops, \\citet{Reale:1997} developed the uniform cooling loop (UCL) model, linking solar flare X-ray decay slopes to spatially resolved lengths of the loops confining the emitting gas. The UCL model was applied to data from the Chandra Orion Ultradeep Project \\citep[COUP,][]{Getman:2005}, and it was discovered that evidently the post-flare loops of TTS can reach up to multiple stellar radii in extent \\citep[][results summarised in Figure \\ref{F-f05loops}]{Favata:2005}. This result is surprising: T Tauri coronae are typically believed to be compact, as evidenced by the common detection of rotationally modulated coronal X-ray emission \\citep{Flaccomio:2005}. X-ray bright regions are generally thought to be confined within the corona, with loops generally not reaching more than a stellar radius in height. In the mechanical equilibrium model of \\citet{Jardine:2005}, prominences form after reconnection occurs in the open field, and it is a straightforward matter to show that cool, dense loops can reach heights well beyond corotation (the radius at which the orbital velocity equals the stellar rotational velocity) and the closed stellar coronal field. Hot loops such as those inferred by the COUP, however, appear to need the support of the stellar field, and thus should be limited to heights within the closed coronal magnetic field: a fraction of the heights inferred by UCL models. It was initially suggested that perhaps these hot, large-scale loops are anchored to circumstellar disc material for stability. Testing this idea, followup studies of the COUP objects that searched for discs were unable to find conclusive evidence for dusty disc material within reach of the loops \\citep[e.g.,][]{Getman:2008,Aarnio:2010}. Furthermore, the apparent anti-correlation of large-scale loops and close-in disc material suggests it is the absence of inner disc material that allows the loops to reach greater extents than they would were close-in disc material present \\citep[as suggested by][close-in circumstellar discs could strip away these stars' outer coronae]{Jardine:2006}. Potentially, this indicates some missing physics in earlier models, and it raises the question of how such large loops remain stable for the duration of the X-ray flare decay phase (in some cases, multiple rotation periods). \\begin{figure} \\begin{center} \\includegraphics[width=85mm,clip=true,trim=30.0 0 10. 20.]{f05loops.eps} \\end{center} \\caption{A summary of the results of \\citet{Favata:2005}: inferred loop height as a function of measured loop temperature. Vertical lines connect the loop height to that star's corotation radius. We have scaled the figure for ease of viewing; the tallest loop (its vertical corotation line shown) is $\\sim$63 R$_{*}$ in height. Filled points denote cases in which any parameters in calculation of the corotation radius were assumed (i.e., any values of R$_{*}$, M$_{*}$, or P$_{rot}$ not drawn from surveys of the ONC; fiducial values were used as discussed in section \\ref{SS-fiducial}). The grey horizontal line at 3 R$_{*}$ indicates the location of the source surface (the extent of the closed coronal magnetic field) in our models. There is a pileup of points at 270 MK because, to be conservative, the authors adopted a maximum best-fit temperature and applied it to all cases above a given threshold (beyond this temperature, the fits to their X-ray data are statistically equivalent, and a higher temperature simply yields a shorter loop).} \\vspace{-15pt} \\label{F-f05loops} \\end{figure} The potential ramifications of highly energetic events on stellar and circumstellar evolution are far-reaching: large-scale magnetic structures, in their formation and eventual disruption, could regulate stellar angular momentum evolution as well as circumstellar disc evolution, and thus planet formation. The combination of the high frequency with which large-scale magnetic structures have been inferred and the great heights at which they form above the stellar surface indicates circumstances are ideal for the generation of substantial torques opposing stellar rotation. Large-scale magnetic reconnection, in addition to shedding mass and angular momentum, also bombards circumstellar material with high energy photons and particles. This could potentially play a pivotal role in planet formation by stimulating grain growth \\citep[cf. flash-heated chondrule formation,][]{Miura:2007}. High energy radiation and particle fluxes can influence disc chemistry \\citep[cf., solar nebula isotopic abundances,][]{Feigelson:2002} and structure: \\citet{Glassgold:1997} showed that hard stellar X-rays could potentially ionise the disc sufficiently for accretion to occur via magnetohydrodynamic instability. The large heights inferred for these loops would put X-ray emitting plasma closer to the disc, reducing the column depth through which the hard X-rays would have to travel before ionising disc material. Given these potential impacts of large scale magnetic loops (and their destabilization) on early star and disc evolution, we aim to ascertain whether the existence of hot, large, magnetic loops, as inferred from the analysis of the brightest, most energetic events in X-ray surveys of TTS, is physically plausible. To do this, we address an evident buoyancy defecit in the previous models by including a stellar wind in the model physics. We do not distinguish between the single loop UCL model or the possibility of multiple loop events; we only seek to demonstrate that it is possible to find hot loops beyond the typical ``quiescent'' X-ray emitting corona, and in some cases, beyond corotation. ", "conclusions": "\\label{S-discus} We have addressed the issue of mechanical equilibrium for large, hot loops. Since cases have been found in which hot loops several stellar radii in height seem to not be disc-supported, we propose that these loops reach mechanical equilibria beyond the closed coronal magnetic field due to the stellar wind. We have shown that, for a generic T Tauri star, our equilibrium loop heights are consistent with the inferred loop lengths for the most powerful flares observed by the COUP. We also recover solutions for hot loops within the range of heights below the source surface anticipated by previous models, suggesting that large and small stable post-flare loops can occur under similar conditions. In the physics of previous models, which were hydrostatic and did not consider the role of stellar winds, loops hotter than the quiescent corona were confined to heights within the closed coronal field, needing its support. Our present work indicates that, given a stellar wind, hot loops can be found at heights well beyond the source surface, and even beyond corotation. Young stars are known to be very magnetically active in the T Tauri phase, with frequent flaring and high X-ray luminosities. The apparent abundance of large scale loops, favourable conditions for their formation in TTS coronae and winds, and potential for consequence on pre-main sequence stellar (and circumstellar) evolution makes understanding these large-scale loops very important. Having verified that such large-scale magnetic structures can indeed exist, interesting subsequent work will be to determine how the loops' disruption can impact circumstellar disc material as well as the star itself. Large-scale disruptions of this nature could effect stellar rotational evolution and activity as well as circumstellar disc evolution and planet formation." }, "1112/1112.5483_arXiv.txt": { "abstract": "We investigate non-spherical behavior of gas accreting onto a central supermassive black hole. Assuming optically thin conditions, we include radiative cooling and radiative heating by the central X-ray source. Our simulations are performed using the 3D SPH code GADGET-3 and are compared to theoretical predictions as well as to 1D simulations performed using the grid code ZEUS. As found in earlier 1D studies, our 3D simulations show that the accretion mode depends on the X-ray luminosity ($L_X$) for a fixed density at infinity and accretion efficiency. In the low $L_X$ limit, gas accretes in a stable, spherically symmetric fashion. In the high $L_X$ limit, the inner gas is significantly heated up and expands, reducing the central mass inflow rate. The expanding gas can turn into a strong enough outflow capable of expelling most of the gas at larger radii. For some intermediate $L_X$, the accretion flow becomes unstable developing prominent non-spherical features. Our detailed analysis and tests show that the key reason for this unstable non-spherical nature of the flow is thermal instability (TI). Small perturbations of the initially spherically symmetric accretion flow that is heated by the intermediate $L_X$ quickly grow to form cold and dense clumps surrounded by overheated low density regions. The cold clumps continue their inward motion forming filamentary structures; while the hot infalling gas slows down because of buoyancy and can even start outflowing through the channels in between the filaments. We measured various local and global properties of our solutions. In particular, we found that the ratio between the mass inflow rates of the cold and hot gas is a dynamical quantity depending on several factors: time, spatial location, and $L_X$; % and ranges between $0$ and $4$. We briefly discuss astrophysical implications of such TI-driven fragmentation of accreting gas on the formation of clouds in narrow and broad line regions of AGN, the formation of stars, and the observed variability of the AGN luminiosity. ", "introduction": "\\label{sec-intro} Active galaxies are believed to be powered by accretion of matter onto the central supermassive black holes (SMBHs) \\citep[e.g.,][]{Salpeter64, LyndenBell69, Blandford76, Ozernoi78, Balick82, Rees84, Fabian90, Kormendy95, Kauffmann00, Ferrarese05}. Active Galactic Nuclei (AGN) are argued to influence cosmological galaxy formation in the form of feedback, whereby the overall properties of a galaxy can be regulated by its central BH \\citep[e.g.,][]{Silk98, King03, Wyithe03, Granato04, Murray05, Begelman05, Best07, Ciotti07, Pipino09, Ostriker10, DiMatteo11}. Observationally AGN are characterized as a compact continuum source around a central SMBH, surrounded by a much larger emission line region consisting of multiphase gas clouds. The continuum emission shows a % relatively flat spectral energy distribution over a broad wave-band from radio to gamma rays, and is modeled as coming from the central engine composed of a gas accretion disk of size $10^{-4} - 10^{-3}$ pc. The broad-line region (BLR) extends up to $0.01 - 1$ pc and is composed of higher-density, higher-velocity gas; while the outer narrow-line regions (NLR) are considered to be located at distances up to $10 - 1000$ pc from the central SMBH. Modeling gas accretion onto SMBHs and resulting AGN feedback in a cosmological context is computationally challenging, because a large dynamic range of length scales is involved: BH accretion on sub-pc scale, to galaxy physics on hundreds of kpc scale. Therefore SMBH accretion and feedback is usually treated by a sub-grid model in cosmological simulations \\citep[e.g.,][]{DiMatteo08, Booth09}. At the same time, gas accretion flow within the Bondi radius of a SMBH has just started to be resolved in observations \\citep{Wong11}. Hence numerical studies resolving the Bondi radius are required for a meaningful comparison, and it would be % useful to improve a sub-grid model of AGN feedback for cosmological simulations based on the results of small-scale simulations. In \\citet[][hereafter Paper~I]{Barai11}, we studied spherically-symmetric Bondi accretion of gas onto a SMBH, % and started to explore the effects of radiative heating and cooling. A number of recent studies resolving the Bondi radius have focused on accretion onto intermediate-mass BHs regulated by feedback from X-ray and UV radiation emitted isotropically near the BH. Analytical work \\citep{Milosavljevic09b} as well as 1D and 2D simulations \\citep{Milosavljevic09, Park11, Park12} find that photoionization heating and radiation pressure significantly reduce the steady-state accretion rate to a fraction of the Eddington rate, or over 2 orders of magnitude below the Bondi rate. However, in 1D simulations by \\citet{Li11} the self-gravity of the gas eventually overcomes the radiative feedback effects, and boosts the accretion to the Eddington rate after one free-fall timescale. \\citet{Park11, Park12} also find the development of Rayleigh-Taylor instabilities in their 2D simulations. In these work the BH luminosity is made to be proportional to the rate at which gas accretes across the inner boundary, whereas in our study the central luminosity has a constant value in each simulation because we want to have a simplified setup with a non-variable source to enable us cleanly investigate the physical processes influencing the accretion flow. In Paper~I we discovered indications of thermal instability (TI) in some runs, which we investigate in detail in the current work. Studying TI in a dynamical flow in 3D is challenging, because the effects of a gravitational field significantly complicate the development of TI \\citep[e.g.,][]{Balbus89}. The foundations for understanding astrophysical thermal instability were laid in the classical paper by \\citet{Field65}. The general criterion for TI to occur is: \\begin{equation} \\label{eq-TI-criterion} \\left [ \\frac{\\partial {\\cal{L}}} {\\partial S} \\right ]_A < 0, \\end{equation} where ${\\cal{L}}$ is the net cooling-heating rate, $S$ is the specific entropy, and $A$ is a thermodynamic variable that is kept constant for a given perturbation. Subsequently the theory of TI has been studied over decades, including analytical developments: local TI in static and dynamical systems \\citep[e.g.,][]{Balbus86, Balbus89, Balbus95}, global TI by linear perturbation analysis \\citep[e.g.,][]{Yamada01}, and numerical hydrodynamical calculations to investigate the perturbations in the non-linear regime \\citep[e.g.,][]{Hattori90}. TI has been invoked to explain observed features in various astrophysical domains: solar prominences \\citep[e.g.,][]{Parker53, Karpen88}; formation of cold gas clouds and clumpy substructure in the interstellar medium \\citep[e.g.,][]{Schwarz72, Parravano87, Hennebelle99, Burkert00, Koyama02, Sanchez-Salcedo02, Inutsuka05}; origin of globular clusters during the collapse of a protogalaxy \\citep[e.g.,][]{Fall85}; formation of galaxies \\citep[e.g.,][]{Gold59, Sofue69}; general cooling flows in galaxies and galaxy cluster cores \\citep[e.g.,][]{Mathews78, Nulsen86, Malagoli87, Meiksin88, Yoshida91, Guo08}; the multi-phase structure of cold filaments condensing out of the hot intracluster medium in galaxy clusters and groups \\citep[e.g.,][]{Sharma10, McCourt11}. More relevant for our work are the applications of TI in black hole accretion systems. Spherically symmetric accretion flow onto a compact object is preheated by radiation from the central X-ray source. \\citet{Ostriker76} showed the preheating suppresses accretion, such that above a X-ray luminosity $0.005$ of the Eddington limit steady flow is impossible. \\citet{Cowie78} found time-dependent behavior of the preheated accretion for a much wider range of source parameters, and the presence of instabilities. However, \\citet{BisnovatyiKogan80} found that numerical solutions exist for any luminosity below the Eddington limit, but they did not investigate the stability of the solutions in detail. Performing stability analyses \\citet{Stellingwerf82} found globally stable solutions, but those become subject to a drift instability at luminosities over the Eddington limit exceeding $\\sim 0.03$. Hydrodynamical simulations by \\citet{Proga07} show that with an accretion luminosity of $0.6$ of the Eddington value, the gas flow settles into a steady state with two components: an equatorial inflow, and a bipolar inflow/outflow with the outflow leaving the system along the disk rotational axis. There the temperature at the outer radius was fixed at a relatively high value ($2 \\times 10^{7}$ K), so the gas did not go through a thermally unstable phase. Also \\citet{Park12} find that a quasi-steady BH accretion is possible at the Eddington luminosity when the gas density is above a critical value inversely proportional to the BH mass. Investigating spherical accretion of gas irradiated by a quasar-like continuum, \\citet{Krolik83} found specific criteria for when TI may disrupt single-phase steady flow, in the luminosity -- efficiency (of mass to energy conversion) plane. \\citet{Krolik88} argued that broad emission line clouds in AGN originates in cooler condensations forming within a hot medium via finite-amplitude TI. However, \\citet{Mathews90} concluded that broad line-emitting clouds are unlikely to arise from thermally unstable condensations in a hotter medium, but are likely composed of higher-density gas ejected from nearby stars or from a dense accretion disk. \\citet{Wang12} investigated the dynamics of clumps, considered to exist by thermal instability, embedded in and confined by advection-dominated accretion flows onto black holes. We find that a spherically accreting gas distribution acted upon by a central gravitational potential, undergoing radiative heating and cooling, can be thermally unstable, cool non-spherically and become clumpy. Hence results from our simulations show that fragmentation can occur in a simple system with minimal physical processes. Our study can give insight into the formation and evolution of clouds % near AGN. TI operating in such a way can be responsible for the origin of two phase medium in the NLR, support star-formation near AGN, contribute to the observed variability of the luminosity. % This paper is organized as follows. We describe our simulations in \\S\\ref{sec-numerical}. We present and explain the results in \\S\\ref{sec-results}, and discuss in \\S\\ref{sec-discussion}. We summarize and conclude in \\S\\ref{sec-conclusion}. ", "conclusions": "\\label{sec-conclusion} Gas accretion onto a central SMBH was studied in Paper~I by performing simulations of spherical Bondi accretion using the 3D SPH code GADGET-3. The simulations included radiative heating by a central X-ray corona and radiative cooling. All the results presented in Paper~I were of spherically-symmetric gas properties. However, for a certain parameter range of X-ray luminosity $L_{X}/L_{\\rm Edd} = 0.01, 0.02, 0.05$ with a fixed density at the outer boundary $\\rho_{\\infty} = 10^{-23}$\\,g/cm$^3$ (i.e., within the unstable $\\xi$-range, \\S\\ref{sec-TI}), we observe non-spherical, multi-phase gas motion. In the current work we present detailed analyses of the inhomogeneous gas behavior. The main results are summarized below: (1) The gas goes through various modes of accretion for different $L_{X}$. There is a stable, spherically-symmetric inflow with $L_{X} = 0.005$ (Run 25). The gas is heated up and expands as $L_X$ is increased, which decreases the central mass inflow at a varying rate. For some intermediate $L_X$ the accretion flow undergoes unstable motion developing non-spherical features. With high enough $L_X$ a strong outflow is produced, ejecting most of the gas outside. (2) With $L_X = 0.01$ and $0.02$, the gas cools and becomes denser in a non-spherical manner, fragments, and forms multiple clumps. The clumping occurs over $r < 30$\\,pc for $L_{X} = 0.01$, and $r \\sim 40 - 100$\\,pc for $L_{X} = 0.02$. It starts to develop from $\\sim 1.6 - 1.7$ Myr, and continues up to the simulation end. The clumps are filamentary in structure, supersonic, and have a lower photoionization. Hotter, lighter, subsonic gas exist in-between the filaments, and start to flow outward at later times. The clumps undergo a dynamic evolution: become stretched, fragment, merge with each other, finally accreting into $r_{\\rm in}$. The gas particle properties versus radius show large scatters (few $100$s to $1000$) at the radial ranges where fragmentation occurs, corresponding to the multi-phase structure of the gas. The mass inflow rate shows large fluctuations in the run with $L_X = 0.01$ after $1.6$\\,Myr, and in the run with $L_X = 0.02$ after $2.3$\\,Myr, because of the accretion of multi-phase gas, with the spikes corresponding to the infall of cold dense clumps. (3) As they move inward the cooled clumps remain denser, however heat up to above $10^5$\\,K. The central heating at $r < 1$\\,pc is mostly by adiabatic compression reaching $T > 10^7$\\,K at $r_{\\rm in}$, while radiative heating would dominate at a few pc for some particles. The photoionization parameter increases while moving to smaller radii, and the clumps remain supersonic. Artificial viscosity heating increases in magnitude toward $r_{\\rm in}$, however it always remains sub-dominant compared to adiabatic heating. This is in contrast to Paper~I due to the different conditions here because of the higher $L_{X}$. (4) The cooling and clumping is caused by the interplay of thermal instability in the inflowing gas acted upon by radiative processes, as well as being adiabatically compressed and heated. Gas particles in our simulations undergo TI within the unstable $\\xi$-range, where the $\\Trad - \\xi$ equilibrium curve has the steepest slope (between $\\xi \\sim 100 - 500$), and most physically relevant processes (those with constant-pressure or free falling) can give rise to perturbations with a shallower slope. We detect a distinct mode of cooling/heating in the $[T - \\xi]$ plane. A particle while being dominated by adiabatic process moves up to higher-$T$ along the unstable part of $\\Trad - \\xi$. This happens because the adiabatic free-fall relation of $T$ versus $\\xi$ ($T_{ff, a} \\propto \\xi_{ff, a}^2$) has the same slope as the radiative equilibrium curve in the unstable regions. At a later time the particle has a transition to being dominated by radiative process. Then classical TI operates, and it undergoes catastrophic cooling or heating (whichever dominates), and moves away from the unstable $\\Trad - \\xi$ part. It finally settles on one of the stable branches where $\\Trad - \\xi$ has a shallower slope. (5) Typically the amplitude of temperature and density perturbations versus time shows that the perturbations have an exponential growth at first, and then come to a saturation for the remaining simulation run. The minimum temperature has an exponential decrease initially, indicating the growth of cooling instability. It remains almost constant after that at the temperature of the lower stable $\\Trad - \\xi$ branch, where $T$ is a weak function of $\\xi$. In terms of density the maximum value rises indicating the growth of cooled, dense clumps caused by TI. (6) We have multi-phase medium (cold gas with $2 \\times 10^4$\\,K and hot gas with $10^6 - 10^7$\\,K) at the same radius, forming self-consistently in the simulation volume. Furthermore we resolve the Bondi radius of the accretion flow. This can be used to constrain the $\\alpha$-parameter that is multiplied to the Bondi-Hoyle-Lyttleton mass accretion rate in the sub-grid model of SMBH feedback in large-scale galaxy simulations. Considering a limiting temperature of $10^5$\\,K between cold and hot phases, we compute the ratio ($\\chi$) of mass inflow rates of cold gas over hot. We find that $\\chi$ is a dynamical quantity depending on several factors: time, spatial location, and X-ray luminosity. The maximum value of $\\chi$ varies between $1 - 5$. At a fixed radius, $\\chi$ increases with time initially, and later remains almost-constant at a value $\\chi \\sim 3 - 4$. Thus we do not find a factor as large as $\\alpha \\sim 100$, since our $\\chi \\leq 1 - 5$. In other words, in our simulations we do not find such a large fraction of gas hidden in the cold phase that is unresolved in cosmological simulations. (7) With $L_X = 0.02$ (Run 27) we see both fragmented cold clumps and buoyantly rising hot bubbles. Initially there is a central spherical outflow at $r < 40$\\,pc, which decreases the mass inflow rate at $r_{\\rm in}$ by $\\sim 1000$ times between $1.4 - 2.2$\\,Myr. The surrounding inflowing gas between $[40 - 70]$\\,pc cools and fragments by TI. The dense clumps grow with time, and continue to move inward, disrupting the spherical symmetry of the central hotter outflow. The clumps start accreting into $r_{\\rm in}$ at $t \\geq 2.2$\\,Myr, causing the central mass inflow rate to rise again. The gas motion becomes inhomogeneous with cooler, denser clumps falling in to the center but heating up as they move in; and hotter, lighter gas moving out. There are also few hot bubbles buoyantly rising from the center. (8) With a high enough value of the X-ray luminosity ($L_X = 0.05$ in Run 28), the gas is heated up strongly, creating a spherically symmetric outflow. This expels a large fraction of gas out of $r_{\\rm out}$, reduces the central mass inflow rate drastically, and produces a gas density profile that decreases towards center. There is also the formation of an anisotropic, hot bubble buoyantly rising from the center. It has a narrow elongated structure, while its head has a bow shock-like morphology, where the interior gas collides with the relatively cold ambient medium and produces a shock region larger than the jet. The bubble has a greater outward velocity, lower density, higher temperature and higher entropy compared to the gas at the same radius." }, "1112/1112.2430_arXiv.txt": { "abstract": "A spectrum of cosmic rays within energy range $10^{15} - 3 \\times 10^{17}$~eV was derived from the data of the small Cherenkov setup, which is a part of the Yakutsk complex EAS array. In this, work a new series of observation is covered. These observations lasted from 2000 till 2010 and resulted in increased number of registered events within interval $10^{16} - 10^{18}$~eV, which in turn made it possible to reproduce cosmic ray spectrum in this energy domain with better precision. A sign of a thin structure is observed in the shape of the spectrum. It could be related to the escape of heavy nuclei from our Galaxy. Cosmic ray mass composition was obtained for the energy region $10^{16} - 10^{18}$~eV. A joint analysis of spectrum and mass composition of cosmic rays was performed. Obtained results are considered in the context of theoretical computations that were performed with the use of hypothesis of galactic and meta-galactic origin of cosmic rays. ", "introduction": "Energy spectrum of cosmic rays (CR) in energy range $3 \\times (10^{15} - 10^{18})$~eV could not be studied in detail with compact arrays due to their small acceptance at energy above $10^{17}$~eV. At the same time this area of the spectrum is of a great interest, since local irregularities are manifested there: production of kinks (thin structure at $3 \\times 10^{15} - 10^{17}$~eV) arising from non-uniform distribution of heavier CR components in our Galaxy. On the other hand, this effect is smoothed by addition of a new component (of meta-galactic or other origin) to the cosmic ray flux near Earth. As a result, presence/absence of significant irregularities in spectra measured by various compact arrays allows one to speculate on the CR origin and propagation in our Galaxy~\\cite{Knurenko2011, Kalmykov2008}. The Yakutsk array in this sense appears as a unique scientific tool. It is related to medium-sized arrays, capable of effective measuring of cosmic rays flux in a wide energy range ($10^{15} - 10^{19}$~eV). Other important traits of the array are its model-independent technique of energy estimation of extensive air showers (EAS) and the ability to track longitudinal EAS development by detecting the Cherenkov light emission. Factors mentioned above enable adopting the unique method, combining the studies of CR spectrum and mass composition aimed at exploration of astrophysical aspect of cosmic rays~\\cite{Knurenko2006, Knurenko2007}. ", "conclusions": "" }, "1112/1112.6025_arXiv.txt": { "abstract": "In the past decade or so observations of supernovae, Large Scale Structures (LSS), and the Cosmic Microwave Background (CMB) have confirmed the presence of what is called dark energy - the cause of accelerating expansion of the Universe. They have also measured its density as well as the value of other cosmological parameters according to the concordance $\\Lambda$CDM model with few percent uncertainties. Next generation of surveys should allow to constrain this model with better precisions, or distinguish between a $\\Lambda$CDM and alternative models such as modified gravity and (interacting)-quintessence models. In this work we parametrize both homogeneous and anisotropic components of matter density in the context of interacting dark energy models with the goal of discriminating between $f(R)$ modified gravity and its generalizations, and interacting dark energy models, for which we also propose a phenomenological description of energy-momentum conservation equations inspired by particle physics. It is based on the fact that the simplest interactions between particles/fields are elastic scattering and decay. The parametrization of growth rate proposed here is nonetheless general and can be used to constrain other interactions. As an example of applications, we present an order of magnitude estimation of the accuracy of the measurement of these parameters using Euclid and Planck surveys data, and leave a better estimation to a dedicated work. ", "introduction": "\\label{sec:frieman} Apriori the measurement of the equation of state of dark energy is simple. It is enough to measure the expansion rate of the Universe $H(z) \\equiv \\dot {a}(z)/a$, or a quantity related to it such as the luminosity distance $D(z)$ at different redshifts. Then, by modeling known constituents of the Universe as non-interacting perfect fluids, one can fit the data and measure the effective equation of state of dark energy $w_{eff}(z)$, defined as $P_{eff}(z)/\\rho_{eff}(z)$. The suffix {\\it ``eff''} is used to remind that pressures and densities obtained in this way can be effective quantities rather than physical pressure and density of constituents, because we have neglected any interaction between components. Therefore, from now on {\\it effective quantities} mean quantities determined from data by considering a null hypothesis. In practice however the life is not so simple. The density of a perfect fluid changes with redshift as $(1+z)^{3\\gamma}$ ($\\gamma$ is defined in (\\ref{gammade})). Therefore, at low redshifts when $z \\rightarrow 0$, the total density is not very sensitive to the value of $\\gamma$ or equivalently $w(z)$ and their variation with $z$, see Appendix \\ref{app:a} for more details. This statement is independent of the type of data or proxy used for determining $H(z)$ or $D(z)$. On the other hand, at high redshifts where $H(z)$ is more sensitive to the equation of state, dark energy is subdominant. Moreover, it is more difficult to measure $H(z)$ and $D(z)$ at higher redshifts and measurement uncertainties can make the estimation of $w(z)$ and its evolution unusable for discrimination between models. If constituents of the Universe do not interact with each other, Friedman equation which determines the evolution of expansion function $a(t)$ can be written as: \\bea &&\\frac{H^2}{H_0^2} = \\frac {\\rho (z)}{\\rho_0} = \\Omega_m (1+z)^3 + \\Omega_h (1+z)^4 + \\Omega_K (1+z)^2 + \\Omega_{de} (1+z)^{3\\gamma(z)}, \\quad {\\rho_c (z)} \\equiv \\frac{3H^2}{8\\pi G} \\label{friedmannoint} \\\\ &&\\gamma (z) = \\frac{1}{\\ln (1+z)} \\int_0^z dz'\\frac{1 + w(z')}{1+z'}, \\quad P_{de} (z) \\equiv w(z) \\rho_{de} \\label{gammade} \\\\ && \\text{\\it m = cold dark matter, b = baryons, h = hot matter, k = curvature, and de = dark energy} \\nonumber \\eea In this class of models matter and radiation densities evolve only due to the expansion. This is a good approximation for all redshifts $z < z_{cmb} \\sim 1100$. In interacting dark energy models matter and radiation terms in the right hand side of the Friedman equation (\\ref{friedmannoint}) can contain an additional redshift-dependent factor: \\be \\frac{H^2}{H_0^2} = \\frac {\\rho_c (z)}{\\rho_{c0}} = \\sum_i \\Omega_i {\\mathcal F}_{i}(z) (1+z)^{3\\gamma_i} \\quad \\quad i=\\text{{\\it m, b, h,k,} and {\\it de}} \\label{friedmanintde} \\ee Without lack of generality we assume that ${\\mathcal F}_{de} = 1$ and all redshift dependent terms are included in $\\gamma (z)$. In quintessence models the coefficient of the curvature term also is constant because it is assumed to be related to geometry/gravity and independent of the behaviour of other components. At present observations are consistent with only gravitational interaction between various components in (\\ref{friedmanintde}), thus additional interactions must be very weak. By definition and without lack of generality we consider ${\\mathcal F}_i(z=0) = 1$. Observations also show that $\\Omega_k \\approx 0$, therefore throughout this work we assume $\\Omega_k = 0$ unless it is explicitly mentioned. Note that in the case of modified gravity models, a parametrization similar to (\\ref {friedmanintde}) can be defined both in Einstein and Jordan frames. A simple example for which an approximate expression for ${\\mathcal F}_i(z)$ coefficients can be found is a model with a cosmological constant as dark energy and a slowly decaying dark matter. The decay remnants are assumed to be visible relativistic particles~\\cite{houridecay}. In this case: \\bea && \\frac{H^2}{H_0^2} \\approx {\\Omega}_m (1 + z)^3 \\exp (\\frac {{\\tau}_0 - t}{\\tau}) + {\\Omega}_b (1 + z)^3 + {\\Omega}_h (1 + z)^4 + {\\Omega}_m (1 + z)^4 \\biggl (1 - \\exp (\\frac {{\\tau}_0 - t}{\\tau}) \\biggr ) + {\\Omega}_{\\Lambda} \\nonumber \\\\ && \\label {totdens} \\\\ && {\\mathcal F}_m(t) \\approx \\exp (\\frac {\\tau_0 - t}{\\tau}) + (1+z) \\biggl (1 - \\exp (\\frac {t_0 - t}{\\tau}) \\biggr ), \\quad \\tau \\gg \\tau_0, \\quad {\\mathcal F}_b = {\\mathcal F}_h = 1 \\quad \\gamma (z) = 0 \\label{dmdecaycoeff} \\eea where $\\tau$ is the lifetime of dark matter and $\\tau_0$ is the age of the Universe. It is demonstrated that in this example, if the decay/interaction of dark matter is not considered in the data analysis, a $w_{eff} < -1$ can be obtained for dark energy, see~\\cite{houriustat,das05} for more details about the set up and the proof. Note that in (\\ref{dmdecaycoeff}), we have included the contribution of relativistic remnants in ${\\mathcal F}_m$. However, as this component has a redshift dependence similar to hot matter, it also makes sense to consider it as hot matter and add it to hot component. It is even possible to add this term to dark energy contribution, as long as it is small and induces only a slight deviation from a cosmological constant. In this case, one can show that the {\\it effective dark energy} will have $w_{eff} < -1$~\\cite{houriustat}. The reason for such freedom is that we do not measure or take into account the decay remnants. This example clearly shows that parametrization (\\ref{friedmanintde}) is not unique when all the components and their interactions are not know. Therefore, one has to be very careful about degeneracies when data are analyzed and interpreted. In particular, prior assumptions such as stability of matter and radiation components can affect measurements and conclusions. This example also show that for ruling out $\\Lambda$CDM model, it is enough to prove that at least one of ${\\mathcal F}_i \\neq 1$, or $\\gamma_{de} \\neq 0$\\footnote{This statement is true if baryon pressure is negligible. Future surveys can be sensitive to small baryon pressure. In this case it must be taken into account before any conclusion about $\\Lambda$CDM model is made.}. Extension of this example to quintessence models without coupling to matter is straightforward and one simply needs to consider $\\gamma (z) \\neq 0$. A more interesting extension is to assume that the quintessence scalar field is one of the remnants of the decay of dark matter, which during cosmological time condensates and makes a classical quintessence field. In this case, it has been shown~\\cite{houriquin,houriquin1,houricondens} that coefficients ${\\mathcal F}_m$, ${\\mathcal F}_h$, and equation of state of dark energy $w(z)$ (or equivalently $\\gamma (z)$) are not independent. However, their relations are too sophisticated and cannot be described in an analytical form and numerical techniques should be employed~\\cite {houriquin}. According to (\\ref{dmdecaycoeff}): \\be {\\mathcal F}_m(z) > {\\mathcal F}_m(z=0) \\label{quinfdmtode} \\ee and because $\\tau \\gg \\tau_0$, ${\\mathcal F}_i(z)$ coefficients are close to 1 at all redshifts. In general, for an interaction which transfers energy from dark matter to other components, the inequality (\\ref{quinfdmtode}) is applied because at high redshifts one expects a larger contribution of dark matter in the total density than in a non-interacting model. Inversely, if energy is transferred from other components, for instance from dark energy, to dark matter: \\be {\\mathcal F}_m(z) < {\\mathcal F}_m(z=0) \\label{quinfdetodm} \\ee An example of such models is {\\it scaling dark energy}~\\cite{scalingde,scalingde1} in which at early times dark energy has a much larger contribution in the total energy density, but it gradually decays to dark matter and only recently its equation of state approaches $w \\sim -1$. Another example is the class of models called {\\it early dark energy}. Although the original model~\\cite{earlyde,earlyde1,earlyde2} is a pure quintessence/k-essence, there are variants of this model in which, there is an interaction in the dark sector~\\cite{earlydeint} or between dark energy and visible sector~\\cite{earlydeintvarpalpha}. In models with elastic interaction between two sectors, no energy is transferred between them, and ${\\mathcal F}_m(z) = {\\mathcal F}_h(z) = 1$. Nonetheless, the phase space of matter and dark energy in these models can change and thereby $w_{de}$ can depend on $z$. For $\\mathrm {f}(R)$ modified gravity models homogeneous Einstein equations and energy conservation equation in Jordan frame are~\\cite{frbean}:\\footnote{When equations apply to both dark matter and baryons, we indicate them collectively with subscript $m$.} \\bea && (1+\\mathrm {f}_R) H^2 + \\frac{1}{6} \\mathrm {f} - \\frac{a''}{a^3}\\mathrm {f}_R + \\frac{\\mathrm {f'}_R}{a} H = \\frac {8\\pi G}{3} \\sum_i \\rho_i \\label{grmg0} \\\\ && \\frac {a''}{a^3} = -\\frac{4\\pi G}{3} \\sum_i (\\rho_i + 3P_i) + (1+\\mathrm {f}_R) H^2 + \\frac{\\mathrm {f}}{6} - \\frac {H\\mathrm {f}'_R}{a} - \\frac{\\mathrm {f}''_R}{2a^2} \\label{grmg1} \\\\ && \\dot{\\rho}_i + 3H\\rho_i = -\\frac{\\dot{f}_R}{2 (1+\\mathrm {f}_R)}(\\rho_i -3 P_i), \\quad i=\\text{\\it m, h, k} \\label{grgmener} \\eea where $a' = a\\dot{a}$.\\footnote{Here we have written Einstein and conservation equations in Jordan frame because they lead to expressions for ${\\mathcal F}_i$ coefficients which are explicitly very different from quintessence case.} Dot and prime mean derivation with respect to comoving and conformal time, respectively. Subscript $R$ means derivation with respect to scalar curvature $R \\equiv R_{\\mu\\nu}g^{\\mu\\nu}$. We remind that at linear order the effect of matter perturbations on $R$ is zero, thus $R$ only depends on $z$ and the effect of $\\mathrm {f}(R) \\neq 0$ on the evolution of perturbations manifests itself by changing the background cosmology. After solving density conservation equation (\\ref{grgmener}), Friedman equation (\\ref{grmg0}) can be written as the following: \\bea \\rho_i (z) &=& \\rho_i (z=0) (1+z)^{3\\gamma_i} \\biggl (\\frac{{1+\\mathrm {f}_R (z=0)}} {1+\\mathrm {f}_R (z)} \\biggr)^{-\\frac{1-3w_i}{2}} \\label{rhosolmg} \\\\ \\frac{H^2}{H_0^2} &=& \\frac {\\rho_c (z)}{\\rho_{c0}} = \\sum_i \\Omega_i {\\mathcal F}_{i}(z) (1+z)^{3\\gamma_i} \\label{friedmanmg} \\\\ \\rho_{de} &=& \\frac{3}{8\\pi G}~\\frac{1}{1+\\mathrm {f}_R}(-\\frac{\\mathrm {f}(R)}{6} + \\frac{a''}{a^3}\\mathrm {f}_R - H\\dot{\\mathrm {f}}_R) \\label{mgdedens} \\\\ {\\mathcal F}_i(z) &=& \\biggl (\\frac{1+\\mathrm {f}_R(R(z=0))}{1+\\mathrm {f}_R(R(z))}\\biggr)^ {-\\frac{1-3w_i}{2}}, \\quad w_m = 0, \\quad w_h = \\frac{1}{3}, \\quad w_k = -\\frac{1}{3} \\label{fmgdef} \\eea Equation (\\ref{mgdedens}) is the energy density of {\\it effective dark energy} in $\\mathrm {f}(R)$ gravity models. Similar to quintessence models we can assume ${\\mathcal F}_{de} = 1$. The only explicit difference between (\\ref{friedmanmg}) and the same equation for an interacting quintessence model is the presence of a nontrivial coefficient for the curvature term if $\\Omega_k \\neq 0$. Nonetheless, the evolution of coefficients ${\\mathcal F}_i(z)$ with redshift is different from their counterparts in interacting quintessence models, in particular from models in which energy is transferred to dark energy at low redshifts, see equation (\\ref{quinfdmtode}). In fact, the function $\\mathrm {f}(R)$ is not completely arbitrary and must satisfy a number of constraints. Notably, $\\mathrm {f}(R)|_{|R| \\gg 0} \\rightarrow 0$ to make the model consistent with Einstein theory of gravity in mild or strong gravity fields, and $\\mathrm {f}_R > 0$ due to stability constraint~\\cite{modgrstabil}. Under these conditions: \\be {\\mathcal F}_i(z) > {\\mathcal F}_i(z=0) \\label{mgfz} \\ee Comparing (\\ref{mgfz}), (\\ref{quinfdmtode}), and (\\ref{quinfdetodm}) one can immediately conclude that the measurement of ${\\mathcal F}_m(z)$ and its evolution with redshift can discriminate between dark energy models in which energy is transferred from dark energy to dark matter such as scaling models, and $\\mathrm {f}_R$ modified gravity models. But it cannot discriminate modified gravity from models in which energy is transferred from dark matter to dark energy such as the model discussed in~\\cite{houriquin,houriquin1,houricondens}. To discriminate the latter and other models of this category from $\\mathrm {f}_R$ modified gravity, the coefficient of relativistic (hot) component ${\\mathcal F}_h(z)$ and its evolution must be measured. Evidently, such measurements are very difficult. For instance, one has to measure very precisely the temperature of CMB at high redshifts or $H(z)$ at a large number of redshift bins and fit the data with ${\\mathcal F}_h \\neq 1$. In Einstein frame the evolution of matter density is the same as in equation (\\ref{rhosolmg})~\\cite{frbean}, but evolution equation of hot matter is similar to $\\Lambda$CDM. In what concerns the discrimination from interacting quintessence what is discussed from Jourdan is applicable. \\subsection {Model-independent discrimination of interacting dark energy models} \\label{sec:discrim} In this section we show that if $\\Lambda$CDM or a simple quintessence are considered as null hypothesis, measurements of effective dark energy density and effective equation of state from $H (z)$ and the function $A(z)$ defined in Appendix \\ref {app:a} separately, give different values for these quantities if dark energy interacts with matter. Similarity of ${\\mathcal F}_m(z)$, specially if the curvature of the Universe is zero, means that we cannot distinguish between interacting quintessence and modified gravity models in a model-independent manner - except for the cases explained above. For this reason in this section we only study the discrimination between interacting dark energy models parametrized as in equation (\\ref{friedmanmg}) and a cosmological constant and/or non-interacting quintessence. For analyzing cosmological data, $\\Lambda$CDM with a stable and non-interacting dark matter is usually used as null hypothesis. Nonetheless, the methodology explained below is not sensitive to redshift dependence of $\\gamma_{de}$, and we can consider the more general case of non-interacting quintessence as the null hypothesis. The expansion of the Universe for such cosmologies is ruled by equation (\\ref{friedmannoint}). Therefore, we rearrange terms in equation (\\ref{friedmanmg}) such that it looks similar to equation (\\ref{friedmannoint}). Then, we determine effective quantities which are measured by fitting a $\\Lambda$CDM or a non-interacting quintessence model to data: \\be \\frac{H^2}{H_0^2} = \\sum_i \\Omega_i (1+z)^{3\\gamma_i} + \\sum_i \\Omega_i ({\\mathcal F}_{i}(z) - 1) (1+z)^{3\\gamma_i} + \\Omega_{de} (1+z)^{3\\gamma_{de} (z)} \\label{friedmanmgequiv} \\ee In null hypothesis model only $\\gamma_{de}$ is redshift dependent and $\\gamma_i,~i=m,~h,~k$ are constant. By comparing (\\ref{friedmanmgequiv}) with (\\ref{friedmannoint}) the {\\it effective} contribution of dark energy is expressed as: \\be \\Omega_{eff}^{(H)} (1+z)^{3\\gamma_{eff}^{(H)}(z)} = \\sum_i \\Omega_i ({\\mathcal F}_i(z) - 1) (1+z)^{3\\gamma_i} + \\Omega_{de} (1+z)^{3\\gamma_{de} (z)} \\label{deeffdef} \\ee In both interacting quintessence and modified gravity models coefficients $F_i$'s are defined such that $F_i (z=0) = 1$, therefore at $z = 0$ the first term in (\\ref{deeffdef}) is null and we can separate $\\Omega_{eff}$ and $\\gamma_{de} (z)$: \\bea && \\Omega_{eff}^{(H)} = \\Omega_{de}, \\quad \\quad \\gamma_{eff}^{(H)}(z=0) = \\gamma_{de}(z=0) \\label{omegaffh} \\\\ && \\gamma_{eff}^{(H)}(z) = \\frac{\\log \\biggl (\\sum_i \\frac{\\Omega_i}{\\Omega_{de}} ({\\mathcal F}_i (z) -1)(1+z)^{3\\gamma_i} + (1+z)^{3\\gamma_{de} (z)}\\biggr )}{3 \\log(1+z)} \\label{deeffexp} \\eea where superscript $(H)$ means measured from Hubble constant $H$. Suppose we can also measure $A(z)$ defined in (\\ref{azdef}), and use it to determine the effective density and equation of state of dark energy. For an interacting dark energy model parametrized according to (\\ref{friedmanmgequiv}) quantities $B(z)$ and $A(z)$ are: \\bea B(z) &\\equiv& \\frac{1}{3 (1+z)^2 \\rho_0} \\frac {d\\rho}{dz} = \\nonumber \\\\ && \\sum_{i=m,h,k} \\Omega_i \\biggl (\\gamma_i {\\mathcal F}_i (z) + (1+z) \\frac{d{\\mathcal F}_i}{dz} \\biggr )(1+z)^{3 (\\gamma_i - 1)} + \\Omega_{de}(w(z)+1) (1+z)^{3(\\gamma_{de}(z) - 1)} \\label{bzintdmde} \\\\ A(z) &\\equiv& B(z) - \\sum_{i=m,h,k} \\Omega_i \\gamma_i (1+z)^{3(\\gamma_i - 1)} = \\nonumber \\\\ && \\sum_{i=m,h,k} \\Omega_i \\biggl (\\gamma_i ({\\mathcal F}_i (z) - 1) + (1+z) \\frac{d{\\mathcal F}_i}{dz} \\biggr ) (1+z)^{3 (\\gamma_i - 1)} + \\Omega_{de}(w(z)+1) (1+z)^{3(\\gamma_{de}(z) - 1)} \\label{azintdmde} \\eea Using (\\ref{azdef}) in Appendix \\ref{app:a} as the definition of $A(z)$, we find the following expression for its parameters: \\bea \\Omega_{eff}^{(A)} (w_{eff}^{(A)}(z) + 1) (1+z)^{3\\gamma_{eff}^{(A)}(z)} &=& \\sum_i \\Omega_i \\biggl (\\gamma_i ({\\mathcal F}_i (z) - 1) + (1+z) \\frac{d{\\mathcal F}_i}{dz} \\biggr ) (1+z)^{3\\gamma_i} + \\nonumber \\\\ && \\Omega_{de}(w(z)+1)(1+z)^{3\\gamma_{de}(z)} = (1+z) A(z) \\label{deeffdefaz} \\eea where superscript $(A)$ means measured from $A(z)$. Equations (\\ref{deeffdef}) and (\\ref{deeffdefaz}) are fundamentally different. In particular: \\be \\Omega_{eff}^{(A)} = \\frac{\\sum_i \\Omega_i \\frac{d{\\mathcal F}_i (z=0)}{dz} + \\Omega_{de} (w(z=0)+1)}{w_{eff}^{(A)}(z=0) + 1} \\label{omegaedeffaz} \\ee which in contrast to $\\Omega_{eff}^{(H)}$, in general is not equal to $\\Omega_{de}$. Equality arises only when ${\\mathcal F}_i$ do not vary with redshift i.e. ${\\mathcal F}_i = 1$ at all redshifts. This condition is satisfied by the null hypothesis $\\Lambda$CDM and by non-interacting quintessence models. Therefore, assuming that $\\Omega_m$ and $\\Omega_k$ are known (e.g. from CMB), simultaneous measurements of $H(z)$ and $A(z)$ at even one $z > 0$ is apriori enough for testing the presence of an interaction between dark matter and dark energy independent of the underlying model. Evidently, in practice the measurements must be performed at many redshift bins to improve statistics and to compensate for measurement errors. Apriori one can use other quantities such as angular diameter distance $D_A$ or luminosity distance $D_L$ which are easier to measure rather than $A(z)$. However, both these quantities are functional of $H(z)$ - through integration of $1/H^{1/2}(z)$. Thus, in general they do not have an analytical expression. Besides, their derivatives depend on ${\\mathcal F}_i$'s only, in contrast to (\\ref{deeffdefaz}) that depends on both ${\\mathcal F}_i$'s and their derivatives. Therefore, $\\Omega_{eff}$ and $\\gamma_{eff}$ obtained from $dD_A/dz$ or $dD_L/dz$ will be equal to ones determined from $H(z)$ irrespective of the underlying cosmology. This shows that the function $A(z)$ (or equivalently $B(z)$) introduced in~\\cite{houriaz} has special properties and is well suited for discriminating between dark energy models. It can be measured from supernovae data, see~\\cite{houriaz} for the methodology. As for LSS data, one needs to determine both $H(z)$ and its evolution $dH(z)/dz$ to be able to calculate $A(z)$, for instance from the BAO and the power spectrum of matter fluctuations~\\cite{psobs}. This is not an easy task. As an example consider supernovae observations that measure the luminosity distance $D_L$ to a supernova from its standardized apparent magnitude. The angular luminosity distance $D_A$ is related to the luminosity distance, see (\\ref{dah}). To determine $dD_A/dz$ apriori one can use the measured $D_A$, and determine its derivative (slope). However, due to scattering and discreteness of data, such a measurement will have large uncertainties. The same problem arises for $dH(z)/dz$ or $A(z)$ because they depend on derivatives of $D_L$, see equations (\\ref{rhoderdl}) and (\\ref{rhoderda}). Nonetheless, there are various methods such as binning of data, using a fit in place of discrete data, etc. that allow to improve the estimation. Near future large area surveys such as Euclid~\\cite{euclid}, BigBOSS~\\cite{bigboss}, LSST~\\cite{lsst} will be able to determine these quantities with relatively good precision, see also Sec. \\ref{sec:forecast} for measurement methodology. In particular, large surface spectroscopic and lensing surveys such as Euclid are able to determine the variation of total density with redshift $d\\rho/dz \\propto B(z)$ with good precision. In Appendix \\ref{app:b} we obtain the Fisher matrix for dark energy parameters without considering a specific parametrization for the equation of state $w(z)$. \\subsection{Discrimination precision} \\label{sec:errordiscrim} Measurements of cosmological parameters show that $w^{obs}_{de} \\sim -1$ irrespective of which proxy or measurement method - supernovae, CMB, or LSS has been used. This means that $|{\\mathcal F}_i(z) - 1| \\approx 0$ and $d{\\mathcal F}_i (z)/dz \\approx 0$. Moreover, addition of ${\\mathcal F}_i(z)$ to the model increases the number of parameters. Giving the fact that we have essentially two observables: $H(z)$ and one of $D_A(z)$, $D_L(z)$ or $B(z)$, greater number of parameters means also greater degeneracy, thus more uncertainty for discrimination between $\\Lambda$CDM, a non-interacting quintessence, and interacting dark energy models. One way of measuring the presence of interaction without having to fit data to the large number of parameters in equations (\\ref{friedmanmgequiv}) and (\\ref{azintdmde}), is to measure how different $\\Omega_{eff}^{(H)},~\\Omega_{eff}^{(A)},~\\gamma_{eff}^{(H)},$ and $\\gamma_{eff}^{(A)}(z)$ are, because as we discussed in the previous section, when ${\\mathcal F}_i \\neq 1$ these effective quantities are not the same. To this end, a natural criteria is: \\be \\Theta (z) \\equiv \\frac{\\Omega_{eff}^{(A)} (w_{eff}^{(A)}(z) + 1) (1+z)^{3\\gamma_{eff}^{(A)}(z)} - \\Omega_{eff}^{(H)} (w_{eff}^{(H)}(z) + 1) (1+z)^{3\\gamma_{eff}^{(H)}(z)}}{\\Omega_{eff}^{(H)} (w_{eff}^{(H)}(z) + 1) (1+z)^{3\\gamma_{eff}^{(H)}(z)}} \\label{deftheta} \\ee This quantity can be explained explicitly as a function of $\\Omega_i,~{\\mathcal F}_i,~\\gamma_i$, and is zero when $F_i= 1,~dF_i/dz = 0$. Note that we have chosen expression (\\ref{deeffdefaz}) for comparison rather than (\\ref{deeffdef}) because it is not possible to determine $\\Omega_{eff}^{(A)}$ in a model independent manner, see equation (\\ref{omegaedeffaz}). By contrast $\\Omega_{eff}^{(H)} = \\Omega_{de}$, thus $\\gamma_{eff}^{(H)}$ and thereby $w_{eff}^{(H)}$ can be determined without any reference to ${\\mathcal F}_i$ coefficients. In~\\cite {houriaz} we suggested to use the sign and evolution of $A(z)$ to discriminate between dark energy with $\\gamma (z) \\neq 0$ and a cosmological constant. Here $\\Theta (z)$ plays a similar role for discriminating between interacting or non-interacting dark energy. Assuming that $\\Omega_m$ and $\\Omega_h$ are determined independently and with very good precision, for instance from CMB anisotropies with marginalization over $\\gamma_{de}$, $\\Theta$ can be determined from the measurement of $H(z)$ and $B(z)$. The latter can be measured from whole sky or wide area spectroscopic surveys data such as Euclid, or multi-band photometric surveys such as DES. Evidently determination of $B(z)$ that depends on $dH/dz$ is very difficult. However, it is easy to see that there is no other quantity which can be measured more easily and discriminates between $\\Lambda$CDM and dynamical dark energy models with a better precision. For instance, the BAO method determines $H(z)$ and $D_A(z)$ directly. But, $D_A(z)$ depends on $w(z)$ or equivalently $\\gamma (z)$ through an integral, see equation (\\ref{dah}). Therefore, it is less sensitive to the variation of $\\gamma (z)$ with redshift. This is analogue to binning a data. Evidently, a binned data is less noisy and has a smaller uncertainty. But, if the goal is to measure the variation of data, the binning can completely smear out small variations. Therefore, irrespective of methods and measured proxies, we are limited by inherent properties of the physical system. In this respect, the precision with which $\\Theta (z)$ can be measured gives the ultimate sensitivity of an observation/data set to deviation from $\\Lambda$CDM. ", "conclusions": "" }, "1112/1112.1096_arXiv.txt": { "abstract": "We use SDSS spectra and optical to far-infrared photometry for a sample of 31 FeLoBAL QSOs to study the relationship between the AGN-driven outflows, and obscured star formation in their host galaxies. We find that FeLoBAL QSOs invariably have IR luminosities exceeding $10^{12}$L$_{\\odot}$. The AGN supplies 75\\% of the total IR emission, on average, but with a range from 20\\% to 100$\\%$. We find a clear anticorrelation between the strength of the AGN-driven outflows and the contribution from star formation to the total IR luminosity, with a much higher chance of seeing a starburst contribution in excess of 25\\% in systems with weak outflows than in systems with strong outflows. Moreover, we find no evidence that this effect is driven by the IR luminosity of the AGN. We conclude that radiatively driven outflows from AGN act to curtail obscured star formation in the host galaxies of reddened QSOs to less than $\\sim 25\\%$ of the total IR luminosity. This is the most direct evidence yet obtained for `quasar mode' AGN feedback. ", "introduction": "Over the last decade, several problems have arisen in modelling the cosmological evolution of galaxies. These include (1) the difficulties that models faced in explaining the observed galaxy luminosity function at low and high redshifts simultaneously (e.g. \\citealt{benson03}), (2) the prediction that rich galaxy clusters should harbour cooling flows when few are observed (e.g. \\citealt{peterson01}), and (3) the difficulties that models face in reproducing the number of IR-luminous galaxies observed at high redshift (e.g. \\citealt{baugh05}). One of the most promising solutions to these issues is AGN feedback. AGN feedback is the exertion of influence of an SMBH on $\\gtrsim$kpc scales to curtail star formation in the host galaxy, and/or accretion onto the SMBH itself. Galaxy evolution models usually assume this feedback to occur in one or both of two simplified modes; `quasar' mode and `radio' mode. Quasar mode feedback occurs via radiation from an accretion disk, while radio mode feedback occurs via a relativistic jet that transfers momentum to the ISM. Both these feedback paradigms have led to improvements in the ability of models to reproduce observations (e.g. \\citealt{bower08,somer08}), but observational evidence for either paradigm remains sparse. Our group has been looking for evidence of quasar mode feedback by examining systems in which such feedback may be ongoing. To do so, we have been studying the `FeLoBAL' QSOs \\citep{hazard87,hall02}. We selected this population for two reasons. First, their UV absorption troughs are unambiguous signatures of radiatively driven outflows powered by an AGN. Second, FeLoBAL QSOs are invariably reddened and have high IR luminosities \\citep{farrah07,farrah10}. We here present results for 31 FeLoBAL QSOs, in which we compare the strength of their outflows as estimated from their UV spectral properties, to the luminosity of obscured star formation in their host galaxies as measured from optical through far-IR photometry. We define \"IR luminosity\" as the luminosity integrated over 1-1000$\\mu$m in the rest-frame. ", "conclusions": "We find that FeLoBAL QSOs are luminous in the IR. All 31 of our sample have total IR luminosities (L$_{Tot}$) in excess of $10^{12}$L$_{\\odot}$, with nine objects exceeding $10^{13}$L$_{\\odot}$. These luminosities are comparable to those of the wider population of BAL QSOs \\citep{gallagher07}, red QSOs \\citep{geo09}, and ULIRGs \\citep{farrah03,farrah09}. The dominant power source is usually an AGN. A pure AGN is either the most likely power source, or consistent within the 90\\% confidence interval, for $35\\%$ of the sample. A starburst component is required for the remaining objects, but in only twelve objects is the starburst more luminous than $10^{12}$L$_{\\odot}$, and in only three objects is the starburst more luminous than the AGN. The mean starburst contribution to the total IR luminosity (f$_{SB}$) is $\\sim24\\%$, comparable to that of local ULIRGs with `warm' IR colours, but lower than those of PG QSOs \\citep{veill09}. \\begin{figure} \\includegraphics[width=52mm,angle=90]{farrah_fig1a.ps} \\includegraphics[width=52mm,angle=90]{farrah_fig1b.ps} \\caption{Absorption strength vs (left) starburst (stars) and AGN (squares) IR luminosities, and (right) starburst contribution to the total IR luminosity.}\\label{luminosities} \\end{figure} We now examine whether or not there is a relationship between the AGN-driven outflows and the obscured star formation. To quantify the strength of the outflows we use the same, single species (MgII$\\lambda$2799\\AA) across the whole sample, and adopt the BI$_{0}$ parameter of \\citet{gibson09}. We measured the MgII absorption strengths using the methods described in \\citet{f2ms}. We first examine if absorption strength depends on AGN or starburst IR luminosity (Fig \\ref{luminosities}, left). There is no correlation between absorption strength and L$_{SB}$ (a Spearman rank test gives $\\rho = -0.10$, $P = 0.58$) and at best a weak correlation between absorption strength and L$_{AGN}$ ($\\rho = 0.39$, $P = 0.04$). Conversely, we see a weak but clear anticorrelation between absorption strength and f$_{SB}$ ($\\rho = -0.49$, $P = 0.005$, Fig \\ref{luminosities} right). Moreover, all the systems with BI$_{0}\\gtrsim 3500$km s$^{-1}$ have $f_{SB}<25\\%$, while the systems with BI$_{0}\\lesssim 3500$km s$^{-1}$ have a wide dispersion in starburst contributions, from $0\\%$ to $\\sim80\\%$. A two-sided Kolmogorov-Smirnov test for the objects above and below BI$_{0} = 3500$km s$^{-1}$ yields a difference in distributions at 99.84\\% confidence, though the number of objects in the two subsamples is too low for this test to be robust. We examine this result in more detail by constructing probability distribution functions (using all the fit solutions) for f$_{SB}$ for two subsamples divided at BI$_{0}=3500$km s$^{-1}$ (Figure \\ref{pdfs}, left). We see a clear difference. The low absorption strength subsample shows a much higher chance of a higher starburst contribution than the high absorption strength subsample. We quantify this by extracting the probabilities of obtaining $f_{SB}>25\\%$. For the whole sample we find P($f_{SB}>25\\%$)$=50.3^{+5.3}_{-5.4}$\\%, for the low absorption strength sample we find P($f_{SB}>25\\%$)$=67.3^{+4.5}_{-4.1}$\\%, while for the high absorption strength sample we find P($f_{SB}>25\\%$)$=17.8^{+3.7}_{-6.5}$\\%. This anticorrelation between absorption strength and f$_{SB}$ is straightforwardly interpreted as the outflow from the AGN curtailing star formation in the host galaxy. In this scenario, the systems with $f_{SB}>25\\%$ are those in which an outflow has yet to curtail star formation, so the outflows always have BI$_{0}\\lesssim 3500$km s$^{-1}$. Conversely, the systems with $f_{SB}<25\\%$ are those in which an outflow has curtailed star formation, {\\itshape and} those in which such an outflow has subsequently waned, making the dispersion in absorption strengths wide. There are however four other ways that this anticorrelation could arise. The first is that the starburst `suppresses' AGN outflows, so when the starburst wanes (via a cause unrelated to the AGN) an AGN driven outflow can appear, possibly because the ISM density has been reduced by the starburst. We cannot test this scenario, but it would likely require a serendipitous conjunction of ISM and SMBH parameters, so we do not consider it further. The second is if a high starburst contribution meant that the Mg II troughs were {\\it observed} to be weaker than they really are. Again, we cannot test this, but the rest-frame UV continua of starbursts in ULIRGs are at least an order of magnitude too weak to provide this effect, and can show absorption in the same species \\citep{farrah05}. The third is if BAL QSOs with strong starbursts preferentially drop out of the SDSS QSO selection compared to BAL QSOs with weak starbursts, and were thus not included in \\citet{trump06}. This possibility is also not testable by us, but the SDSS is now turning up FeLoBAL QSOs in large numbers, and the SDSS QSO followup colour selections are fairly relaxed, so we do not consider this possibility likely either. The fourth possibility is that stronger outflows reflect an increase in the IR emission from the AGN, but have no effect on the starburst; if this is the case we would see a decline in the contribution from star formation to the total IR luminosity with increasing absorption strength, but with no direct relationship behind the decline. This is a possibility we can test, as follows. If outflow strength is just a proxy for AGN luminosity, then we should see a bigger difference between the starburst contribution PDFs for subsamples divided by AGN luminosity than between subsamples divided by absorption strength. In the right panel of Figure \\ref{pdfs} we show starburst contribution PDFs for two subsamples divided by AGN luminosity at L$_{AGN}=10^{12.5}$L$_{\\odot}$. The difference between the PDFs divided by AGN luminosity is {\\it weaker} than the difference between the PDFs divided by absorption strength. Furthermore, we are more likely (albeit only at $\\sim$2$\\sigma$) to obtain a smaller starburst contribution by selecting high absorption strength systems than we are by selecting high AGN luminosity systems; for P($f_{SB}>25\\%$): the BI$_{0}>3500$km s$^{-1}$ subsample is $17.8^{+3.7}_{-6.5}$\\% while the L$_{AGN}>10^{12.5}$L$_{\\odot}$ subsample is $38.7^{+8.5}_{-10.0}$\\%). In other words, we are more successful in finding systems with a large starburst contribution to the total IR emission by selecting on weak outflows than we are by selecting on low AGN luminosity. Also, we are more likely to find star formation with a lower absolute luminosity in the {\\itshape lower} luminosity AGN subsample (e.g. for P(L$_{Sb}<10^{12}$L$_{\\odot})$: the L$_{AGN}<10^{12.5}$L$_{\\odot}$ subsample is $69.4^{+5.2}_{-4.6}$\\% while the L$_{AGN}>10^{12.5}$L$_{\\odot}$ subsample is $43.2^{+6.9}_{-10.4}$\\%) Overall therefore, we find that radiatively driven outflows from an AGN with absorption strengths $\\gtrsim3500$ km s$^{-1}$ act to curtail star formation in their host galaxies. We also find that this effect is (at least largely) relative; such outflows reduce the contribution from star formation to the total IR luminosity to less than $\\sim25\\%$. We also propose that the {\\itshape infrared} luminosity of the AGN is not a good proxy for the degree of AGN feedback that is taking place. \\begin{figure} \\includegraphics[width=52mm,angle=90]{farrah_fig2a.ps} \\includegraphics[width=52mm,angle=90]{farrah_fig2b.ps} \\caption{Probability Distribution Functions for the starburst contribution to the total IR luminosity. In both panels, the solid line is the PDF for the whole sample. The other lines are for subsamples divided according to two criteria. {\\itshape Left panel}: by absorption strength. {\\itshape Right panel}: by AGN IR luminosity.}\\label{pdfs} \\end{figure}" }, "1112/1112.3093_arXiv.txt": { "abstract": " ", "introduction": "With the advent of the \\textit{Square Kilometre Array} (SKA; \\citealt{dewdney2009}) and its precursors and pathfinders, including the \\textit{Australian SKA Pathfinder} (ASKAP; \\citealt{deboer2009}), the \\textit{Karoo Array Telescope} (Meer\\-KAT; \\citealt{jonas2009}), and the \\textit{Aperture Tile In Focus} (APERTIF; \\citealt{oosterloo2009}), the prospect of deep radio continuum and \\ion{H}{i} surveys of large areas on the sky demands for new strategies in the areas of data reduction and analysis, given the sheer volume of the expected data streams, in particular for spectroscopic surveys. Of particular importance is the automatic and accurate identification and parametrisation of sources with high completeness and reliability. Due to the large data volumes to be searched, source finding algorithms must be fully automated, and the once common practice of source finding `by eye' will no longer be feasible. Moreover, accurate source para\\-metrisation algorithms need to be developed to generate reliable source catalogues free of systematic errors, as otherwise the integrity of scientific results based on the survey data could be compromised. In this paper we will take a closer look at the \\textsc{Duchamp} source finder\\footnote{\\textsc{Duchamp} website: http://www.atnf.csiro.au\\slash{}people\\slash{}Matthew.Whiting\\slash{}Duchamp/} \\citep{whiting2011a,whiting2012}. \\textsc{Du\\-champ} has been developed by Matthew Whiting at CSIRO as a general-purpose source finder for three-dimensional data cubes as well as two-di\\-mensional images and will serve as the default source finder in the processing of data from the ASKAP survey science projects. The software identifies sources by searching for regions of emission above a specified flux threshold. To improve its performance, \\textsc{Duchamp} offers several methods of preconditioning and filtering of the input data, including spatial and spectral smoothing as well as reconstruction of the entire image or data cube with the help of wavelets. In addition to source finding, \\textsc{Duchamp} provides the user with basic source parametrisation, including the measurement of position, size, radial velocity, line width, and integrated flux of a source. More information about the capabilities of the software is available from the \\textsc{Duchamp} User Guide \\citep{whiting2011b}. A brief overview of \\textsc{Du\\-champ}'s basic functionality is provided in Section~\\ref{sect_duchamp}. \\begin{table}[ht] \\begin{center} \\caption{Summary of the parameters used to generate the visibility data set and noise image for the point source models.} \\label{tab_model} \\begin{tabular}{lrl} \\hline Parameter (visibility) & Value & Unit \\\\ \\hline Number of antennas & $36$ & \\\\ System temperature & $50$ & $\\mathrm{K}$ \\\\ Declination & $-45^{\\circ}$ & \\\\ Total integration time & $8$ & $\\mathrm{h}$ \\\\ Hour angle range & $\\pm 4$ & $\\mathrm{h}$ \\\\ Cycle time & $5$ & $\\mathrm{s}$ \\\\ Stokes parameters & I & \\\\ Number of channels & $31$ & \\\\ Frequency & $1.42$ & $\\mathrm{GHz}$ \\\\ Channel width & $18.31$ & $\\mathrm{kHz}$ \\\\ & $3.86$ & $\\mathrm{km \\, s}^{-1}$ \\\\ \\hline Parameter (image) & Value & Unit \\\\ \\hline Final image size & $31 \\times 31$ & $\\mathrm{px}$ \\\\ Field diameter & $5$ & $\\mathrm{arcmin}$ \\\\ Pixel size & $10$ & $\\mathrm{arcsec}$ \\\\ Robustness & $0$ & \\\\ Gaussian $uv$ taper & $7.28$ & $\\mathrm{k} \\lambda$ \\\\ & $1.54$ & $\\mathrm{km}$ \\\\ \\textsc{rms} noise & $1.95$ & $\\mathrm{mJy}$ \\\\ Synthesised beam & & \\\\ \\quad major axis & $27.1$ & $\\mathrm{arcsec}$ \\\\ \\quad minor axis & $26.7$ & $\\mathrm{arcsec}$ \\\\ \\quad position angle & $87\\fdeg{}9$ & \\\\ \\hline \\end{tabular} \\end{center} \\end{table} So far, \\textsc{Duchamp}'s source finding and parametrisation capabilities have never been systematically tested on a large set of artificial sources with well-defined parameters. The aim of this paper is to bridge this gap by thoroughly testing the performance of \\textsc{Duchamp} on sets of artificial point sources and galaxy models as well as a data set containing real galaxies and telescope noise. The tests were originally motivated by the need to identify suitable source finding algorithms for the \\textit{Widefield ASKAP L-band Legacy All-sky Blind Survey} (WALLABY; \\citealt{koribalski2009}),\\footnote{Principal investigators: B\\\"{a}rbel Koribalski and Lister Staveley-Smith; public website: http://www.atnf.csiro.au\\slash{}research\\slash{}WALLABY/} one of the large, extragalactic ASKAP survey science projects currently in pre\\-paration \\citep{westmeier2010}. Hence, the tests presented here will focus on the detection of compact and extended \\ion{H}{i}~sources, in particular galaxies, in three-dimensional data cubes with ASKAP characteristics. However, we believe that the results and conclusions presented in this paper will be of interest not only to those involved in SKA precursor science, but to a larger community of astronomers interested in the automatic detection and parametrisation of sources in their data sets, regardless of the wavelength range involved. For a comparison of \\textsc{Duchamp}'s performance with that of other source finding algorithms we refer to the paper by \\citet{popping2011} in this issue. This paper is organised as follows: In Section~\\ref{sect_surveys} we summarise the source finding strategies of other large \\ion{H}{i}~surveys in the past, followed by a brief overview of the \\textsc{Duchamp} source finder in Section~\\ref{sect_duchamp}. In Section~\\ref{sect_pointsources} we present the outcome of our test of \\textsc{Duchamp} on point source models with simple Gaussian line profiles. Section~\\ref{sect_galaxies} describes our testing of \\textsc{Duchamp} on models of disc galaxies with varying physical parameters. In Section~\\ref{sect_whisp} we apply \\textsc{Duchamp} to a data cube containing real galaxies and genuine noise extracted from radio observations. A discussion of our results is presented in Section~\\ref{sect_discussion}. Finally, Section~\\ref{sect_summary} summarises our main results and conclusions. ", "conclusions": "\\label{sect_discussion} In general, \\textsc{Duchamp} does what it promises to do. It is able to reliably detect sources down to low signal-to-noise ratios and accurately determine their position and radial velocity. These are the most fundamental requirements for any source finder. Our tests also demonstrated that by using and fine-tuning the options of `\\`{a} trous' wavelet reconstruction and growing of sources to lower flux levels the performance of \\textsc{Duchamp} can be greatly enhanced. \\subsection{Improving Reliability} \\label{sect_reliability} The reliability figures reported throughout this paper have all been ``raw'' reliabilities, i.e.~reliabilities as achieved by \\textsc{Duchamp} prior to any filtering of the output source catalogue. The user would normally wish to substantially improve these through appropriate filtering of the source catalogue based on the source parameters as measured by \\textsc{Duchamp}. The left-hand panel of Figure~\\ref{fig_false} shows the measured integrated flux plotted against measured line width for all genuine (black data points) and false (red data points) detections found by \\textsc{Duchamp} in the point source models discussed in Section~\\ref{sect_pointsources}. It is obvious that genuine and false detections occupy largely disjunct regions of $F_{\\rm int}$--$w_{50}$ parameter space, with false detections generally occurring near the low end of the integrated flux spectrum. Similar plots can be generated for other combinations of source parameters, but $F_{\\rm int}$ and $w_{50}$ usually provide the best distinction between genuine and false detections. The easiest way to improve the reliability of \\textsc{Du\\-champ}'s source finding results is to simply apply a cut in $F_{\\rm int}$ to exclude most false detections while retaining most of the genuine sources. In our example, applying a cut at $F_{\\rm int} = 40~\\mathrm{mJy \\, km \\, s}^{-1}$ will discard $97.2\\%$ of all false detections while at the same time retaining $96.9\\%$ of all genuine sources, thereby increasing the overall reliability from $77.1\\%$ to $99.2\\%$ while only moderately decreasing the overall completeness from $83.0\\%$ to $80.6\\%$. \\begin{figure}[t] \\begin{center} \\includegraphics[width=\\linewidth]{fig_whisp.eps} \\caption{Completeness (filled circles, solid lines) and reliability (open circles, dashed lines) of \\textsc{Duchamp} on the WSRT model cube with WHISP galaxies for different flux thresholds and input parameters. The colours, as shown in the legend, distinguish the different wavelet reconstruction modes (one-dimensional versus three-dimensional) and wavelet reconstruction thresholds ($3 \\sigma$ versus $4 \\sigma$) used in the tests (see Table~\\ref{tab_input3} for details). The numbers alongside the data points refer to the corresponding runs as listed in Table~\\ref{tab_input3}.} \\label{fig_whisp} \\end{center} \\end{figure} \\begin{figure}[t] \\begin{center} \\includegraphics[width=\\linewidth]{fig_completeness_3.eps} \\caption{Completeness as a function of integrated flux for selected runs (see legend) of \\textsc{Duchamp} on the WSRT model cube with WHISP galaxies. The choice of colours is the same as in Figure~\\ref{fig_whisp}.} \\label{fig_completeness_3} \\end{center} \\end{figure} A similar cut can be applied to the results from run~7 on the test data cube containing artificially redshifted WHISP galaxies, as plotted in the right-hand panel of Figure~\\ref{fig_false}. Again, applying a simple flux threshold of $0.5~\\mathrm{Jy \\, km \\, s}^{-1}$ will improve reliability from $10\\%$ to $84\\%$, while only moderately decreasing completeness from $59\\%$ to $50\\%$. The method is not quite as successful as for the point sources, as we are now dealing with real galaxies and real noise with interference and artefacts, but nevertheless a significant improvement in reliability can be achieved without any severe impact on the number of genuine detections. This simple example illustrates that the ``raw'' reliability figures quoted throughout this paper should not be considered as the final numbers. Reliability can be greatly improved through very basic filtering in parameter space of the \\textsc{Duchamp} output catalogue. In principle, this applies to the output of almost any source finder. Alternatively, instead of removing sources from the output catalogue, it may be desirable to calculate a reliability number for each catalogue entry based on the source's location in parameter space and leave it to the catalogue's users to decide as part of their scientific analysis at which reliability level they wish to make the cut. \\begin{figure*}[t] \\begin{center} \\includegraphics[width=\\linewidth]{fig_false.eps} \\caption{Measured integrated flux, $F_{\\rm int}$, versus measured line width, $w_{50}$, of all genuine (black) and false (red) detections made by \\textsc{Duchamp} in the point source models with Gaussian line profiles (left) and the test cube with artificially redshifted WHISP galaxies (right). The dashed, black lines indicate the flux levels of $0.04$ and $0.5~\\mathrm{Jy \\, km \\, s}^{-1}$ used to filter false detections.} \\label{fig_false} \\end{center} \\end{figure*} \\subsection{Source Parametrisation Issues} When it comes to source parametrisation, the measurements provided by \\textsc{Duchamp} are affected by several systematic errors. These systematic errors are not due to errors in the software itself, but a consequence of the the presence of noise in the data as well as the methods and algorithms used for measuring source parameters. Spectral line widths determined by \\textsc{Duchamp} are generally very accurate and not much affected by noise-induced, systematic errors as far as the $w_{50}$ parameter is concerned. The two other line width parameters calculated by \\textsc{Duchamp}, $w_{20}$ and $w_{\\rm vel}$, appear to be systematically too large over a wide range of signal-to-noise ratios and should not be used unless explicitly required in special, well-defined circumstances. Peak fluxes, as reported by \\textsc{Duchamp}, are in general slightly too large for bright sources and significantly too large (on a relative scale) for faint sources. This is due to the fact that \\textsc{Duchamp} determines the peak flux by simply selecting the value of the brightest pixel encountered. This method introduces a bias towards positive noise peaks sitting on top of the brightest region of a source, and hence, in the presence of noise, peak fluxes measured by \\textsc{Duchamp} will be systematically too high. Integrated fluxes determined by \\textsc{Duchamp} are significantly and systematically too small, in particular for faint sources. This is likely caused by the fact that \\textsc{Duchamp} simply sums over the flux of discrete elements above a given threshold to determine the integrated flux, thereby missing some of the flux from elements below the flux threshold. Hence, the raw integrated flux measurements currently provided by \\textsc{Duchamp} are not useful and need to be corrected to compensate for the systematic offset. This issue is particularly sensitive as many scientific projects, including the ASKAP survey science projects WALLABY and DINGO\\footnote{\\textit{Deep Investigation of Neutral Gas Origins}; principal investigator: Martin Meyer; public website: http:/\\slash{}internal.physics.uwa.edu.au\\slash{}$\\sim$mmeyer\\slash{}dingo/} \\citep{meyer2009}, rely on accurate flux measurements, for example for determining the \\ion{H}{i}~mass function of galaxies. Finally, a particular problem in the case of galaxies is that under certain circumstances galaxies either get broken up into multiple detections or only one half of a galaxy is detected. This problem mainly affects faint, edge-on galaxies with broad spectral profiles that are partly hidden in the noise and results in systematic errors in the measurements of essentially all source parameters, including basic parameters such as position and radial velocity. Such cases of multiple or partial detections must be identified and treated separately to prevent biases in any scientific analysis based on the source finding results. In this paper we present and discuss the results of basic, three-dimensional source finding tests with \\textsc{Du\\-champ}, the standard source finder for the Australian SKA Pathfinder, using different sets of unresolved and extended \\ion{H}{i}~model sources as well as a data set of real galaxies and noise obtained from \\ion{H}{i}~observations with the WSRT. Overall, \\textsc{Duchamp} appears to be a successful, gen\\-eral-purpose source finder capable of reliably detecting sources down to low signal-to-noise ratios and accurately determining their position and velocity. In the case of point sources with simple Gaussian spectral lines we achieve a completeness of about 50\\% at a peak signal-to-noise ratio of~3 and an integrated flux level of about $0.1~\\mathrm{Jy \\, km \\, s}^{-1}$. The latter corresponds to an \\ion{H}{i} mass sensitivity of about $2 \\times 10^{8}~\\mathrm{M}_{\\odot}$ at a distance of $100~\\mathrm{Mpc}$ which is slightly better than what the WALLABY project is expected to achieve for real galaxies \\citep{koribalski2009}. The situation is less ideal for extended sources with double-horn profiles. In this case we achieve 50\\% completeness at an integrated flux level of about $2.5~\\mathrm{Jy \\, km \\, s}^{-1}$ for the model galaxies and $0.7~\\mathrm{Jy \\, km \\, s}^{-1}$ for the WHISP galaxies. The latter is equivalent to an \\ion{H}{i} mass sensitivity of about $1.7 \\times 10^{9}~\\mathrm{M}_{\\odot}$ at a distance of $100~\\mathrm{Mpc}$, illustrating that the performance of \\textsc{Duchamp}, as well as any other source finder, will strongly depend on source morphology. However, these figures may well be improved by carefully optimising the various input parameters offered by \\textsc{Duchamp}. In its current state \\textsc{Duchamp} is not particularly successful in parametrising sources in the presence of noise in the data cube, and other, external algorithms for source parametrisation should be considered instead. It appears, however, that most, if not all, para\\-metrisation issues are due to intrinsic limitations in the implemented algorithms themselves and not due to errors in their implementation, suggesting that most of the problems can in principle be solved by implementing more sophisticated parametrisation algorithms in \\textsc{Duchamp}. Alternatively, corrections would have to be applied to all parameters derived by \\textsc{Duchamp} to compensate for systematic errors. Such corrections, however, would have to be highly specialised and tailored to the particular survey and source type concerned." }, "1112/1112.0679_arXiv.txt": { "abstract": "We provide a quantitative description and statistical interpretation of the optical continuum variability of quasars. The Sloan Digital Sky Survey (SDSS) has obtained repeated imaging in five UV-to-IR photometric bands for 33,881 spectroscopically confirmed quasars. About 10,000 quasars have an average of 60 observations in each band obtained over a decade along Stripe 82 (S82), whereas the remaining \\about25,000 have 2--3 observations due to scan overlaps. The observed time lags span the range from a day to almost 10 years, and constrain quasar variability at rest-frame time lags of up to 4 years, and at rest-frame wavelengths from 1000\\AA\\ to 6000\\AA. We publicly release a user-friendly catalog of quasars from the SDSS Data Release 7 that have been observed at least twice in SDSS or once in both SDSS and the Palomar Observatory Sky Survey, and we use it to analyze the ensemble properties of quasar variability. Based on a damped random walk (DRW) model defined by a characteristic time scale and an asymptotic variability amplitude that scale with the luminosity, black hole mass, and rest wavelength for individual quasars calibrated in S82, we can fully explain the ensemble variability statistics of the non-S82 quasars such as the exponential distribution of large magnitude changes. All available data are consistent with the DRW model as a viable description of the optical continuum variability of quasars on time scales of \\about 5--2000~days in the rest frame. We use these models to predict the incidence of quasar contamination in transient surveys such as those from PTF and LSST. ", "introduction": "The optical continuum variability of quasars has been recognized since their first optical identification (Matthews \\& Sandage 1963), and it has been proposed and utilized as an efficient method for their discovery (van den Bergh, Herbst, Prit-\\\\ chet 1973; Hawkins 1983; Hawkins \\& Veron 1995; Ivezi\\'{c} et al.\\ 2004a; Rengstorf et al.\\ 2006). The observed characteristics of the variability can then be used to constrain the origin of their emission (e.g., Kawaguchi et al.\\ 1998; Trevese, Kron \\& Bunone 2001, and references therein). The variability of quasars has typically been quantified using a structure function (SF) analysis (e.g., Hughes et al.\\ 1992; Collier \\& Peterson 2001; Bauer et al.\\ 2009; Koz{\\l}owski et al.\\ 2010b; Welsh, Wheatley, \\& Neil 2011), where the SF is the root-mean-square (rms) magnitude change (\\dm) as a function of the time lag ($\\Delta t$) between measurements (similar to an auto-correlation function). It is fairly well established that quasar variability properties depend on physical properties such as the quasar luminosity, wavelength, time scale, and the presence of radio emission. However, despite the considerable observational effort invested over last few decades, many conflicting claims about various correlations exist in the literature (see Giveon et al.\\ 1999 for a detailed discussion). The traditional method for studying variability has been to monitor a small, select sample of quasars over a long time baseline (e.g., Hawkins 2002; Giveon et al.\\ 1999; Rengstorf et al.\\ 2004). In this case, it is possible to compute the SF for each quasar, which can later be sample-averaged or studied individually. An alternative, utilized in more recent studies based on Sloan Digital Sky Survey (SDSS; York et al.\\ 2000) and Palomar Observatory Sky Survey (POSS; Minkowski \\& Abell 1963) data, is to compute a single SF for all quasars in a particular wavelength or luminosity range. This approach, mandated by the fact that typically only a few epochs were available per object, only measures {\\it ensemble} properties and {\\it assumes} that all quasars selected from a narrow range of physical properties vary in the same way. Nevertheless, this approach enabled studies of quasar optical variability based on tens of thousands of objects and several hundred thousand photometric observations, as well as explorations of the long-term variability (Vanden Berk et al.\\ 2004, hereafter VB04; Ivezi\\'{c} et al.\\ 2004c, hereafter I04; Wilhite et al.\\ 2005; Mahabal et al.\\ 2005; De~Vries, Becker, White, \\& Loomis 2005, hereafter dV05; Sesar et al.\\ 2006, hereafter Ses06). For example, the size and quality of the sample analyzed by VB04 (two-epoch photometry for 25,000 spectroscopically confirmed quasars) allowed them to constrain how quasar variability in the rest frame optical/UV regime depends upon rest frame time lag (up to $\\sim$2 years), luminosity, rest wavelength, redshift, the detection of radio or X-ray emission, and the presence of broad absorption line systems. By comparing SDSS and POSS measurements for \\about 20,000 quasars spectroscopically confirmed by the SDSS, Ses06 constrained the optical quasar variability on time scales from 10 to 50 years (in the observer's frame). They report that there is a characteristic time scale of order 1 year in the quasar rest frame beyond which the SF flattens to a constant value. The SDSS has also facilitated both individual- and ensemble-based approaches by providing a large multi-epoch sample of quasars over the Northern Galactic Cap and well-sampled light curves in the Southern Stripe 82 (S82) survey. It is reassuring that the two approaches lead to similar SFs, as discussed by dV05. A test of this assumption is also described in MacLeod et al.\\ (2008), who show that indeed the {\\it mean} behavior is the same. With such large samples, the ensemble SF(\\dt) slopes are well-constrained, and the values suggest that accretion disk instabilities are the most likely mechanism causing the observed optical variability (VB04; Kawaguchi et al.\\ 1998; see also Lyubarskii 1997). However, attempts to constrain physical models using the ensemble SF are invalid as soon as one realizes that the ensemble SF(\\dt) is a weighted sum of individual quasars with different structure functions (MacLeod et al.\\ 2008). While studies have traditionally examined ``non-parametric'' statistical measures of variability such as the SF, a major challenge has been to describe the variability of individual quasars in a compact way. Recently, the introduction of a damped random walk (DRW) model has provided a way to mathematically characterize quasar light curves in terms of a characteristic time scale ($\\tau$) and an amplitude ($SF_{\\infty}$) which are then correlated with the physical properties such as luminosity and black hole mass. Kelly et al.\\ (2009, hereafter KBS09) modeled a sample of 100 quasar light curves as a DRW and suggested that thermal fluctuations driven by an underlying stochastic process such as a turbulent magnetic field may be the dominant cause for the optical flux fluctuations. Koz{\\l}owski et al.\\ (2010a; hereafter Koz{\\l}10) applied the DRW model to the well-sampled Optical Gravitational Lensing Experiment (OGLE) light curves (Udalski et al. 1997; Udalski et al. 2008) of mid-infrared-selected quasars behind the Magellanic Clouds from Koz{\\l}owski \\& Kochanek (2009). Their analysis shows that the DRW model is robust enough to efficiently select quasars from other variable sources, despite the large surface density of foreground Magellanic Cloud stars (see also Butler \\& Bloom 2011; MacLeod et al.\\ 2011; Koz{\\l}owski et al.\\ 2011a). In MacLeod et al.\\ (2010, hereafter Mac10), we applied the DRW model to the light curves of \\about 10,000 quasars in S82 and found a correlation between $SF_{\\infty}$ and black hole mass which is independent of the anti-correlations with luminosity and wavelength (see also Ai et al.\\ 2010; Meusinger et al.\\ 2011). We also found that $\\tau$ increases with increasing wavelength, remains nearly constant with redshift and luminosity, and increases with increasing black hole mass (see also KBS09; Koz{\\l}10). In Kelly et al.\\ (2011), it was shown that a similar stochastic model but with multiple time scales for a single object can accurately reproduce the X-ray variability of active galactic nuclei (AGN) and microquasars. An inhomogeneous accretion disk model, where the temperature fluctuations throughout the disk are driven by a DRW process, can explain the disk sizes derived from microlensing light curves (see Morgan et al.\\ 2010) while matching the observed level of optical variability, and predicts SEDs which are in better agreement with observations than standard thin disk models (Dexter \\& Agol 2011). One defining feature of the DRW model for a {\\em single} quasar is that it predicts a Gaussian distribution of magnitude differences \\dm\\ for a given \\dt. On the other hand, the observed \\dm\\ distribution in the optical for an {\\em ensemble} of quasars observed at two times separated by \\dt\\ deviates strongly from a Gaussian but is well fit by an exponential distribution (I04). This conflict represents an important puzzle for understanding the DRW model and its applicability to quasar light curves. Also, the high likelihood of extreme values of \\dm\\ has important implications for the interpretation of observations of transients. For example, Vanden Berk et al.\\ (2002) reported the detection of an orphan gamma-ray burst afterglow based on the 2.5 magnitude decrease in optical flux. Such a large flux change was inconsistent with a quasar based on a Gaussian model for their variability, but the source was nevertheless confirmed to be a highly variable quasar (Gal-Yam et al.\\ 2002). An accurate statistical description of the two-epoch photometry for ensembles of quasars will be important for transient detection in large surveys, where quasars represent a major contaminant. Our goal here is to produce a unified view of ensemble and individual optical variability in the context of the DRW model. In this study, we show that the differences in shape between the ensemble SF and the DRW SF are well-explained by averaging over the properties of individual quasars. We also show that the exponential distributions of magnitude changes for ensembles of quasars at fixed time lag are naturally constructed by summing the intrinsically Gaussian distributions of magnitude changes produced by individual quasars. There are several residual issues which we discuss as part of the comparison. An overview of the SDSS and POSS data used in this study is presented in Section~\\ref{data}. In Section~\\ref{OV}, we describe the observed ensemble quasar variability in terms of the SF as a function of wavelength and time lag in the observer's frame. We then convert to rest-frame quantities and compare the data to a model ensemble SF based on our previous DRW analysis of S82 light curves. In Section~\\ref{LT}, we combine the constraints on short-term quasar variability based on SDSS data with the constraints on long-term variability derived from matching the SDSS and POSS catalogs. In Section~\\ref{FS}, we discuss the implications our results have on transient identification, with a focus on future time-domain surveys such as the Palomar Transient Factory and the Large Synoptic Survey Telescope. Our results are discussed and summarized in Section~\\ref{Disc}. ", "conclusions": "\\label{Disc} We have assembled, organized and publicly released a dataset including \\about 3.5 million photometric measurements for 80,000 spectroscopically confirmed quasars. The available time lags span 0.8~days to almost 20~years in the observer's frame. We have analyzed and quantified the observed variability in the observer's and rest frames. By assuming a DRW model for each quasar in our sample, we reconcile the observed variability of individual quasars in S82 with their ensemble statistics. Our principal results are as follows: \\begin{enumerate} \\item Long-term quasar variability measurements, constrained using SDSS and POSS data for time lags up to 50~years (in the observer's frame), conclusively show that a simple power-law dependence for the structure function cannot be extrapolated beyond a decade, and suggests an average characteristic time scale for quasar variability in the rest frame of $\\sim$2~years and an average long-term dispersion of $\\sim$0.26~mag (for rest wavelengths 2000--3000\\AA). This behavior extrapolates well to the UV results of Welsh, Wheatley \\& Neil (2011), who find that the SF for GALEX NUV data reaches about 0.4~mag and flattens at $\\dt_{RF}>300$~days. This SF limit corresponds to a limiting SF value of 0.33~mag when using our definition of the SF (see Section~\\ref{SF}), and 0.27~mag when also scaling to the $u$-band using the wavelength dependence from Eq.~\\ref{eq:form}. This result is in close agreement with the SF at the shortest wavelengths in our data set. Voevodkin (2011) found that a broken power-law provides a good fit to the S82 ensemble SF with a slope of 0.33 at long time scales steepening to 0.79 below 42~days. Our 2-epoch SDSS data are consistent with the shallower slope of 0.33, but our data do not support the conclusion of a much steeper SF(\\dt) for small \\dt\\ found by Voevodkin (2011). While we cannot rule out a broken power-law dependence with the available data, the observed SF is fully consistent with the form expected for a DRW (Eq.~\\ref{eq:sfdt}). \\item We tested the DRW model results based on SDSS Stripe 82 data on an independent dataset, and confirm that the variability parameters $\\tau$ and SF$_{\\infty}$ correlate with physical parameters as found for individual quasars (e.g., Mac10 and references therein). This is evident from the agreement of our model with the observed ensemble variability of SDSS quasars. However, the results indicate that the measured $\\tau$ and SF$_{\\infty}$ distributions are biased low for the S82 sample by a factor of about 2 and $\\sqrt{2}$, respectively. This bias most likely results from the 10-year length limit of the S82 light curves, although it could also be due to uncertain behavior for long time scales. The best-fit $A$ coefficients from Eq.~\\ref{eq:form} (Mac10) need to be shifted upwards by 0.15 dex in the case of SF$_{\\infty}$, and by 0.30 dex in the case of $\\tau$, in order to explain the long time scale constraints provided by the SDSS--POSS dataset. These shifts leave the shorter time scale variability statistics unchanged. \\item For a given time lag and wavelength, the magnitude difference (\\dm) distribution is exponential rather than Gaussian for large magnitude changes. This is well explained as a cumulative effect of averaging over quasars with a range of different $\\tau$ and ${\\rm SF}_{\\infty}$. This is a remarkable result given that the \\dm\\ distribution of every individual quasar is Gaussian. \\item We made predictions for the incidence of quasar contamination in transient surveys using detailed simulations of quasar light curves from a mock LSST catalog. Due to the exponential nature of the \\dm\\ distributions for quasars, the probability of observing $\\dm>1$~mag reaches 0.02 in the $u$ band (where variability is strongest) for time lags of 300~days, and $6\\times 10^{-4}$ for $\\dm>2$~mag. Assuming Gaussian \\dm\\ distributions will result in erroneous likelihood estimates that are about 10 and 1000 times smaller, respectively. \\end{enumerate} It is clear that a major limitation for the S82 quasars is the quality of light curves in both sampling density and time span. It is also clear that our variability model needs to be better tested given the evidence for a likely bias in the S82 time scale estimates. The best current sample for these improvements is that from the OGLE microlensing survey, since the light curves are more densely-sampled and longer than for S82. Here, the probem is the lack of spectroscopic identification of quasar candidates, although the follow-up confirmation of quasars is rapidly improving (Koz{\\l}owski et al.\\ 2011b). The next-generation surveys will also greatly improve the constraints on the long-term SF both individually and for ensembles of quasars. The best short-term prospects are PTF (Law et al.\\ 2009), Pan-STARRS (Kaiser et al.~2002), and the Dark Energy Survey (DES; Honscheid et al.~2008). In particular, the DES supernova program will greatly expand many of the S82 quasar light curves with $griz$ sampling once per week for $\\sim$3~months per year over 5~years. The combination of SDSS, Pan-STARRS, DES, and LSST will yield well-sampled light curves covering over 25 years for 10,000 quasars in Stripe 82. The success of the model presented here suggests that a range of characteristic time scales exists among an ensemble of quasars, which can be related to physical time scales in the accretion disk. While we assumed a single $\\tau$ per quasar, there is evidence that multiple time scales can exist for a given quasar (Collier \\& Peterson 2001; Kelly et al.\\ 2011). Therefore, the study presented here can be extended to adopt the model in Kelly et al.\\ (2011), which fits more than one $\\tau$ for a given object. This model, also called a mixed Ornstein-Uhlenbeck (OU) process, reproduces a PSD of the form exhibited by the X-ray light curves of galactic black holes and AGN, which is flat below a low-frequency break, decays as $1/f$ above the low-frequency break, and steepens to $1/f^2$ above a high-frequency break. In this case, with two characteristic time scales for each quasar, the long-term ensemble SF can be revisited and possibly explained in the context of a mixed OU process. Note that recent optical data from the Kepler mission (Mushotzky et al.\\ 2011), which have a sampling of 1 data point roughly every 30 minutes and 0.1\\% errors, suggest an additional break to a steeper slope ($\\sim 1/f^3$), but this dependence is seen on time scales shorter than can be resolved in SDSS data. \\vskip 0.4in \\leftline" }, "1112/1112.0509_arXiv.txt": { "abstract": "\\emph{WMAP} data when combined with ancillary data on free-free, synchrotron and dust allow an improved understanding of the spectrum of emission from each of these components. Here we examine the sky variation at intermediate and high latitudes using a cross-correlation technique. In particular, we compare the observed emission in several large partitions of the sky plus 33 selected sky regions to three ``standard'' templates. The regions are selected using a criterion based on the morphology of these template maps. The synchrotron emission shows evidence of steepening between GHz frequencies and the \\emph{WMAP} bands. There are indications of spectral index variations across the sky but the current data are not precise enough to accurately quantify this from region-to-region. The emission correlated with the \\halpha template shows clear evidence of deviation from a free-free spectrum. The emission can be decomposed into a contribution from both free-free and spinning dust in the warm ionised medium of the Galaxy. The derived free-free emissivity corresponds to a mean electron temperature of $\\sim 6000$~K, although the value depends critically on the impact of dust absorption on the \\halpha intensity. The WIM spinning dust emission has a peak emission in intensity in the range 40--50~GHz. The anomalous microwave emission associated with dust is detected at high significance in most of the 33 fields studied. The anomalous emission correlates well with the Finkbeiner et al. (1999) model 8 predictions (FDS8) at 94 GHz, and is well described globally by a power-law emission model with an effective spectral index between 20 and 60~GHz of $\\beta \\approx -2.7$. It is clear that attempts to explain the emission by spinning dust models require multiple components, which presumably relates to a complex mix of emission regions along a given line-of-sight. An enhancement of the thermal dust contribution over the FDS8 predictions by a factor $\\sim 1.2$ is required with such models. Furthermore, the emissivity varies by a factor of $\\sim 50\\%$ from cloud to cloud relative to the mean. The significance of these results for the correction of CMB data for Galactic foreground emission is discussed. ", "introduction": "A major goal of observational cosmology is to determine those parameters that describe the Universe. Observations of the Cosmic Microwave Background (CMB) at frequencies in the range 20 -- 200 GHz provide unique data to achieve this by establishing the statistical properties of temperature (and polarisation) measurements. However, an impediment to such studies arises due to foreground emission in our own Galaxy from at least three sources -- synchrotron, free-free and thermal dust emission. As CMB studies move to ever higher precision it is essential to determine the properties of these components to similarly high accuracy. Indeed, although the combined foreground level reaches a minimum in this frequency range ($\\nu \\approx 70$\\,GHz), it remains the dominant signal over large fractions of the sky. Of particular relevance to this discussion is the fact that each of the foreground components has a spectral index that varies from one line of sight to another so using a single spectral index can lead to significant uncertainties in the corrections required. It is therefore essential to study the nature of the Galactic signal at microwave wavelengths in its own right. All-sky observations by the Wilkinson Microwave Anisotropy Probe \\citep[\\emph{WMAP},][]{Bennett_WMAP1:2003a} at the 5 frequencies of 23, 33, 41, 61 and 94 GHz can provide the basis for improving our understanding of local foregrounds. By comparing these maps with templates for synchrotron, free-free and dust emission, made at frequencies where the specific emission mechanisms dominate, it is possible to clarify important properties of the emission. Indeed, new insights into the nature of the Galactic diffuse emission have arisen from studies of the \\emph{WMAP} data, including the detection of several unexpected new contributions. Anomalous dust-correlated emission \\citep{Leitch:1997} was originally observed in the \\emph{COBE}-DMR data \\citep{Kogut_DMR:1996a} but was thought to be due to free-free emission. Draine \\& Lazarian (1998a,b) shifted attention to the dust itself as the source of emission through a mechanism now referred to as ``spinning dust'', or dipole emission from very rapidly spinning grains. A reanalysis of the intermediate and high Galactic latitude data taken by \\emph{COBE}-DMR and supplemented by 19~GHz observations \\citep{Banday_DMR:2003} led to evidence for dust at intermediate Galactic latitudes emitting a spectrum consistent with the the form expected for spinning dust, specifically a hint of a turnover at frequencies below $\\sim$ 20~GHz. However, study of the \\emph{WMAP} data has allowed further refinement of our understanding of the emission. Cross-correlation of the K-band data with observations at 15~GHz \\citep{dOC:2004} again indicated a plateau or downturn in foreground emission inconsistent with a free-free or synchrotron origin. \\citet{Lagache_WMAP:2003} compared the \\emph{WMAP} data to HI column density measurements,revealing an increase in emission with decreasing density, suggesting that the anomalous emission is connected to small, transient heated grains. In addition, a series of papers \\citep{Finkbeiner_WMAP1:2004, Dobler_WMAP3:2008a, Dobler_WMAP3:2008b} have strongly confirmed the presence of anomalous dust emission in the \\emph{WMAP} data, and claim to have found evidence of such a component from the diffuse warm ionised medium (WIM) of the Galaxy. Specifically, the correlation with a H$_{\\alpha}$ template, commonly utilised as a proxy for the free-free emission, is not consistent with the spectrum expected for ionised gas, and a broad bump is seen peaking towards $\\sim$ 40~GHz. Subsequently, \\citet{Dobler_WMAP5:2009} have attempted to constrain specific spinning dust parameters such as the density and typical electric dipole moment of the grains. Recently, \\cite{Peel:2011} have shown that the K-band dust-correlated component is not strongly affected by the inclusion of a 2.3\\,GHz synchrotron template, reducing the possibility of a flat-spectrum synchrotron component. Finally, the so-called \\emph{WMAP}-haze was identified by \\citet{Finkbeiner_WMAP1:2004} although it was already clearly apparent in the foreground residuals of \\citet{Bennett_WMAP1:2003b} and subsequently in the SMICA analysis of \\citet{Patanchon:2005}. The initial physical interpretation of the haze was that it was associated with free-free emission from hot gas in excess of $10^5$\\,K, but this was refuted on the basis of the lack of associated X-ray emission. The detection of the haze relies upon the use of standard templates to remove known foreground emission utilising spatially independent spectra over the entire (high-latitude) sky. It has been argued that such a crude approximation to the true behaviour of the foreground emission at microwave frequencies may well lead to unphysical results. Indeed, \\citet{Gold_WMAP7:2011} find that a spatial variation of spectral index of order 0.25 between 408 MHz and K-band is sufficient to reproduce the haze amplitude. Furthermore, a corresponding polarised signal was found to be absent from the \\emph{WMAP} data. However, the lack of polarised emission can be explained by the entanglement of the Galactic magnetic field towards the Galactic centre, leading to a lower polarisation fraction there as compared to the outer Galaxy. Nevertheless, the microwave haze remains an active area of research, particularly given its possible association with a gamma-ray counterpart observed in the \\emph{Fermi} data \\citep{Dobler_Fermi:2010}. In this paper, we characterise the spatial variation in the foreground emission in terms of the various emission mechanisms noted above by introducing a new partition of the sky into morphologically selected regions. In previous work \\citep[hereafter D06]{Davies_WMAP1:2006}, our approach was to identify regions away from the Galactic plane which were expected to be dominant in one of the three foreground components, free-free, synchrotron or dust and to derive the spectrum for each component. Five regions covering angular scales of $3\\deg$ to $20\\deg$ were chosen for each component, based on foreground template maps, making 15 in all. Here, we generalise this approach and introduce an algorithm to define a set of 35 regions for further study. The selection is intended to minimise the potential cross-talk between the various physical components and to select regions over which the spectral behaviour is uniform thus supporting the use of a template-based comparison. We use a well-known and understood cross-correlation technique to undertake the analysis. The paper is organised as follows. Section~\\ref{sec:data} describes the \\emph{WMAP} data and foreground templates used in this analysis while Section~\\ref{sec:masks} defines the regions of interest for investigation. The methodology of the cross-correlation analysis is outlined in Section~\\ref{sec:methods} and the corresponding results are presented in Section~\\ref{sec:results}. Model-dependent spectral fits for each component are considered in Section~\\ref{sec:modelfits} and overall conclusions given in Section~\\ref{sec:conclusions}. Appendix~\\ref{app:halpha} discusses in detail issues related to the \\halpha template used in the analysis, Appendix~\\ref{app:regions} defines the detailed method for partitioning the sky, and finally Appendix~\\ref{app:complete_results} tabulates all of the template-fit coefficients for all the regions analysed. ", "conclusions": "\\end{figure*} \\subsection{Synchrotron}\\label{sec:results_sync} \\begin{table} \\scriptsize \\begin{center} \\begin{tabular}{lccc} \\hline & \\multicolumn{3}{c}{\\bf Synchrotron spectral indices}\\\\ \\hline Region & K/408 & Ka/408 & Q/408\\\\ \\hline EBV &$\\bf -2.90 ^{+ 0.01 }_{ -0.01 }$&$\\bf -2.91 ^{+ 0.02 }_{ -0.02 }$&$\\bf -2.92 ^{+ 0.03 }_{ -0.03 }$\\\\ KQ85 &$\\bf -2.91 ^{+ 0.01 }_{ -0.01 }$&$\\bf -2.93 ^{+ 0.02 }_{ -0.02 }$&$\\bf -2.94 ^{+ 0.03 }_{ -0.04 }$\\\\ NPS &$\\bf -3.03 ^{+ 0.03}_{-0.03}$&$ -3.06 ^{+0.08}_{-0.12} $&$ < -2.90 $\\\\ GN &$\\bf -2.90 ^{+ 0.01 }_{ -0.01 }$&$\\bf -2.90 ^{+ 0.02 }_{ -0.02 }$&$\\bf -2.91 ^{+ 0.04 }_{ -0.04 }$\\\\ GN$_{\\rm reduced}$ &$\\bf -2.88 ^{+ 0.01 }_{ -0.01 }$&$\\bf -2.88 ^{+ 0.02 }_{ -0.02 }$&$\\bf -2.88 ^{+ 0.03 }_{ -0.04 }$\\\\ EN &$\\bf -2.95 ^{+ 0.01 }_{ -0.01 }$&$\\bf -2.99 ^{+ 0.03 }_{ -0.03 }$&$ -3.04 ^{+ 0.06 }_{ -0.09 }$\\\\ EN$_{\\rm reduced }$ &$\\bf -2.94 ^{+ 0.01 }_{ -0.01 }$&$\\bf -2.97 ^{+ 0.03 }_{ -0.04 }$&$ -3.02 ^{+ 0.06 }_{ -0.09 }$\\\\ GS &$\\bf -2.91 ^{+ 0.01 }_{ -0.01 }$&$\\bf -2.91 ^{+ 0.02 }_{ -0.02 }$&$\\bf -2.92 ^{+ 0.04 }_{ -0.05 }$\\\\ ES &$\\bf -2.90 ^{+ 0.01 }_{ -0.01 }$&$\\bf -2.90 ^{+ 0.02 }_{ -0.02 }$&$\\bf -2.89 ^{+ 0.03 }_{ -0.03 }$\\\\ \\hline 1 &$ < -2.96 $&$ < -2.88 $&$ < -2.80 $\\\\ 2 &$ \\bf -2.94 ^{+ 0.04 }_{ -0.05 }$&$ < -2.79 $&$ < -2.74 $\\\\ 3 &$ \\bf -3.03 ^{+ 0.03 }_{ -0.03 }$&$ -3.04 ^{+ 0.07 }_{ -0.10 }$&$ < -2.89 $\\\\ 4 &$ < -2.77 $&$ < -2.60 $&$ < -2.50 $\\\\ 5 &$ < -2.75 $&$ < -2.62 $&$ < -2.50 $\\\\ 6 &$ -3.11 ^{+ 0.10 }_{ -0.16 }$&$ < -2.87 $&$ < -2.79 $\\\\ 7 &$ -3.06 ^{+ 0.09 }_{ -0.15 }$&$ < -2.84 $&$ < -2.78 $\\\\ 8 &$ \\bf -2.86 ^{+ 0.01 }_{ -0.01 }$&$ \\bf -2.88 ^{+ 0.04 }_{ -0.04 }$&$ \\bf -2.89 ^{+ 0.06 }_{ -0.09 }$\\\\ 9 &$ -3.15 ^{+ 0.09 }_{ -0.13 }$&$ < -2.91 $&$ < -2.84 $\\\\ 10 &$ \\bf -2.90 ^{+ 0.06 }_{ -0.09 }$&$ -2.76 ^{+ 0.08 }_{ -0.14 }$&$ -2.64 ^{+ 0.08 }_{ -0.14 }$\\\\ 11 &$ \\bf -2.78 ^{+ 0.02 }_{ -0.03 }$&$ -2.88 ^{+ 0.08 }_{ -0.14 }$&$ < -2.76 $\\\\ 12 &$ \\bf -2.92 ^{+ 0.06 }_{ -0.08 }$&$ < -2.72 $&$ < -2.63 $\\\\ 13 &$ < -3.00 $&$ < -2.93 $&$ < -2.85 $\\\\ 15 &$ \\bf -2.84 ^{+ 0.04 }_{ -0.04 }$&$ < -2.78 $&$ < -2.80 $\\\\ 16 &$ \\bf -2.98 ^{+ 0.03 }_{ -0.03 }$&$ -2.97 ^{+ 0.07 }_{ -0.09 }$&$ < -2.82 $\\\\ 17 &$ \\bf -2.94 ^{+ 0.04 }_{ -0.05 }$&$ < -2.85 $&$ < -2.82 $\\\\ 18 &$ \\bf -2.90 ^{+ 0.06 }_{ -0.07 }$&$ < -2.68 $&$ < -2.59 $\\\\ 19 &$ < -2.85 $&$ < -2.67 $&$ < -2.58 $\\\\ 20 &$ < -2.82 $&$ < -2.65 $&$ < -2.53 $\\\\ 21 &$ \\bf -2.98 ^{+ 0.07 }_{ -0.09 }$&$ < -2.78 $&$ < -2.70 $\\\\ 22 &$ -2.98 ^{+ 0.09 }_{ -0.15 }$&$ < -2.69 $&$ < -2.59 $\\\\ 23 &$ < -2.94 $&$ < -2.87 $&$ < -2.81 $\\\\ 24 &$ \\bf -2.80 ^{+ 0.04 }_{ -0.05 }$&$ -2.71 ^{+ 0.07 }_{ -0.10 }$&$ -2.65 ^{+ 0.08 }_{ -0.14 }$\\\\ 25 &$ < -2.92 $&$ < -2.76 $&$ < -2.65 $\\\\ 26 &$ -2.82 ^{+ 0.10 }_{ -0.16 }$&$ < -2.53 $&$ < -2.45 $\\\\ 27 &$ -2.77 ^{+ 0.08 }_{ -0.11 }$&$ < -2.51 $&$ < -2.42 $\\\\ 28 &$ < -2.79 $&$ < -2.66 $&$ < -2.57 $\\\\ 29 &$ -2.80 ^{+ 0.10 }_{ -0.17 }$&$ < -2.50 $&$ < -2.40 $\\\\ 30 &$ < -2.82 $&$ < -2.67 $&$ < -2.57 $\\\\ 31 &$ -2.87 ^{+ 0.08 }_{ -0.13 }$&$ < -2.70 $&$ < -2.67 $\\\\ 32 &$ \\bf -2.95 ^{+ 0.04 }_{ -0.04 }$&$ -2.95 ^{+ 0.08 }_{ -0.14 }$&$ < -2.76 $\\\\ 33 &$ -2.89 ^{+ 0.08 }_{ -0.11 }$&$ < -2.62 $&$ < -2.52 $\\\\ \\hline Average &$ \\bf -2.95 ^{+ 0.01 }_{ -0.01 }$&$ \\bf -2.97 ^{+ 0.03 }_{ -0.03 } $&$ \\bf -3.00 ^{+ 0.05 }_{ -0.06 } $\\\\ \\hline \\end{tabular} \\end{center} \\caption{Synchrotron fits between 408 MHz and \\emph{WMAP }K-, Ka- and Q-bands for various regions of the sky. For those regions that are detected at 2-$\\sigma$ significance, the fit values from Table~\\ref{tab:results_synch_33regions} are converted into a spectral index using the usual power-law relation $\\beta$ ($T_b \\propto \\nu^{\\beta}$). Associated errors are determined by using the fit values plus and minus the 1$\\sigma$ error bars. Otherwise, one-sided 95\\% confidence level upper limits on the index are quoted based on the fit value plus 1.64$\\sigma$. Bold text denotes detections at 3-$\\sigma$ significance or more. For the global fits, NPS -- North Polar Spur, GN -- Galactic North, GN$_{\\rm reduced}$ -- Galactic North with the NPS removed, EN -- Ecliptic North, EN$_{\\rm reduced}$ -- Ecliptic North with the NPS removed, GS -- Galactic South, ES -- Ecliptic South.} \\label{tab:simple_synch_spectra} \\end{table} As expected, the template fit coefficients between the Haslam 408 MHz data and the \\emph{WMAP} sky maps all fall with frequency in a manner consistent with power-law emission. In Table~\\ref{tab:simple_synch_spectra} we define simple pairwise spectral indices between 408 MHz and the K-, Ka- and Q-bands. The global fits exhibit a typical index of order $-2.90$ at K-band, with the results from all masks slightly flatter than the value of $-3.01$ found in the \\citetalias{Davies_WMAP1:2006} analysis of the \\emph{WMAP} first-year data for the then-preferred Kp2 sky-coverage. Analyses of the lower-frequency surveys at 408, 1420 and 2326\\,MHz by \\citet{Giardino_synch_pol:2002} and \\citet{Platania_indices:2003} suggest a spectral index over this lower frequency range of approximately $-2.7$, thus our results support the idea of spectral steepening, continuing beyond K-band. We will consider this further in Section~\\ref{sec:modelfits}. However, there are differences in the coefficients depending on the exact sky coverage that must reflect genuine spectral variations on the sky. The NPS is recognised to be an arc of steep spectrum emission at lower frequencies, thus it is not surprising that it is notably steeper than the rest of the high latitude sky. The spectral index value of $-3.03$ at K-band is quite consistent with the value adopted in \\citet{Finkbeiner_WMAP1:2004} to remove the prominent emission from the \\emph{WMAP} data. The presence of the NPS also impacts measures of the spectral index in both the Galactic and Ecliptic northern sky, resulting in a modest steepening of the index. Interestingly, the northern Ecliptic hemisphere is notably steeper than the other hemispheres, and exhibits an increasingly steep index with frequency. Conversely, the corresponding southern hemisphere hints at spectral flattening, whereas both the north and south Galactic hemispheres are consistent with simple power-law behaviour. From the 33 regions of interest, there are 13 regions where the synchrotron fit coefficients are detected at greater than $3 \\sigma$ confidence at K-band. Most of these regions include contributions close to the Galactic plane, although regions 2, 3 and 32 are mostly at high Galactic latitude. The inferred spectral indices span the range $-2.78$ to $-3.03$, inconsistent with statistical variation alone, and likely representing genuine spectral variations on the sky. We note that region 3 contains the NPS and its behaviour seems to be dominated by that component. Significant emission is detected for region 8 at K-, Ka- and Q-band. Regions 13 and 14 correspond to an area of the sky containing the Gum nebula, and show no evidence for detection of synchrotron emission, with upper limits consistent with a steep spectrum, particularly in the southern region. Interestingly, region 9 has one of the steepest spectral indices on the sky, despite the putative presence of the \\emph{WMAP} haze. However, the overlap between the brightest regions of the haze emission as seen in \\citet{Dobler_WMAP3:2008a} and regions 9, 16 and 17 is very small, and unlikely to affect any studies here. It is interesting that the spectral indices inferred from the mean of the regional fit coefficients are steeper than the typical global fit values. This trend was also seen in our previous work \\citepalias{Davies_WMAP1:2006}, and may be due to a selection effect in that the regional subdivisions partly favour stronger synchrotron emission regions which may exhibit steeper spectra than normal due to synchrotron losses. In addition, evidence of spectral steepening with frequency is again generally seen although regions 10 and 24 show inconsistent behaviour with the other regions in that a flattening of the spectrum is indicated. \\subsection{Free-free}\\label{sec:results_freefree} In this paper, we use an H$_{\\alpha}$ template as a proxy for the free-free emission. We see significant correlation between the \\emph{WMAP} data and the template for the global fits at all frequencies (Table~\\ref{tab:results_freefree_33regions}). As with the synchrotron results, there are interesting variations depending on the exact sky coverage, with the northern Ecliptic hemisphere showing significantly enhanced amplitude, whilst the south indicates a lower emissivity. 14 individual regions are detected at 3$\\sigma$ significance at K-band. Most of these regions lie close to the Galactic plane ($|b|\\lesssim 20^{\\circ}$). \\citetalias{DDD:2003} detail the relationship between the expected free-free brightness temperature and the related H$_{\\alpha}$ intensity, and its dependence on both frequency and electron temperature ($T_e$) in the ionised medium. As can be seen from their Fig.~5, the spectral dependence of the emission shows weak curvature, but over the range of frequencies covered by \\emph{WMAP} a reasonable approximation is a power-law with index $-2.15$. Inspection of the coefficients in Table~\\ref{tab:results_freefree_33regions} indicates that there are departures from this behaviour. Following \\citet{Dobler_WMAP3:2008b} we plot these results in intensity units in Fig.~\\ref{fig:results_summary}. It should be apparent that a bump in emission is seen around 40--50~GHz for the global fits, a feature that \\citet{Dobler_WMAP3:2008b} argue is indicative of contributions from both classical free-free emission plus a spinning dust component in the WIM. The mean spectrum of the regional fits perhaps indicates a slightly broader bump in the spectrum. However, it is also the case that there is a range of behaviour seen amongst the individual regions, some of which are consistent with emission from a single physical component only - either free-free emission or a more steeply falling spectrum as expected from spinning dust. In Section~\\ref{sec:modelfits}, we will undertake a more detailed modelling of the emission in terms of these components. Such a feature in the spectrum of course has implications for the determination of physical parameters such as $T_e$. Nevertheless, we can make some general inferences, particularly by examining the K- and W-band amplitudes that are least affected by a putative WIM spinning dust component. The global fits seem to be consistent with values in the range 6000 -- 8000\\,K, with some dependency on the exact mask used. The average free-free electron temperature inferred from the 33 regions is also in this range, although for individual regions there is a spread of values between 4000 and more than 10000\\,K. These values are somewhat higher than seen previously in \\citet{Davies_WMAP1:2006}, and this is due to the use of 3$^{\\circ}$ smoothed data here, rather than the 1$^{\\circ}$ resolution data used earlier for reasons provided in Appendix~\\ref{app:halpha}. It is also interesting to note that the 7000\\,K temperature inferred from the 31.5 GHz $COBE$-DMR data at 7$^{\\circ}$ resolution in \\citet{Banday_DMR:2003} is quite consistent with the Ka-band values determined here. Moreover, the higher values in this paper are in better agreement with the electron temperatures derived from radio recombination line studies of extended HII regions \\citep{Shaver:1983,Paladini_HII:2004,Alves:2011} which derive an average $T_e$ value of $\\approx 7000$\\,K in the vicinity of the solar neighbourhood. There is therefore no need to invoke a large fraction of scattered \\halpha light to account for this discrepancy (e.g., \\citealt{Witt_Scattered_Halpha:2010}). \\subsection{Impact of Dust Extinction}\\label{sec:dust_extinct} \\begin{figure*} \\begin{tabular}{ccc} \\epsfig{file=figures/fig_41.ps,width=0.3\\linewidth,angle=0,clip=}& \\epsfig{file=figures/fig_42.ps,width=0.3\\linewidth,angle=0,clip=}& \\epsfig{file=figures/fig_43.ps,width=0.3\\linewidth,angle=0,clip=}\\\\ \\epsfig{file=figures/fig_44.ps,width=0.3\\linewidth,angle=0,clip=}& \\epsfig{file=figures/fig_45.ps,width=0.3\\linewidth,angle=0,clip=}& \\epsfig{file=figures/fig_46.ps,width=0.3\\linewidth,angle=0,clip=}\\\\ \\end{tabular} \\caption{Dependence of the correlation coefficients between the \\emph{WMAP} data and the \\halpha template as a function of the absorption correction $f_{d}$. The union mask refers to the combination of all 33 small regions defined in this paper. Thus the results for ``Union + Galactic North'' refers to the inverse-noise-weighted average coefficients for the subset of regions in the northern Galactic hemisphere. The observed dependencies for regions 13, 14 and 23 are representative of that seen for all regions with a significant detection amplitude at K-band.} \\label{fig:coefficients_fd} \\end{figure*} \\citet{Bennett_WMAP1:2003b} summarise various uncertainties in the use of the H$_{\\alpha}$ template to trace free-free emission in the Galaxy. These include uncertainties in the value of the electron temperature $T_e$, and in the value of the dust absorption correction, specified here by the $f_d$ parameter. In principle, there will be variations in both throughout the Galaxy. As noted previously, \\citetalias{DDD:2003} have determined that for local regions such as Barnard's Arc and the Gum nebula, there is little absorption by dust, and this is the default situation that we have assumed in our analyses. However, for mid-Galactic latitudes a value $f_{d} \\sim$ 0.3 is preferred, and \\citet{Finkbeiner:2003} adopted the assumption that the H$_{\\alpha}$ emission is co-extensive with dust emission along the line-of-sight, ie. $f_{d} \\sim$ 0.5. In this section we discuss the impact of varying $f_d$ on the template fit coefficients. In Fig.~\\ref{fig:coefficients_fd} we show this variation as a function of $f_d$, and compare against the behaviour expected for a range of values of $T_e$. In general, we expect that as the amplitude of the H$_{\\alpha}$ template is corrected upwards by increasing $f_d$, then the fit coefficients will decrease, implying a lower value for the electron temperature. This is indeed what is seen, but the extent of the correction depends on the region of sky under consideration. For the largest sky area that we consider, the EBV mask, the use of a template corrected absorption specified by $f_{d} = 0.5$ results in coefficients consistent with values of $T_e$ of order 3000\\,K. However, assuming no dust absorption correction yields values closer to 6000\\,K. A similar result is seen for the southern extension of the Gum nebula (region 14). If it is indeed the case that there is no evidence for dust absorption, then a higher temperature of 6000\\,K is determined. Conversely, regions 15, 18 and 21 yield coefficients at K-band consistent with temperatures closer to 10000\\,K if no absorption correction is applied, whereas values of order 8000\\,K are found for $f_d = 0.5$. Given that these regions exhibit rising spectra that may require the presence of significant emission from a spinning dust component, then the latter value would be more consistent. However, there are also regions of the sky typified by region 23, where apparently acceptable values of the electron temperature are inferred over a range of values of $f_d$. Indeed, it may be that the large spread in coefficients seen for the different regions reflects changes in the fraction of dust mixed with the WIM as much as variations in $T_e$. It does appear, therefore, that there are a range of values for $T_e$ and $f_d$ throughout the Galaxy, and reaching conclusions about their values solely from studies of H$_{\\alpha}$ correlations suffers from degeneracies between the parameters. The adoption of a value for the dust absorption correction of 0.5 in, for example, \\citet{Dobler_WMAP3:2008b}, thus seems to be associated, at least in part, with the low electron temperatures inferred. This uncertainty then has implications for modelling the emission of the diffuse component. A more serious complication would arise if the validity of using H$_{\\alpha}$ as a tracer of free-free emission were questioned. \\citet{Mattila_Scattered_Halpha:2007} have argued that the H$_{\\alpha}$ excess towards the high-latitude interstellar cloud LDN~1780 is the result of scattering of H$_{\\alpha}$ photons produced elsewhere in the Galaxy by dust in the cloud. Earlier computations by \\citet{Wood_Scattered_Halpha:1999} suggested that this contribution would typically be 5--20\\% at high-latitudes, although their model has been criticised due to the assumption of a smooth distribution of material in the ISM. \\citet{Witt_Scattered_Halpha:2010} have used an empirical relation to relate the scattered H$_{\\alpha}$ intensity in the translucent cloud LDN1780 to the \\emph{IRAS} 100~$\\mu$m diffuse background intensity and conclude that this estimate is reasonable for 50\\% of the high-latitude sky, but that the scattered contribution can be highly structured and result in contributions of between 25 and 50\\% of the observed intensity for a further 40\\% of this region of the sky. Such a result would clearly have implications for using an H$_{\\alpha}$ template to trace free-free emission, and a complex relationship between the template and dust would arise both due to the scattering contribution and due to any dust absorption correction applied to the data. Applying a correction for the former would effectively result in an increase of the H$_{\\alpha}$-\\emph{WMAP} correlation coefficients and therefore of the inferred electron temperature, whereas application of the latter results in the opposite behaviour. More recently, \\citet{Brandt_Scattered_Halpha:2011} have estimated the fraction of high-latitude H$_{\\alpha}$ that is scattered to be $19\\pm 4\\%$, a value consistent with that proposed by \\citet{Dong_WIM_Halpha:2011} to reconcile the low ratio of radio free-free to H$_{\\alpha}$. Ultimately, unravelling the degeneracy between the electron temperature, H$_{\\alpha}$ scattering and dust absorption requires additional observations. Detailed RRL surveys in the Galactic plane together with radio-continuum surveys at frequencies of $\\sim$5~GHz, as expected from the C-BASS \\citep{King_CBASS:2010} will be important in this respect. Finally, we would like to make some remarks about region 11, the coefficients of which suggest an exceptionally high temperature of more than 25000\\,K. If this result were considered unphysical, then naively a value of $f_d = 1$ would be required in order to lower the inferred temperature to the 6000\\,K seen in the EBV fit. In fact, a more realistic assessment of the situation is that the EBV mask is not large enough to eliminate some parts of region 11 close to the Galactic plane where the simple dust absorption correction is untrustworthy. If instead we apply the KQ85 mask before analysing the region, then for a more plausible value of $f_{d} = 0.5$ the inferred electron temperature is again consistent with 6000\\,K. \\subsection{Dust}\\label{sec:results_dust} As can be seen in the right-hand panels of Fig.~\\ref{fig:results_summary}, the template fit coefficients determined between the \\emph{WMAP} data and the FDS8 dust template prediction at W-band are consistent with a rising thermal dust contribution at frequencies higher than V-band, and a rising spectrum to lower frequencies below it. The latter corresponds to the now widely identified anomalous microwave emission (AME). It appears that the FDS8 template underpredicts the W-band amplitude by approximately 30\\% for the global fits, consistent with the results of \\citetalias{Davies_WMAP1:2006}. A broad range of values for the individual regions is seen, but only three detect emission at a statistically significant level. The mean emissivity at W-band of all regions shows a more modest enhancement, but is nevertheless consistent with the FDS8 predictions, as indeed are the three significant regions. At K-band, there are variations in the global fit amplitudes depending on sky coverage. The KQ85 mask indicates a lower emissivity compared to the EBV as might be expected. The Southern Galactic and Ecliptic hemispheres have the highest amplitudes. All coefficients are higher than those determined for the Kp2 sky coverage in \\citetalias{Davies_WMAP1:2006}. 26 of the individual regions detect emission at 3$\\sigma$ significance or higher. The mean emission amplitude lies between that for EBV and KQ85, and again somewhat higher than in \\citetalias{Davies_WMAP1:2006}. The regions indicate a variation around the mean of approximately 50\\% of its amplitude, inconsistent with statistical errors alone and indicating genuine spatial variations in the AME emissivity. The overall spectrum would appear to be well-described by a superposition of two power-law emissivities. The thermal dust emission described by the FDS8 model is adequately represented by an emissivity index $\\beta_d = 1.55$ over the \\emph{WMAP} range of frequencies. We fit the power-law AME spectrum with the thermal dust index fixed to this value, and find AME spectral indices of order $-2.7$ for both the global fits and regional mean. \\citet{Draine_spinning:1998} first proposed that the AME could be explained by electric dipole radiation from rotationally excited small interstellar grains, or spinning dust. Our results have ramifications for such models of the emission. In particular, given that the spinning dust spectra typically fall off steeply with frequency beyond their peak, then it is unlikely that a single such spectrum could account adequately for the effective power-law emission. Indeed, the observed spectrum is presumably formed from a superposition of components with varying spectra as a consequence of their differing physical environments along a given line-of-sight. We will discuss detailed fits of the observed emission in Section~\\ref{sec:modelfits}, and their implications for more refined models of the AME. The following analytic forms are used to fit the synchrotron coefficients. \\begin{itemize} \\item Model SI : Given the power law energy distribution of cosmic ray electrons, we assumed a power-law emissivity in terms of brightness temperature over the WMAP frequencies as \\begin{equation} T_A (\\nu) = A_{s} \\times \\left( \\frac{\\nu}{23} \\right )^{\\beta_{s}}_{\\text{GHz}} \\end{equation} where $\\beta_s$ is the spectral index and $A_s$ is the normalised amplitude expressed in $\\mu \\text{K}$ with respect to the frequency $\\nu_0 = 23$\\,GHz. \\item Model SII : A power-law emissivity is assumed to extend from 408~MHz up to and through the WMAP frequencies. Since we use the 408~MHz survey as a template for the synchrotron emission, the amplitude at the low frequency must be reproduced perfectly. This results in an effective constraint to be applied to the analytical form above, and we then fit for the spectral index $\\beta_{s}$ only. \\begin{equation} T_A (\\nu) = 10^6 \\times \\left( \\frac{\\nu}{0.408} \\right )^{\\beta_{s}}_{\\text{GHz}} \\end{equation} \\item Model SIII : The cosmic ray electron energy spectrum is expected to steepen with time due to radiation energy loss. A review of cosmic-ray propagation including electrons can be found in \\citet{Strong_etal:2007}, whilst \\citet{Strong_etal:2011} directly test propagation models based on cosmic-ray and gamma-ray data against synchrotron data from 22~MHz to 94~GHz as averaged over mid-latitude regions ($10^{\\circ} < \\mid b \\mid < 45^{\\circ}$). The latter analysis confirms the need for a low-energy break in the cosmic-ray electron injection spectrum to account for the steepening synchrotron spectrum. Since we do not include synchrotron information at frequencies intermediate to 408~MHz and the \\emph{WMAP} data, we follow the treatment of \\citet{Gold_WMAP5:2009}. Specifically, the emissivity is assumed to follow a power-law from 408~MHz until K-band and then to exhibit spectral curvature as follows, \\begin{align*} T_A (\\nu) &= A_{s} \\times \\left( \\frac{\\nu}{23} \\right )_{\\text{GHz}}^{\\beta_{s}} &\\nu < \\nu_\\text{K} \\\\ &= A_{s} \\times \\left( \\frac{\\nu}{23} \\right )_{\\text{GHz}}^{\\beta_{s} + c_s ln(\\frac{\\nu}{\\nu_\\text{K}})} & \\nu > \\nu_\\text{K} \\end{align*} As above, the 408~MHz point is fixed, thus $A_s$ can be written in terms of $\\beta_s$, and we are left to fit this spectral index and the curvature $c_s$. For a WMAP frequency point, the above equation reduces to the form: \\begin{equation} T_A (\\nu) = 10^6 \\times \\left( \\frac{\\nu}{0.408} \\right )^{\\beta_s}_{\\text{GHz}} \\times \\left( \\frac{\\nu}{23} \\right )^{ c_s ln(\\frac{\\nu}{\\nu_K})}_{\\text{GHz}} \\end{equation} \\end{itemize} \\begin{table} \\scriptsize \\begin{center} \\begin{tabular}{lcccccccc} \\hline &\\multicolumn{3}{c}{SI} &\\multicolumn{2}{c}{SII} &\\multicolumn{3}{c}{SIII}\\\\ \\hline Region & $A_{s}$ & $\\beta_{s}$ & $\\chi^2$ & $\\beta_{s}$ & $\\chi^2$ & $\\beta_{s}$ & $c_s$ & $\\chi^2$ \\\\ \\hline EBV & 7.96 $\\pm$ 0.18 & -2.99 $\\pm$ 0.14 & 0.047 & -2.91 $\\pm$ 0.01 & 0.106 & -2.91 $\\pm$ 0.01 & -0.17 $\\pm$ 0.28 & 0.011 \\\\ KQ85 & 7.64 $\\pm$ 0.20 & -3.11 $\\pm$ 0.19 & 0.246 & -2.92 $\\pm$ 0.01 & 0.470 & -2.92 $\\pm$ 0.01 & -0.44 $\\pm$ 0.39 & 0.071 \\\\ NPS & 4.68 $\\pm$ 0.55 & -3.43 $\\pm$ 0.92 & 0.021 & -3.05 $\\pm$ 0.03 & 0.069 & -3.04 $\\pm$ 0.03 & -0.91 $\\pm$ 2.18 & 0.003 \\\\ GN & 8.16 $\\pm$ 0.25 & -2.94 $\\pm$ 0.19 & 0.007 & -2.91 $\\pm$ 0.01 & 0.013 & -2.91 $\\pm$ 0.01 & -0.06 $\\pm$ 0.35 & 0.006 \\\\ EN & 6.41 $\\pm$ 0.26 & -3.55 $\\pm$ 0.33 & 1.077 & -2.97 $\\pm$ 0.01 & 1.914 & -2.96 $\\pm$ 0.01 & -1.38 $\\pm$ 0.81 & 0.628 \\\\ GS & 7.84 $\\pm$ 0.26 & -2.99 $\\pm$ 0.21 & 0.082 & -2.92 $\\pm$ 0.01 & 0.093 & -2.92 $\\pm$ 0.01 & -0.20 $\\pm$ 0.42 & 0.031 \\\\ ES & 7.95 $\\pm$ 0.22 & -2.74 $\\pm$ 0.16 & 0.337 & -2.91 $\\pm$ 0.01 & 0.504 & -2.91 $\\pm$ 0.01 & 0.37 $\\pm$ 0.23 & 0.075 \\\\ \\hline 2 & 6.82 $\\pm$ 1.25 & -3.05 $\\pm$ 1.23 & 0.018 & -2.95 $\\pm$ 0.04 & 0.015 & -2.95 $\\pm$ 0.05 & -0.34 $\\pm$ 2.59 & 0.012 \\\\ 3 & 4.66 $\\pm$ 0.53 & -3.03 $\\pm$ 0.76 & 0.006 & -3.04 $\\pm$ 0.03 & 0.005 & -3.05 $\\pm$ 0.03 & 0.08 $\\pm$ 1.30 & 0.005 \\\\ 8 & 9.23 $\\pm$ 0.53 & -3.01 $\\pm$ 0.38 & 0.002 & -2.88 $\\pm$ 0.01 & 0.037 & -2.87 $\\pm$ 0.01 & -0.25 $\\pm$ 0.72 & 0.005 \\\\ 10 & 7.67 $\\pm$ 2.04 & -0.56 $\\pm$ 0.42 & 0.076 & -2.86 $\\pm$ 0.05 & 2.305 & -2.87 $\\pm$ 0.05 & 1.69 $\\pm$ 0.28 & 0.700 \\\\ 11 & 12.61 $\\pm$ 1.24 & -4.33 $\\pm$ 1.16 & 0.560 & -2.81 $\\pm$ 0.02 & 1.403 & -2.79 $\\pm$ 0.03 & -3.67 $\\pm$ 3.07 & 0.441 \\\\ 12 & 7.33 $\\pm$ 1.88 & -2.09 $\\pm$ 1.03 & 0.028 & -2.92 $\\pm$ 0.06 & 0.121 & -2.93 $\\pm$ 0.06 & 1.01 $\\pm$ 0.88 & 0.010 \\\\ 15 & 9.74 $\\pm$ 1.62 & -5.38 $\\pm$ 2.78 & 1.284 & -2.88 $\\pm$ 0.04 & 1.801 & -2.85 $\\pm$ 0.04 & -5.99 $\\pm$ 7.08 & 1.194 \\\\ 16 & 5.88 $\\pm$ 0.62 & -2.96 $\\pm$ 0.69 & 0.002 & -2.99 $\\pm$ 0.03 & 0.002 & -2.99 $\\pm$ 0.03 & 0.01 $\\pm$ 1.32 & 0.002 \\\\ 17 & 6.42 $\\pm$ 1.26 & -5.08 $\\pm$ 3.08 & 0.574 & -2.98 $\\pm$ 0.05 & 0.822 & -2.95 $\\pm$ 0.05 & -4.99 $\\pm$ 7.95 & 0.527 \\\\ 18 & 7.95 $\\pm$ 1.96 & -1.62 $\\pm$ 0.77 & 0.070 & -2.89 $\\pm$ 0.05 & 0.373 & -2.90 $\\pm$ 0.06 & 1.26 $\\pm$ 0.55 & 0.055 \\\\ 21 & 5.76 $\\pm$ 1.69 & -3.18 $\\pm$ 1.99 & 0.001 & -2.99 $\\pm$ 0.07 & 0.003 & -2.99 $\\pm$ 0.07 & -0.40 $\\pm$ 4.20 & 0.001 \\\\ 22 & 5.65 $\\pm$ 2.39 & -1.44 $\\pm$ 1.17 & 0.031 & -2.97 $\\pm$ 0.09 & 0.195 & -2.98 $\\pm$ 0.09 & 1.40 $\\pm$ 0.83 & 0.038 \\\\ 24 & 11.98 $\\pm$ 2.20 & -1.36 $\\pm$ 0.48 & 0.104 & -2.78 $\\pm$ 0.04 & 1.056 & -2.79 $\\pm$ 0.04 & 1.27 $\\pm$ 0.35 & 0.231 \\\\ 32 & 6.64 $\\pm$ 0.98 & -2.91 $\\pm$ 0.93 & 0.007 & -2.96 $\\pm$ 0.04 & 0.006 & -2.96 $\\pm$ 0.04 & 0.05 $\\pm$ 1.63 & 0.007 \\\\ \\hline \\end{tabular} \\end{center} \\caption{Model fits to the synchrotron coefficients determined between the 5 WMAP frequencies and the Haslam 408 MHz template for large sky areas (upper part of table) and for those regions that indicate a 3$\\sigma$ significant amplitude at K-band (lower part). The models SI, SII and SIII are described in Section~\\ref{sec:discussion_synch}. $A_{s}$ represents the normalisation amplitude at K-band, $\\beta_{s}$ the synchrotron spectral index and $c_{s}$ the spectral curvature. The key for the global fits is as for Table~\\ref{tab:simple_synch_spectra}. \\label{tab:synch_model_results}} \\end{table} \\begin{center} \\begin{figure} \\begin{tabular}{ccc} \\epsfig{file=figures/fig_51.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_52.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_53.ps,width=0.3\\linewidth,angle=0,clip=} \\\\ \\epsfig{file=figures/fig_54.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_55.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_56.ps,width=0.3\\linewidth,angle=0,clip=} \\\\ \\end{tabular} \\caption{Synchrotron spectral fits for various regions of the sky. A comparison is made of the observed template fits amplitudes derived from the 5 \\emph{WMAP} frequency bands and the 408~MHz survey with the 3 models SI, SII and SIII as defined in Section~\\ref{sec:discussion_synch}. The observed spectral dependencies of regions 3, 10 and 15 are representative of those seen for all regions with a significant detection at K-band.\\label{fig:synch_model_plots}} \\end{figure} \\end{center} The results are summarised in Table~\\ref{tab:synch_model_results}. Fig.~\\ref{fig:synch_model_plots} presents a comparison of the model fits for three of the global masks, plus three of the regions that can be considered representative of the general results seen. The $\\chi^2$ results generally should be considered indicative rather than definitive, as would be expected given the small number of data points and the comparable number of model parameters. The synchrotron fit coefficients obtained for the global masks that include the 408~MHz datum as a reference point (model SII) indicate a typical power-law spectral index of $\\beta_s=-2.9\\pm0.1$. Model SI fits to the 5 \\emph{WMAP} frequencies are generally steeper, however, only the Ecliptic North region shows evidence of such behaviour at a significant level (a 2$\\sigma$ shift in the spectral index). Indeed, a model with negative spectral curvature is the best fit to the data, indicating steepening due to cosmic ray energy loss mechanisms. Curiously, the Ecliptic South indicates spectral flattening, although at much lower significance. The SII fit to the NPS region finds a spectral index $\\beta_s=-3.05\\pm0.03$ which is steeper than the average spectral index variation over the remaining sky, as expected \\citep{Lawson:1987}. For the individual regions, the results are generally similar to the global fits. The spectral behaviour is consistent with emission from power-law cosmic ray electron spectra with only hints of a steepening between 408~MHz and K-band. Region 11 might be considered to show weak evidence for negative spectral curvature. More interestingly, three regions (10, 12 and 24) indicate positive curvature that might be understood as due to the presence of multiple emission regions with varying spectral behaviour along the line-of-sight. However, region 10 has a very flat spectrum between K- and W-band that is strongly favoured over a steeper spectrum extending to 408~MHz, and perhaps even mildly inconsistent with the curvature model adopted here. Whether this indicates some problem with the analysis in this region, due to problems with the templates or cross-talk between components, is difficult to determine. It is apparent that the region lies close to the North Celestial Pole where the 408~MHz template still retains considerable striations from the original survey, whereas we certainly consider that the template fitting methodology has by now been extensively tested. However, the spectral flattening seen in other analyses is less dramatic. For example, \\citet{Gold_WMAP7:2011} find a range of values of $c_{s}$ between $\\pm 0.7$ with a variance of $\\sim 0.4$ for their MCMC analysis of the \\emph{WMAP} data combined with the 408~MHz data. \\subsection{Free-free}\\label{sec:discussion_freefree} The free-free spectral index is almost constant over the \\emph{WMAP} frequency range and changes only slightly with the the electron temperature. Over this frequency range, the spectral dependence of the corresponding \\halpha emission can then be approximated with a power-law model of fixed spectral index $-0.15$. However, \\citet{Dobler_WMAP3:2008a} observed deviations from such power-law behaviour, which they attributed to the presence of spinning dust emission in the WIM also traced by \\halpha. We investigate our template fit coefficients in terms of the following three models. \\begin{itemize} \\item Model FI : We consider that the emission is due entirely to the free-free mechanism, ie. a power-law model emissivity with a fixed spectral index of $-0.15$ is assumed, \\begin{equation} I(\\nu) = A_{f}\\times \\left( \\frac{\\nu}{23} \\right )^{-0.15}_{\\text{GHz}} \\end{equation} Thus only the free-free amplitude ($A_{f}$) needs to be estimated from the template fit coefficients. \\\\ \\item Model FII : The coefficients are fitted with an empirically motivated power-law model emissivity. This is particularly illustrative as to the extent that the standard free-free emissivity index is a poor fit to the data. \\begin{equation} I(\\nu) = A_{PL} \\times \\left( \\frac{\\nu}{23} \\right )^{\\beta_{PL}}_{\\text{GHz}} \\end{equation} In this case, both the foreground amplitude ($A_{PL}$) and spectral index ($\\beta_{PL}$) are to be estimated.\\\\ \\item Model FIII : The coefficients are fitted with a combination of free-free emission and a spinning dust model computed by the {\\tt{SPDUST2}} code \\citep{SpDust2:2011} for typical WIM conditions. This model peaks around 30~GHz, but we allow a simple shift $\\Delta \\nu_{\\text{WIM}}$ to be applied to the spectrum, as an approximation that represents the effect of varying the WIM physical parameters to match the model spectrum to the data. \\begin{equation} I(\\nu) = A_{f} \\times\\left( \\frac{\\nu}{23} \\right)^{-0.15}_{\\text{GHz}} + A_{\\text{WIM}} \\times D_{\\text{WIM}}(\\nu - \\Delta \\nu_{\\text{WIM}}) \\end{equation} Here, $D_{\\text{WIM}}(\\nu)$ represents the normalised spinning dust spectral model for the WIM at a given frequency. Clearly, we must now fit 3 parameters -- the free-free amplitude ($A_{f}$), WIM amplitude ($A_{\\text{WIM}}$ ) and WIM frequency shift ($\\Delta \\nu_{\\text{WIM}}$). We have also considered fits of this model to the coefficients derived using the \\halpha after correction for dust absorption ($f_{d}=0.5$).\\\\ \\end{itemize} \\begin{table} \\scriptsize \\begin{center} \\begin{tabular}{lccccccccc} \\hline & \\multicolumn{2}{c}{FI} & \\multicolumn{3}{c}{FII} & \\multicolumn{4}{c}{FIII} \\\\ \\hline Region & $A_{f}$ & $\\chi^2$ & $A_{PL}$ & $\\beta_{PL}$ & $\\chi^2$ & $A_{f}$ & $A_{WIM}$ & $\\Delta \\nu_{\\text{WIM}}$ & $\\chi^2$ \\\\ \\hline EBV & 0.156 $\\pm$ 0.001 & 6.374 & 0.153 $\\pm$ 0.002 & 0.010 $\\pm$ 0.036 & 1.633 & 0.150 $\\pm$ 0.004 & 2.094 $\\pm$ 0.429 & 13.575 $\\pm$ 4.727 & 0.340 \\\\ KQ85 & 0.166 $\\pm$ 0.003 & 2.449 & 0.163 $\\pm$ 0.003 & 0.061 $\\pm$ 0.067 & 0.205 & 0.161 $\\pm$ 0.004 & 2.822 $\\pm$ 1.204 & 15.930 $\\pm$ 6.463 & 0.439 \\\\ GN & 0.163 $\\pm$ 0.002 & 2.057 & 0.161 $\\pm$ 0.003 & -0.073 $\\pm$ 0.065 & 2.264 & 0.141 $\\pm$ 0.017 & 3.483 $\\pm$ 1.979 & 5.343 $\\pm$ 3.317 & 0.704 \\\\ EN & 0.210 $\\pm$ 0.004 & 2.177 & 0.207 $\\pm$ 0.004 & 0.031 $\\pm$ 0.070 & 0.730 & 0.203 $\\pm$ 0.008 & 3.442 $\\pm$ 1.245 & 14.290 $\\pm$ 7.196 & 0.062 \\\\ GS & 0.156 $\\pm$ 0.002 & 4.622 & 0.153 $\\pm$ 0.002 & 0.039 $\\pm$ 0.044 & 0.296 & 0.152 $\\pm$ 0.002 & 2.386 $\\pm$ 0.806 & 16.354 $\\pm$ 4.503 & 1.163 \\\\ ES & 0.146 $\\pm$ 0.001 & 3.499 & 0.144 $\\pm$ 0.002 & -0.003 $\\pm$ 0.041 & 1.031 & 0.141 $\\pm$ 0.005 & 1.775 $\\pm$ 0.491 & 12.932 $\\pm$ 6.538 & 0.202 \\\\ \\hline 7 & 0.126 $\\pm$ 0.029 & 1.003 & 0.141 $\\pm$ 0.030 & -2.043 $\\pm$ 2.257 & 0.601 & 0.000 $\\pm$ 0.000 & 15.936 $\\pm$ 3.512 & 0.000 $\\pm$ 0.000 & 1.381 \\\\ 8 & 0.111 $\\pm$ 0.021 & 0.086 & 0.110 $\\pm$ 0.024 & -0.043 $\\pm$ 0.831 & 0.109 & 0.053 $\\pm$ 0.152 & 8.168 $\\pm$ 18.511 & 3.214 $\\pm$ 10.241 & 0.024 \\\\ 9 & 0.162 $\\pm$ 0.004 & 3.619 & 0.156 $\\pm$ 0.004 & 0.205 $\\pm$ 0.101 & 1.046 & 0.152 $\\pm$ 0.006 & 5.470 $\\pm$ 1.890 & 15.818 $\\pm$ 5.131 & 0.115 \\\\ 11 & 0.388 $\\pm$ 0.011 & 0.298 & 0.386 $\\pm$ 0.012 & -0.105 $\\pm$ 0.128 & 0.358 & 0.388 $\\pm$ 0.011 & 0.000 $\\pm$ 0.000 & 0.000 $\\pm$ 0.000 & 0.596 \\\\ 12 & 0.135 $\\pm$ 0.034 & 0.731 & 0.150 $\\pm$ 0.036 & -1.693 $\\pm$ 2.087 & 0.481 & 0.000 $\\pm$ 0.000 & 17.099 $\\pm$ 4.139 & 0.000 $\\pm$ 0.000 & 0.909 \\\\ 13 & 0.145 $\\pm$ 0.005 & 0.037 & 0.145 $\\pm$ 0.006 & -0.119 $\\pm$ 0.169 & 0.039 & 0.137 $\\pm$ 0.038 & 1.238 $\\pm$ 4.593 & 4.601 $\\pm$ 20.905 & 0.003 \\\\ 14 & 0.156 $\\pm$ 0.003 & 2.336 & 0.153 $\\pm$ 0.003 & 0.060 $\\pm$ 0.066 & 0.096 & 0.152 $\\pm$ 0.003 & 2.858 $\\pm$ 1.513 & 17.556 $\\pm$ 5.663 & 1.060 \\\\ 15 & 0.200 $\\pm$ 0.011 & 0.836 & 0.191 $\\pm$ 0.012 & 0.227 $\\pm$ 0.192 & 0.051 & 0.192 $\\pm$ 0.012 & 6.819 $\\pm$ 6.145 & 20.334 $\\pm$ 7.279 & 0.627 \\\\ 18 & 0.235 $\\pm$ 0.027 & 0.772 & 0.216 $\\pm$ 0.029 & 0.620 $\\pm$ 0.358 & 0.018 & 0.219 $\\pm$ 0.031 & 19.115 $\\pm$ 17.644 & 21.162 $\\pm$ 7.240 & 0.612 \\\\ 20 & 0.135 $\\pm$ 0.012 & 1.495 & 0.123 $\\pm$ 0.013 & 0.685 $\\pm$ 0.276 & 0.116 & 0.127 $\\pm$ 0.014 & 14.890 $\\pm$ 9.172 & 23.658 $\\pm$ 4.864 & 1.101 \\\\ 21 & 0.250 $\\pm$ 0.043 & 1.159 & 0.212 $\\pm$ 0.044 & 1.062 $\\pm$ 0.411 & 0.014 & 0.222 $\\pm$ 0.048 & 42.774 $\\pm$ 29.924 & 23.582 $\\pm$ 5.653 & 0.893 \\\\ 23 & 0.164 $\\pm$ 0.023 & 0.157 & 0.168 $\\pm$ 0.025 & -0.447 $\\pm$ 0.708 & 0.134 & 0.066 $\\pm$ 0.170 & 12.174 $\\pm$ 20.912 & 0.000 $\\pm$ 0.000 & 0.144 \\\\ 24 & 0.118 $\\pm$ 0.005 & 0.340 & 0.116 $\\pm$ 0.006 & 0.074 $\\pm$ 0.195 & 0.048 & 0.114 $\\pm$ 0.015 & 1.977 $\\pm$ 1.888 & 13.646 $\\pm$ 21.049 & 0.093 \\\\ 32 & 0.164 $\\pm$ 0.041 & 0.374 & 0.144 $\\pm$ 0.042 & 0.914 $\\pm$ 0.657 & 0.019 & 0.150 $\\pm$ 0.046 & 19.687 $\\pm$ 27.751 & 22.391 $\\pm$ 10.970 & 0.368 \\\\ \\hline \\end{tabular} \\end{center} \\caption{Model fits to the free-free coefficients determined between the 5 WMAP frequencies and the DDD \\halpha template for large sky areas (upper part of table) and for those regions that indicate a 3$\\sigma$ significant amplitude at K-band (lower part). The models FI, FII and FIII are fully defined in Section~\\ref{sec:discussion_freefree}. $A_{f}$ represents the normalisation amplitude of the free-free emission at K-band, $A_{PL}$ represents the normalisation amplitude for the power-law model emission at K-band, $\\beta_{PL}$ the corresponding power-law spectral index, $A_{WIM}$ is the amplitude of the WIM spinning dust model in units of 10$^{20}$ R cm$^{-2}$, and $\\Delta \\nu_{\\text{WIM}}$ the shift in frequency of the peak of the dust model to better the fit the data. The key for the global fits as for Table~\\ref{tab:simple_synch_spectra}. \\label{tab:freefree_model_hemispheres}} \\end{table} \\begin{center} \\begin{figure} \\begin{tabular}{ccc} \\epsfig{file=figures/fig_61.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_62.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_63.ps,width=0.3\\linewidth,angle=0,clip=} \\\\ \\epsfig{file=figures/fig_64.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_65.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_66.ps,width=0.3\\linewidth,angle=0,clip=} \\\\ \\end{tabular} \\caption{Free-free spectral fits for various regions of the sky. A comparison is made of the observed template fits amplitudes derived from the 5 \\emph{WMAP} frequency bands and the DDD \\halpha template with the 3 models FI, FII and FIII as defined in Section~\\ref{sec:discussion_freefree}. Also shown in the plot are the separate free-free (FF) and WIM spinning dust components that constitute model FIII. The observed spectral dependencies of regions 7, 13 and 14 are representative of those seen for all regions with a significant detection at K-band. \\label{fig:freefree_model_hemispheres}} \\end{figure} \\end{center} The results are summarised in Table~\\ref{tab:freefree_model_hemispheres}. Fig.~\\ref{fig:freefree_model_hemispheres} presents a comparison of the model fits for three of the global masks, plus three of the regions that can be considered representative of the general results seen. The global fits all indicate a deviation from the free-free emission law (FI) at high significance. Unconstrained power-law fits (FII) are generally flatter, and models containing both free-free and spinning dust emission (FII) are typically preferred. For the EBV mask, the ratio of free-free to spinning dust emission (at the shifted peak frequency) is of order 7, but this drops to 4 for the Galactic North. All regional fits are consistent with a 15\\,GHz shift to higher frequencies of the peak intensity for the WIM spinning dust component with the exception of the Galactic North region which favours a smaller value of 5\\,GHz. The fits to individual regions are generally consistent with the global mask results, and only region 9 shows a clear rejection of model FI. The typical frequency shift is again of order 15~GHz, with a larger dispersion with several regions preferring no shift at all. Of those, regions 7 and 12 can be explained by a single component - either free-free, power-law emission or spinning dust alone with no frequency shift. However, it should be noted that the \\halpha fit coefficients for these regions are only significant at K-band. Region 11 is consistent with free-free only emission, and yields no evidence for spinning dust, but the region itself has been flagged as anomalous as discussed previously in section \\ref{sec:dust_extinct}. Region 13 corresponds to the Northern Gum nebula, and is dominated by free-free emission with a small contribution from spinning dust at a lower peak frequency, $\\sim 35$~GHz, than is typical. The Southern part of the nebula is contained in region 14. This is the most clearly detected structure on the sky, significant at all frequencies, as traced by the \\halpha template, and shows significant evidence for a spinning dust contribution with a free-free to spinning dust ratio of $\\sim$ 5. Otherwise, this ratio varies considerably from region to region. Some regions that have a rising spectrum in terms of the \\halpha coefficients and naturally favour models with a spinning dust contribution over pure free-free are better fitted still by a flatter/rising power-law emission model. This might be alleviated with more detailed spinning dust modelling (beyond the scope of our paper), or by including physical effects that increase the spinning dust amplitude at frequencies higher than the peak. Such modifications have been investigated by \\citet{Hoang:2010}. The $A_{f}$ fit coefficients can, of course, be converted to estimates of the thermal electron temperature in the ionised medium. Since the global mask fits require the presence of a spinning dust component, we only consider the $T_e$ results from model FIII, as presented in Table~\\ref{tab:freefree_electron_temperatures}. The table also includes results derived from fits to an \\halpha template corrected for dust absorption assuming $f_d=0.5$. We do not include the detailed coefficient results here since the interpretation presented above remains essentially unchanged, and only conclusions about $T_{e}$ are affected. The global masks are consistent with values of the electron temperature of order 6000\\,K without any dust absorption correction, falling to 3000\\,K when $f_d=0.5$ is assumed. These values are for guidance only -- since the regions are not independent, an average is meaningless. It is interesting to note that the Ecliptic North shows an enhanced temperature some 50\\% higher than these typical values. Whether this reflects some property of the \\halpha template is difficult to say, but the coverage is dominated by measurements from \\emph{WHAM} data. Conversely, the Ecliptic South global fit gives a lower temperature that may also suggest issues with the template, given that it is largely comprised of the \\emph{SHASSA} fields. New observations from the southern extension of the \\emph{WHAM} survey \\citep{Haffner_WHAM-S:2011} should help resolve this issue in the near future. Nevertheless, problems may still remain near the ecliptic poles given difficulties with removing the geocoronal \\halpha contribution. For the individual regions (ignoring region 11 which is considered to be anomalous), we find a weighted average of $6300 \\pm 200$\\,K (dropping to 5900\\,K if region 11 is excluded) without a correction for dust absorption, and $2900\\pm 100$\\,K otherwise. These values are in good agreement with the global averages as might be expected. However, note that the dispersion of values is considerably larger than the quoted uncertainty, implying true variations in temperature on the sky. Moreover, as discussed previously, it is likely that some of the dispersion seen reflects the existence of a range of values for both $T_{e}$ and $f_{d}$ throughout the Galaxy. \\begin{table} \\scriptsize \\begin{center} \\begin{tabular}{lcc} \\hline & \\multicolumn{2}{c}{Inferred $T_{e}$ (K) for model FIII} \\\\ Region & $f_{d} = 0.0$ & $f_{d} = 0.5$ \\\\ \\hline EBV & 5900 $\\pm$ 300 & 2600 $\\pm$ 200 \\\\ KQ85 & 6600 $\\pm$ 300 & 3200 $\\pm$ 200 \\\\ GN & 5300 $\\pm$ 1000 & 2500 $\\pm$ 500 \\\\ EN & 9500 $\\pm$ 600 & 5300 $\\pm$ 600 \\\\ GS & 6000 $\\pm$ 100 & 2600 $\\pm$ 100 \\\\ ES & 5300 $\\pm$ 300 & 2200 $\\pm$ 200 \\\\ \\hline 7$^{(a)}$ & 4400 $\\pm$ 1700 & 5600 $\\pm$ 1100 \\\\ 8$^{(a)}$ & 3600 $\\pm$ 1200 & 2300 $\\pm$ 900 \\\\ 9 & 6000 $\\pm$ 400 & 3200 $\\pm$ 300 \\\\ 11 & 25500 $\\pm$ 1100 & 11800 $\\pm$ 500 \\\\ 12$^{(a)}$ & 5000 $\\pm$ 2100 & 3300 $\\pm$ 1300 \\\\ 13 & 5100 $\\pm$ 2200 & 3500 $\\pm$ 1500 \\\\ 14 & 6000 $\\pm$ 200 & 2700 $\\pm$ 100 \\\\ 15 & 8600 $\\pm$ 800 & 5900 $\\pm$ 600 \\\\ 18 & 10700 $\\pm$ 2300 & 7700 $\\pm$ 1700 \\\\ 20 & 4500 $\\pm$ 800 & 2400 $\\pm$ 500 \\\\ 21 & 10900 $\\pm$ 3600 & 4000 $\\pm$ 1800 \\\\ 23$^{(a)}$ & 6800 $\\pm$ 1400 & 5400 $\\pm$ 1000 \\\\ 24 & 3700 $\\pm$ 600 & 2300 $\\pm$ 700 \\\\ 32 & 5900 $\\pm$ 2900 & 5600 $\\pm$ 2500 \\\\ \\hline \\end{tabular} \\end{center} \\caption{Inferred free-free electron temperature $T_{e}$ in Kelvins corresponding to model fit FIII with dust corrections $f_d=0.0$ and $0.5$. $^(a)$ These regions show an effective degeneracy between free-free only or WIM spinning dust only solutions, thus FIII solutions have no contribution from free-free emission. We have therefore used the FI results to compute $T_{e}$ in these cases.} \\label{tab:freefree_electron_temperatures} \\end{table} \\subsection{Dust}\\label{sec:discussion_dust} The total dust emission is modelled as a combination of the relatively well-understood thermal dust emission and the AME. The former is assumed to have a fixed spectral index relative to the FDS8 94~GHz template over the \\emph{WMAP} frequency range as determined directly from the FDS8 model. We consider the following two models in order to fit the dust coefficients. \\begin{itemize} \\item Model DI : The dust coefficients are fitted with a combination of thermal (vibrational) dust and a power law dust-correlated AME. \\\\ \\begin{equation} T_A (\\nu) = A_{PLD} \\times \\left( \\frac{\\nu}{23.} \\right)^{\\beta_{PLD}}_{\\text{GHz}} + A_{TD}\\times \\left( \\frac{\\nu}{94} \\right)^{1.55}_{\\text{GHz}} \\end{equation} Three parameters -- the thermal dust amplitude ($A_{TD}$), the power law dust amplitude ($A_{PLD}$) and power law dust spectral index ($\\beta_{PLD}$) -- are fitted to the coefficients. \\item Model DII : The dust coefficients are fitted with a combination of thermal dust and two spinning dust components (CNM and WNM). The two spectra are generated using the {\\tt{SPDUST2}} code assuming typical CNM and WNM conditions \\citep{Draine_spinning:1998}. Given that both spectra peak at approximately the same frequency ($\\approx 30$\\,GHz), and the limited number of degrees of freedom available in the fit, it is only possible to apply a frequency shift to one component in order to match observations. \\citet{Hoang:2011} found that modifying the CNM properties to increase its peak frequency yields a closer match to the \\emph{WMAP} observations, and therefore we elect to allow a frequency shift of this component. \\\\ \\begin{equation} T_A (\\nu) = A_{\\text{WNM}} \\times D_{\\text{WNM}}(\\nu) + A_{\\text{CNM}} \\times D_{\\text{CNM}}(\\nu - \\Delta \\nu_{\\text{CNM}}) + A_{TD} \\times \\left( \\frac{\\nu}{94} \\right )^{1.55}_{\\text{GHz}} \\end{equation} We fit four parameters to the template fit coefficients: the thermal dust amplitude ($A_{TD}$), the WNM amplitude ($A_{\\text{WNM}}$) normalised at 23~GHz, the CNM amplitude ($A_{\\text{CNM}}$) normalised at 41~GHz, and the CNM peak frequency shift ($\\Delta \\nu_{\\text{CNM}}$). \\end{itemize} \\begin{table} \\scriptsize \\begin{center} \\begin{tabular}{lccccccccc} \\hline Region &\\multicolumn{4}{c}{Model DI} &\\multicolumn{5}{c}{Model DII}\\\\ \\hline & $A_{PLD}$ &$\\beta_{PLD}$ & $A_{TD}$ & $\\chi^2$ & $A_{WNM}$ & $A_{CNM}$ & $\\Delta {\\nu_{\\text{CNM}}}$ & $A_{TD}$ & $\\chi^2$ \\\\ \\hline EBV & 9.15 $\\pm$ 0.09 & -2.74 $\\pm$ 0.06 & 1.01 $\\pm$ 0.08 & 1.710 & 9.35 $\\pm$ 0.10 & 0.67 $\\pm$ 0.09 & 22.42 $\\pm$ 1.22 & 1.22 $\\pm$ 0.08 & 0.825 \\\\ KQ85 & 7.61 $\\pm$ 0.08 & -2.78 $\\pm$ 0.07 & 1.14 $\\pm$ 0.08 & 0.667 & 7.78 $\\pm$ 0.09 & 0.53 $\\pm$ 0.08 & 23.34 $\\pm$ 1.33 & 1.29 $\\pm$ 0.08 & 1.208 \\\\ GN & 8.91 $\\pm$ 0.14 & -2.86 $\\pm$ 0.11 & 1.00 $\\pm$ 0.13 & 0.769 & 9.12 $\\pm$ 0.15 & 0.50 $\\pm$ 0.14 & 23.35 $\\pm$ 2.51 & 1.18 $\\pm$ 0.13 & 0.829 \\\\ EN & 8.32 $\\pm$ 0.13 & -2.83 $\\pm$ 0.11 & 0.99 $\\pm$ 0.13 & 0.617 & 8.51 $\\pm$ 0.14 & 0.49 $\\pm$ 0.14 & 23.35 $\\pm$ 2.42 & 1.16 $\\pm$ 0.13 & 0.679 \\\\ GS & 9.73 $\\pm$ 0.13 & -2.64 $\\pm$ 0.08 & 1.02 $\\pm$ 0.13 & 0.699 & 9.93 $\\pm$ 0.13 & 0.85 $\\pm$ 0.13 & 21.99 $\\pm$ 1.41 & 1.28 $\\pm$ 0.12 & 0.114 \\\\ ES & 9.82 $\\pm$ 0.13 & -2.69 $\\pm$ 0.08 & 0.99 $\\pm$ 0.12 & 1.162 & 10.01 $\\pm$ 0.13 & 0.78 $\\pm$ 0.12 & 22.07 $\\pm$ 1.47 & 1.24 $\\pm$ 0.11 & 0.285 \\\\ \\hline 2 & 8.03 $\\pm$ 1.60 & -4.14 $\\pm$ 2.16 & 0.00 $\\pm$ 0.00 & 0.093 & 8.17 $\\pm$ 1.63 & 0.00 $\\pm$ 0.00 & 0.00 $\\pm$ 0.00 & 0.00 $\\pm$ 0.00 & 0.378 \\\\ 3 & 8.73 $\\pm$ 0.99 & -2.60 $\\pm$ 0.66 & 2.03 $\\pm$ 0.95 & 0.016 & 8.94 $\\pm$ 1.04 & 0.88 $\\pm$ 0.97 & 23.97 $\\pm$ 8.94 & 2.20 $\\pm$ 0.91 & 0.001 \\\\ 6 & 7.49 $\\pm$ 2.33 & -3.06 $\\pm$ 2.31 & 0.84 $\\pm$ 2.20 & 0.007 & 7.68 $\\pm$ 2.44 & 0.17 $\\pm$ 2.50 & 25.27 $\\pm$ 100.72 & 0.99 $\\pm$ 2.23 & 0.002 \\\\ 7 & 9.79 $\\pm$ 0.63 & -2.56 $\\pm$ 0.40 & 1.19 $\\pm$ 0.64 & 0.034 & 9.96 $\\pm$ 0.67 & 0.98 $\\pm$ 0.65 & 21.14 $\\pm$ 6.79 & 1.49 $\\pm$ 0.60 & 0.008 \\\\ 8 & 6.70 $\\pm$ 0.63 & -2.73 $\\pm$ 0.62 & 1.41 $\\pm$ 0.62 & 0.019 & 6.86 $\\pm$ 0.66 & 0.56 $\\pm$ 0.68 & 24.81 $\\pm$ 8.69 & 1.48 $\\pm$ 0.61 & 0.018 \\\\ 9 & 13.09 $\\pm$ 0.45 & -2.61 $\\pm$ 0.22 & 1.09 $\\pm$ 0.52 & 0.121 & 13.34 $\\pm$ 0.48 & 1.20 $\\pm$ 0.48 & 21.58 $\\pm$ 3.83 & 1.49 $\\pm$ 0.50 & 0.003 \\\\ 10 & 6.55 $\\pm$ 0.71 & -3.16 $\\pm$ 0.76 & 0.67 $\\pm$ 0.60 & 0.051 & 6.69 $\\pm$ 0.74 & 0.11 $\\pm$ 0.70 & 25.45 $\\pm$ 44.79 & 0.78 $\\pm$ 0.61 & 0.110 \\\\ 11 & 4.20 $\\pm$ 0.63 & -2.45 $\\pm$ 0.84 & 0.81 $\\pm$ 0.64 & 0.061 & 4.21 $\\pm$ 0.86 & 0.44 $\\pm$ 0.59 & 16.51 $\\pm$ 25.45 & 1.02 $\\pm$ 0.59 & 0.006 \\\\ 12 & 5.29 $\\pm$ 0.36 & -3.79 $\\pm$ 0.68 & 0.35 $\\pm$ 0.38 & 0.304 & 5.40 $\\pm$ 0.37 & 0.00 $\\pm$ 0.00 & 0.00 $\\pm$ 0.00 & 0.34 $\\pm$ 0.35 & 1.691 \\\\ 13 & 5.61 $\\pm$ 0.71 & -3.16 $\\pm$ 0.88 & 0.78 $\\pm$ 0.60 & 0.048 & 5.73 $\\pm$ 0.75 & 0.09 $\\pm$ 0.69 & 24.66 $\\pm$ 62.57 & 0.88 $\\pm$ 0.61 & 0.091 \\\\ 14 & 9.70 $\\pm$ 0.45 & -2.53 $\\pm$ 0.28 & 1.20 $\\pm$ 0.43 & 0.033 & 9.87 $\\pm$ 0.48 & 1.03 $\\pm$ 0.46 & 21.51 $\\pm$ 4.49 & 1.50 $\\pm$ 0.40 & 0.000 \\\\ 15 & 8.25 $\\pm$ 0.54 & -2.41 $\\pm$ 0.38 & 1.64 $\\pm$ 0.55 & 0.015 & 8.40 $\\pm$ 0.58 & 1.06 $\\pm$ 0.56 & 21.79 $\\pm$ 5.14 & 1.92 $\\pm$ 0.50 & 0.037 \\\\ 16 & 6.29 $\\pm$ 0.62 & -4.72 $\\pm$ 1.36 & 0.00 $\\pm$ 0.00 & 0.327 & 6.38 $\\pm$ 0.63 & 0.00 $\\pm$ 0.00 & 0.00 $\\pm$ 0.00 & 0.00 $\\pm$ 0.00 & 3.088 \\\\ 17 & 6.59 $\\pm$ 1.79 & -3.90 $\\pm$ 2.82 & 0.17 $\\pm$ 1.62 & 0.055 & 6.73 $\\pm$ 1.83 & 0.00 $\\pm$ 0.00 & 0.00 $\\pm$ 0.00 & 0.15 $\\pm$ 1.52 & 0.148 \\\\ 18 & 6.99 $\\pm$ 0.90 & -3.41 $\\pm$ 1.10 & 0.57 $\\pm$ 0.78 & 0.053 & 7.14 $\\pm$ 0.91 & 0.00 $\\pm$ 0.00 & 0.00 $\\pm$ 0.00 & 0.67 $\\pm$ 0.72 & 0.127 \\\\ 19 & 9.29 $\\pm$ 1.88 & -1.87 $\\pm$ 0.99 & 2.81 $\\pm$ 2.23 & 0.050 & 9.48 $\\pm$ 2.01 & 2.19 $\\pm$ 1.97 & 22.96 $\\pm$ 7.81 & 3.36 $\\pm$ 1.81 & 0.062 \\\\ 20 & 9.85 $\\pm$ 0.72 & -2.90 $\\pm$ 0.52 & 0.87 $\\pm$ 0.71 & 0.070 & 10.09 $\\pm$ 0.76 & 0.45 $\\pm$ 0.76 & 23.50 $\\pm$ 14.84 & 1.08 $\\pm$ 0.71 & 0.024 \\\\ 21 & 6.92 $\\pm$ 0.45 & -2.98 $\\pm$ 0.47 & 0.96 $\\pm$ 0.40 & 0.049 & 7.08 $\\pm$ 0.47 & 0.28 $\\pm$ 0.47 & 24.34 $\\pm$ 12.08 & 1.08 $\\pm$ 0.41 & 0.124 \\\\ 22 & 8.73 $\\pm$ 0.81 & -2.40 $\\pm$ 0.54 & 1.69 $\\pm$ 0.87 & 0.019 & 8.87 $\\pm$ 0.87 & 1.08 $\\pm$ 0.85 & 20.69 $\\pm$ 8.30 & 2.03 $\\pm$ 0.80 & 0.023 \\\\ 23 & 8.96 $\\pm$ 0.36 & -2.54 $\\pm$ 0.22 & 1.12 $\\pm$ 0.32 & 0.079 & 9.12 $\\pm$ 0.38 & 0.95 $\\pm$ 0.30 & 21.38 $\\pm$ 3.44 & 1.40 $\\pm$ 0.30 & 0.005 \\\\ 24 & 7.75 $\\pm$ 0.54 & -3.56 $\\pm$ 0.63 & 0.04 $\\pm$ 0.49 & 0.207 & 7.92 $\\pm$ 0.55 & 0.00 $\\pm$ 0.00 & 0.00 $\\pm$ 0.00 & 0.10 $\\pm$ 0.45 & 0.867 \\\\ 28 & 8.46 $\\pm$ 1.80 & -2.03 $\\pm$ 1.09 & 2.47 $\\pm$ 2.01 & 0.038 & 8.64 $\\pm$ 1.91 & 1.73 $\\pm$ 1.89 & 23.34 $\\pm$ 9.27 & 2.83 $\\pm$ 1.71 & 0.031 \\\\ 30 & 9.73 $\\pm$ 3.14 & -2.00 $\\pm$ 1.62 & 2.30 $\\pm$ 3.47 & 0.006 & 9.85 $\\pm$ 3.36 & 2.09 $\\pm$ 3.18 & 21.92 $\\pm$ 14.67 & 2.81 $\\pm$ 2.91 & 0.002 \\\\ 31 & 12.00 $\\pm$ 3.23 & -1.93 $\\pm$ 1.32 & 2.70 $\\pm$ 3.66 & 0.005 & 12.13 $\\pm$ 3.46 & 2.67 $\\pm$ 3.27 & 21.47 $\\pm$ 12.17 & 3.45 $\\pm$ 3.00 & 0.011 \\\\ 32 & 8.15 $\\pm$ 0.81 & -2.87 $\\pm$ 0.70 & 0.85 $\\pm$ 0.81 & 0.003 & 8.33 $\\pm$ 0.85 & 0.53 $\\pm$ 0.87 & 24.26 $\\pm$ 12.86 & 0.95 $\\pm$ 0.81 & 0.098 \\\\ 33 & 8.53 $\\pm$ 1.88 & -3.08 $\\pm$ 1.60 & 0.78 $\\pm$ 1.70 & 0.025 & 8.74 $\\pm$ 1.99 & 0.12 $\\pm$ 1.91 & 23.77 $\\pm$ 135.62 & 0.99 $\\pm$ 1.72 & 0.004 \\\\ \\hline \\end{tabular} \\end{center} \\caption{Model fits to the dust coefficients determined between the 5 WMAP frequencies and the FDS8 template for large sky areas (upper part of table) and for those regions that indicate a 3$\\sigma$ significant amplitude at K-band (lower part). The models DI and DII are fully defined in Section~\\ref{sec:discussion_dust}. $A_{PLD}$ and $\\beta_{PLD}$ represent the normalisation amplitude at K-band and spectral index respectively of a power-law anomalous dust emission component. $A_{WNM}$ is the amplitude of the WNM spinning dust model at K-band, $A_{CNM}$ is the amplitude of the CNM spinning dust model normalised at 41~GHz, and $\\Delta \\nu_{CNM}$ is the shift in frequency of the peak of the CNM dust model to better the fit the data. $A_{TD}$ is the amplitude of the thermal dust emission with an assumed spectral index $\\beta_{TD} = 1.55$ as determined directly from the FDS8 dust model. The key for the global fits as for Table~\\ref{tab:simple_synch_spectra}.} \\label{tab:dust_model_hemispheres} \\end{table} \\begin{center} \\begin{figure} \\begin{tabular}{ccc} \\epsfig{file=figures/fig_71.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_72.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_73.ps,width=0.3\\linewidth,angle=0,clip=} \\\\ \\epsfig{file=figures/fig_74.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_75.ps,width=0.3\\linewidth,angle=0,clip=} & \\epsfig{file=figures/fig_76.ps,width=0.3\\linewidth,angle=0,clip=} \\\\ \\end{tabular} \\caption{Dust spectral fits for various regions of the sky. A comparison is made of the observed template fit amplitudes derived from the 5 \\emph{WMAP} frequency bands and the FDS8 template with the 2 models DI and DII as defined in Section~\\ref{sec:discussion_dust}. Also shown in the plot are the separate power-law anomalous dust (PLD) and vibrational (thermal) dust (VD) that constitute model DI, and the cold neutral medium (CNM) and warm neutral medium (WNM) components that, together with the vibrational dust, constitute model DII. The observed spectral dependencies of regions 12, 14 and 21 are representative of those seen for all regions with a significant detection at K-band.\\label{fig:dust_model_hemispheres}} \\end{figure} \\end{center} The results are summarised in Table~\\ref{tab:dust_model_hemispheres}. Fig.~\\ref{fig:dust_model_hemispheres} presents a comparison of the model fits for three of the global masks, plus three of the regions that can be considered representative of the general results seen. It should be apparent from the table that both models DI and DII provide adequate fits to the global masks. However, in order to explain the emission at V-band in terms of spinning dust models, two contributions are required in combination with an enhanced thermal dust normalisation some 20\\% larger than that predicted at 94\\,GHz by FDS8. Such models also require a shift in the peak CNM emission of $\\sim$22\\,GHz and a ratio of WNM to CNM spinning dust emission of order 4:1. In fact, if the CNM emission were to be omitted, or the thermal dust amplitude kept at the canonical FDS8 value, then the V- and/or W-band amplitudes would be substantially underpredicted. In practise, given the low number of data points sensitive to the thermal dust emission, there is an effective degeneracy between the CNM and vibrational dust components, that may be alleviated somewhat if the spinning dust models can be enhanced in amplitude at frequencies higher than their peak. Indeed, spinning dust can contribute significantly to emission in the $\\sim$60 to 90 GHz range depending on the local physical conditions. However, one might then expect to see a much flatter spectrum around the Q-, V-, and W-bands than is actually observed. It may also be that some modification of the assumed thermal dust spectral index is required. The fits to individual regions are less informative, as expected given the lack of clear detections for the template fit coefficients at V- and W-band. Given this, the analysis might be criticised for overfitting the data points. Nevertheless, both a power-law and a WNM spinning dust component are good candidates to explain the lower frequency points. In almost all cases, a significant contribution from CNM or thermal dust emission is not required. This discrepancy between the regional fits and global analysis may simply reflect a signal-to-noise issue at higher frequencies than the former, or the fact that the latter are essentially averages over many regions which exhibit some variation in dust properties. Regions 12 and 16 are interesting in that the power-law model is clearly preferred over model DII, although the analysis is really only constrained by the data points at K- and Ka-band. For those regions in which a significant vibrational dust contribution is required, the amplitude is systematically higher than the canonical FDS8 value, although still consistent with it. However, the required DI enhancement is lower, thus the DII results more likely again point to a deficiency in the specific spinning dust models used in this analysis. The FDS8 model value for the dust spectral index used here is quite flat, whereas early \\emph{Planck} papers \\citep{Planck_ERXXIV_ISM:2011,Planck_ERXXV_Molecular:2011} studying emission at frequencies higher than $\\sim$100~GHz favour values closer to 1.8. However, \\citet{Planck_ERXIX_DarkGas:2011} suggests that the dust SED flattens in the millimetre wavelength range. At lower frequencies, there is considerable uncertainty, and detailed modelling for candidate AME regions (including Perseus and $\\rho$-Ophiucus) of the foreground spectra for all components and across the \\emph{Planck} frequencies suggests values in the range 1.5--2 \\citep{Planck_ERXX_AME:2011}. We have repeated the calculations above but imposing a thermal dust spectral index of 1.7 on the fits. For model DI, there is a general increase in amplitude of the power-law dust component, with an associated flattening of the spectral index at levels below the 1-$\\sigma$ error bar, and no impact on the thermal dust amplitude. Model DII shows similarly modest increases in the amplitudes of the two spinning dust components, with little change in the CNM peak frequency shift, and again no change in the thermal dust amplitude. However, adopting the FDS8 thermal dust model at W-band and then imposing a steeper index to lower frequencies is not self-consistent. Therefore, a more detailed treatment is required including specific modelling of all dust components combined with higher frequency measurements. An obvious potential error in our analysis is the selective use of the CNM and WNM spectra for typical conditions in those phases of the ISM. What is clear is that the frequency range of 60--100\\,GHz is a remarkably interesting regime for dust astrophysics." }, "1112/1112.2817_arXiv.txt": { "abstract": "Detection and characterization of exoplanets by direct imaging requires a coronagraph designed to deliver high contrast at small angular separation. To achieve this, an accurate control of low order aberrations, such as pointing and focus errors, is essential to optimize coronagraphic rejection and avoid the possible confusion between exoplanet light and coronagraphic leaks in the science image. Simulations and laboratory prototyping have shown that a Coronagraphic Low Order Wave-Front Sensor (CLOWFS), using a single defocused image of a reflective focal plane ring, can be used to control tip-tilt to an accuracy of $10^{-3}$ $\\lambda{/D}$. This paper demonstrates that the data acquired by CLOWFS can also be used in post-processing to calibrate residual coronagraphic leaks from the science image. Using both the CLOWFS camera and the science camera in the system, we quantify the accuracy of the method and its ability to successfully remove light due to low order errors from the science image. We also report the implementation and performance of the CLOWFS on the Subaru Coronagraphic Extreme AO (SCExAO) system and its expected on-sky performance. In the laboratory, with a level of disturbance similar to what is encountered in a post Adaptive Optics beam, CLOWFS post-processing has achieved speckle calibration to 1/300 of the raw speckle level. This is about 40 times better than could be done with an idealized PSF subtraction that does not rely on CLOWFS. ", "introduction": "Since the discovery of 51 Pegasi by \\citet{1995Natur.378..355M}, several hundreds extrasolar planets have been uncovered, for the most part via indirect detection methods such as radial velocity measurements of the reflex motion of their host star and photometric transit. Direct imaging was recently able to produce the first high contrast images of \\emph{solid} planetary candidates orbiting nearby main sequence stars: Fomalhaut \\citep{2008Sci...322.1345K}, Beta Pictoris \\citep{2009A&A...493L..21L} and HR 8799 \\citep{2008Sci...322.1348M, 2009ApJ...694L.148L}. Following up on these direct detections, \\citet{2010ApJ...710L..35J} obtained the first direct extrasolar planet spectrum, opening the way to better characterization of planetary atmospheres. These direct detections are currently limited to planets at large orbital separations of several dozens of AU, and only probe a small fraction of the distribution of known extrasolar planets, for which the median orbital separation is likely closer to 1 AU. At near infrared wavelength, an 8-meter class telescope provides sufficient angular resolution (40 milli-arcseconds at $\\lambda$=$\\sim$1.6 ${\\mu}$m) to be able to detect companions in the Habitable Zone of nearby stars (d $<$ 30 pc). High contrast imaging near the diffraction limit however requires very good control and calibration of the wavefront aberrations in the optical system. From the ground, this task is complicated by the rapidly changing atmospheric wavefront creating speckled images. While Adaptive Optics (AO) offers a major improvement, performance is still limited and direct imaging of planets often requires post-processing techniques such as Angular Differential Imaging (ADI) \\citep{2006ApJ...641..556M}, which is most efficient at large ($\\gtrapprox$ 10 $\\lambda/D$) angular separations. In the near future, high contrast imaging systems employing new coronagraphic and wavefront control techniques will greatly improve our ability to image exoplanets. These Extreme-AO systems include the Gemini Planet Imager (GPI) \\citep{2006SPIE.6272E..18M}, ESO's SPHERE \\citep{2008SPIE.7014E..41B} and Subaru's SCExAO \\citep{2009SPIE.7440E..20M}. They will use efficient coronagraphs and high-speed high-order wavefront corrections to reduce speckles in the coronagraphic image. The images of the HR 8799 planetary system obtained by \\citet{2010AAS...21537706S} on a well-corrected 1.5 meter aperture using a high-performance coronagraph demonstrate the relevance of this approach. Imaging companions close to the edge of the occulting mask in a coronagraph (that is $\\sim$40 mas on SCExAO) however remains an unachieved feat. The current state of the art for PSF calibration is an optimized version of ADI called LOCI introduced by \\citet{2007ApJ...660..770L}. LOCI can calibrate out high-order aberrations to a very high level of contrast (12 magnitudes and higher), but like any technique relying on ADI, only for angular separations greater than 10 $\\lambda/D$. Detection at the edge of a $\\sim$ 1 $\\lambda/D$ occulter is extremely sensitive to low-order aberrations such as pointing and focus. Of these aberrations, pointing is especially critical, since near the occulter, a tip-tilt excursion along a given direction mimics the signal of a true companion in a coronagraphic image. This issue, first identified for high contrast space borne coronagraphs, has been addressed by \\cite{2009ApJ...693...75G}, with the Coronagraphic Low-Order Wavefront Sensor (CLOWFS), a scheme using the light occulted by a modified focal plane mask as an accurate pointing tracker. The idea of using the light otherwise lost in the coronagraphic focal plane for tracking is not a novelty. It was for instance successfully implemented on the LYOT project \\citep{2006ApJ...650..484D}. The calibration unit of GPI also uses the light occulted by the focal plane mask to measure low order aberrations, after re-imaging the pupil in a Shack Hartmann type wavefront sensor \\citep{2010SPIE.7736E.179W}: pointing performance with this scheme reaches 2 mas for typical expected observing conditions. Maximum sensitivity to pointing errors is reached when the light from opposite edges of the pupil is allowed to interfere, which naturally happens in the focal plane. In this respect, while robust to a wide range of errors, a Shack-Hartman doesn't appear optimal to measure pointing: because it splits the pupil into sub-pupils, this capability is lost, resulting in lesser performance than a focal-plane based wavefront sensing scheme \\citep{2005ApJ...629..592G}. The originality of the CLOWFS design resides in its dual-zone focal plane mask, designed to suppress a strong offset to the signal, carrying most of the power but no information, in a manner reminiscent of strioscopy. The suppression of this offset turns the otherwise imperceptible changes due to small pointing errors into a macroscopic change of the CLOWFS image. Using this scheme, \\citet{2009ApJ...693...75G} were able to stabilize tip-tilt at the level of $10^{-3} \\lambda/D$ in a closed-loop system for $\\lambda=633$ nm, in a laboratory experiment. The current implementation of CLOWFS on SCExAO exhibits pointing residuals $<$0.2 mas, with a 50 Hz frame rate. While this level of performance is quite remarkable, we demonstrate in this work that additional calibration can be achieved in post-processing, and lead to an improved subtraction of coronagraphic leaks due to low-order aberrations in a long exposure. Using the SCExAO system as a testbed, we experimentally demonstrate a 40 times improvement of the detection limit over a classical calibration procedure at angular separation of a few $\\lambda/D$. This paper is organized as follows : in Section~\\ref{sec:method}, we introduce how to use CLOWFS for post processing of coronagraphic images. In Section~\\ref{sec:exp}, we describe the implementation of the concept on the SCExAO experiment of the Subaru telescope, and we present our results in Section~\\ref{Sec:results}. In Section~\\ref{sec:discussion}, we summarise our results and discuss possible updates to the presented CLOWFS configuration. \\begin{figure*}[htb!] \\centerline{\\includegraphics[scale=0.4]{f1.eps}} \\caption{Optical layout of SCExAO used in Lyot-coronagraph mode. A mask lit by an IR ($\\lambda=1.55$ ${\\mu}m$) laser diode emulates the Subaru Telescope pupil. The beam is focused on the dual-zone focal-plane occulting mask described in the text. The light reflected by the focal-plane mask feeds the (slightly defocused) CLOWFS detector, used to characterize pointing errors. The light that is not intercepted by the mask is then re-imaged on the \"science\" detector. Examples of images obtained on both cameras are presented in Fig.~\\ref{fig:clowfs_img} and \\ref{fig:sci_img}.} \\label{fig:bench} \\end{figure*} ", "conclusions": "\\label{sec:discussion} The results presented in Sec~\\ref{Sec:results} demonstrate that, used in post-processing, CLOWFS improves the sensitivity of a coronagraph by a factor 40, in comparison to more conventional -yet idealized- type of calibration. The method described here for this proof of concept is yet still quite rudimentary, and could benefit from several improvements. Instead of relying on an exhaustive search by a MMA-like algorithm on a massive database, necessarily containing redundant information, the dictionary could be simplified using approaches like Principal Component Analysis, in which CLOWFS images would be projected onto a smaller sub-set of CLOWFS modes. In addition, an extended knowledge of several experimental parameters, such as the telescope elevation, star color or magnitude, could be implemented in the dictionary so as to improve on the actual calibration. The approach is nevertheless powerful as it requires neither modelling of the coronagraph, nor assumptions such as linearity of the response in order to work. The absence of such requirements makes it extremely robust, and able to handle multiple situations and types of coronagraphic leaks. The Subaru Coronagraphic Extreme AO Project implements a CLOWFS, used in close-loop. The system architecture however allows for a simple implementation of the CLOWFS post-processing, with no impact on the hardware, to calibrate residual low-order modes induced coronagraphic leaks. CLOWFS on SCExAO currently generates a steady 4 Mbits/s data stream. While not prohibitively large for an experiment, the storage of entire nights of CLOWFS data will quickly prove impractical as SCExAO enters a more aggressive observing phase. Eventually, the calibration of coronagraphic leaks using CLOWFS will require the matching to be performed on-the-fly, so that only the final coronagraphic leak image is actually saved. Because it makes no assumption on targets and the current state of the optical bench, CLOWFS used both for live and post-processing is likely to play a significant part to the success of SCExAO, focused on the detection of planetary companions at small angular separation." }, "1112/1112.5430_arXiv.txt": { "abstract": "For a narrow band of values of the top quark and Higgs boson masses, the Standard Model Higgs potential develops a false minimum at energies of about $10^{16}$ GeV, where primordial Inflation could have started in a cold metastable state. A graceful exit to a radiation-dominated era is provided, {\\it e.g.}, by scalar-tensor gravity models. We pointed out that if Inflation happened in this false minimum, the Higgs boson mass has to be in the range $126.0 \\pm 3.5$ GeV, where ATLAS and CMS subsequently reported excesses of events. Here we show that for these values of the Higgs boson mass, the inflationary gravitational wave background has be discovered with a tensor-to-scalar ratio at hand of future experiments. We suggest that combining cosmological observations with measurements of the top quark and Higgs boson masses represents a further test of the hypothesis that the Standard Model false minimum was the source of Inflation in the Universe. ", "introduction": " ", "conclusions": "" }, "1112/1112.1025_arXiv.txt": { "abstract": "We present $K$-band integral field observations of the circumstellar envelope of the evolved star \\ohtto. Spatial and spectral information were simultaneously acquired using the {\\sc sinfoni} integral field unit, with adaptive optics, on the Very Large Telescope. The observations reveal the discovery of H$_2$ emission (1) around the centre of the nebula and (2) located in clumps along the Western side of the Northern lobe, presumably associated with the strong shocks that stimulate the previously reported \\halpha~emission at the same location. An observed H$_2$ \\rto~line ratio of ~8.3$\\pm$1.9 was calculated for the central field, a value consistent with shock excitation. ", "introduction": "\\label{sec:Intro} OH~231.8+4.2 (hereafter OH231) is an O-rich late spectral type (M) central star (Mira variable, QX Pup) with bipolar high-velocity dust and gas outflows~\\citep{2001A&A...373..932A}, filamentary structures observed in scattered and molecular line emission, and large angular size (10$^{\\prime\\prime}$$\\times$60$^{\\prime\\prime}$). Often labeled as a post-AGB object or pre-planetary nebula, the presence of both a Mira central star and a main-sequence companion of spectral type A \\citep{2004ApJ...616..519S} seems to contradict this classification. OH231 is more likely a D-type bipolar symbiotic system \\citep[][]{2010PASA...27..129F}. However in some cases, morphological similarities do exist between post-AGB and symbiotic objects, most strikingly the presence of highly collimated bipolar nebulae. It is via fast-collimated outflows that these stars shape their surrounding nebula. Understanding the development and origin of these fast outflows is critical for advancing hydrodynamical models of wind interaction. Recent work by \\citet{2009ApJ...696.1630L} attempting to reproduce the high velocity molecular emission in AFGL~618 using collimated fast wind models, emphasises the need for further observations and model development in this area. OH231 has been the subject of many studies spanning multiple wavelength ranges, for example: \\citet{1985ApJ...297..702C} were first to propose the existence of a binary companion; \\citet{2002A&A...389..271B} imaged the shape of the shocks using H${\\alpha}$ (reproduced in Fig.~\\ref{fig:oh231_naco_sin_wfpc2} [a]) detected with the {\\em Hubble Space Telescope (HST)}; \\citet{2003ApJ...585..482M} report {\\em HST}/NICMOS NIR images of the dust distribution and hence a high resolution map of the extinction through the nebula. \\citet{2006ApJ...646L.123M} using the MIDI and NACO instruments on the Very Large Telescope (VLT) detected a compact circumstellar disc. The envelope of OH231 is also known to be rich in molecular species (e.g. H$_2$O, OH, and SiO) however previous studies in the NIR have all returned null detections of H$_2$ \\citep[e.g.][]{1998ApJ...509..728W,2006ApJ...646L.123M}. In this Letter, we present the results of preliminary observations of OH231 at NIR ($K$-band) wavelengths showing for the first time the presence of shock-excited H$_2$. Throughout this work we assume OH231 is a member of the open cluster M46 at a distance of 1.3~kpc~\\citep{1985ApJ...292..487J}. The origin of the coordinate system used in Figures~1,3, and 4 is given by the location of the SiO maser emission at RA=07$^h$42$^m$16$^s$.93, Dec=-14$^\\circ$42$^{\\prime}$50$^{\\prime\\prime}$.2 (J2000)~\\citep{2002A&A...385L...1S}, and the inclination angle of the bipolar axis is 36$^\\circ$~to the plane of the sky \\citep{1992ApJ...398..552K}. ", "conclusions": "\\label{sec:discuss} We have presented VLT/{\\sc sinfoni} integral field observations of OH231, revealing the presence of several ro-vibrational H$_2$ lines. The main conclusions are: \\begin{list}{\\makelabel{-}}{\\itemsep=-0.2em \\leftmargin=1em} \\item The discovery of H$_2$ emission near the centre of OH231, possibly located at the edge of an axisymmetric shell or an incomplete disc. \\item A \\rto~value of 8.3$\\pm$1.9 is found for the equatorial H$_2$, suggesting a collisional excitation mechanism. \\item Our observations of the central shell/disc of H$_2$ show no velocity structure. However, higher S/N and/or velocity resolution data are needed to accurately probe the kinematics in this region. \\item We detect fast-moving H$_2$ (\\simm 110 \\kms, along the bipolar axis) via the \\ozso transition along the North-Western tip of the nebula, a region where a strong \\halpha~bow shock exists. Due to the small FOV of our observations, the full extent of this H$_2$ is unknown. \\end{list}" }, "1112/1112.3020_arXiv.txt": { "abstract": "Recent measurements of cosmic-ray spectra of several individual nuclear species by the CREAM, TRACER, and ATIC experiments indicate a change in the spectral index of the power laws at TeV energies. Possible explanations among others include non linear diffusive shock acceleration of cosmic-rays, different cosmic-ray propagation properties at higher and lower energies in the Galaxy and the presence of nearby sources. In this paper, we show that if supernova remnants are the main sources of cosmic rays in our Galaxy, the effect of the nearby remnants can be responsible for the observed spectral changes. Using a rigidity dependent escape of cosmic-rays from the supernova remnants, we explain the apparent observed property that the hardening of the helium spectrum occurs at relatively lower energies as compared to the protons and also that the spectral hardening does not persist beyond $\\sim (20-30)$ TeV energies. ", "introduction": "Recently, cosmic-ray (CR) measurements by the new-generation balloon-borne experiments such as the ATIC (Panov et al. 2007), CREAM (Yoon et al. 2011), and TRACER (Ave et al. 2008) seem to indicate that the CR spectrum deviates from a single power law. The spectra of all individual elements seem to be harder at TeV energies than at lower energies. Such a hardening is not easy to explain under the standard models of CR acceleration and their propagation in the Galaxy. Under the standard theory, CRs below the knee ($\\sim 3$ PeV) are considered to be produced by supernova remnant (SNR) shock waves by diffusive shock acceleration mechanism (Bell 1978, Blandford $\\&$ Eichler 1987). Such a mechanism naturally predicts a power law spectrum of $E^{-\\gamma}$ with the index $\\gamma=2$ for strong shocks. On the other hand, CR propagation in the Galaxy is considered to be of diffusive nature which is due to scattering by magnetic field irregularities and the CR self excited Alfven and hydromagnetic waves present in the Galaxy. Measurements of CR secondary-to-primary ratios indicate that the diffusion is energy dependent with the diffusion coefficient $D(E)\\propto E^a$ with $a\\approx (0.3-0.7)$. Under these considerations, the CR spectrum in the Galaxy is expected to follow a single power law with index $(\\gamma+a)$ up to the knee, which do not seem to agree quite easily with the observed hardening at TeV energies. The observed data can be explained if either the source spectrum or the diffusion index flattens at higher energies. Non linear diffusive shock acceleration theories where CRs modify the shock structure predict concave spectra (flatter at higher energies) at the shocks. But, the total spectrum injected into the interstellar medium which is the sum of the instantaneous spectra over the SNR life time is very close to a pure power-law (Caprioli et al. 2010). The concave signature can be even more diluted when summed over an ensemble of SNRs (Ptuskin $\\&$ Zirakashvili 2005). From the propagation point of view, there are models which assume a harder or constant CR diffusion coefficient at higher energies in the Galaxy (Ave et al. 2009). Such models are motivated not only by the apparent flattening of the observed boron to carbon ratio above $\\sim 100$ GeV energies, but also by the observed CR anisotropy which is almost independent of energy. Recently, it has also been proposed that dispersion in the spectral indices of CR source spectrum from many sources can also be responsible for the observed spectral hardening (Yuan et al. 2011). Another possible explanation, as also pointed out in Ahn et al. 2010, is the presence of nearby sources. Erlykin $\\&$ Wolfendale 2011 suggested that an extra component of CRs with a steep spectrum could be contributing below $\\sim 200$ GeV/n while above that, the spectrum is entirely determined by a harder CR background. They proposed that the sources of the extra component could be in OB associations in the Local Bubble. Recently, Ohira $\\&$ Ioka 2011 proposed that the hardening could be due to decreasing Mach number in hot superbubbles with multiple supernovae. In another recent work, Vladimirov et al. 2011 investigated several possible interpretations (including local source effect) for the observed spectral features at low and high energies using the GALPROP propagation code. They also presented the possible effects on other observed properties such as CR anisotropy, isotopic ratios and the Galactic diffuse $\\gamma$-ray emissions. In our present study, we investigate whether the spectral hardening observed at TeV energies could be an effect of the nearby SNRs. Although there has not been any direct detection of CRs from any sources, SNRs remain the most favorable candidates both theoretically as well as observationally. At least the presence of high energy particles up to few TeVs inside SNRs have been confirmed by the detections of non-thermal X-rays (Parizot et al. 2006) and TeV $\\gamma$-rays from several SNRs (Aharonian et al. 2006, 2008a). Moreover, the detection of TeV electrons by the HESS experiment (Aharonian et al. 2008b) indicates the presence of one or more CR sources within a distance of $\\sim 1$ kpc from us. If these sources produce both, electrons as well as nuclei, we expect to see some effects on the spectra of CR nuclei observed at the Earth. \\section {Model} The diffusive propagation of CRs in the Galaxy neglecting the effects due to nuclear spallation can be described by the following equation, \\begin{equation} \\nabla\\cdot(D\\nabla N)+Q=\\frac{\\partial N}{\\partial t} \\end{equation} where $N(\\textbf{r},E,t)$ is the differential number density at a distance $\\textbf{r}$ from the source at time $t$, $E$ is the kinetic energy/nucleon and $Q(\\textbf{r},E,t)$ is the source term. The diffusion coefficient is taken as $D(\\Re)=D_0(\\Re/\\Re_0)^a$ for $\\Re>\\Re_0$, where $\\Re$ denotes the particle rigidity which is given by $\\Re=AE/Z$ for charge $Z$ and mass number A. For our study, we consider two sets of values for $(D_0, \\Re_0, a)$: one based on purely diffusion model (hereafter Model A) and the other based on models including CR reacceleration due to interstellar turbulence (hereafter Model B). We choose $(D_0, \\Re_0, a)=(2.9, 3, 0.6)$ for Model A (Thoudam 2008) and $(5.8, 4, 0.33)$ for Model B (Strong et al. 2010), where $D_0$ is in units of $10^{28}$ cm$^2$ s$^{-1}$ and $\\Re_0$ is in GV. Under diffusive shock acceleration theory, CRs are confined within the remnant due to the magnetic turbulence generated by the CRs themselves. They can escape when their upstream diffusion length defined as $l_{diff} = D_s(E)/u_s$ is greater than the escape length from the shock front which is usually taken as $l_{esc}\\approx 0.1 R_s$, where $u_s$ and $R_s$ denote the shock velocity and the shock radius respectively. In the Bohm diffusion limit, the upstream diffusion coefficient scales linearly with energy as $D_s(E)\\propto E$ which implies that higher energy particles can escape the remnant at early times followed later by the lower energy ones. But, the exact energy dependence of $D_s$ is still not well understood and depends on some poorly known quantities which include the spectral distribution of the CR self-excited turbulence waves, the level of magnetic field amplification by the CRs and the dynamical reaction of CRs on the shock structure. Therefore, we follow a simple but reasonable parameterization for the CR escape time similar to that adopted by Gabici et al. 2009 as given below, \\begin{equation} t_{esc}(\\Re)=t_{sed}\\left(\\frac{\\Re}{\\Re_{max}}\\right)^{-1/\\alpha} \\end{equation} where $t_{sed}$ denotes the start of the Sedov phase, $\\Re_{max}$ denotes the maximum CR rigidity and $\\alpha$ is a positive constant. We assume that the maximum CR energy accelerate by an SNR scales with the charge number $Z$ as $ZU_{max}$, where $U_{max}$ denotes the maximum kinetic energy of the protons which is taken as $1$ PeV for our study (Berezhko 1996). This scaling gives $\\Re_{max}=1$ PV. In units of energy/nucleon, the maximum energy for helium is $E_{max}=0.5$ PeV/n. Eq. (2) assumes that the highest energy CRs of all the species start escaping at the onset of the Sedov phase itself. Writing Eq. (2) in terms of total kinetic energy, it is easy to check that for the same kinetic energy, the escape time of CRs scales with the charge number as $Z^{1/\\alpha}$, i.e, higher charged particles escape at relatively later stages of the SNR evolution. Thus, our escape model takes into account the general understanding of diffusive shock acceleration theory that higher charged particles can be confined for relatively longer duration within the remnant. In terms of energy/nucleon, we can write Eq. (2) as \\begin{equation} t_{esc}(E)=t_{sed}\\left(\\frac{AE}{Z\\Re_{max}}\\right)^{-1/\\alpha} \\end{equation} Eq. (3) shows that for the same energy/nucleon, all nuclei with charge $Z>1$ escape earlier than the protons by a factor of $(A/Z)^{-1/\\alpha}$. We further assume that no particles remain confined after the shock completely dies out which we assume to occur when the SNR age $10^5$ yr. Taking this into account, the CR escape time for our study is taken as $T_{esc}(E)= \\mathrm{min}\\left[t_{esc}(E), 10^5 \\mathrm{yr}\\right]$. For detailed studies on particle escape from SNRs, see e.g., Ptuskin $\\&$ Zirakashvili 2005, Caprioli et al. 2009 and Ohira et al. 2010. The corresponding escape radius of CRs is calculated using the age-radius Sedov relation for SNRs as given below, \\begin{equation} R_{esc}(E)=2.5u_0\\;t_{sed}\\left[\\left(\\frac{T_{esc}}{t_{sed}}\\right)^{0.4}-0.6\\right] \\end{equation} where $u_0$ represents the initial shock velocity, i.e the velocity at $t=t_{sed}$. The source term in Eq. (1) is taken as, \\begin{equation} Q(\\textbf{r},E,t)=\\frac{q(E)}{A_{esc}}\\delta(t-T_{esc})\\delta(r-R_{esc}) \\end{equation} where $A_{esc}=4\\pi R_{esc}^2$ denote the surface area of the SNR at the time when CRs of energy $E$ escape the remnant. It should be noted that our consideration of the rigidity dependent escape time and the finite source size are different from the commonly adopted burst-like point source approximation where CRs of all rigidities are assumed to escape at the same time from a point source. For CR study near the sources, the point source approximation can break down and it looks more realistic to take their sizes into account (Thoudam $\\&$ H\\\"orandel 2011). Recently, such importance has also been highlighted in Ohira et al. 2011 in the study of $\\gamma$-ray emission from SNRs interacting with molecular clouds. The source spectrum in Eq. (5) is taken as $q(E)=Aq(U)$ with $q(U)$ given by, \\begin{equation} q(U)=k(U^2+2Um)^{-(\\gamma+1)/2}(U+m) \\end{equation} where $U=AE$ represents the particle total kinetic energy, $m$ is the rest mass energy and $k$ is the normalization constant which is related to the CR injection efficiency. Solving Eq. (1), the spectrum at a distance $r_s$ from the SNR follows, \\begin{eqnarray} N(r_s,E,t)=\\frac{q(E)\\,R_{esc}}{r_sA_{esc}\\sqrt{\\pi D(t-T_{esc})}}\\mathrm{exp}\\left[-\\frac{\\left(R_{esc}^2+r_s^2\\right)}{4D(t-T_{esc})}\\right]\\nonumber\\\\ \\times\\;\\mathrm{sinh}\\left(\\frac{r_sR_{esc}}{2D(t-T_{esc})}\\right)\\; \\end{eqnarray} For high energy particles for which the diffusion radius defined as $r_{diff}=\\sqrt{D(t-T_{esc})}$ is much larger than $(r_s, R_{esc})$, Eq. (7) follows a power-law of the form $N(E)\\propto E^{-\\left(\\Gamma+\\frac{3}{2}a\\right)}$. Eq. (7) can be used to calculate the CR spectra from the nearby SNRs. We choose proton and helium for our study and consider only those SNRs with distances $<1$ kpc from the Earth and ages $< 2\\times 10^5$ yr. From the available literature, we found 10 SNRs listed as follows with their distances (kpc) and ages (yr) respectively given in parentheses: Cygnus Loop $(0.54, 10^4)$, HB21 $(0.8, 1.9\\times 10^4)$, HB9 $(0.8, 6.6\\times 10^3)$, S147 $(0.8, 4.6\\times 10^3)$, Vela $(0.3, 1.1\\times 10^4)$, G299.2-2.9 $(0.5, 5\\times 10^3)$, SN185 $(0.95, 1.8\\times 10^3)$, Monogem $(0.3, 1.1\\times 10^5)$, G114.3+0.3 $(0.7, 4.1\\times 10^4)$ and Vela Junior $(0.75, 3.5\\times 10^3)$. In addition to the contributions from the nearby SNRs, we assume that there exists a steady CR background in the Galaxy which dominates the overall CR spectrum. For the CRs observed at the Earth, we assume that this background component consists of contributions from distant SNRs plus any other possible sources in the Galaxy. For our study, we obtain the background by fitting the observed CR spectrum between $(20-200)$ GeV/n. This is the energy region where the contamination from the nearby sources is expected to be less and at the same time, not much affected by the Solar modulation. In fact, it has been shown in Thoudam 2008 that the presence of nearby sources can produce stronger density fluctuations at higher energies than at lower energies because of the energy dependent nature of CR diffusion. Therefore, we believe that it is reasonable to assume that the low energy CRs that we observe at the Earth are not much affected by the presence of nearby SNRs and they largely represent the averaged background spectrum in the Galaxy. We will show in the following that this is indeed the most likely case. ", "conclusions": "" }, "1112/1112.3957_arXiv.txt": { "abstract": "Approximately $14$\\% of known Galactic open clusters possess absolute errors $\\le 20$\\% as evaluated from $n\\ge3$ independent distance estimates, and the statistics for age estimates are markedly worse. That impedes such diverse efforts as calibrating standard candles and constraining masses for substellar companions. New data from the VVV survey may be employed to establish precise cluster distances with comparatively reduced uncertainties ($\\le10$\\%). This is illustrated by deriving parameters for Pismis~19 and NGC~4349, two pertinent open clusters which hitherto feature sizable uncertainties ($60$\\%). Fundamental parameters determined for Pismis~19 from new VVV $JHK_s$ photometry are $d=2.40\\pm0.15$ kpc, $<{\\it E_{J-H}}>=0.34\\pm0.04$, and $\\log{\\tau}=9.05\\pm0.10$, whereas for NGC~4349 the analysis yielded $d=1.63\\pm0.13$ kpc, ${\\it E_{J-H}}=0.09\\pm0.02$, $\\log{\\tau}=8.55\\pm0.10$. The results exhibit a significant ($\\ge 5 \\times$) reduction in uncertainties, and indicate that: $i$) existing parameters for the substellar object NGC~4349 127b require revision, in part because the new cluster parameters imply that the host is $20$\\% less-massive (${\\cal M_{*}/M_{\\sun}} \\sim 3.1$); $ii$) R Cru is not a member of NGC~4349 and should be excluded from period-Wesenheit calibrations that anchor the distance scale; $iii$) and results for Pismis~19 underscore the advantages gleaned from employing deep VVV $JHK_s$ data to examine obscured ($A_V \\sim 4$) and differentially reddened intermediate-age clusters. ", "introduction": "} Approximately $30$\\% of the 395 open clusters featuring $n\\ge3$ independent distance estimates exhibit absolute errors $\\ge 20$\\% \\citep[][their Fig.~2]{pa06}. There are $\\ge 2 \\times 10^3$ cataloged Galactic open clusters \\citep{di02}, implying that merely $\\sim 14$\\% of the known sample possess errors $\\le 20$\\% as evaluated from three distance estimates. The uncertainties permeate into analyses which rely on the cluster zero-point, such as the calibration of any constituent standard candles or substellar companions \\citep{lm07,ma11b}. Consider that published parameters for NGC~4349 span $d=900-2200$ pc and $\\tau=0.1-0.6$ Gyr (\\S \\ref{s-ngc4349}). Yet physical parameters for the substellar companion to TYC 8975-2601-1 \\citep{lm07,ka08} rely on those inferred for the host from cluster membership (NGC~4349). Furthermore, the nearer distance and younger age for NGC 4349 potentially imply cluster membership for the classical Cepheid R Cru, which lies within the cluster's corona. Establishing cluster membership would enable the subsequent calibration of Cepheid period-luminosity and period-Wesenheit relations \\citep{tu10}. Such functions bolster efforts to establish extragalactic distances and zero-point the SNe Ia scale \\citep[e.g.,][]{pg04}. The aforementioned examples underscore the broad ramifications of an uncertain cluster scale. Admittedly, age estimates for open clusters are less reliable since a third exhibit absolute errors $>50$\\% \\citep[][$n\\ge3$]{pa06}, and presumably the statistics worsen for obscured clusters near the Galatic plane. In this study, new VVV (VISTA Variables in the V\\'{i}a L\\'actea) $JHK_s$ photometry is employed to illustrate the marked improvement that can be achieved \\textit{vis \\`a vis} open cluster distances. Two important clusters featuring particularly discrepant published parameters are examined, namely Pismis~19 and NGC~4349. Distances for the clusters display a $\\sim 60$\\% spread and individual uncertainties of $\\sim 30$\\%. Efforts to secure precise parameters for Pismis~19 via optical photometry have been complicated by differential reddening and $A_V\\sim4$ magnitudes of obscuring dust. Parameters for Pismis~19 and NGC~4349 derived here exhibit a marked ($>5 \\times$) reduction in uncertainties (\\S \\ref{s-analysis}), and highlight the advantages of using VVV data to determine reliable cluster distances and compliment existing efforts. ", "conclusions": "} VVV $JHK_s$ observations may be employed to help establish precise cluster distances that feature comparatively reduced uncertainties ($\\le 10$\\%). That is illustrated by deriving fundamental parameters for Pismis~19 and NGC~4349, two important clusters which hitherto exhibit sizable uncertainties ($60$\\%, \\S \\ref{s-pismis19} and \\ref{s-ngc4349}). A precise distance determination for Pismis~19 from optical photometry was hampered in part by significant reddening (Fig.~\\ref{fig-dr}, $A_V \\sim 4$). The existing ambiguity surrounding the distance to NGC~4349 ensured that the pertinence of invaluable putative constituents were mitigated (i.e., the classical Cepheid R Cru and a substellar companion for the member TYC 8975-2601-1). Parameters derived for Pismis~19 are: $d=2.40\\pm0.15$ kpc, $<{\\it E(J-H)}>=0.34\\pm0.04$, $\\log{\\tau}=9.05\\pm0.10$ (Fig.~\\ref{fig-dr}), whereas NGC~4349 exhibits $d=1.63\\pm0.10$ kpc, ${\\it E(J-H)}=0.09\\pm0.02$, $\\log{\\tau}=8.55\\pm0.10$ (Fig.~\\ref{fig-cm}). The nature of the VVV survey ensured that the revised results, which have pertinent ramifications, compliment existing estimates and display a marked improvement ($\\ge 5 \\times$) in precision. New VVV $JHK_s$ for stars in NGC~4349 and Pismis~19 imply that: existing physical parameters derived for NGC~4349 127b need to be redetermined in part since the mass for the host star was revised downward to ${\\cal M_{*}/M_{\\sun}}\\sim 3.1$ (\\S \\ref{s-ngc4349}); the classical Cepheid R Cru is not a member of NGC~4349 (\\S \\ref{s-ngc4349}); and VVV $JHK_s$ photometry is particularly suited for constraining parameters of obscured and differentially reddened intermediate-age clusters (e.g., Pismis~19, $A_V \\sim 4$, Fig.~\\ref{fig-dr}). The VVV and UKIDSS surveys \\citep{lu08,mi10} may be employed to achieve significant gains toward strengthening the open cluster distance scale. Yet considerable work remains, and unknown systematic errors may be discovered given the nascent nature of the aforementioned surveys. Consequently, obtaining independent multiband observations are desirable to corroborate derived cluster parameters \\citep[e.g.,][]{ca11}. \\subsection*{{\\rm \\scriptsize ACKNOWLEDGEMENTS}} \\scriptsize{DM is grateful to the following individuals and consortia whose efforts lie at the foundation of the research: 2MASS, P. Stetson (DAOPHOT), W. Lohmann, WEBDA (E. Paunzen), DAML (W. Dias), C. Lovis, M. Mayor, CDS, arXiv, and NASA ADS. We gratefully acknowledge use of data from the ESO Public Survey programme ID 179.B-2002 taken with the VISTA telescope, the Cambridge Astronomical Survey Unit, and funding from the FONDAP Center for Astrophysics 15010003, the BASAL CATA Center for Astrophysics and Associated Technologies PFB-06, the MILENIO Milky Way Millennium Nucleus from the Ministry of Economics ICM grant P07-021-F, and Proyecto FONDECYT Regular 1090213. WG, CMB, and DG are grateful for support from the Chilean Center for Astrophysics FONDAP 15010003 and the BASAL Centro de Astrofisica y Tecnologias Afines (CATA) PFB-06/2007. RK acknowledges support from Proyecto DIUV23/2009, Universidad de Valparaiso. RS acknowledges financial support from CONICYT through GEMINI Project Nr. 32080016.}" }, "1112/1112.4746_arXiv.txt": { "abstract": "The Ly-${\\alpha}$ line of He {\\sc ii} at 304 \\AA\\ is one of the spectral lines of choice for EUV channels of narrow-band imagers on board space telescopes, which provide spectacular intensity images of the outer solar atmosphere. Since the magnetic field information is encoded in the polarization of the spectral line radiation, it is important to investigate whether the He {\\sc ii} line radiation from the solar disk can be polarized, along with its magnetic sensitivity. Here we report some theoretical predictions concerning the linear polarization signals produced by scattering processes in this strong emission line of the solar transition region, taking into account radiative transfer and the Hanle effect caused by the presence of organized and random magnetic fields. We find that the fractional polarization amplitudes are significant (${\\sim}1\\%$), even when considering the wavelength-integrated signals. Interestingly, the scattering polarization of the Ly-${\\alpha}$ line of He {\\sc ii} starts to be sensitive to the Hanle effect for magnetic strengths $B{\\gtrsim}100$ G (i.e., for magnetic strengths of the order of and larger than the Hanle saturation field of the hydrogen Ly-${\\alpha}$ line at 1216 \\AA). We therefore propose simultaneous observations of the scattering polarization in both Ly-${\\alpha}$ lines to facilitate magnetic field measurements in the upper solar chromosphere. Even the development of a narrow-band imaging polarimeter for the He {\\sc ii} 304 \\AA\\ line alone would be already of great diagnostic value for probing the solar transition region. ", "introduction": "The understanding of the complex interface region between the photosphere and corona of the Sun is a very important challenge in astrophysics. In this highly structured and dynamic region of the outer solar atmosphere, where magnetic and hydrodynamic forces compete for dominance, most of the non-thermal energy that creates the corona and solar wind is released. Novel measurements of key physical quantities, like the magnetic field, would improve our understanding of this boundary region. Spectroscopic observations are needed for determining temperatures, flows and waves, while the magnetic field information is encoded in the spectral line polarization (e.g., Stenflo 1994; Landi Degl'Innocenti \\& Landolfi 2004). The aim of this paper is to propose a new technique, based on spectro-polarimetric measurements, which may be particularly useful for determining the magnetic field vector in the solar transition region. The spectral lines that originate in the outer solar atmosphere (chromosphere, transition region and corona) are mainly located in the FUV (912-3000 \\AA) and EUV (100-912 \\AA) spectral ranges, such as the Ly-${\\alpha}$ lines of H {\\sc i} and He {\\sc ii} at 1216 \\AA\\ and 304 \\AA, respectively. Their intensity $I(\\lambda)$ profiles are practically insensitive to the magnetic fields one may expect in the plasma of the outer solar atmosphere, so they cannot be used to obtain quantitative information on the strength ($B$), inclination ($\\theta_B$) and azimuth ($\\chi_B$) of the magnetic field vector. A similar drawback applies to the circular polarization (quantified by the Stokes $V(\\lambda)$ profile) produced by the longitudinal Zeeman effect, because the Stokes-$V$ amplitude scales with the ratio, ${\\cal R}$, between the Zeeman splitting and the Doppler line width. For such solar spectral lines ${\\cal R}{\\ll}1$, especially outside sunspots (${\\cal R}={{1.4{\\times}10^{-7}{\\lambda}B}{/}{\\sqrt{1.663{\\times}10^{-2}T/\\alpha+{\\xi}^2}}}$, where $\\lambda$ is the spectral line wavelength in \\AA, $T$ the kinetic temperature in K, $\\xi$ the microturbulent velocity in ${\\rm km}\\,{\\rm s}^{-1}$, $\\alpha$ the atomic weight of the atom under consideration, and $B$ is in gauss; see Landi Degl'Innocenti \\& Landolfi 2004). The situation is even worse for the linear polarization (quantified by the Stokes $Q(\\lambda)$ and $U(\\lambda)$ profiles) produced by the transverse Zeeman effect because their amplitude scales with ${\\cal R}^2$. In order to obtain quantitative information on the magnetic field of the outer solar atmosphere we need to measure the polarization caused by scattering processes in the spectral lines that form in such regions, ideally using two or more spectral lines with different sensitivities to the Hanle effect (e.g., the review by Trujillo Bueno 2010; see also Stenflo, Keller \\& Gandorfer 1998; Manso Sainz, Landi Degl'Innocenti \\& Trujillo Bueno 2004). We recall that the Hanle effect is the modification of the linear polarization produced by scattering processes in a spectral line, caused by the presence of a magnetic field inclined with respect to the symmetry axis of the incident radiation field. Approximately, the emergent linear polarization is sensitive to magnetic strengths between 0.2$B_H$ and 5$B_H$, where $B_H{=}\\,{1.137{\\times}10^{-7}}/{t_{\\rm life}g}$ is the critical Hanle field for which the Zeeman splitting of the line's level under consideration is equal to its natural width ($t_{\\rm life}$ is the level's radiative lifetime, in seconds, and $g$ its Land\\'e factor). In a recent paper we showed that the hydrogen Ly-${\\alpha}$ line is expected to show measurable scattering polarization when observing the solar disk and that via the Hanle effect the line-center polarization amplitude is sensitive to the presence of magnetic fields in the solar transition region, with good sensitivity to magnetic strengths between 10~G and 100~G (see Trujillo Bueno, \\v{S}t\\v{e}p\\'an \\& Casini 2011). The observational signatures of the Hanle effect might however be confused with the symmetry breaking effects caused by the presence of horizontal atmospheric inhomogeneities (e.g., Manso Sainz \\& Trujillo Bueno 2011). Although for the hydrogen Ly-${\\alpha}$ line such symmetry breaking effects can often be distinguished from the Hanle effect (e.g., \\v{S}t\\v{e}p\\'an \\& Trujillo Bueno 2012), it is of great interest to find another transition region line with measurable scattering polarization but such that is practically immune to the weak magnetic fields expected for the quiet regions of the upper solar chromosphere ($B{<}100$ G). The aim of this Letter is to show some theoretical predictions concerning the linear polarization produced by scattering processes in the Ly-${\\alpha}$ line of He {\\sc ii}, whose significant emission originates in the solar transition region. As we shall see below, the line-center fractional polarization signals are significant (${\\sim}1\\%$) and have a very interesting magnetic sensitivity, which lead us to argue that the development of a space-based instrument capable of obtaining high-resolution $I$, $Q/I$ and $U/I$ images in the He {\\sc ii} Ly-${\\alpha}$ line would be very useful to determine the three-dimensional magnetic structure of the solar transition region. ", "conclusions": "The Stokes $I(\\lambda)$ profile of the Ly-${\\alpha}$ line of He {\\sc ii} at 304 \\AA\\ is about 10 times narrower than that of the Ly-${\\alpha}$ line of H {\\sc i} at 1216 \\AA. While only the core of the hydrogen Ly-${\\alpha}$ line originates in the solar transition region, the Ly-${\\alpha}$ line of He {\\sc ii} is emitted mostly in the solar transition region. Noteworthy is also that the observed line-center intensities in the Ly-${\\alpha}$ lines of H {\\sc i} and He {\\sc ii} are of the same order of magnitude (e.g., Fontenla et al. 1993). The radiative transfer investigation reported in this paper indicates that the line-center fractional linear polarization amplitude of the Ly-${\\alpha}$ line of He {\\sc ii} should be significantly larger than that expected for the Ly-${\\alpha}$ line of H {\\sc i} (e.g., a factor three larger at $\\mu\\,{\\approx}\\,0.3$ in the unmagnetized case). Moreover, the fractional linear polarization amplitude that results from the wavelength-integrated Stokes profiles of the He {\\sc ii} line turns out to be similar to the line-center signal\\footnote{In a forthcoming publication we will show that the fractional linear polarization amplitude obtained from the wavelength-integrated Stokes profiles is even larger when partial frequency redistribution effects are taken into account.}. These results partially compensate the fact that the total number of photons emerging from any solar disk position in the Ly-${\\alpha}$ line of He {\\sc ii} is significantly smaller than within a small (0.1 \\AA) wavelength interval around the hydrogen Ly-${\\alpha}$ line center. We have shown also that for magnetic strengths $B{<}100$ G the Ly-${\\alpha}$ line of He {\\sc ii} is nearly immune to magnetic fields\\footnote{The fact that with a narrow-band instrument, designed to measure the intensity and polarization of the Ly-${\\alpha}$ line of He {\\sc ii}, we may have a small contribution from the nearby Si {\\sc XI} emission at 303.3 \\AA\\ should not be a problem because the critical Hanle field of this line is also very large (i.e., $B_H{\\approx}730$ G).}. This is particularly interesting because for magnetic strengths between 10 and 100 G the Ly-${\\alpha}$ line of H {\\sc i} is indeed sensitive to the Hanle effect. Therefore, outside active regions the Ly-${\\alpha}$ line of He {\\sc ii} can be used as a reasonable reference line for facilitating magnetic field ``measurements\" via the Hanle effect in the Ly-${\\alpha}$ line of H {\\sc i}. All these results encourage the development of the following instruments for a space telescope: \\begin{itemize} \\item A spectropolarimeter for measuring the line-core polarization of the Ly-${\\alpha}$ line of H {\\sc i} with a spectral resolution of at least 0.1 \\AA. \\item A narrow-band polarimeter for obtaining intensity and linear polarization images of the solar transition region in the Ly-${\\alpha}$ light of He {\\sc ii}. \\end{itemize} Although the combined use of the two Ly-$\\alpha$ lines opens up a new diagnostic window for magnetic-field measurements in the upper solar chromosphere, the interpretation of such spectro-polarimetric observations will still require radiative transfer calculations in realistic atmospheric models because the two spectral lines are not formed in exactly the same way (e.g., Jordan 1975; Fontenla, Avrett \\& Loeser 2002; Pietarila \\& Judge 2004). In spite of such a complication the comparison between spectro-polarimetric observations in the two Ly-$\\alpha$ lines can provide unique insights into the physics and geometry of the transition region. Finally, it is of interest to note that off-limb observations of resonant scattering polarization in the He {\\sc ii} 304 \\AA\\ line may be also useful for exploring the geometry and magnetic field structure of spicules, prominences and of the solar corona. {\\bf Acknowledgments} Financial support by the Spanish Ministry of Science and Innovation through projects AYA2010-18029 (Solar Magnetism and Astrophysical Spectropolarimetry) and CONSOLIDER INGENIO CSD 2009-00038 (Molecular Astrophysics: The Herschel and Alma Era) is gratefully acknowledged." }, "1112/1112.0414_arXiv.txt": { "abstract": "We present $UBV$ photometry of the highly reddened and poorly studied open cluster Berkeley~55, revealing an important population of B-type stars and several evolved stars of high luminosity. Intermediate resolution far-red spectra of several candidate members confirm the presence of one F-type supergiant and six late supergiants or bright giants. The brightest blue stars are mid-B giants. Spectroscopic and photometric analyses indicate an age $50\\pm10$~Myr. The cluster is located at a distance $d\\approx4$~kpc, consistent with other tracers of the Perseus Arm in this direction. Berkeley~55 is thus a moderately young open cluster with a sizable population of candidate red (super)giant members, which can provide valuable information about the evolution of intermediate-mass stars. ", "introduction": "\\label{introduction} After exhaustion of H in their cores, stars evolve toward lower $T_{{\\rm eff}}$, becoming, according to their masses, red giants or supergiants (RSGs). Both high- and intermediate-mass stars are subject to complex physical processes in their later evolution, which result in important changes in their observable characteristics ($T_{{\\rm eff}}$, $L_{{\\rm bol}}$). As a consequence, their evolutionary tracks trace loops in the HR diagrams \\citep[e.g.,][]{chiosi}. The shape of these loops depends on the physics of the stellar interior, generally modeled via poorly understood parameters \\citep[e.g.,][]{chiosi,mf94,salas99,mm00}. The extent of semi-convection and overshooting has very important consequences on many aspects of stellar evolution, most notably the ratio of initial mass to white dwarf mass \\citep[e.g.,][]{jeff97,weide00} and the boundary between stars that leave white dwarfs as remnants and those that explode as supernovae \\citep[SNe; e.g.,][]{poel08}. M-type supergiants represent the final evolutionary stage of moderately massive stars, with typical initial masses in the 8--25$\\,M_{\\odot}$ range \\citep{levesque05}. Such objects are the immediate progenitors of Type II-P SNe, the most frequent type of supernova explosion in the Local Universe \\citep{smartt09,smith11}. Most of the explosions come from low-mass RSGs, stars with initial masses $M_{*}\\la12\\,M_{\\sun}$. The lower mass limit for a star to produce a supernova has been estimated at $\\approx 8.5^{+1}_{-1.5}\\,M_{\\sun}$ \\citep{smarttal}. Nevertheless, stars with lower masses are also classified as supergiants, generally K\\,Ib objects, showing that the morphological separation between red supergiants and red giants does not coincide exactly with the boundary between high- and intermediate-mass stars. Since stars become more luminous as they experience blue loops and later enter the AGB branch, a given star may appear first as a bright giant and later in its evolution as a supergiant. Open clusters with large populations of evolved stars can help constrain the inputs of models and therefore improve our understanding of such basic questions. Identification and study of such clusters is thus an important astrophysical issue. To this aim, we have analyzed a sample of poorly-studied, cataloged, optically-visible clusters in the Northern Hemisphere, using 2MASS data \\citep{skru06}. We made use of the $Q_{{\\rm IR}}$ index, a reddening-free parameter, defined as $(J-H)-1.70(H-K_{{\\rm S}})$, which has been proved to be a very useful tool for identification of intrinsically blue stars \\citep[see, e.g.,][]{cp05,ns07} and separation of luminous red stars from red dwarfs \\citep{neg11}. Berkeley~55 (Be~55) was readily identified as containing an obvious clump of red luminous stars associated with a sequence of intrinsically blue stars. Be~55 is a faint, compact open cluster in the constellation Cygnus. The WEBDA database\\footnote{At {\\tt http://www.univie.ac.at/webda/}} provides coordinates RA:~21h 16m 58s, Dec:~$+51^{\\circ}\\:45\\arcmin\\:32\\arcsec$ ($\\ell=93\\fdg03$, $b=+1\\fdg80$). \\citet{macie07} presented $BV$ photometry of an extended field around the obvious cluster core and found it to be extremely compact, with $r_{{\\rm core}}=0\\farcm7$. They derived a distance $d=1.2$~kpc and an age $\\log t=8.5$ (315~Myr), which was later used by \\citet{tadross} to calibrate a fit to the 2MASS data in the region. However, the cluster looks too compact for this age and distance, while the isochrone fit of \\citet{tadross} would suggest that most of the evolved cluster members are AGB stars, an extremely unusual situation. In this paper, we present $UBV$ photometry of the cluster area and a spectroscopic survey of likely members, showing that its population is much younger than implied by the single-color fit of \\citet{macie07}. From our new data, we show that Be~55 is a young ($\\log t\\approx7.7$) open cluster with a rich population of evolved stars. ", "conclusions": "\\label{discussion} We have presented spectroscopy of a sizable sample of stars in the open cluster Berkeley~55, revealing a population of 7 low-luminosity supergiants or bright giants (one of them, not a certain member), 4 mid-B giants and a main sequence starting around B4\\,V. \\subsection{Cluster parameters} Only two previous studies of the cluster have been published. \\citet{macie07} derive an age of $\\sim315$~Myr. \\citet{tadross} used their result to calibrate his fitting procedure and therefore derives a similar age of 300~Myr, from a fit to the 2MASS data alone. This age is completely incompatible with the observed population, as stars earlier than B5 are only seen in clusters younger than 100~Myr. Clusters with $\\sim100$~Myr have stars of spectral type $\\sim$B6 around the main-sequence turn-off (e.g., NGC~7790, \\citealt{matthews95}; Berkeley~58, \\citealt{turner08}). The difference between our age determination and that of \\citet{macie07} is almost certainly due to the identification of the evolved population. The decontaminated $V/(B-V)$ in \\citet{macie07} does not show a clump of red stars, as in our Fig.~\\ref{rawcmd}. Some of the (super)giants seem to be blended in their images and do not appear in their original photometry. The others are not conspicuous when a wide field is considered, because of contamination by field interlopers with similar $(B-V)$ color. The stars forming this clump, however, cannot be overlooked when the infrared data are considered. S1 is the brightest star in $K_{{\\rm S}}$ within $10\\arcmin$ of the cluster center, and there are only 5 field stars with $K_{{\\rm S}}$ magnitudes comparable to the 7 red (super)giants. The 7 candidate members form a clearly separated clump in the $(J-K_{{\\rm S}})$/$K_{{\\rm S}}$, $H-K_{{\\rm S}}$/$K_{{\\rm S}}$ and $Q$/$K_{{\\rm S}}$ diagrams, and have similar (and unusual) spectral types. Though accurate radial velocities will be needed to confirm individual memberships, there can be no doubt that they represent the population of evolved stars in the cluster. The small systematic offset in $(B-V)$ between our photometry and that of \\citet{macie07} is unlikely to have biased their age determination. It is simply very difficult to obtain an accurate fit when photometry in only two bands is available. At Galactic longitude $\\ell=93\\fdg0$, a distance $\\sim4$~kpc (corresponding to our $DM=13.0$) seems compatible with tracers of the Perseus arm, though some Galactic models suggest a higher distance for this feature. Spiral tracers in this direction are scarce. NGC~7067 ($\\ell=91\\fdg2$) has distance estimates of 3.6~kpc \\citep{ys04} and 4.4~kpc \\citep{hassan73}. NGC~7128 ($\\ell=97\\fdg3$), with $DM=13.0\\pm0.2$ (4~kpc), is another likely tracer at some distance from Be~55 \\citep{balog}. As the age of Berkeley~55 is $\\sim50$~Myr, the red stars in the cluster should have initial masses $M_{*}\\ga7.0\\,M_{\\odot}$ \\citep{marigo}, or slightly higher if they started their lives rotating quickly \\citep[cf.][]{mm00}. Such stars define the boundary between intermediate-mass and massive stars, and therefore can provide important constraints for stellar evolution models. Milky Way clusters with similar ages and rich populations of RSGs (or bright giants) are relatively scarce. NGC~6649 ($\\sim50$~Myr; \\citealt{turner81}) with one Cepheid and four red (super)giants is one of the few examples. NGC~6664 (also $\\sim$50~Myr; \\citealt{schmidt82}) has a similar population, but is less well studied. As a consequence, the sample of Galactic RSGs studied by \\citet{levesque05} contains almost no stars in the $6\\,M_{\\odot}$ 340 km\\,s$^{-1}$), collimated bipolar outflow nature. The detection of the northwestern knot in POSS red plates allows us to carry out a proper motion analysis by combining three POSS red plates and two narrow-band H$\\alpha$+[N\\,{\\sc ii}] CCD images, with a time baseline of $\\simeq$ 57 yr. A proper motion of 20 $\\pm$ 6\\,mas\\,yr$^{-1}$ along position angle 312$^{\\circ}$ $\\pm$ 15$^{\\circ}$, and a dynamical age of 1375$^{+590}$\\llap{$_{-320}$}\\,yr are obtained for the bipolar outflow. The measured proper motion and the spatio-kinematical properties of the bipolar outflow yield a lower limit of 2.7\\,kpc for the distance to Hu\\,1-2. ", "introduction": "Planetary nebulae (PNe) represent the final evolutionary stages of low-- and intermediate--mass stars. They are formed when the envelope ejected during the previous Asymptotic Giant Branch phase is ionized by the hot central star. In the last years the presence of collimated bipolar outflows in PNe has been widely recognized. These outflows probably play a crucial role in the shaping and dynamical evolution of PNe \\citep{b25}. Therefore, the knowledge of their dynamical properties is critical to understand how PNe form. The kinematics of the outflows has been studied by means of high-resolution spectroscopy that permits to obtain the radial component of the velocity (e.g., Miranda et al. 1999, 2001; Guerrero et al. 2000). The measurement of the tangential component, i.e., the proper motion of collimated outflows, is restricted to a very few cases and it has been obtained mostly from {\\it HST} observations (NGC\\,7009: Fern\\'andez et al. 2004; He\\,2-90: Sahai et al. 2002; Hen\\,3-1475: Borkowski \\& Harrington 2001; Riera et al. 2003) or {\\it VLA} radio continuum data (NGC\\,7009: Rodr\\'{\\i}guez \\& G\\'omez 2007) with the noticeable exceptions of the works by Liller (1965) and Meaburn (1997) who used photographic material to measure proper motions of the collimated outflows in NGC\\,7009 and KjPn\\,8, respectively. The measurement of proper motions is important because it allows us to obtain the distances to the PN, if the expansion velocity of the outflows is known (e.g., Meaburn 1997), or the expansion velocity vector, if the distance is known (e.g., Fern\\'andez et al.2007). Therefore, further measurements of proper motions of collimated outflows in PNe are highly desirable. Collimated outflows may exist in the PN Hu\\,1-2 (PN\\,G086.5$-$08.8) that has been classified as elliptical with ansae \\citep{b4}. The kinematics of the inner nebular regions ($\\simeq$ 10$''$) of Hu\\,1-2 was analyzed by Sabbadin et al. (1983, 1987). These authors identified a toroid and bipolar lobes but a reconstruction of the spatio-kinematical structure was not possible because of the peculiar velocity field. In addition, one single knot can be hinted in the images by Manchado et al. (1996) northwestwards and well outside the main nebula, while a possible southeastern counterpart is superimposed by a field star. Remarkably, the northwestern knot can be identified in POSS red plates of the Digitized Sky Surveys. In this respect, Hu\\,1-2 constitutes a rare case (together with NGC\\,7009 and KjPn\\,8) because knots/collimated outflows in PNe usually are weak and/or located close to the much brighter main nebula, which do not favor their detection in POSS plates. Moreover, the H$\\alpha$ and [N\\,{\\sc ii}] emission lines are by far the dominant emissions from knots/collimated outflows in PNe in the optical (e.g., Balick 1993, 1994), as it is the case of the northwestern knot of Hu\\,1-2. In consequence, the emission from the northwestern knot of Hu\\,1-2 detected in the POSS red plates can be attributed to these two emission lines. Therefore, Hu\\,1-2 offers an excellent opportunity to attempt a proper motion analysis of knots/collimated outflows in PNe by combining POSS red plates with modern H$\\alpha$+[N\\,{\\sc ii}] imagery. In this paper we present a new H$\\alpha$+[N\\,{\\sc ii}] image of Hu\\,1-2 obtained under subarsecond conditions, and a high resolution, long-slit spectrum that allow us to identify the southeastern counterpart of the northwestern knot and to establish that these two knots constitute a high velocity, collimated bipolar outflow. In addition, we carry out an analysis of five images obtained at different epochs, including three images from the POSS, to measure the proper motion of the northwestern knot and to constrain the distance to Hu\\,1-2. ", "conclusions": "We have presented a subarsecond H$\\alpha$+[N\\,{\\sc ii}] image and a high resolution long-slit spectrum of Hu\\,1-2 covering its bipolar knots. In addition, we have analyzed five images of Hu\\,1-2 with a time baseline of $\\simeq$ 57 yr, including three POSS red plates in which one of the bipolar knots can be identified, to measure its proper motion and to constrain the distance to the nebula. The main conclusions of this paper are: \\begin{itemize} \\item The subarcsecond image allows us to identify the two bipolar knots of the pair, while in previous images only one of them is clearly detected. The knots present bow-shock-like morphology, are located at 27$\\farcs$5 from the central star and oriented at PA 320$^{\\circ}$ that coincides with the orientation of the main nebular axis. \\item The radial velocity of the knots is $\\pm$ 60 km\\,s$^{-1}$. If the knots moved along the main nebular axis (tilted $<$ 10$^{\\circ}$ with respect to the plane of the sky), their expansion velocity would be $>$ 340\\,km\\,s$^{-1}$. Therefore, morphology and kinematics demonstrate that these knots constitute a true high velocity, collimated bipolar outflow and, most probably, represent bow-shocks associated to high velocity bullets. \\item A proper motion of 20 $\\pm$ 6 mas\\,yr$^{-1}$ along PA 312$^{\\circ}$ $\\pm$ 15$^{\\circ}$ is obtained for the bipolar knots. The corresponding dynamical age is 1375$^{+590}$\\llap{$_{-320}$}\\,yr. \\item A lower limit of 2.7\\,kpc for the distance to Hu\\,1-2 is required to make compatible the measured proper motion with the spatio-kinematical properties of the bipolar outflow. This lower limit rules out a large fraction of the statistical distances and the extinction distance previously determined for Hu\\,1-2. \\end{itemize}" }, "1112/1112.0622_arXiv.txt": { "abstract": "We investigate the feasibility of implementing a system that will coordinate ground-based optical telescopes to cover the Fermi GBM Error Circle (EC). The aim of the system is to localize GBM detected GRBs and facilitate multi-wavelength follow-up from space and ground. This system will optimize the observing locations in the GBM EC based on individual telescope location, Field of View (FoV) and sensitivity. The proposed system will coordinate GBM EC scanning by professional as well as amateur astronomers around the world. The results of a Monte Carlo simulation to investigate the feasibility of the project are presented. ", "introduction": "Gamma-ray bursts (GRB) are bursts of gamma-rays that arguably signal the birth of a black hole somewhere in the universe. Based on the duration and spectrum, two classes of bursts have been observed~\\citep{Kouveliotou1993}: those that last less than two seconds and have on the average hard spectra (short GRBs), and those that last longer than two seconds and are spectrally softer (long GRBs). The exact nature of the GRB progenitors is unknown, although it is possible that long GRBs come from the collapse of massive, rapidly rotating stars~\\citep{Woosley2006, WoosleyBloom2006} and short GRBs result from the merger of compact objects~\\citep{Eichler1989, Narayan1992}. Regardless of the progenitor system, accretion onto the resulting compact object is thought to create a highly relativistic jet. The prompt gamma-ray emission from the GRBs may arises from the internal shocks due to collisions of faster shells with slower ones ejected earlier by the central engine. The subsequent softer multi-wavelength emission, referred to as the afterglow, may be due to the collision of the fireball with the extra-stellar material~\\citep{Rees1994,Piran1999}. Our understanding of GRBs progressed very rapidly after the detection of multi-wavelength afterglows. Well localized, favorably positioned GRBs get fairly good multi-wavelength afterglow coverage. Currently, the leading GRB afterglow detection mission is $Swift$ which detects 90--100 GRBs annually. Most of the $Swift$ GRBs get observed by various instruments around the world because of its rapid arc-minute localization capability. Compared to $Swift$, $Fermi$ Gamma-ray Burst Monitor (GBM) detects about 250 burst per year but with poor localization. The Error Circle (EC) of $Fermi$ GBM detected bursts is too large for a single telescope to observe effectively. The typical statistical uncertainty of the GBM burst location is about 3.3 degrees. However, when combined with the systematic uncertainty of 3.8 degrees~\\citep{Briggs2009}, the total burst location uncertainty is $\\sim$ 5.0 degrees (i.e., 5.0 degree error radius). Naturally, a brighter burst will have a smaller GBM EC than a weaker burst. Even though the localization is poor, GBM detected bursts have very good timing and spectral information including crucial $E_{\\rm peak}$ measurements ($E_{\\rm peak}$ is the peak energy of the GRB $\\nu F_{\\nu}$ spectrum). If there is a method to localize GBM detected GRBs to a few arc-seconds uncertainty, then large telescopes can do deeper follow--up observations to determine the redshift of the burst and also potentially identify any emerging supernova. In addition, $Swift$ can also slew quickly to the GBM burst in order to observe the X-ray afterglow and obtain its light curve in X-ray wavelengths. Based on $Swift$ observations about $\\sim$ 60\\% of GRBs have optical counterparts~\\citep{Gehrels2009}. These optical counterparts are detected by various observatories with R magnitudes ranging from 14 to 22 within few hours after the burst~\\citep{Fiore2007}. Thus, it is reasonable to assume about 60\\% of the GBM detected GRBs also have optical counterparts with similar brightness distribution. If we were able to cover the entire GBM EC within about 24 hours after the burst it is conceivable that we would be able to find optical afterglows of $\\sim$150 GRBs per year, which is more than the total number of burst $Swift$ detects per year. Due to the small energy range (15-150 keV) of the $Swift$ Burst Alert Telescope (BAT), $Swift$ measurements alone cannot constrain the $E_{\\rm peak}$ of all BAT detected bursts~\\citep{Sakamoto2009}. In contrast, due to the wide energy range (8 keV - 40 MeV) of GBM, all GRBs detected by GBM have fairly good $E_{\\rm peak}$ measurements. Hence, addition of possibly another $\\sim$100 bursts per year with good $E_{\\rm peak}$ and redshift measurements may allow us to explore the validity of various GRB luminosity relations and to conduct detailed GRB Hubble diagram studies. We have investigated the feasibility of using a system to do coordinated monitoring of the BAT field-of-view (FoV) for prompt optical emission from GRBs~\\citep{Ukwatta2011}. The study showed that with the current instrumentation, performing such a coordinated monitoring is not practical mainly due to the BAT's very large FoV. However, a similar coordinated observing campaign can be used to find the optical afterglow of GBM detected bursts. The GBM EC is much smaller than the BAT FoV and observers do not need to continuously monitor the field to detect the optical afterglow. This enables a given observatory to perform multiple observations inside the GBM EC and thereby increase the chance of a afterglow detection. ", "conclusions": "We investigate the feasibility of implementing a system that will coordinate ground based telescopes (both amateur and professional) to scan the GBM EC in order to localize GBM bursts. Unlike the GCN system, proposed system will send individual customized messages to telescopes to observe certain patches in the GBM EC. The scientific objective of the system is by localizing GBM detected burst, we will be able to increase the number of GBM bursts with mutil-wavelength followups potentially with redshifts measurements. These measurements are scientifically very important because there are hints that GBM bursts may represent significantly different burst population. Based on our simulation, we can detect about 25 GRB afterglows per year using just 10 telescopes with $1.0^{0} \\times 1.0^{0}$ field-of-view. With more telescopes participating in the program, we should be able to detect many more afterglows and study a potentially interesting burst population that is currently inaccessible to the GRB community." }, "1112/1112.0308_arXiv.txt": { "abstract": "We demonstrate a new technique for detecting radio transients based on interferometric closure quantities. The technique uses the bispectrum, the product of visibilities around a closed-loop of baselines of an interferometer. The bispectrum is calibration independent, resistant to interference, and computationally efficient, so it can be built into correlators for real-time transient detection. Our technique could find celestial transients anywhere in the field of view and localize them to arcsecond precision. At the Karl G. Jansky Very Large Array (VLA), such a system would have a high survey speed and a 5$\\sigma$\\ sensitivity of 38 mJy on 10 ms timescales with 1 GHz of bandwidth. The ability to localize dispersed millisecond pulses to arcsecond precision in large volumes of interferometer data has several unique science applications. Localizing individual pulses from Galactic pulsars will help find X-ray counterparts that define their physical properties, while finding host galaxies of extragalactic transients will measure the electron density of the intergalactic medium with a single dispersed pulse. Exoplanets and active stars have distinct millisecond variability that can be used to identify them and probe their magnetospheres. We use millisecond time scale visibilities from the Allen Telescope Array (ATA) and VLA to show that the bispectrum can detect dispersed pulses and reject local interference. The computational and data efficiency of the bispectrum will help find transients on a range of time scales with next-generation radio interferometers. ", "introduction": "Since the discovery of the first pulsar, radio observations have steadily revealed a wide range of fast ($\\lesssim1$s) transient phenomena. Pulsars \\citep{1968Natur.217..709H,1982Natur.300..615B,2006Natur.442..892C}, planets and active stars \\citep{2007ApJ...663L..25H,2007AA...475..359G}, and peculiar new transients \\citep{2006Natur.439..817M,2007Sci...318..777L} all have distinct radio phenomenology related to the physics of the host and the medium through which the signal propagates. While large, single-dish telescopes have pioneered the study of fast radio transients, the nature of their design has several limits. Single-dish telescopes can only localize an individual pulse to somewhere in its field of view \\citep[typically arcminutes;][]{2001MNRAS.328...17M}. Timing of periodic emission can localize very precisely, but individual pulses can't be localized well enough to easily identify optical or X-ray counterparts. Since interferometers simultaneously have a large collecting area and a large field of view, they survey more efficiently than single-dish telescopes. Survey speed scales roughly the square of the product of the number of dishes and their diameter; this predicts that the VLA surveys almost 40 times faster than the Green Bank Telescope, assuming all other factors are equal. Despite this potential, next-generation radio interferometers (VLA, PAPER, LOFAR, ASKAP, and MeerKAT) have fast transients observing modes that are essentially identical to the single-dish concept: point one or more beams on the sky and search for pulses \\citep{2011ApJ...734...20M,2011arXiv1109.2659S}. In principle, interferometers have much more information available to them because they detect the electromagnetic wave over a distributed array of receivers. Instead of forming a beam, interferometers can measure visibilities by correlating signals across the array. Visibilities contain spatial information that can reject interference and localize a source to \\emph{arcsecond} precision over a large field of view. Traditionally, visibilities haven't been used in this way because it was difficult to run correlators this fast and the data volume was too large to handle or efficiently search for pulses \\citep{2011ApJ...742...12L}. As a result, a standard radio interferometer averages data to second time scales, losing nearly all information about millisecond variability in the sky. Even on these longer time scales, imaging algorithms are complex and subject to subtle statistical biases \\citep{1998AJ....115.1693C}. The challenge of data rates and volumes is expected to grow with next-generation interferometers, which will have more antennas, wider bandwidths, and larger fields of view \\citep{2011ApJ...739L...1P}. Here, we present a technique that reduces the challenge of detecting transients in visibilities to the much simpler single-dish problem, while maintaining the utility of visibilities. The technique is based on the ``bispectrum'', a quantity formed from visibilities that is sensitive to a point source anywhere in the field of view \\citep{1987AA...180..269C,1989AJ.....98.1112K,1995AJ....109.1391R}. Differencing visibilities makes it possible to interpret the bispectrum as a statistic for transients anywhere in the field of view on time scales from milliseconds to seconds. We describe an algorithm for real-time, arcsecond localization of transients using raw visibilities from a radio interferometer. The theory of closure quantities and their application to transients is presented in \\S \\ref{theory}. In \\S \\ref{demo} we test the theoretical predictions with millisecond time scale visibilities from the VLA and the ``Pocket Correlator'' (PoCo) instrument at the ATA \\citep{2011ApJ...742...12L}. As shown in \\S \\ref{demand}, this technique has a moderate computational demand and can reduce the flow of data to a manageable size. We conclude with thoughts about new science accessible with this technique and how it can address the growing challenge of big data in the study of transients with radio interferometers. ", "conclusions": "We demonstrate how to use the bispectrum to detect and localize transients with radio interferometers. The technique produces a single statistic proportional to the brightness of celestial transients, thus reducing the problem of interferometric pulse detection to that of single-dish pulse detection. For a small cost in sensitivity, this technique makes an interferometer into a massive multibeam receiver with a proportionately large survey speed. Using coherent beamforming to survey a similar area with next-generation radio interferometers would require $10^2$\\ to $10^5$\\ times more computational power. The bispectrum technique is efficient enough to build into a correlator for real-time, millisecond transient detection. This system could save visibility data associated with a candidate and apply standard calibration and imaging techniques to localize it to arcsecond precision. The growing capacity of correlators for interferometers will make it possible to run such a system in real time to commensally search all data for transients. The technique can also be used to find transients offline and on longer time scales. In that case, the bispectrum is useful because it is computationally simple, resistant to interference, and requires no calibration. A real-time, commensal system would probe an unprecedented volume of data, giving it sensitivity to rare transients. Each pulse could be localized in real time to arcsecond precision, opening a range of new science. Pulses from rotating radio transients \\citep{2006Natur.439..817M} could be associated with X-ray counterparts to help address their relation to the normal pulsar population. The dispersion of extragalactic transients would measure the electron density of the intergalactic medium for the first time \\citep{2003ApJ...596..982M,2007Sci...318..777L}; localization will help find optical counterparts and redshifts to host galaxies for each pulse. A fast transients survey of the Galactic center could detect individual pulses from pulsars that would probe the dispersion toward the region and potentially find a pulsar in orbit around a supermassive black hole \\citep{2004ApJ...615..253P,2010ApJ...715..939M,2011arXiv1111.4216W}. Magnetospheres of exoplanets and active stars can emit bright, complex bursts that could uniquely identify them \\citep{2004ApJ...612..511L,2006ApJ...653..690H}. The strength of the bispectrum technique is its real-time ability to find transients throughout the field of view and reduce the flow of data to select times. This shows how algorithms will be key to efficiently extracting the best science in the era of big data. The ability to find pulses with interferometers may have application to communications problems, such as developing smart antennas for the efficient use of spectrum. If so, this technique would follow in the footsteps of previous transients searches that contributed to the development of Wi-Fi \\citep{1978Natur.276..590O}." }, "1112/1112.3598_arXiv.txt": { "abstract": "The rapid, irreversible change of the photospheric magnetic field has been recognized as an important element of the solar flare process. This Letter reports such a rapid change of magnetic fields during the 2011 February 13 M6.6 flare in NOAA AR 11158 that we found from the vector magnetograms of the Helioseismic and Magnetic Imager with 12-min cadence. High-resolution magnetograms of Hinode that are available at \\sm$-$5.5, $-$1.5, 1.5, and 4 hrs relative to the flare maximum are used to reconstruct three-dimensional coronal magnetic field under the nonlinear force-free field (NLFFF) assumption. UV and hard X-ray images are also used to illuminate the magnetic field evolution and energy release. The rapid change is mainly detected by HMI in a compact region lying in the center of the magnetic sigmoid, where the mean horizontal field strength exhibited a significant increase by 28\\%. The region lies between the initial strong UV and hard X-ray sources in the chromosphere, which are cospatial with the central feet of the sigmoid according to the NLFFF model. The NLFFF model further shows that strong coronal currents are concentrated immediately above the region, and that more intriguingly, the coronal current system underwent an apparent downward collapse after the sigmoid eruption. These results are discussed in favor of both the tether-cutting reconnection producing the flare and the ensuing implosion of the coronal field resulting from the energy release. ", "introduction": "It has been known that the long-term evolution of photospheric magnetic field (PMF) driven by new flux emergence and surface flows plays important roles in building up free energy in the corona, and that this free magnetic energy powers flares and coronal mass ejections \\citep[CMEs;][]{priest02}. On the other hand, a short-term variation of the PMF associated with flares has not been considered because PMFs are strongly line-tied to the dense high-$\\beta$ photosphere and thus are thought unlikely to be altered by any flare-related disturbances created in the tenuous low-$\\beta$ corona. Only recently, a back reaction of the coronal magnetic field (CMF) on the PMF during the reconfiguration of the CMF has been seriously considered from the theoretical point of view. The idea is that the CMF should contract inward, as the magnetic energy of the CMF decreases after flares and/or CMEs \\citep[called ``implosion'';][]{hudson00}. This may make the PMF be oriented more horizontally, resulting in a Lorentz force acting downwardly on the solar surface \\citep{fisher10,hudson08}. In fact, a few theoretical flare models predict the appearance of more horizontal PMFs after flares/CMEs in the form of newly formed, low-lying field lines close to the surface \\citep[e.g.,][]{melrose97,moore01}. The phenomenon of coronal implosion has been reported in several recent observational studies \\citep{liur+implosion09,liur+wang09,liur+wang10}. The related phenomenon of the rapid change of the PMF to incline more horizontally after flares had also been reported almost two decades ago \\citep{wang92,wang94}. Namely, the transverse field near the magnetic polarity inversion line (PIL) at the core flaring region is enhanced rapidly and irreversibly, which is often accompanied by an increase of magnetic shear. Such a rapid change of vector PMF change closely associated with flares/CMEs has been consistently found later on using ground-based \\citep{schmieder94,wang02b,wang04,wang+liu05,liu05,wang07a,wang10} and \\hinode\\ observations \\citep{jing08,li09}. Indirect evidence includes the unbalanced flux evolution of the line-of-sight magnetic field \\citep{wang10}, the variation of the sunspot white-light structure in flaring regions \\citep{wang04a,deng05,liu05,chen07b}, and the change in the pattern of the penumbral Evershed flow \\citep{deng11}. Additional support also comes from the three-dimensional (3D) MHD simulations of an emerging twisted flux tube \\citep{fan10}, the eruption of which induces similar impact on the low-altitude magnetic fields \\citep{li10}. \\begin{figure*}[t] \\epsscale{1} \\plotone{f1.eps} \\caption{$B_v$ maps are superimposed with arrows (for clarity, those in negative and positive fields are coded in different colors) representing horizontal plasma flows (a) and magnetic field vectors (b--d). The NLFFF lines in (e) stem from the four flare kernels FP1--FP4 in 1700~\\AA\\ (f) at the flare onset. The white contours in (e) outline the double J-shaped flare ribbons in Ca~{\\sc ii}~H. Contours (with levels of 38\\%, 48\\%, 58\\%, 68\\%, 78\\%, 88\\% of the maximum flux) in (f) denote \\hsi\\ CLEAN images reconstructed with detectors 2--8 (except that 17:32:08~UT image is made without the detector 2 due to the relatively low counts at this time). The white box overplotted on all panels denotes the region R under study with enhanced $B_h$ after the flare. \\label{f1}} \\end{figure*} While the evolution of vector PMFs closely associated with flares were mostly studied using ground-based observations subject to seeing variations, \\citet{wang+shuo11} and \\citet{sun11} used data from the \\hmi~\\cite[HMI;][]{schou11} on board the recently launched \\Sdo\\ (\\sdo) to find a rapid enhancement of transverse field associated with the 2011 February 15 X2.2 flare. Since solar activity is now increasing, this is an opportune moment to advance the study of the flare-related evolution of the PMF using the state-of-the-art, seeing-free photospheric vector magnetograms acquired by space missions. \\begin{figure*}[t] \\epsscale{.8} \\plotone{f2.eps} \\caption{Temporal evolution of $\\langle B_h \\rangle$ of the region R in Figure~\\ref{f1}, in comparison with the flare light curves of soft and hard X-rays. The mean error of $B_h$ of the region R at each time instance measured by HMI is plotted as error bars. The error estimates for \\hinode\\ data are currently not usable$^2$. \\label{f2}} \\end{figure*} In this Letter, we investigate the PMF evolution associated with an M6.6 flare on 2011 February 13 that was well covered by both HMI and \\hinode\\ (see Section~\\ref{data}), in which the transverse field enhancement at the flaring PIL is unequivocally detected. We will discuss the rapid change of the 3D CMF associated with that of the PMF using nonlinear force-free field (NLFFF) extrapolations from the active region. ", "conclusions": "We have used the vector magnetograms from HMI and \\hinode\\ to analyze the change of the PMF associated with the 2011 February 13 M6.6 flare, and interpreted the results with the aid of NLFFF modeling based on the \\hinode\\ data under the context of 3D magnetic reconnection. Major results are summarized as follows. \\begin{enumerate} \\item A compact region R at the central flaring PIL shows a rapid and permanent enhancement of $\\langle B_h \\rangle$ by \\sm320~G (\\sm28\\% of the preflare magnitude as observed by HMI) closely associated with the flare. A similar trend of $\\langle B_h \\rangle$ evolution is also detected by \\hinode. Overall, the PMF at the region R becomes more inclined to the surface after the flare. \\item The magnetic fields lying immediately above R, which correspond to the central part of the sigmoidal region, are highly sheared and have strong horizontal electric currents. The most intense flare kernels are initially located at the two ends of this channel-like current system, implying that the start of electron acceleration could be closely associated with the strong horizontal currents \\citep{leka93}. The configuration of the NLFFF lines is favorable for a subsequent tether-cutting reconnection \\citep{moore01}, when disturbed by converging surface flows. This ends up with the short and low-lying loops, which explains the enhanced $B_h$ and $\\tilde{S}$ at the region R. \\item Across the flare energy-release time, the nonpotentiality as represented by $\\tilde{S}$ increases in the height range from the surface up to \\sm1~Mm and then decreases at higher altitudes, which is closely correlated with the evolution of $J_h$. This is in accordance with the reconnection of current-carrying loops modeled by \\citet{melrose97}. Remarkably, the vertical profiles of $J_h$ as well as $\\varphi$ before and after the flare clearly demonstrate a downward collapse of the CMF by \\sm1.3~Mm toward the surface, which is probably related to coronal implosion \\citep{hudson00}. \\end{enumerate} In summary, the change of the PMF and the corresponding change of the CMF observed in this study are not caused by the long-term development of the active region but occurred as rapid as the flare. The magnetic field change may involve two closely related physical processes: the tether-cut reconnection producing the flare and the ensuing collapse of the CMF resulting from the energy release." }, "1112/1112.4976_arXiv.txt": { "abstract": "{Very-high-energy (VHE, E $>$ 100 GeV) \\g\\ radiation has already been detected from several supernova remnants (SNRs). These objects, which are well-studied in radio, optical and X-ray wavelengths, constitute one of the most intriguing source classes in VHE astronomy. H.E.S.S., an array of four imaging atmospheric Cherenkov telescopes in Namibia, has recorded an extensive dataset of VHE \\gr\\ observations covering the central region of the Milky Way, both from pointed observations as well as from the Galactic Plane Survey conducted in the inner region of the Galaxy. From radio observations, several hundred SNRs are known in the Milky Way, but until now only few of them have been identified as VHE \\gr\\ emitters. Using the H.E.S.S. dataset and a large ensemble of radio SNRs localized in the inner region of the Galaxy, the standard framework that links the origin of cosmic rays to the \\gr\\ visibility of SNRs can now be tested. Here we present the ensemble of investigated SNRs and discuss constraints on the parameter space used within a theoretical model of hadronic VHE \\gr\\ production.} ", "introduction": "SNRs are widely believed to be the main sources of Galactic Cosmic Rays (CRs) with energies of up to \\unit[$10^{15}$]{eV}. The amount of energy released during a supernova (SN) explosion, the rate of Galactic SNe and the energy density of CRs together with diffusive shock acceleration as a possible acceleration mechanism make SNRs the prime candidates for the acceleration of CRs. In radio, optical and X-ray wavelengths deep observations on SNRs have been performed and several SNRs have been discovered in the relatively young field of VHE \\gr\\ astronomy . With H.E.S.S. the shell-type morphology of individual SNRs such as RX~J1713.7-3946~\\cite{HESS:1713I}, SN~1006~\\cite{HESS:SN1006}, Vela Junior~\\cite{HESS:VelaJunior} and HESS~J1731-347~\\cite{HESS:1731} has been resolved. These discoveries support the hypothesis of particle acceleration in the shell of SNRs. Using the large dataset of H.E.S.S. accumulated during the Galactic plane survey (GPS)~\\cite{ICRC2011:Survey} in recent years we study SNRs in the inner Galactic region, ranging from 275$^\\circ$ to 60$^\\circ$ in Galactic longitude and from -3$^\\circ$ to 3$^\\circ$ in Galactic latitude. Here, the VHE \\gr\\ signal of known radio SNRs listed in Green's radio SNR catalog~\\cite{GREEN:RadioSNRs} is investigated. For each SNR within the GPS region, an upper limit on the VHE \\gr\\ flux has been calculated and can be compared to the model predictions. In this work we test the commonly used model of the hadronic \\gr\\ emission expected from SNRs presented by Drury~\\etal~\\cite{Drury:SNRVisibility}. They calculate the integral \\gr\\ flux from hadronic interactions based on intrinsic SN parameters. VHE \\g\\ rays result from the decay of neutral pions generated in proton--proton interactions in the expanding shell of the SNR. For the first time, the high sensitivity of H.E.S.S. and the large dataset of the GPS allow us to derive upper limits that are low enough to test model parameters. Using the upper limit on the VHE \\gr\\ flux represents a conservative test since an additional leptonic component might be present in the flux. ", "conclusions": "\\label{sec:paramUL} Supernova remnants are the most plausible sites of cosmic-ray acceleration. As a consequence, they are predicted to be emitters of VHE $\\gamma$ radiation by $\\pi^0$ decay during the early stage of their evolution. The good sensitivity of H.E.S.S.~and the large dataset accumulated in its Galactic Plane Survey now allow us to compare quantitative estimates for this emission to observations, and to study the SNR population in the Inner Galaxy as a whole. From Green's catalog of radio SNRs, candidates were selected whose estimated age makes them predicted sources of VHE \\g\\ radiation and the existance of a distance measurement allows for the expected flux to be calculated. Fig.~\\ref{log_p_dist} shows the distribution of the upper limits on the product $p\\equiv \\theta{}E_\\mathrm{SN}n/\\unit[10^{51}]{erg}\\,\\mathrm{cm}^{-3}$ as calculated from the H.E.S.S.~flux upper limit in the framework of the model of Drury \\etal. The vertical line marks a value of $p=0.1$ which would be expected for a supernova explosion energy of $E_\\mathrm{SN}=10^{51}\\,\\mathrm{erg}$, an ambient number density of $n=1\\,\\mathrm{cm}^{-3}$, and an acceleration efficiency of $\\theta=0.1$. The H.E.S.S.~upper limits fall within the same order of magnitude of this value. SNRs with an upper limit on the $p$ parameter above 0.1 can be divided into two groups: those with low exposure and those with detection in VHE \\g\\ rays. In the first case the H.E.S.S. coverage of the region is not deep enough to put a constraining upper limit on the SNR, whereas in the latter case VHE \\gr\\ emission has been detected, which can originate from the shell itself (e.g. G347.3-0.5, HESS~J1713-397), from a possible bright pulsar wind nebulae (PWN, e.g. G0.9+0.1, HESS~J1747-281) or can result from an overlapping bright source (e.g. G27.4+0.0, HESS~J1841-055). There are a few SNRs that appear to be underluminous, and they are interesting because, in principle, they have the potential of falsifying the model of hadronic \\gr\\ production. But the uncertainties in the input parameters to the model are large. In particular, the ambient density is usually uncertain by an order of magnitude or more and the acceleration efficiency is also not well known. In summary, the present study constitutes a first falsification test for a commonly used model of the \\gr\\ visibility of supernova remnants. Further observations and future instruments like the Cherenkov Telescope Array (CTA) with its vastly increased sensitivity and good survey capabilities will contribute to a better understanding of acceleration mechanisms of cosmic rays. {\\small \\subsubsection*{Acknowledgements} The support of the Namibian authorities and of the University of Namibia in facilitating the construction and operation of H.E.S.S. is gratefully acknowledged, as is the support by the German Ministry for Education and Research (BMBF), the Max Planck Society, the French Ministry for Research, the CNRS-IN2P3 and the Astroparticle Interdisciplinary Programme of the CNRS, the U.K. Science and Technology Facilities Council (STFC), the IPNP of the Charles University, the Polish Ministry of Science and Higher Education, the South African Department of Science and Technology and National Research Foundation, and by the University of Namibia. We appreciate the excellent work of the technical support staff in Berlin, Durham, Hamburg, Heidelberg, Palaiseau, Paris, Saclay, and in Namibia in the construction and operation of the equipment.}" }, "1112/1112.0002_arXiv.txt": { "abstract": "We present the results of a multiplicity survey of $212$ T~Tauri stars in the Chamaeleon~I and Taurus-Auriga star-forming regions, based on high-resolution spectra from the Magellan Clay $6.5$\\,m telescope. From these data, we achieved a typical radial velocity precision of $\\sim\\!80$\\,m\\,s$^{-1}$ with slower rotators yielding better precision, in general. For $174$ of these stars, we obtained multi-epoch data with sufficient time baselines to identify binaries based on radial velocity variations. We identified eight close binaries and four close triples, of which three and two, respectively, are new discoveries. The spectroscopic multiplicity fractions we find for Cha~I ($7\\%$) and Tau-Aur ($6\\%$) are similar to each other, and to the results of field star surveys in the same mass and period regime. However, unlike the results from imaging surveys, the frequency of systems with close companions in our sample is not seen to depend on primary mass. Additionally, we do not find a strong correlation between accretion and close multiplicity. This implies that close companions are not likely the main source of the accretion shut down observed in weak-lined T~Tauri stars. Our results also suggest that sufficient radial velocity precision can be achieved for at least a subset of slowly rotating young stars to search for hot Jupiter planets. ", "introduction": "\\label{sec:Introduction} Most stars, both in the solar neighborhood and in young clusters are members of binary or multiple systems. Yet, the formation and early evolution of binary and multiple stars is poorly constrained observationally, and not well understood theoretically. For instance, the fraction of wide binaries in dense star-forming regions such as the Orion Nebula Cluster and IC~348 are comparable to that of field stars \\citep{1999AA...343..831D,2006AA...458..461K} whereas the frequency of binaries is much higher among young stars in dispersed T~associations like Taurus-Auriga \\citep[e.g.,][]{1995ApJ...443..625S,1997ApJ...481..378G,2003AJ....126.2009B}; for reviews see \\citet{2000prpl.conf..703M} and \\citet{2007prpl.conf..379D}. Furthermore, high-order multiples are more common in nearby star-forming regions than for solar-type main-sequence stars in the solar neighborhood \\citep{2006AA...459..909C}. Some simulations suggest that stars usually form in triples and higher-order multiple systems only to be dispersed later, with the fraction of stars in multiple systems decreasing from $80\\%$ down to $40\\%$ by about $10$\\,Myrs \\citep{2004MNRAS.351..617D}. Predictions are that the binary fraction is higher among higher mass stars, and that brown dwarfs are never close companions to stars \\citep{2004AA...414..633G}. While past multiplicity surveys, using speckle imaging and adaptive optics on $4$-m class telescopes, have drawn attention to the ubiquity of binaries in star-forming regions \\citep[e.g.,][]{1993AJ....106.2005G,1993AA...278..129L}, their limited contrast and angular resolution have left many key questions unanswered or only partially answered. For instance, the frequency of higher-order multiples is uncertain, and so is the frequency of very low-mass stellar and sub-stellar companions \\citep{2007ApJ...671.2074A}. Multiple systems are probably common. With adaptive optics on $8$-m class telescopes, it has become straightforward to detect all stellar and even all brown dwarf companions down to the deuterium-burning limit with separations of tens of AU for nearby young stars \\citep[e.g.,][]{2008ApJ...683..844L}. Like \\citet{2008MNRAS.385.2210M}, our radial velocity study is complementary and covers the close separations. We present the results of a high-resolution spectroscopic survey of $212$ stars spanning $\\sim\\!0.2$--$3\\,M_{\\sun}$ in the nearby $\\sim\\!2$\\,Myr old star-forming regions Chamaeleon~I (hereafter Cha~I) and Taurus-Auriga (hereafter Tau-Aur). Cha~I and Tau-Aur are at distances of $\\sim\\!160$\\,pc and $\\sim\\!140$\\,pc, respectively \\citep{1997AA...327.1194W, 1994AJ....108.1872K}. Previously, we presented a study of rotation, disk, and accretion signatures for a subsample of $144$ stars showing no evidence of spectroscopic companions and broadening functions showing only a single source \\citep{2009ApJ...695.1648N}; we use the derived projected rotational velocities ($v\\,\\sin\\,i$) and accretion signatures (H$\\alpha\\,10\\%$\\,width) for this work. Furthermore, we also analyzed the variability in accretion-related emission lines for a subsample of $40$ classical T~Tauri stars \\citep{2009ApJ...694L.153N}. \\citet{2008ApJ...683..844L} presented a census of wide binaries in Cha~I that encompasses our sample while \\citet{2007ApJ...670.1337D} investigated circumstellar disks, including the effect of companions on disks, for a subsample of Cha~I targets with available near-IR data. Among the issues we address in this work is the dependence of multiplicity on primary mass, i.e., whether higher mass stars are more likely to be in binaries and multiples than their lower mass counterparts. An increase in wide binaries with increasing mass has been observed for both young stars and field dwarfs \\citep[e.g.,][]{2007ApJ...662..413K,2008ApJ...683..844L}. Furthermore, we look at how close-in multiplicity varies between different star-forming regions and the field; the multiplicity of $\\sim\\!0.1$--$2\\,M_{\\sun}$ field dwarfs in the solar neighborhood has been studied extensively \\citep[e.g.,][]{1991AA...248..485D,1992ApJ...396..178F,2010ApJS..190....1R}. We also explore whether close companions contribute to the observed difference between classical T~Tauri stars (CTTS), which are accreting, and weak-lined T~Tauri star (WTTS), which are not accreting based on weak H$\\alpha$ emission. Although the source of dichotomy between CTTS and WTTS is currently unknown, the presence of non-accreting $2$\\,Myr old stars is surprising. It has been suggested that the inner disks around weak-lined objects may have been truncated by close binary companions \\citep[e.g.,][and references therein]{2000prpl.conf..703M}. ", "conclusions": "\\label{sec:SummaryAndConcludingRemarks} Binary and higher-order multiple systems are prevalent in young clusters, and studying the statistical properties of these systems offers indirect constraints on star formation. Here, we present a spectroscopic study of mulitiplicity for young stars in the Chamaeleon~I and Taurus-Auriga star-forming regions. Our main results are as follows. \\begin{enumerate} \\item The spectroscopic multiplicity fractions for Cha~I and Tau-Aur are similar to each other, and to those of field stars in the same mass and period regime (for details, see \\S\\ref{sec:BinaryPopulations}). This finding implies that the overall fraction of short period stellar companions could stablize after initial formation. \\item Close multiplicity is not seen to depend on primary mass. The mass of the host star can become important in the separation regime of imaging surveys. For discussion, see \\S\\ref{sec:MassDependence}. \\item There is no strong correlation between accretion and the frequency of systems with close companions, and thus, close stellar companions are unlikely the principal source of the accretion cutoff observed in WTTS (see \\S\\ref{sec:BinaryPopulations}). Moreover, if close companions are responsible solely for the accretion difference between CTTSs and WTTSs then there should be a disparity in the proportion of spectroscopic binaries between these two populations, but this is not seen in our survey of Cha~I and Tau-Aur. \\end{enumerate} By undertaking this extensive spectroscopic survey of T~Tauri stars, we have gained some insight that could be beneficial to future efforts. First and foremost, we now know the main limiting factor for a radial velocity study of young stars is the strong intrinsic noise present in some objects. This noise would need to be reduced in some way in order both to detect lower mass companions more effectively, and to extend the measurement to longer orbital periods. Our radial velocity precision is sufficient to detect, at the time of this writing, a few dozen of the known hot Jupiter planets, which have radial velocity semi-amplitudes of a few $100{\\rm\\,m\\,s}^{-1}$. If such planets were to be detected around pre-main sequence stars, it would provide direct constraints on their formation and migration timescales. For young stars, one way to reduce the intrinsic noise is to observe at infrared wavelengths \\citep{2006ApJ...644L..75M,2008AA...489L...9H}. Another option is to average out the noise by observing over its characteristic period, i.e., the rotation period of the noisy star. To some extent, we have already done this by taking multiple spectra of the same targets during an observing run. Ideally, one would multiplex observations, e.g., by using multiple fibers as in VLT/FLAMES. This solution may not be suitable for star-forming regions like Taurus-Auriga, which span a large area of the sky, but it would probably be quite useful for compact regions like Chamaeleon. We have also learned that our high resolution and S/N were extremely beneficial. It was probably a major reason we could find so many SB2 candidates; being able to resolve the stellar rotation helped us to determine the stability of the broadening functions. Less clear is whether one would really need the large spectral range, or whether a few well-chosen echelle orders would suffice. This consideration is relevant since it would determine whether or not multiplexing is feasible." }, "1112/1112.0234_arXiv.txt": { "abstract": "With the discovery of massive neutron stars such as PSR J1614-2230, the question has arisen whether exotic matter such as hyperons can exist in the neutron star core. We examine the conditions under which hyperons can exist in massive neutron stars. We consistently investigate the vector meson-hyperon coupling, going from SU(6) quark model to a broader SU(3) symmetry. We propose that the maximum neutron star mass decreases linearly with the strangeness content $f_s$ of the neutron star core as $M_{max}(f_s) = M_{max}(0) - 0.6 M_{\\odot} (f_s/0.1)$, which seems to be independent of the underlying nuclear equation of state and the vector baryon-meson coupling scheme. Thus, pulsar mass measurements can be used to constrain the hyperon fraction in neutron stars. ", "introduction": "The discovery of the massive neutron star PSR J1614-2230 has raised new challenges for theories of dense matter beyond nuclear saturation density. Shapiro delay measurements from radio timing observations of the binary millisecond pulsar indicate a large mass of 1.97$\\pm$0.04$M_{\\odot}$ of the neutron star \\cite{Demorest10}. The core of a neutron star harbors a dense matter environment, which could be the site for strangeness containing matter, such as hyperons. Though nuclear interactions in the saturation regime are well understood, one has to utilize neutron star observations to find clues about the physics of cold and dense matter beyond several times saturation density. Any theory of ultra dense matter has to explain the recently observed large neutron star mass. According to existing models of dense matter, the presence of hyperons leads to a considerable softening of the equation of state (EoS), resulting in a corresponding reduction of the maximum mass of the neutron star \\cite{Glendenning92derivativecoupling,Baldo00,Vidana00,Djapo}. Then the existing theories involving hyperons are in conflict with the large pulsar masses \\cite{Lattimer10}. On including hyperons, most relativistic models obtain maximum neutron star masses in the range $1.4-1.8 M_{\\odot}$ \\cite{Glen85,GM1,Knorren,BalbergGal,PalHanauske,Hanauske00,Zschiesche,Long12}. However, in exceptional cases, neutron stars with maximum masses larger than 2$M_{\\odot}$ have been obtained, either by pushing the threshold for appearance of hyperons to higher densities, or due to strong hyperon vector repulsion \\cite{Long12,Huber,Hofmann,Rikovska,Dhiman07,Dexheimer08,Bombaci08,Cavagnioli11}. Taurines et al. \\cite{Taurines} achieved large neutron star masses including hyperons by considering a model with density dependent coupling constants, which were varied nonlinearly with the scalar field. Recently, Bednarek et al. \\cite{Bednarek2011} also achieved a stiffening of the EoS by using a non-linear relativistic mean field (RMF) model with quartic terms involving the hidden strangeness vector meson. In addition to the inclusion of such a meson into a density dependent RMF model, Lastowiecki et al. \\cite{Lastowiecki12} assumed a quark matter core in order to obtain massive stars. Bonanno and Sedrakian \\cite{Bonanno12} also modeled massive neutron stars with a hyperon and quark core using a fairly stiff EoS and vector repulsion among quarks. In several studies, the maximum neutron star masses obtained when including hyperons were not very different from those containing nucleons only \\cite{Hofmann,Rikovska,Dexheimer08}. In more sophisticated models such as the Brueckner-Hartree-Fock model, the maximum neutron star masses were generally found to be lower than $1.6M_{\\odot}$ which is in contradiction with observed pulsar masses \\cite{Baldo00,Vidana00,Djapo,Baldo98,Nishizaki,SchulzeVidana,SchulzeRijken,Logoteta12}.\\\\ From the studies cited in the previous paragraph, it seems that the possible presence of hyperons in massive neutron stars is in many cases reconciled by incorporating large vector repulsion in an ad hoc way. In contrast, we investigate the role of vector repulsion starting from symmetry arguments. Assuming SU(3) symmetry, we perform a controlled parameter study and constrain the parameters using the observed mass of PSR J1614-2230. This procedure is in line with modern microscopic models for realistic baryon-baryon potentials such as the Bonn potentials \\cite{Machleidt87} and the Nijmegen potentials \\cite{Rijken99} which adopt SU(3) symmetry to describe the baryon interactions for the baryon octet. For our investigations, we employ a RMF model, in which the parameters are calibrated around nuclear saturation density \\cite{Glen85,Knorren,Schaffner96}. However, the extrapolation of such properties to supranuclear densities presents uncertainties. In a previous paper \\cite{WsbChtt12}, we investigated how the uncertainty in nuclear saturation properties, such as effective nucleon mass or nuclear compressibility, or hypernuclear properties, such as potential depths of hyperons in nuclear matter, could influence our conclusions about the presence of hyperons in the core of massive neutron stars. In this work, we question the fundamental assumption of SU(6) symmetry, which relates the hyperon couplings to the nuclear couplings.\\\\ \\indent This paper is organized in the following way. In Sec. \\ref{model}, we describe the model to calculate the EoS. The parameters of the model are listed in Sec. \\ref{parameters}. The results of our calculations are discussed in Sec. \\ref{results}, and a summary of our conclusions is given in Sec. \\ref{summary}. \\\\ ", "conclusions": "\\label{results} \\subsection{Varying the $g_8/g_1$ Ratio $z$} We probe the effects of the $g_8/g_1$ ratio $z$ on the stiffness of the hadronic EoS. We plot the coupling constants as functions of $z$ in Fig. \\ref{figurezfree}, for the explicit formulae given in equation (\\ref{zfreecouplings}). For $z=0$, all coupling constants $g_{\\it{B\\omega}}$ are the same and similarly are all coupling constants $g_{\\it{B\\phi}}$ equal. This is due to the fact that $z=0$ corresponds to $g_8=0$, which results in the equality of the corresponding baryon-meson couplings as the baryons couple only to the flavor singlet state. With increasing $z$, i.e. with increasing contribution from the coupling to the octet $g_8$, the resulting couplings all become smaller except for $g_{\\it{\\Xi\\phi}}$. At $z=1/\\sqrt{6}\\approx0.4082$ the SU(6) case is reached where the $\\phi$ does not couple to the nucleon. Thereafter, for $z>1/\\sqrt{6}$ the coupling constant $g_{\\it{N\\phi}}$ changes its sign so that it is not a repulsive but now an attractive interaction. Note, that the $\\omega$ coupling constants for $\\Lambda$ and $\\Sigma$ hyperons are equal for all values of $z$. As anticipated in Sec. \\ref{beyondsu6}, we restrict $z$ to the interval $z\\in[0:2/\\sqrt{6}]$. \\\\ \\begin{figure} \\includegraphics[height=8.6cm,angle=270]{fig1_color} \\caption{Relative vector meson coupling constants as functions of the $g_8/g_1$ ratio $z$ for fixed $\\alpha_V=1$. The value $z=1/\\sqrt{6}$ corresponds to the SU(6) case.} \\label{figurezfree} \\end{figure} \\begin{figure} \\includegraphics[width=8.6cm]{fig2_color} \\caption{EoS for different $g_8/g_1$ ratios $z$ within a nonlinear $\\sigma-\\omega$ model with additional $\\phi$ meson and the full baryon octet for GM1 parameter set. The EoS get stiffer with decreasing $z$.} \\label{zfreeeosfigure} \\end{figure} \\indent The EoS for $z=0.1,0.2,...,0.8$ are plotted in Fig. \\ref{zfreeeosfigure}. At first glance it becomes clear that the EoS stiffens with decreasing $z$. This can be explained with the help of Fig. \\ref{figurezfree} where we had plotted the $z$ dependence of the vector meson coupling constants. We noticed that with increasing $z$, all couplings except $g_{\\it{\\Xi\\phi}}$ decrease. Since the $\\Xi$ hyperons only play a subordinate role compared to the neutrons, the increase of $g_{\\it{\\Xi\\phi}}$ does not prevent that part of the overall interaction between the baryons, which is mediated by the vector mesons, to become less repulsive. Therefore, the EoS must soften with increasing $z$. Together with further decreasing coupling strengths between the vector mesons and the other baryons (except again $g_{\\it{\\Xi\\phi}}$) the EoS becomes even softer until $z=2/\\sqrt{6}$ is reached. \\\\ \\begin{figure} \\includegraphics[width=08.6cm]{fig3_color} \\caption{Mass-radius relations for the EoS displayed in Fig. \\ref{zfreeeosfigure}. The maximum mass is obtained for the case $z=0$, where all baryons couple to the vector mesons with equal strengths.} \\label{zfreemrfigure} \\end{figure} \\indent In Fig. \\ref{zfreemrfigure} we plot the mass-radius relations for the various EoS we just discussed in the context of Fig. \\ref{zfreeeosfigure}. As expected from the influence of $z$ on the stiffness of the EoS, the lowest maximum mass is obtained for $z=2/\\sqrt{6}\\approx0.8165$, or in the case of Fig. \\ref{zfreemrfigure} at $z=0.8$, namely $M=1.49M_{\\odot}$. The maximum mass grows up to the value $M=2.36 M_{\\odot}$ for $z=0$. We notice that the maximum mass of a neutron star in our RMF model reacts rather strongly to the variation of $z$: over the whole $z$ range the change in the maximum mass is $\\Delta_M$=0.87$M_{\\odot}$.\\\\ \\indent After varying $z$ in rather big steps, we now plot in Fig. \\ref{zfreedifferentmodelsfigure} the maximum neutron star mass as a continuous function of $z$ for the models GM1, NL3 and TM1 and for nuclear matter as well as for baryonic matter. \\begin{figure} \\includegraphics[width=8.6cm]{fig4_color} \\caption{Maximum masses as functions of the $g_8/g_1$ ratio $z$ for NL3, GM1 and TM1 parameter sets. For each parameterization the case of pure nucleonic matter is also displayed.} \\label{zfreedifferentmodelsfigure} \\end{figure} As already analyzed in the discussion of Fig. \\ref{figurezfree}, we see that the branch for nucleonic matter is insensitive to the changes in $z$ for the NL3 and GM1 models. For TM1, the quartic self interaction term in the Lagrangian spoils this property since we had to keep the corresponding coupling constant $d$ from the SU(6) value of $z$ fixed for the whole $z$ range: the maximum masses for pure nucleonic stars therefore depend on the actual $z$ value, showing a minimum for the SU(6) case $z\\approx0.4082$, namely $M=2.18M_{\\odot}$. \\\\ \\begin{figure} \\includegraphics[width=8.6cm]{fig5_color} \\caption{Particle fractions for the GM1 parameter set for 3 different values of $z$: $z=0.8$ (a), $z=0.408$ (b) corresponding to the SU(6) case, and $z=0$ (c). The threshold for the appearance of the hyperons $\\Lambda$, $\\Xi^-$ and $\\Xi^0$ is pushed to higher densities with decreasing $z$.} \\label{GM1frac} \\end{figure} Considering the neutron stars containing hyperons, we see in Fig. \\ref{zfreedifferentmodelsfigure} that the maximum masses depend on $z$ as already observed in Fig. \\ref{zfreemrfigure}: for the largest $z$ values the maximum masses are the smallest and continually grow with decreasing $z$. It is interesting to see that towards $z=0$ the maximum masses of these stars seem to approach the maximum masses of the corresponding pure nucleonic stars for all parameter sets studied. A look at the particle number fractions for the GM1 parameter set at $z=0.8$ (which we plot in Fig. \\ref{GM1frac}(a)) shows that the first hyperons to appear in the hadronic matter are the $\\Xi^-$ and the $\\Lambda$ at total baryon number densities of $n_b\\approx$ 0.28 ${fm}^{-3}$ and $n_b\\approx$ 0.29 ${fm}^{-3}$ respectively, while the $\\Xi^0$ appears much later at $n_b\\approx$ 0.76 ${fm}^{-3}$. On increasing $z$ to its SU(6) value (Fig. \\ref{GM1frac}(b)), the $\\Lambda$ hyperon appears first at $n_b\\approx$ 0.36 ${fm}^{-3}$, followed by $\\Xi^-$ at $n_b\\approx$ 0.4 ${fm}^{-3}$ and $\\Xi^0$ at $n_b\\approx$ 0.89 ${fm}^{-3}$. At $z=0$ (Fig. \\ref{GM1frac}(c)), the threshold of appearance of hyperons is pushed to even higher densities: $n_b\\approx$ 0.73 ${fm}^{-3}$ for $\\Lambda$, $n_b\\approx$ 0.74 ${fm}^{-3}$ for $\\Xi^-$ and $n_b\\approx$ 1.38 ${fm}^{-3}$ for $\\Xi^0$. Thus, for $z=0$ the neutron stars consist mainly of nuclear matter which is why the maximum masses are so close to those of the pure nucleonic stars. For the case of the parameter sets NL3 and TM1, the particle fractions are qualitatively the same as in the GM1 case. In the case of NL3 parameterization, a well-known instability occurs at high densities when the effective nucleon mass becomes zero \\cite{Schaffner96}. The critical density for the appearance of the instability depends on the value of the hyperon coupling constants. However for the present investigation, this instability plays no role as it appears beyond the maximum densities reached in the neutron star interior. \\subsubsection{Combining $m_N^*$ and $z$ variations} The impact of z on the maximum mass of neutron stars is as comparably large as the influence of the effective nucleon mass at saturation $m_N^*$ as investigated in our previous study \\cite{WsbChtt12}. We therefore combine both parameters in a single plot, Fig. \\ref{figuremeffz}, where we show the maximum neutron star mass as a function of $m_N^*/m_N$ for different $z$ values. The incompressibility in this case is fixed to $K=240$ MeV, but the exact value is irrelevant as shown in our previous paper \\cite{WsbChtt12}. We see in Fig. \\ref{figuremeffz} that the effective mass has basically the same effect for all $z$ values and $z$ the same effect for all effective masses: for fixed $z$, the maximum masses decrease drastically for increasing effective mass. For low $z$ values, where the EoS is stiffer than for higher $z$ values, the dependence of the maximum mass on the effective mass is slightly larger: the difference along the whole range of $m_N^*/m_N$ is $\\approx$ 0.6${M}_\\odot$ for $z=0.8$ while for $z=0$ it is $\\approx$ 0.9${M}_\\odot$. The influence of $z$ on the maximum masses is slightly more pronounced than that of the effective masses: the difference in the maximum mass between $z=0$ and $z=0.8$ is $\\approx$ 0.65${M}_\\odot$ for $m_N^*/m_N=0.8$ and about 1${M}_{\\odot}$ for $m_N^*/m_N=0.55$. For comparison, we also plot the maximum masses of purely nucleonic neutron stars, and also mark in the figure the points corresponding to the SU(6) case, for several other RMF sets fitted to properties of nuclei e.g., TM1, NL3 or NL-SH \\cite{NL3,Sugahara94,NL-Z,PL-Z,NL-SH}. In Fig. \\ref{zfreedifferentmodelsfigure} the masses of baryonic stars were found to approach the limit of purely nucleonic stars for decreasing values of z. With Fig. \\ref{figuremeffz} we can now visualize that the maximum masses in these two cases differ slightly ($\\Delta_M<0.01M_\\odot$).\\\\ \\indent We close this section by concluding that a maximum mass of at least 1.97 $\\pm$ 0.04${M}_{\\odot}$ requires very small values of $z$ for large effective masses ($z\\leq0.1$ at $m_N^*/m_N=0.8$), $z$ values around SU(6) for effective masses close to $m_N^*/m_N \\approx 0.7$ and very low effective masses for large $z$ values ($m_N^*/m_N < 0.58$ for $z=0.7$). We note, that $z$ values close to the maximum of $z=2/\\sqrt{6}$ are now allowed configurations within the plot range: the maximally allowed $z$ value for the investigated model is $z < 0.77$ at $m_N^*/m_N=0.55$. \\begin{figure} \\includegraphics[width=8.6cm]{fig6_color} \\caption{Maximum masses of hyperonic neutron stars as functions of effective nucleon mass $m_N^*$ for different values of the $g_8/g_1$ ratio $z$. For comparison, a line for nucleonic stars and points to mark RMF sets (e.g. TM1, NL3) corresponding to the SU(6) case are also given.} \\label{figuremeffz} \\end{figure} \\subsection{Varying the F/(F+D) ratio $\\alpha_V$} After systematically investigating a wider $z$ range we repeat the formalism for the $\\alpha_V$-ratio in the present section. The $F/(F+D)$ ratio is by definition restricted to the interval $\\alpha_V\\in[0;1]$, where the lower bound corresponds to a pure D-type coupling and the upper limit (i.e. the SU(6) value) corresponds to a pure F-type coupling. Analogous to the case studied above, we adopt ideal mixing as well as a $g_8/g_1$ ratio fixed to its SU(6) value $z=1/\\sqrt{6}$ and we allow for the $\\phi$ meson to couple to the nucleon. We plot the coupling strengths in Fig. \\ref{figurealphafreeB} where we vary $\\alpha_V$ between 0 and 1. Note that the coupling constants $g_{\\omega \\Lambda}$ for $\\Lambda$ hyperons do not change considerably and that all vector couplings remain repulsive (do not change their sign).\\\\ \\begin{figure} \\includegraphics[height=8.6cm,angle=270]{fig7_color} \\caption{Vector meson coupling constants as functions of the $F/(F+D)$ ratio $\\alpha_V$. The $g_8/g_1$ ratio is fixed to its SU(6) value $z=1/\\sqrt{6}$ and ideal mixing is assumed. The SU(6) case is given by $\\alpha_V=1$.} \\label{figurealphafreeB} \\end{figure} \\indent The ratio of the F- and D-type couplings can be continuously varied between the two extremes of a pure F- and a pure D-type coupling. We see that from $\\alpha_V=1$ down to $\\alpha_V=0$ all couplings become stronger except for $g_{\\it{\\Sigma\\phi}}/g_{\\it{N\\omega}}$. Since the $\\Sigma$ hyperons have but very little influence on the EoS up to neutron star densities, we can expect that the EoS become stiffer with decreasing $\\alpha_V$. This is exactly what we find in Fig. \\ref{figurealphaeosB} where we plot the EoS for several values of $\\alpha_V$ using \"model $\\sigma \\omega \\rho \\phi$\" and GM1 parameter set.\\\\ \\begin{figure} \\includegraphics[width=8.6cm]{fig8_color} \\caption{EoS for ``model $\\sigma \\omega \\rho \\phi$`` in GM1 parameterization for different values of $\\alpha_V$. $z$ is fixed to its SU(6) value $z=1/\\sqrt{6}$ and ideal mixing is assumed. The EoS become stiffer with decreasing $\\alpha_V$.} \\label{figurealphaeosB} \\end{figure} \\indent The stiffness depends monotonously on $\\alpha_V$ and we get the softest EoS for the SU(6) case $\\alpha_V=1$, i.e. a pure F-type coupling, while the stiffest EoS is obtained for $\\alpha_V=0$ which corresponds to the pure D-type coupling of the baryon and meson multiplets. The EoS for $\\alpha_V\\leq0.2$ appear to be indistinguishable at neutron star densities. This is evident in Fig. \\ref{figurealphamrB}, where we plot the mass-radius relations corresponding to the EoS from Fig. \\ref{figurealphaeosB}: for the values $\\alpha_V=0.0-0.2$ the maximum masses but also the radii of the corresponding stars coincide (M$_{\\it{max}}=2.36$M$_\\odot$, R$=11.8$ km). We note, that this value of the maximum mass is also obtained for the purely nucleonic case (compare e.g. Fig. \\ref{zfreedifferentmodelsfigure}). Thus, for the \"model $\\sigma \\omega \\rho \\phi$\" the nuclear matter limit is reached below $\\alpha_V < 0.2$ in the case of GM1, GM3 and NL3 parameter sets, while for the very stiff PL-Z EoS (having effective mass $m^*_N/m_N \\simeq 0.55$) \\cite{PL-Z} pure nucleonic stars are already obtained for $\\alpha_V < 0.3$. \\\\ \\begin{figure} \\includegraphics[width=8.6cm]{fig9_color} \\caption{Mass-radius relations obtained from the EoS in Fig. \\ref{figurealphaeosB}.} \\label{figurealphamrB} \\end{figure} \\indent In this way, for hyperonic stars the limit of nucleonic stars is continuously approached for decreasing values of the $g_8/g_1$ ratio $z$, or for decreasing values of the $F/(F+D)$ ratio $\\alpha_V$ away from the SU(6) value, respectively. \\\\ \\\\ \\begin{figure} \\includegraphics[width=8.6cm]{fig10_color} \\caption{Maximum masses of neutron stars as functions of strangeness fraction $f_s$ in the neutron star core for four different EoS. The maximum mass decreases linearly with the strangeness fraction approximately as $M_{max}(f_s) = M_{max}(0) - 0.6 M_{\\odot} (f_s/0.1)$.} \\label{mmax_fs} \\end{figure} \\indent To generalize our findings, we plot in Fig. \\ref{mmax_fs} the maximum masses of neutron stars as a function of strangeness fraction $f_s$ (the number of strange quarks divided by the total number of quarks) for four different EoS: for $m_N^*/m_N$ = 0.55, 0.75 and for the NL3 ($m^*_N/m_N = 0.6$) and GM1 ($m^*_N/m_N = 0.7$) parameter sets, by varying $\\alpha_V$ (solid lines) and $z$ (dashed lines). It is evident from the figure that on decreasing $\\alpha_V$ or $z$, the strangeness fraction in the core decreases, and there is a corresponding increase in the maximum mass of the star. At zero strangeness fraction the nucleonic limit is reached, and this corresponds to the highest value of the maximum mass. For $m^*_N/m_N \\lessapprox 0.7$, the relation between the maximum mass of the star and its strangeness fraction can be fitted linearly according to the formula: \\begin{equation} \\frac{M_{max}}{M_{\\odot}} = \\frac{M_{max} (f_s = 0)}{M_{\\odot}} - c \\left(\\frac{f_s}{0.1}\\right)~, \\end{equation} where $c\\approx0.6M_{\\odot}$. This has interesting consequences when we use these results to predict the maximally allowed strangeness fraction in maximum mass neutron stars. In an associated study \\cite{TolosSagert}, we applied the results from measurements of sub-threshold kaon production in heavy-ion collisions (HIC) to study the implications on neutron star properties. It was found that the heavy-ion data and causality imply a firm upper limit on maximum mass of compact stars of 3 solar masses. Substituting this value into the derived formula above, we can show that the strangeness fraction in a 2 $M_{\\odot}$ star cannot be more than: \\begin{eqnarray} f_s^{max} &=& \\frac{M_{max}^{HIC} - M_{max}^{obs}}{6 M_{\\odot}} \\nonumber\\\\ &=& \\frac{3 M_{\\odot} - 2 M_{\\odot}}{6 M_{\\odot}} = 0.17 ~. \\end{eqnarray} \\indent We also point out that for the EoS compatible with the observed mass limit, the hyperon fraction in a canonical $1.4M_{\\odot}$ star is zero. Only for small effective nucleon masses and values of $z$ above the SU(6) value can a hyperon fraction of less than $0.5\\%$ be reached.\\\\ \\indent We have now found that we can reach the nuclear matter limit starting from a RMF model including hyperons and continuously changing the model parameters, instead of making a discrete ``on/off'' decision about whether to include or exclude hyperons. Instead, we should take the position of saying that one can only exclude a RMF parameter set as soon as the corresponding maximum mass for nucleonic stars is incompatible with observations. As long as the observational mass limit is below the nucleonic mass limit of the model, it is possible to have hyperons in the core of maximum mass neutron stars." }, "1112/1112.1697_arXiv.txt": { "abstract": "We present continued radio observations of the tidal disruption event Swift\\,J164449.3+573451 extending to $\\delta t\\approx 216$ days after discovery. The data were obtained with the EVLA, AMI Large Array, CARMA, the SMA, and the VLBA+Effelsberg as part of a long-term program to monitor the expansion and energy scale of the relativistic outflow, and to trace the parsec-scale environment around a previously-dormant supermassive black hole (SMBH). The new observations reveal a significant change in the radio evolution starting at $\\delta t\\approx 1$ month, with a brightening at all frequencies that requires an increase in the energy by about an order of magnitude, and an overall density profile around the SMBH of $\\rho\\propto r^{-3/2}$ ($0.1-1.2$ pc) with a significant flattening at $r\\approx 0.4-0.6$ pc. The increase in energy cannot be explained with continuous injection from an $L\\propto t^{-5/3}$ tail, which is observed in the X-rays. Instead, we conclude that the relativistic jet was launched with a wide range of Lorentz factors, obeying $E(>\\Gamma_j)\\propto \\Gamma_j^{-2.5}$. The similar ratio of duration to dynamical timescale for \\sw\\ and GRBs suggests that this result may be applicable to GRB jets as well. The radial density profile may be indicative of Bondi accretion, with the inferred flattening at $r\\sim 0.5$ pc in good agreement with the Bondi radius for a $\\sim {\\rm few}\\times 10^6$ M$_\\odot$ black hole. The density at $\\sim 0.5$ pc is about a factor of 30 times lower than inferred for the Milky Way galactic center, potentially due to a smaller number of mass-shedding massive stars. From our latest observations ($\\delta t\\approx 216$ d) we find that the jet energy is $E_{\\rm j,iso}\\approx 5\\times 10^{53}$ erg ($E_j\\approx 2.4\\times 10^{51}$ erg for $\\theta_j=0.1$), the radius is $r\\approx 1.2$ pc, the Lorentz factor is $\\Gamma_j\\approx 2.2$, the ambient density is $n\\approx 0.2$ cm$^{-3}$, and the projected angular size is $r_{\\rm proj}\\approx 25$ $\\mu$as, below the resolution of the VLBA+Effelsberg. Assuming no future changes in the observed evolution and a final integrated total energy of $E_j\\approx 10^{52}$ erg, we predict that the radio emission from \\sw\\ should be detectable with the EVLA for several decades, and will be resolvable with VLBI in a few years. ", "introduction": "\\label{sec:into} The discovery of the unusual $\\gamma$-ray/X-ray transient Swift\\,J164449.3+573451 (hereafter, \\sw), which coincided with the nucleus of an inactive galaxy at $z=0.354$, has opened a new window into high-energy transient phenomena, with potential implications to our understanding of relativistic outflows in systems such as gamma-ray bursts (GRBs) and active galactic nuclei (AGN). The prevailing interpretation for this event is the tidal disruption of a star by a dormant supermassive black hole (SMBH) with a mass of $M_{\\rm BH}\\sim 10^6-10^7$ M$_\\odot$ (\\citealt{bgm+11,bkg+11,ltc+11,zbs+11}; but see \\citealt{kp11,osj11,qk11} for alternative explanations). The argument for a tidal disruption origin is based on: (i) a positional coincidence ($\\lesssim 0.2$ kpc) with the host galaxy nucleus; (ii) rapid time variability in $\\gamma$-rays and X-rays ($\\lesssim 10^2$ s), which requires a compact source of $\\lesssim 0.15$ AU, a few times the Schwarzschild radius of a $\\sim 10^6$ M$_\\odot$ black hole; (iii) high $\\gamma$-ray and X-ray luminosity of $\\sim 10^{47}$ erg s$^{-1}$, which exceeds the Eddington limit of a $\\sim 10^6$ M$_\\odot$ black hole by $2-3$ orders of magnitude; (iv) a lack of previous radio to $\\gamma$-ray activity from this source to much deeper limits than the observed outburst, pointing to a rapid onset; and (v) long-term X-ray luminosity evolution following $L_X\\propto t^{-5/3}$, as expected from the fallback of tidally disrupted material (e.g., \\citealt{ree88,sq09}). Equally important, \\sw\\ was accompanied by bright radio synchrotron emission, with an initial peak in the millimeter band ($F_\\nu\\approx 35$ mJy) and a steep spectral slope at lower frequencies indicative of self-absorption (\\citealt{zbs+11}; hereafter, ZBS11). The properties of the radio emission established the existence of a relativistic outflow with a Lorentz factor of $\\Gamma\\sim {\\rm few}$ (ZBS11, \\citealt{bgm+11}). The spectral energy distribution also demonstrated that the lack of detected optical variability required significant rest-frame extinction ($A_V\\gtrsim 5$ mag; ZBS11, \\citealt{ltc+11}), and that the X-rays were produced by a distinct emission component, rather than inverse Compton scattering by the radio-emitting relativistic electrons (ZBS11). Finally, the evolution of the radio emission on a timescale of $\\delta t\\sim 5-22$ d pointed to an ambient density with a radial profile of roughly $\\rho\\propto r^{-2}$, as well as a mild increase in the energy of the outflow (ZBS11). The formation of a relativistic jet with dominant X-ray and radio emission were not predicted in standard tidal disruption models (e.g., \\citealt{ree88,sq09}), which instead focused on the thermal optical/UV emission from the long-term accretion of the stellar debris. A signature of the latter process is a mass accretion rate that evolves as $\\dot{M}\\propto t^{-5/3}$, presumably leading to emission with the same temporal dependence (e.g., \\citealt{kg99,gbm+08,vfg+11}). Shortly before the discovery of \\sw, \\citet{gm11} investigated the potential signature of a putative relativistic outflow, and concluded that the interaction of the outflow with the ambient medium will lead to radio emission on a timescale of $\\delta t\\sim 1$ yr (for typical off-axis observers). While the mechanism for the radio emission from \\sw\\ is interaction with an external medium, the actual light curves differ from the off-axis prediction. To address this issue, in a follow-up paper \\citet{mgm11} (hereafter, MGM11) reconsidered the model for a relativistic jet interacting with an ambient medium. They draw on the inferences from the early radio emission described in ZBS11 to infer the properties of the environment and the jet kinetic energy, and use this information to predict the future evolution of the radio emission. This long-term radio evolution is of great interest because it can provide several critical insights: \\begin{itemize} \\item The integrated energy release in the relativistic outflow, including the anticipated injection from on-going accretion. \\item The density profile around a previously-dormant SMBH on $\\sim 0.1-10$ pc scales, which cannot be otherwise probed in AGN. \\item The potential to spatially resolve the outflow with very long baseline interferometry (VLBI), and hence to measure the dynamical evolution (expansion and potentially spreading) of a relativistic jet. \\item Predictions for the radio emission from tidal disruption jets as viewed by off-axis observers on timescales of months to years to decades. \\end{itemize} The energy scale and jet dynamics are of particular importance since the total energy input and the structure of the jet may also have implications for relativistic jets in GRBs and AGN. The ability to trace the environment on parsec scales provides a unique probe of gas inflow or outflow around an inactive SMBH on scales that cannot be probed outside of the Milky Way. Finally, the long-term radio emission from \\sw\\ will inform future radio searches for tidal disruption events (TDEs) that can overcome the low detection rate in $\\gamma$-rays/X-rays (due to beaming), and obscuration due to extinction in the optical/UV (as in the case of \\sw). To extract these critical properties we are undertaking long-term monitoring of the radio emission from \\sw\\ using a wide range of centimeter- and millimeter-band facilities. Here we present radio observations of \\sw\\ that extend to $\\delta t\\approx 216$ d, and use these observations to determine the evolution of the total energy and ambient density. We find that the evolution of both quantities deviates from the behavior at $\\delta t\\lesssim 1$ month (presented in ZBS11), thereby providing crucial insight into the structure of the relativistic outflow and the ambient medium. This paper is the first in a series that will investigate the long-term radio evolution of \\sw\\ and the implications for relativistic jets and parsec-scale environments around supermassive black holes, including efforts to resolve the source with VLBI and to measure polarization. The current paper is organized as follows. We describe the radio observations in \\S\\ref{sec:obs}, and summarize the radio evolution at $\\delta t\\approx 5-216$ d in \\S\\ref{sec:prem}. In \\S\\ref{sec:model} we present our modeling of the radio emission, which utilizes the formulation of MGM11. The implications for the energy scale and ambient density are discussed in \\S\\ref{sec:energy} and \\S\\ref{sec:density}, respectively, and we finally consider the implications for relativistic jets and the parsec-scale environments of SMBHs in \\S\\ref{sec:implic}. ", "conclusions": "\\label{sec:conc} We presented radio observations of \\sw\\ extending to $\\delta t\\approx 216$ d and spanning a wide range of frequencies. The evolution of the radio emission changes dramatically at $\\delta t \\gtrsim 1$ month, requiring an increase in the total energy by about an order of magnitude, a density profile of $\\rho\\propto r^{-3/2}$ ($0.1-1.2$ pc), and a flattening at $r\\approx 0.4-0.6$ pc. A comparison of the model to optical limits and near-IR detections indicates a cooling break at $\\nu_c\\sim 10^{13}$ Hz and host galaxy extinction of $A_{\\rm V,host} \\gtrsim 3.5$ mag. The increase in energy cannot be explained by injection from an $L\\propto t^{-5/3}$ tail that is expected in tidal disruption events and which matches the evolution of the X-ray emission. We conclude that a natural explanation is a structured outflow with $E(>\\Gamma_j)\\propto \\Gamma_j^{-2.5}$. The inferred density profile and the radial scale of the density enhancement are in rough agreement with the expectation for Bondi accretion from a circumnuclear medium. The jet energetics and structure, as well as the detailed density profile on $\\sim 0.1-1$ pc scale are a testament to the important insight that can be gained from continued radio observations of \\sw. In particular, the radial density profile is traced in greater detail than even the inner parsec of the Milky Way. Continued radio observations will probe the environment to a scale of $\\sim 10$ pc in the coming decade. Using the results of our analysis we can predict the future evolution of the radio emission (modulo any future unpredictable changes in energy and/or density as we have found here). We use the evolution of $L_{\\rm j,iso}$ and $n_{\\rm CNM}$ as inferred from the data at $\\delta t\\lesssim 216$ d, and assume that the density will continue to evolve as $\\rho\\propto r^{-1.5}$ and that the energy will increase to a maximum beaming-corrected value of $E_j=L_{\\rm j,iso}t_j[1-{\\rm cos} (\\theta_j)]$ with $E_j=10^{52},\\,3\\times 10^{52},\\,10^{53}$ erg. The resulting light curves at 6 and 22 GHz are shown in Figure~\\ref{fig:predlc}. The long-term evolution is marked by a break when $E_j$ achieves its maximum value, corresponding to about $5$, $28$, and $180$ yr for our three choices of maximum energy. Using the $5\\sigma$ sensitivity of the EVLA in an observation of a few hours\\footnotemark\\footnotetext{At 22 GHz we use the sensitivity for the full 8 GHz bandwidth that will become available some time in 2012.}, we find that the emission at 22 and 6 GHz should be detectable for at least $\\sim 40$ yr and $\\sim 80$ yr, respectively. Indeed, any significant upgrades to the EVLA or the construction of more sensitive radio facilities in the coming decades may extend the range of detectability to centuries\\footnotemark\\footnotetext{Significant budget cuts to radio facilities in the future may lead to the opposite effect.}. The same is true if the total energy scale is $\\sim 10^{53}$ erg. An equally important question is whether the jet will be resolvable with VLBI in the future. The projected radius is $r_{\\rm proj}\\approx r\\theta_j$, as long as the jet maintains its collimation. In Figure~\\ref{fig:predrad} we plot the predicted future evolution of $r$ using the prescription described above. We find that for $\\theta_j=0.1$ and a best-case VLBI angular resolution of $\\approx 0.2$ mas (FWHM), the source should become resolvable at $\\delta t\\approx 6$ yr. On this timescale the 22 GHz flux density is expected to be only $\\approx 2$ mJy (Figure~\\ref{fig:predlc}), still accessible with VLBI. While the flux density at 6 GHz is expected to be larger by about a factor of $2.6$, the angular resolution at this frequency is poorer by about a factor of 3.7, making it less competitive than 22 GHz. Thus, we conclude that the radio emission from \\sw\\ may be marginally resolved in a few years. On the other hand, if the jet undergoes significant spreading on the timescale at which it becomes non-relativistic (as expected for GRB jets: e.g., \\citealt{lw00}) it is possible that it will become resolvable at $\\delta t\\sim 1-2$ yr when the expected 22 GHz flux density is still $\\sim 10$ mJy. We are undertaking continued multi-frequency radio monitoring of \\sw\\ to follow the long-term evolution of the relativistic outflow and the radial profile of the ambient medium. Even in the absence of any future dramatic changes relative to the current evolution, we expect that in the next few years we may be able to determine the total energy of the relativistic outflow, measure the spreading of the jet, and study the radial density profile to a scale of $\\sim 10$ pc. Future papers in this series will detail these results." }, "1112/1112.2668.txt": { "abstract": "V\\,393 Scorpii is a bright Galactic Double Periodic Variable showing a long photometric cycle of $\\approx$ 253 days. We present new $VIJK$ photometric time series for V\\,393 Scorpii along with the analysis of ASAS $V$-band photometry. We disentangled all light curves into the orbital and long cycle components. The ASAS $V$-band {\\it orbital} light curve was modeled with two stellar components plus a circumprimary optically thick disc assuming a semidetached configuration. We present the results of this calculation, giving physical parameters for the stars and the disc, along with general system dimensions. Our results are in close agreement with those previously found by Mennickent et al. (2010) from IR spectroscopy and the modeling of the spectral energy distribution. The stability of the orbital light curve suggests that the stellar $+$ disc configuration remains stable during the long cycle. Therefore, the long cycle should be produced by an additional variable and not-eclipsed emitting structure. We discuss the evolutionary stage of the system finding the best match with one of the evolutionary models of van Rensbergen et al. (2008). According to these models, the system is found to be after an episode of fast mass exchange that transferred 4 M$_{\\odot}$ from the donor to the gainer in a period of 400.000 years. We argue that a significant fraction of this mass has not been accreted by the gainer but remains in an optically thick {\\it massive} ($\\sim$ 2 M$_{\\odot}$) disc-like surrounding pseudo-photosphere whose luminosity is not driven by viscosity but probably by reprocessed stellar radiation. Finally, we provide the result of our search for Galactic Double Periodic Variables and briefly discuss the outliers $\\beta$ Lyr and $RX$ Cas. ", "introduction": "V\\,393 Scorpii is one of the Galactic Double Periodic Variables (DPVs), a group of interacting binaries showing a long photometric cycle lasting roughly 33 times the orbital period (Mennickent et al. 2003, Mennickent \\& Ko{\\l}aczkowski 2009, Michalska et al. 2009, Poleski et al. 2010). DPVs have been interpreted as semi-detached interacting binaries with ongoing cyclic episodes of mass loss into the interstellar medium (Mennickent et al. 2008, Mennickent \\& Ko{\\l}aczkowski 2010). The 253-d long photometric cycle of V\\,393\\,Scorpii was discovered by Pilecki \\& Szczygiel (2007) after inspection of the ASAS catalogue for eclipsing binaries with additional variability. The IUE-UV properties of V\\,393 Scorpii were studied by Peters (2001) who found evidence for a hot temperature region produced by the tangential impact of the gas stream into the gainer photosphere. This region should be the origin of the superionized lines observed in the UV, like N\\,V, C\\,IV and Si\\,IV, that are likely produced by resonance scattering in a plasma of temperature T $\\sim$ 10$^{5}$ K and electron density $N_{e} \\sim$ 10$^{9}$ cm$^{-3}$ (Peters \\& Polidan 1984). The star was also studied by means of multi-epoch high-resolution IR spectroscopy by Mennickent et al. (2010, hereafter M10), who also studied broad-band photometry and IUE ultraviolet spectra. After summarizing the available literature of this object, these authors argued for a semidetached B3 + F0 binary with masses 8 M$_{\\odot}$ and 2 M$_{\\odot}$ for the gainer and donor (hereafter also called primary and secondary, respectively) and orbital separation of 35 $R_{\\odot}$. Most remarkably, M10 found evidence for large mass loss through the Lagrangian L3 point during epochs of long cycle minimum and claim that their observations suggest that the mass loss producing the long cycle is probably concentrated in equatorial regions. %None detailed spectroscopic study of the long cycle of V\\,393\\,Sco has been still published. In this paper we refine stellar and system parameters of V\\,393 Scorpii by fitting the light curve with a multicomponent model including a stationary circumprimary disc. %and present new insights on the nature of the long cycle based on a multi-site and multi-epoch high-resolution optical spectroscopic study. A detailed report of the observations used in this paper is given in Section 2, our results are presented in Section 3, a detailed discussion of these results is given in Section 4 and our conclusions are presented in Section 5. %Preliminary reports of this research were recently published in conference proceedings (Mennickent et al. 2010x, 2010x). %showing larger H$\\alpha$ emission during the long cycle maximum and a modulation of the radial velocity and full width of half maximum (FWHM) of the He\\,I\\,5875 line during the long cycle. ", "conclusions": "In this paper we have modeled the {\\it orbital} light curve of the intermediate-mass interacting binary \\var to obtain stellar and system parameters. We also disentangled the long-term light curve at optical and infrared photometric bands. We have found insights on the system evolutionary stage and long cycle nature. The main results of our research are:\\\\ \\begin{itemize} \\item The orbital period change, if present, is shorter than 0.5 sec per year. % $\\dot{P_{o}} < $ 0.5 sec yr$^{-1}$. \\item The long-term light curves are characterized by a smooth oscillation in a time scale of 253 days with larger amplitudes in redder bandpasses. \\item The best fit to the orbital light curve requires a non-stellar component that was modeled with an optically thick disc model. The disc radius is about half of the Roche lobe radius of the gainer and two bright spots are required to fit the observations. \\item We found the stellar and system parameters that best match the observations, which are given in Table 2 along with parameters for the disc and the bright spots. \\item The stability of the orbital light curve suggests that the stellar $+$ disc configuration remains stable during the long cycle. Variability of the optically thick disc is not the main source for long cycle. \\item Therefore and in order to fit the redder color at long maximum, we argue that the long cycle is produced by free-free emission in a variable structure, probably visible perpendicular to the orbital plane, something reminiscent of the jets found in $\\beta$ Lyr (Harmanec et al. 1996). The suggestion of equatorial mass loss as the cause of the long cycle by M10 is probably biased by detection of equatorial outflows not necessarily related to the long cycle. \\item A comparison with published evolutionary tracks provides an estimate for the age of the system, namely log t = 7 $\\times$ 10$^{7}$ yr. We find the system after a mass exchange episode, where 4 M$_{\\odot}$ were transferred from the donor to the gainer in a period of 400.000 years. \\item The evolutionary model with initial stellar masses of 6 and 3.6 M$_{\\odot}$ reproduces relatively well the present donor parameters and orbital period, but overestimates the gainer temperature and luminosity, a fact that could be ascribed to the optically thick disc not considered in the evolutionary tracks. In order to explain these features, we argue that part of the mass, maybe up to 2 M$_{\\odot}$ has not been accreted by the gainer, but remains in the {\\it massive} optically-thick disc. \\item We find a discrepancy between values of $\\dot{M}$ derived from theoretical model fitting of observationally-derived parameters and those derived from accretion-theory analysis of observationally-derived parameters, the former being smaller by 2 orders of magnitude. If the optically thick disc is a representation of a massive disc-like pseudo-photosphere, then its luminosity could not be accretion-driven. The constancy of the orbital period supports the view that the disk luminosity is not driven by viscosity, but probably by reprocessed stellar radiation. %Alternatively, V\\,393\\,Sco could be at the end of the first mass transfer burst or starting the second one (i.e. 3\\% younger and 7\\% older, respectively). These models fit $\\dot{M}$ obtained from accretion-disc theory, but their $\\chi^{2}$ are larger by a factor of {\\bf 10} than the best fit model. For both models, our result of a massive disc-like pseudo-photosphere remains unaltered. \\item We present the results of our search for Galactic DPVs and make a comparison with hot emission-line binaries with cyclic long-term brightness changes. 13 Galactic DPVs show a similar correlation between $P_{o}$ and $P_{long}$ to that observed in LMC and SMC DPVs. The systems $\\beta$ Lyr and RX Cas deviate from this tendency. These systems could be in an earlier evolutionary stage compared with DPVs. \\end{itemize}" }, "1112/1112.4598_arXiv.txt": { "abstract": "The existence of 10$^9$ M$_{\\odot}$ black holes (BH) in massive galaxies by $z \\ sim$ 7 is one of the great unsolved mysteries in cosmological structure formation. One theory argues that they originate from the black holes of Pop III stars at $z \\sim$ 20 and then accrete at the Eddington limit down to the epoch of reionization, which requires that they have constant access to rich supplies of fuel. Because early numerical simulations suggested that Pop III stars were $\\gtrsim$ 100 M$_{\\odot}$, the supermassive black hole seeds considered up to now were 100 - 300 M$_{\\odot}$. However, there is a growing numerical and observational consensus that some Pop III stars were tens of solar masses, not hundreds, and that 20 - 40 M$_{\\odot}$ black holes may have been much more plentiful at high redshift. However, we find that natal kicks imparted to 20 - 40 M$_{\\odot}$ Pop III BHs during formation eject them from their halos and hence their fuel supply, precluding them from Eddington-limit growth. Consequently, supermassive black holes are far less likely to form from low-mass Pop III stars than from very massive ones. ", "introduction": "The existence of 10$^9$ M$_{\\odot}$ black holes (BH) in massive galaxies by $z \\sim 7$, only a billion years after the Big Bang \\citep[e.g.][]{mort11}, poses one of the great unsolved problems in cosmological structure formation. In the $\\Lambda$CDM paradigm, early structure formation is hierarchical, with small objects at high redshifts evolving into ever more massive ones by accretion and mergers through cosmic time. For this reason it is generally supposed that the supermassive black holes (SMBH) that power the $z \\sim 7$ \\textit{Sloan Digital Sky Survey} (\\textit{SDSS}) quasars grow from much smaller seeds at earlier epochs. The origin of SMBH and how they reach such large masses in such short times is a subject of ongoing debate. Three modes of formation have been proposed for SMBH seeds: the collapse of Pop III stars into 100 - 300 M$_{\\odot}$ black holes at $z \\sim 20$ \\citep{awa09}, baryon collapse in 10$^8$ M$_{\\odot}$ dark matter halos that have somehow bypassed previous star formation into 10$^4$ - 10$^6$ M$_{\\odot}$ BH at $z \\sim 15$ \\citep{wta08,rh09,sbh10}, and more exotic pathways like the relativistic collapse of dense primeval star clusters into 10$^4$ - 10$^6$ M$_{\\odot}$ BH \\citep[see section 3.3 of][for a recent review]{brmvol08}. Stellar-mass SMBH seeds form at $z \\sim 20$ when Pop III stars die in either core-collapse supernovae (SNe, 15 - 45 M$_{\\odot}$) or by direct collapse to a BH (45 - 100 M$_{\\odot}$, $\\gtrsim 260$ M$_{\\odot}$) \\citep{hw02}. This formation channel is favored by some because most dark matter halos will form a Pop III star at this epoch if they reach masses of $\\sim$ 10$^5$ M$_ {\\odot}$ \\citep{abn02, bcl02}. However, these BH have such low initial masses that they must continuously accrete at the Eddington limit to reach 10$^9$ M$_{\\odot}$ by $z \\sim 7$. This is problematic for several reasons. First, numerical simulations have shown that Pop III stars usually evaporate the halos that give birth to them, so the BH are 'born starving' \\citep[e.g.][]{ wan04,ket04,wet08a}. Filamentary inflows and mergers later restore baryons to the halo but only after 50 - 100 Myr \\citep{yet07}, during which crucial e-foldings in mass are lost. Second, preliminary studies indicate that once accretion commences, the BH itself emits ionizing radiation that disperses its own fuel supply, limiting its growth rate to a fraction of the Eddington limit \\citep{ milos09,pm11,pm12} \\citep[but see][]{li11}. Furthermore, if the seed BH is not confined to the halo, its duty cycle as it meanders through cosmological density fields is intermittent, which also curtails its growth \\citep{awa09}. Until now, 20 - 40 M$_{\\odot}$ Pop III BH \\citep{zwh08,wet08b} have been overlooked as candidates for SMBH seeds because previous studies assume that primordial stars are $\\gtrsim$ 100 M$_{\\odot}$. However, there is a growing numerical and observational consensus that some Pop III stars are tens of solar masses, not hundreds. More recent, much larger ensembles of numerical simulations found many halos with central collapse rates consistent with 20 - 60 M$_{\\odot}$ for the final mass of the star \\citep{on07} and that a fraction of the halos form binaries in this mass range \\citep{turk09}. Furthermore, new simulations of the formation of Pop III protostellar accretion disks suggest that they were prone to fragmentation into as many as a dozen smaller stars \\citep{ stacy10,clark11,sm11,get11}. Very preliminary calculations of I-front breakout from these disks indicate that ionizing UV radiation may terminate accretion onto the nascent star at $\\sim$ 40 M$_{\\odot}$ \\citep{hos11,stacy12}. On the observational side, recent attempts to reconcile the nucleosynthetic yields of Pop III supernovae with the chemical abundances found in ancient, dim extremely metal-poor stars in the Galactic halo suggest that 15 - 40 M$_{\\odot}$ primordial stars may have been responsible for most of the heavy elements expelled into the primeval IGM \\citep{jet09b}. The failure to detect the distinctive 'odd-even' nucleosynthetic signature of 140 - 260 M$_{\\odot}$ pair-instability SNe in metal-poor stars to date reinforces the fact that some Pop III stars might not be very massive, but this pattern may have been masked by selection effects in the observations \\citep{karl08}. Low-mass Pop III BH are crucially different from more massive BH because they are born in supernova explosions rather than by direct collapse. Asymmetries in the core-collapse engine can impart kicks of 200 - 1000 km/s to 20 - 40 M$_{ \\odot}$ BH, ejecting them from the halos that gave birth to them. In this Letter we examine the implications of natal kicks for low-mass Pop III black holes as candidates for SMBH seeds. In $\\S \\, 2$ we review the formation pathways for low-mass Pop III BH. In $\\S \\, 3$ we calculate their post-supernova kinematics and retention fractions in halos. In $\\S \\, 4$ we conclude. ", "conclusions": "The number of stars that 10$^5$ - 10$^7$ M$_{\\odot}$ halos typically form is not well constrained. The studies of Pop III protostellar disk fragmentation performed thus far do not follow the evolution of the disk for enough dynamical times to determine the ultimate fate of the fragments, which may later merge with the central object or be destroyed by gravitational torques before becoming distinct stars. Ionizing UV radiation from one star-forming fragment or even from a nearby halo can also prematurely halt the collapse of other fragments in the disk, lowering the number of stars that eventually form in the halo \\citep[e.g.][]{su06,wet08b,suh09,wet10}. We also note that while the evolution of the fragments in the disk is expected to be roughly coeval, their 5 - 10 Myr quasistatic collapse times raise the possibility that the first star to form in the halo may explode and pre-empt the collapse of other fragments \\citep[e.g.][]{sak09}. Consequently, the number of low-mass Pop III stars that occupy the halo likely ranges from one to at most ten. Ejecta-driven natal kicks will evict most 20 - 32 M$_{\\odot}$ BH from their host halos, neutrino-driven kicks can drive more than 90\\% of 32 - 40 M$_{\\odot}$ BH from their halos, as we show in the right panel of Figure \\ref{fig:BH_MR}. This guarantees that on average all the BH will vacate the halo even if ten stars originally formed in it. Post-supernova kinematics thus strongly discourages 20 - 40 M$_{\\odot}$ Pop III BH from becoming supermassive because they are ejected from their fuel supply and deprived of crucial early e-foldings in mass. This process greatly reduces the parameter space in stellar mass from which SMBH can originate \\citep[e.g.][]{th09, lfh09}, especially if Pop III stars were mostly less than 50 M$_{\\odot}$. Also, if a given halo is capable of supporting early continuous Eddington rate accretion, a 20 - 40 M$ _{\\odot}$ BH is much less likely to become supermassive than a 100 M$_{\\odot}$ BH, either because it is ejected from the halo at birth or because it must undergo additional e-folding times to reach large masses. If most low-mass Pop III black holes were ejected from their halos at $z \\sim$ 20, where are they today? If on average they depart their host halos at $\\sim$ 500 km/s, they are unlikely to encounter another halo capable of capturing them in less than a Hubble time, and so many of these BH were exiled to the voids between galaxies. Over time, they may have gradually gained mass as they encountered high-density regions. In contrast, Pop III BH above 40 M$_{\\odot}$ are unlikely to be born with kicks and remain in the halo, intermittently accreting and growing over cosmic time. These black holes are much more likely to reside in the galaxies into which their host halos were taken, a few of which may have become the supermassive black holes found in the SDSS quasars today." }, "1112/1112.3147_arXiv.txt": { "abstract": "We analyze a high-resolution spectrum of the A3m star HD\\,27411. We compare abundances derived from ATLAS9 model atmospheres with those using the more computationally-intensive ATLAS12 code. We found very little differences in the abundances, suggesting that ATLAS9 can be used for moderate chemical peculiarity. Our abundances agree well with the predictions of diffusion theory, though for some elements it was necessary to calculate line profiles in non-thermodynamic equilibrium to obtain agreement. We investigate the effective temperatures and luminosities of Am/Fm stars using synthetic Str\\\"{o}mgren indices derived from calculated spectra with the atmospheric abundances of HD\\,27411. We find that the effective temperatures of Am/Fm stars derived from Str\\\"{o}mgren photometry are reliable, but the luminosities are probably too low. Caution is required when deriving the reddening of these stars owing to line blanketing effects. A comparison of the relative proportions of pulsating and non-pulsating Am stars with $\\delta$~Scuti stars shows quite clearly that there is no significant decrease of helium in the driving zone, contrary to current models of diffusion. ", "introduction": "The ``metallic-lined'' or Am stars are A-type stars which have strong absorption lines of some metals such as Zn, Sr, Zr and Ba and weaker lines of other metals such as Ca and/or Sc relative to their spectral type as determined by the strength of the hydrogen lines \\citep{preston74}. The strong metallic lines are more typical of an F star rather than an A star. The work of \\citet{michaud70} established radiative diffusion in a strong magnetic field as the likely cause of the chemical peculiarities in Ap stars. When the magnetic field is absent, diffusion leads to the Am/Fm stars \\citep{watson71}. The presence of magnetic fields in Am stars has been investigated, but with negative results, (e.g. \\citet{fossati07}). A peculiarity of Am stars is that their projected rotational velocities are generally much smaller than normal A stars and they are nearly always members of close binary systems. Rotational braking by tidal friction in a binary system is regarded as a possible explanation for the low rotational velocities in Am stars. Slow rotation further assists the segregation of elements by diffusion. The abundance anomalies predicted by the diffusion hypothesis are usually much larger than observed. \\citet{richer00} developed detailed models of the structure and evolution of Am/Fm stars using OPAL opacities, taking into account atomic diffusion and the effect of radiative acceleration. These models develop a convective zone due to ionization of iron-group elements at a temperature of approximately 200,000~K. In addition to this convective zone, these stars also have a thin superficial convective zone in which H and He{\\sc i} are partially ionized. By assuming sufficient overshoot due to turbulence, these separate convective zones become one large convective zone. The resulting mixing dilutes the large abundance anomalies predicted by previous model, leading to abundances which closely resemble those observed in Am/Fm stars. A detailed abundance analysis of eight Am stars belonging to the Praesepe cluster \\citep{fossati07} show good agreement with the predictions of \\citet{richer00} for almost all the common elements except for Na and possibly S. The models of \\citet{richer00} assume a certain ad-hoc parametrization of turbulent transport coefficients which are adjusted to reproduce observations. Other parameterizations of turbulence have been proposed for other types of stars. \\cite{talon06} have investigated to what extent these are consistent with the anomalies observed on Am/Fm stars. They find that the precision of current abundances is insufficient to distinguish between models. More recently, \\citet{michaud11} have studied the abundance anomalies of the mild Am star Sirius A. They find that except for B, N and Na, there is good agreement with the predicted anomalies but turbulent mixing or mass loss is required. It is not clear whether it is turbulence or mass loss which competes with diffusion to lower the abundance anomalies. For example, \\citet{vick11} find that diffusion in the presence of weak mass loss can explain the observed abundance anomalies of pre-main-sequence stars. This is in contrast to turbulence models which do not allow for abundance anomalies to develop on the pre-main-sequence. Most of the pulsational driving in $\\delta$~Scuti stars is caused by the $\\kappa$~mechanism operating in the He{\\sc ii} ionization zone. Diffusion tends to drain He from this zone and therefore pulsational driving may be expected to be weaker or absent in Am/Fm stars \\citep{baglin72}. In fact, for many years it was thought that classical Am/Fm stars did not pulsate, though claims were made for some stars \\citep{kurtz89}. Recently, intensive ground-based observations by SUPER-WASP \\citep{smalley11}, and also from the {\\it Kepler} mission \\citep{balona11} have shown that many Am/Fm stars do pulsate. \\citet{smalley11}, for example, found that about 200 Am/Fm stars out of a total of 1600 (12.5 percent) show $\\delta$~Sct pulsations, but with generally lower amplitudes. They found that the pulsating Am/Fm stars are confined between the red and blue radial fundamental edges, in agreement with \\citet{balona11}. While there are many $\\delta$~Sct stars hotter than the fundamental blue edge, this does not seem to be the case for pulsating Am/Fm stars. The significance of this result remains to be evaluated. The effect of draining of He from the He{\\sc ii} ionization zone is to reduce the width of the instability strip, the blue edge moving towards the red edge, eventually leading to the disappearance of the instability strip when He is sufficiently depleted \\citep{cox79}. \\citet{turcotte00} has discussed the effect of diffusion on pulsations in Am/Fm stars using the models by \\citet{richer00}. One significant difference with earlier models is that a substantial amount of He remains in the He{\\sc ii} ionization zone. The blue edge of the instability strip for Am/Fm stars is sensitive to the magnitude of the abundance variations and is thus indicative of the depth of mixing by turbulence. \\citet{turcotte00} predict that pulsating Am/Fm stars should lie in a confined region of the HR diagram close to the red edge of the $\\delta$~Sct instability strip. However, \\citet{balona11} show that there is no relationship between the predicted Am/Fm instability strip and the actual location of these stars in the HR diagram. A particularly interesting result of the pulsation analysis of \\citet{turcotte00} is the prediction of long-period g modes in A-type stars. As the star evolves, the driving regions shift deeper into the star and the g modes become gradually more and more excited. Whereas p modes are stabilized through diffusion, g modes tend to be excited as a result of that process. It appears that diffusion may act to enhance driving of long-period g modes due to a significant increase in opacity due to iron-group elements. This may have a bearing on the fact that nearly all A-type stars observed by {\\it Kepler} have unexplained low-frequencies \\citep{balona11a}. When dealing with objects with non-standard chemical composition, such as Am stars, it is crucial that the opacities are correctly calculated. This question has been investigated by several authors in recent years. These studies show that a non-standard chemical composition of the stellar atmosphere alters the flux distribution of the star or modifies the profiles of the Balmer lines (\\citet{leo97}, \\citet{cat04}). Therefore a determination of T$_{\\rm eff}$ and $\\log g$ based on a comparison between observed and computed Balmer-line profiles will not be correct unless one takes into account the metallicity of the star. Thus, even estimates based on standard analysis of the spectra may be in error when applied to Am/Fm stars. In this paper, we investigate the determination of effective temperature and surface gravity of the Am star HD\\,27411 (HR~1353, A3m) using spectra in the ESO archives. The purpose is to determine whether the stellar parameters of this star agree with those obtained from Str\\\"{o}mgren photometry and hence to test the reliability of the effective temperature calibration applied to Am/Fm stars. The star was used by \\citet{ryab08} as a comparison in their study on the calcium stratification in Ap stars. HD\\,27411 is not known to pulsate. However, as we know from {\\it Kepler} observations, pulsations in A and F stars with amplitudes too low to be detected from the ground are common. Atmospheric models obtained with ATLAS9 \\citep{kur93} use precomputed line opacities in the form of opacity distribution functions (ODFs). These are tabulated for multiples of the solar metallicity and for various microturbulent velocities. This approach allows very fast computation of model atmospheres, but with very little flexibility in choice of chemical profile and microturbulent velocity. While this is satisfactory for most applications, it fails for chemically peculiar stars where a non-standard chemical composition profile is required. This can be done with ATLAS12 \\citep{kur97}, which is essentially identical to ATLAS9, but uses the opacity sampling (OS) method to evaluate line opacities. In this study we compare the abundances of HD\\,27411 obtained with both codes to determine if the use of ATLAS12 is essential. \\begin{figure*} \\centering \\includegraphics[width=8.8cm]{balmer1.ps} \\includegraphics[width=8.8cm]{balmer2.ps} \\caption{Comparison between the observed (crosses) and computed (solid red line) hydrogen line profiles. From top to bottom: the Balmer line profiles from H$_{\\delta}$ to H$_{\\alpha}$. The synthetic profiles were computed with SYNTHE using an ATLAS12 model atmosphere with T$_{\\rm eff}$\\,=\\,7400~$\\pm$~150~K, $\\log g$\\,=\\,4.0~$\\pm$~0.1, $\\xi$\\,=\\,4.2~$\\pm$~0.3~km~s$^{-1}$, $v_e \\sin i$\\,=\\,20.5\\,$\\pm$\\,0.5~km~s$^{-1}$ and individual abundances shown in Table~\\ref{abund}.} \\label{balmer} \\end{figure*} The result that Am stars are not confined to particular region of the $\\delta$~Sct instability strip depends, to a large extent, on effective temperatures and luminosities estimated from Str\\\"{o}mgren photometry \\citep{smalley11}. It is not clear whether the calibration, derived from normal AF stars, can be applied to Am/Fm stars. In this paper we use synthetic Str\\\"{o}mgren photometry applied to models of Am/Fm stars to investigate the reliability of fundamental parameters estimated from the photometry. Finally, we discuss the relative numbers of pulsating and non-pulsating Am stars and compare these to the relative numbers of $\\delta$~Scuti and constant stars in the instability strip. From this comparison, one can deduce the effectiveness of pulsational driving in the He{\\sc ii} ionization zone and compare the He abundance to that expected from diffusion calculations. \\begin{table} \\centering \\caption{Ions used to determine the microturbulent velocity in HD\\,27411. The number of spectral lines used, the microturbulent velocity, $\\xi$, the derived abundance and the radial velocity, RV, are listed.} \\begin{tabular}{lcccc} \\hline \\hline Elem & N & $\\xi$ & Abundance & RV \\\\ & & km s$^{-1}$ & $\\log N_{\\rm el}/N_{\\rm Tot}$ & km s$^{-1}$ \\\\ \\hline Fe{\\sc i} & 71 & 3.9\\,$\\pm$\\,0.3 & $-$3.80\\,$\\pm$\\,0.03 & 40.4\\,$\\pm$\\,0.6\\\\ Fe{\\sc ii} & 15 & 4.4\\,$\\pm$\\,0.2 & $-$3.84\\,$\\pm$\\,0.04 & 40.9\\,$\\pm$\\,0.6\\\\ Ni{\\sc i} & 24 & 4.4\\,$\\pm$\\,0.6 & $-$4.63\\,$\\pm$\\,0.03 & 40.5\\,$\\pm$\\,1.3\\\\ \\hline Adopted & -- & 4.2\\,$\\pm$\\,0.3 & -- & 40.6\\,$\\pm$\\,0.3\\\\ \\hline \\end{tabular} \\label{xi_vr} \\end{table} ", "conclusions": "We analyzed the spectrum of the Am star HD\\,27411 from the UVES POP archive. Our aim was to investigate if the ATLAS12 model atmosphere code provides more reliable results than the ATLAS9 code for chemically peculiar stars. We found that there is very little difference in the abundances derived from ATLAS9 and from ATLAS12. Since ATLAS12 demands considerably greater resources, it seems safe to use ATLAS9, at least for moderate metallic enhancement. We find that the derived abundances in HD\\,27411 are in good agreement with the predictions of diffusion models by \\citet{richer00}. There were discrepancies for Na, Al, Si, S, and Cr, but these are resolved by using NLTE model atmospheres. We investigated the reliability of effective temperatures and luminosities of Am/Fm stars determined by Str\\\"{o}mgren photometry by synthesizing spectra having the abundances of HD\\,27411 for a range of effective temperatures. The resulting synthetic colours indicate that effective temperatures can be reliably determined from photometry, but owing to line blanketing in the $c_1$ passband, the resulting surface gravities are systematically to high, leading to lower luminosities. This result appears to be verified by comparing luminosities of Am/Fm stars obtained from their parallaxes and from photometry. {\\bf Determination of reliable luminosities for Am stars remains a difficult problem. At this stage, parallaxes offer the best results, but this can only be done for very few stars. As we have seen, luminosities obtained from Str\\\"{o}mgren photometry are subject to a systematic bias which depends on the overabundances of metals. The error in the surface gravity from high-resolution spectroscopy is typically 0.1 in $\\log g$. For A--F main sequence and giants, this translates into an error of about 0.12 in $\\log L/L_\\odot$ when using the calibration of \\citet{torres10}. For HD\\,27411, for example, we derive $\\log g = 4.0 \\pm 0.1$ from spectroscopy, whereas the value derived from the parallax is $\\log g = 3.8 \\pm 0.1$ (Table\\,2), Although the two values only differ by two standard deviations, this is enough to cause a difference of 0.36 in $\\log L/L_\\odot$. Although spectroscopic determinations of luminosities may lead to quite large errors in the luminosity, they should at least not be biased.} By far the most serious problem confronting the diffusion model is that there seems to be no appreciable settling of He in the He{\\sc ii} ionization zone, as predicted by the models. This is demonstrated by the fact that pulsating Am/Fm stars occur throughout the $\\delta$~Scuti instability strip, though they tend to be cooler than normal $\\delta$~Sct stars. In fact, the relative proportions of pulsating Am stars to non-pulsating Am stars is no different from the proportion of $\\delta$~Sct stars to constant stars in the $\\delta$~Sct instability strip. There is clearly a need to revise current ideas of diffusion to explain the Am phenomenon." }, "1112/1112.1967_arXiv.txt": { "abstract": "\\baselineskip=16pt We present the holographic predictions for cosmological 3-point correlators, involving both scalar and tensor modes, for a universe which started in a non-geometric holographic phase. Holographic formulae relate the cosmological 3-point functions to stress tensor correlation functions of a holographically dual three-dimensional non-gravitational QFT. We compute these correlators at 1-loop order for a theory containing massless scalars, fermions and gauge fields, and present an extensive analysis of the constraints due to Ward identities showing that they uniquely determine the correlators up to a few constants. We define shapes for all cosmological bispectra and compare the holographic shapes to the slow-roll ones, finding that some are distinguishable while others, perhaps surprisingly, are not. ", "introduction": "In a recent series of papers \\cite{McFadden:2009fg,McFadden:2010na, McFadden:2010jw,McFadden:2010vh,Easther:2011wh,McFadden:2011kk} we put forward a holographic framework for inflationary cosmology and discussed a novel class of models describing a universe that started in a non-geometric phase, which is described holographically via a large-$N$ three-dimensional QFT. The power spectra and the scalar bispectrum were computed in \\cite{McFadden:2009fg,McFadden:2010na} and \\cite{McFadden:2010jw}, respectively, and in this paper we complete this program by computing the non-Gaussianities that involve tensors. Non-Gaussianities involving tensors are not expected to be measurable in near-future experiments. Nevertheless, they are still interesting theoretically and their structure has been the topic of several recent papers \\cite{Maldacena:2011nz,Soda:2011am,Shiraishi:2011st,Gao:2011vs}. The holographic model is specified by providing the dual QFT and the holographic dictionary that relates QFT correlation functions to cosmological observables. We worked out the holographic dictionary for non-Gaussianities involving tensors in \\cite{McFadden:2011kk}, and in this paper we compute the relevant QFT correlation functions. The models we discuss are based on perturbative three-dimensional QFTs that admit a large-$N$ limit and have a generalised conformal structure \\cite{Jevicki:1998ub,Kanitscheider:2008kd}. An example of such a theory is $SU(N)$ Yang-Mills theory coupled to massless scalars and fermions, with all fields transforming in the adjoint of $SU(N)$. The non-Gaussianities are extracted from the 3-point function of the stress tensor of this theory. The leading order 1-loop computation of this 3-point function is independent of the interactions of the QFT, and thus our main task is to compute this 3-point function for free QFTs. Since all leading order results depend only on the free theory, let us briefly discuss the case in which the holographic model is a free QFT. In such a model, the spectrum is the exactly scale-invariant Harrison-Zel'dovich spectrum and the bispectrum is given exactly by the results reported here, {\\it i.e.}, the leading order results are the exact answer in the free theory. The shapes associated with the bispectrum may thus be considered as the analogue of the exact scale-invariant spectrum for higher point functions. We have seen in \\cite{McFadden:2010vh} that the scalar bispectrum shape for this model is indeed special: it is exactly equal to the factorisable equilateral shape\\footnote{ The holographic model is also the only model that yields exactly this shape.} originally introduced in \\cite{Creminelli:2005hu}. One may thus anticipate that shapes associated with the other 3-point functions will also have special properties. Possible shapes for the bispectrum involving only tensors have been discussed recently in \\cite{Maldacena:2011nz} and here we will define and discuss shapes for the bispectrum involving both tensors and scalars. The computation of the 3-point function of the stress tensor at 1-loop is a non-trivial task even for free QFTs. We discuss and develop several methods for evaluating the relevant Feynman diagrams. The 1-loop result is constrained by Ward identities and these provide a very non-trivial check of the expression we obtained by a direct computation. As mentioned above, to 1-loop order only the free part of the QFT enters. The QFT consists of gauge fields, fermions, minimal and conformal scalars. Conformal scalars and fermions are conformal field theories and their 3-point functions are constrained by conformal Ward identities. As is well known (from a position space analysis) \\cite{Osborn:1993cr}, the 3-point function of the stress tensor in $d=3$ is uniquely fixed by conformal invariance to be a linear combination of two conformal invariants, and is thus parametrised by two constants. (We assume parity is preserved). Our computation is done in momentum space and we thus provide the most general such 3-point functions in momentum space, where the two parameters are the number of conformal scalars and the number of fermions. The same computation was also recently reported in \\cite{Maldacena:2011nz}\\footnote{Our results agree with the ones in v2 of \\cite{Maldacena:2011nz}. Relative to \\cite{Maldacena:2011nz}, we also computed semi-local terms.}. We have explicitly verified that our results satisfy the conformal Ward identities. Let us now turn to minimal scalars and gauge fields. In three dimensions vectors are dual to scalars, so one may expect that gauge fields contribute the same as minimal scalars at 1-loop order. We will indeed verify that this is the case. Note that beyond 1-loop the two are expected to contribute differently. Minimal scalars differ from conformal scalars in the way they couple to gravity, which in flat spacetime is reflected in their having a different stress tensor. More precisely, the stress tensor $T^\\phi_{ij}$ for a minimal scalar may be decomposed into a part $\\wT^\\phi_{ij}$ corresponding to the stress tensor for a conformal scalar plus an ``improvement term'': \\[ \\label{minconfdecomp} T_{ij}^\\phi = \\wT_{ij}^\\phi - \\frac{1}{8} \\left( \\delta_{ij} \\partial^2 - \\partial_i \\partial_j \\right) \\phi^2. \\] It follows that the 3-point function of $T_{ij}^\\phi$ may be computed from the 3-point functions involving $\\wT_{ij}^\\phi$ and the dimension one operator ${\\cal O}_1= \\phi^2$. In turn, these 3-point functions are uniquely determined by conformal invariance, up to constants \\cite{Osborn:1993cr}. Thus, effectively all 3-point functions are determined by conformal 3-point functions at this order, even though the underlying theory is not conformal. In \\cite{Maldacena:2011nz}, the 3-point functions for tensors were computed in a de Sitter background. The de Sitter isometries act as the conformal group at late times, and the 3-point functions are then constrained by conformal invariance to be specific linear combinations of the 3-point functions of conformal scalars and fermions. Note that the de Sitter result is the leading order approximation for slow-roll inflation. In general one expects that (broken) conformal invariance would constrain cosmological correlators in asymptotically de-Sitter slow-roll inflation, see also \\cite{Antoniadis:2011ib, Creminelli:2011mw}. Our holographic results are for a very different universe, but we have seen that all relevant 3-point functions are essentially determined by conformal 3-point functions. One may then wonder how our results compare with those of slow-roll inflation. The 3-point functions involving only tensors are determined by the 3-point functions of conformal scalars and fermions, and, as in the discussion of \\cite{Maldacena:2011nz}, they agree exactly with slow-roll inflation if the field content of the dual QFT is appropriately chosen. The other 3-point functions (involving both scalar and tensor perturbations) are different but, perhaps surprisingly, they are rather similar. To quantify the differences we define (and plot) shape functions for all correlators, generalising the notion of shape functions for 3-point functions of only scalars or only tensors. This paper is organised as follows. In Section \\ref{sec:dual} we discuss the dual QFT and in Section \\ref{sec:dict} we present the holographic dictionary. In Section \\ref{sec:eva} we compute all relevant QFT correlation function by direct evaluation of the relevant Feynman integrals, and in Section \\ref{sec:Wid} we explain their structure using Ward identities. The holographic predictions for the cosmological observables are presented in Section \\ref{sec:HolPredictions} and these results are compared with the slow-roll ones in Section \\ref{sec:slow-roll}. We discuss our results in Section \\ref{sec:disc}. Several technical results are presented in four appendices: in Appendix \\ref{sec:hel} we summarise our notation and conventions for the helicity tensors, in Appendix \\ref{methods_app} we present three different methods for evaluating the relevant diagrams, in Appendix \\ref{App_ghosts} we show that ghosts and gauge fixing terms do not contribute in correlators of the stress tensor and in Appendix \\ref{app_WI} we present the conformal and diffeomorphicm Ward identities. ", "conclusions": "\\label{sec:disc} In this paper we computed the complete set of bispectra (and defined and extracted the corresponding shapes\\footnote{ To our knowledge, shapes other than those relevant for purely scalar or purely tensor bispectra have not been discussed before.}) for a class of holographic models of the very early universe % based on perturbative QFT. The leading 1-loop result actually depends only on the free part of the QFT, so in particular our results are also the complete answer when the dual QFT is free. The field content of the dual theory includes gauge fields, massless fermions, massless minimal and conformal scalars and thus the parameters that can appear in the results are the number of species for each type of field. The bispectra could, a priori, depend on these in a complicated way, but it turns out that we get instead (nearly) universal results that are independent of all details of the dual QFT, within the class of the theories we consider. Thus, % these models make clean and precise predictions. One can trace this universality to the specific form of the holographic map, the fact that to leading order the QFT is free, symmetry considerations and properties of $d=3$ theories. Let us explain this. Firstly, in three dimensions, vectors are dual to scalars so one may anticipate that the contribution due to gauge fields (at 1-loop order) is equal to that of the contribution due to mininal scalars, and we indeed find this to be the case. Taking this into account, the answer could then depend on three parameters, the number of conformal scalars, ${\\cal N}_\\chi$, the number of fermions, ${\\cal N}_\\psi$ and the total number of gauge fields plus minimal scalars, ${\\cal N}_{(B)}$. The trace Ward identity of the dual QFT and the specific form of the holographic formulae then imply that, in all correlators involving at least one factor of $\\zeta$, the field content appears only as a multiplicative factor and is such that the corresponding shape functions are completely independent of the field content. Let us now turn to correlators involving only tensors: these are effectively determined by the 3-point function of the stress tensor of a CFT. In three dimensions, this 3-point function is parametrised by two constants, which in our case are related to the field content. Indeed, the shape corresponding to three positive helicity gravitons does depend on the field content, but surprisingly the shape for two positive and one negative helicity graviton is independent of the field content. This can be explained in part by the fact that the $\\$ correlator (at separated points) is actually uniquely fixed by conformal invariance up to a single constant. We emphasize however that this by itself is not sufficient to explain the independence of the corresponding shape function from the field content, as the specific form of the semi-local terms (both in the holographic map and in $\\$) is crucial for this to happen. Our calculations carefully include all such semi-local contributions. In the holographic formulae for the bispectra, these contributions appear as terms in the numerator that are non-analytic in only one of the three momenta. Since the denominator of the holographic formulae is however non-analytic in all three momenta, the net contribution of these semi-local terms to the bispectra is in fact non-analytic in {\\it two} of the three momenta. Semi-local terms in the holographic formulae may thus contribute, for example, to `local'-type non-Gaussianity behaving as $1/q_1^3 q_2^{3} + \\mathrm{perms}$. Contributions of this nature therefore play a crucial role in allowing different cosmological shapes to be distinguished. To get a feeling for our results we also computed the corresponding slow-roll results and compared them with the holographic results. Firstly, comparing the power spectra one obtains a relation between the parameters $N^2$, ${\\cal N}_{(A)}$ and ${\\cal N}_{(B)}$ of the QFT and the parameters $\\kappa^2$, $H_*^2$ and $\\epsilon_*$ of the slow-roll model. Comparing the 3-point functions, we find that the $\\bg^{(+)}\\bg^{(+)}\\bg^{(-)}$ correlators agree exactly, while the $\\bg^{(+)}\\bg^{(+)}\\bg^{(+)}$ correlators can be made to agree if one imposes that that the field content satisfies the relation $2\\mathcal{N}_\\psi = \\mathcal{N}_\\phi+\\mathcal{N}_A+\\mathcal{N}_\\chi$. As explained in \\cite{Maldacena:2011nz}, these slow-roll correlators are constrained by the late-time de Sitter isometries to satisfy conformal Ward identities, and thus at separated points they should be expressible in terms of the 3-point functions of conformal scalars and free fermions. Indeed, the linear combination found in v2 of \\cite{Maldacena:2011nz} is the same as the one we find (setting $\\mathcal{N}_\\phi=\\mathcal{N}_A=0$ in our relation). By taking into account the contribution from semi-local terms, however, we are further able to correctly recover every individual term appearing in the graviton bispectra. There is no apparent reason for the remaining slow-roll and holographic correlators to agree. Nevertheless we find rather similar results. To quantify the difference we used the cosine orthogonality measure of \\cite{Fergusson:2008ra} to obtain a first indication of the distinguishability of the corresponding shapes. We find that the shapes for $\\z \\z \\bg^{(s)}$ are nearly indistinguishable, while for $\\z \\bg^{(+)} \\bg^{(+)}$, the two shapes may be distinguished (as a consequence of differing behaviour in the squeezed limit where the momentum associated with the $\\z$ goes to zero), with the case of $\\z \\bg^{(+)} \\bg^{(-)}$ lying in between. All in all, we have a rather complete understanding of this class of models and their phenomenology. There are still a few things to be understood better: what constrains the semi-local contributions to the tensor correlators, and why are the holographic results apparently close to slow-roll ones? One can presumably also understand the squeezed limit of the correlators using Ward identities. On a whole, however, the structure of these models is reasonably firmly understood. It would be interesting to arrive at a similar level of understanding for the class holographic models that are based on deformations of conformal field theories." }, "1112/1112.4290_arXiv.txt": { "abstract": "{The physical origin of the strong magnetic activity in T Tauri stars and its relation to stellar rotation is not yet well-understood. } {We investigate the relation between the X-ray activity, rotation, and Rossby number for a sample of young stars in the $\\approx 3$~Myr old cluster IC~348. } {We use the data of four \\textit{Chandra} observations of IC~348 to derive the X-ray luminosities of the young stars. Basic stellar parameters and rotation rates are collected from the literature. This results in a sample of 82 X-ray detected stars with known rotation periods. We determine the Rossby numbers (i.e.~the ratio of rotation period to convective turnover time) of 76 of these stars from stellar structure- and evolution-models for pre-main sequence stars. } {The young stars in IC~348 show no correlation between X-ray activity and rotation period. For the Rossby numbers, nearly all IC~348 stars are in the saturated regime of the activity--rotation relation defined by main-sequence stars. Searching for possible super-saturation effects, we find a marginal (but statistically in-significant) trend that the stars with the smallest Rossby numbers have slightly lower X-ray activity levels. There are no significant differences in the X-ray activity level for stars of different spectral types and no relation between spectral type and Rossby number is seen. In addition, for stars belonging to different IR-classes, no significant differences are present for the X-ray activity level as well as for their Rossby numbers. We compare the dispersion in the fractional X-ray luminosities of the stars in the saturated rotation regime in IC~348 to that seen in younger and older stellar populations. The scatter seen in the $\\approx 3$~Myr old IC~348 [$\\sigma\\left(\\log\\,(L_{\\rm X}/L_{\\rm bol}) \\right) = 0.43$] is considerably smaller than for the $\\approx 1$~Myr old Orion Nebula Cluster [\\,$\\sigma\\left(\\log (L_{\\rm X}/L_{\\rm bol}) \\right) = 0.63$], but, at the same time, considerably larger than the dispersion seen in the $\\approx 30$~Myr old cluster NGC~2547 [$\\sigma\\left(\\log (L_{\\rm X}/L_{\\rm bol}) \\right) = 0.24$] and in main-sequence stars. } { The results of our X-ray analysis of IC~348 show that neither the rotation rates nor the presence/absence of circumstellar disks are of fundamental importance for determining the level of X-ray activity in TTS. Our results suggest that the scatter in the X-ray activity levels for the rapidly rotating members of young clusters decreases with the age of the stellar population. We interpret this as a signature of the changing interior structure of pre-main sequence stars and the consequent changes in the dynamo mechanisms that are responsible for the magnetic field generation.} ", "introduction": "Young stellar objects (YSOs) in all evolutionary stages, from class~I protostars, to T Tauri stars (TTS) to zero-age main-sequence stars, have highly ($\\sim 10^3 - 10^4$ times) elevated levels of X-ray activity (for reviews of the X-ray properties of YSOs and stellar coronal astronomy in general, see Feigelson \\& Montmerle 1999 and Favata \\& Micela 2003). Despite many years of research, the physical origin of this X-ray activity remains poorly understood. Although there is strong evidence that in most TTS the X-ray emission originates from magnetically confined coronal plasma \\citep[e.g.][]{Preibisch05}, it is unclear what kind of dynamo processes create the required magnetic fields. For main-sequence stars, the level of the magnetic activity (and thus the strength of the X-ray emission) is mainly determined by their rotation rate. Observations have revealed a clear rotation--activity relation of the form $L_{\\rm X}/L_{\\rm bol} \\propto P_{\\rm rot}^{-2.6}$ \\citep[e.g.][]{Pallavicini81,Pizzolato03}, which is in good agreement with the expectations of solar-like $\\alpha\\!-\\!\\Omega$ dynamo models \\citep[e.g.][]{Maggio87,Ossendrijver03}. The solar dynamo is thought to be anchored in the ``tachocline'', a thin zone between the inner radiative core and the outer convection zone. However, the connection between the observed surface magnetic activity and the properties of the dynamo generating the magnetic flux is not yet fully understood \\citep[e.g.,][]{Isik11}. For main-sequence stars, the increase in magnetic activity towards shorter rotation periods stops for periods shorter than $\\sim 2-3$ days, where the activity saturates around $\\log\\left(L_{\\rm X}/L_{\\rm bol}\\right) \\approx -3$. The physical reasons for this saturation effect are poorly understood \\citep[see, e.g.,][]{Jardine99}. For extremely rapidly rotating stars, the activity level seems to decrease slightly with increasing rotation rate \\citep{Prosser96,Randich00,James00,Jeffries11}, a phenomenon that is denoted as ``super-saturation'' and also not well-understood. Studies of stellar populations with different ages show that that there is a continuous evolution from the very high X-ray activity levels in the youngest stages to the much lower activity seen in older (more than a few hundred Myr old) stars \\citep[e.g.,][]{Guedel97,PF05,Telleschi05}. This evolution can be explained by the temporal decrease in the stellar angular momentum \\citep{Bouvier97,Herbst07}. Furthermore, the presence of strong magnetic fields on the surface of T Tauri stars has been clearly established \\citep[e.g.][]{Johns-Krull07}. These pieces of evidence suggest that the X-ray activity of YSOs originates in dynamo processes similar to those present in our Sun. However, the relation between rotation and X-ray activity in TTS remained unclear until recently, since in most studies of star forming regions the number of X-ray detected TTS with known rotation periods was too small to draw sound conclusions. A few years ago, the $Chandra$ Orion Ultradeep Project (COUP), a ten-day long observation of the Orion Nebula Cluster (ONC) with $Chandra$/ACIS \\citep[for details of the observation and data analysis, see][]{Getman05} and the XMM-Newton Extended Survey of the Taurus Molecular Cloud \\citep[XEST, see][]{Guedel07} provided very sensitive X-ray data sets for large samples of TTS. The COUP and the XEST data have both shown that the TTS in Orion and Taurus do \\textit{not} follow the relation between rotation period and X-ray luminosity for main-sequence stars \\citep{Preibisch05,Briggs07}. In addition the TTS spin up during the first $\\sim$10-30 Myr \\citep{Herbst05} does not lead to an increase in the X-ray luminosity. This places doubt on the solar-like dynamo activity scenario for TTS. Another argument against solar-like dynamos in young TTS comes from theoretical considerations: at ages of $\\le 2$~Myr, most TTS are expected to be fully convective, and thus should not possess a tachocline. The conventional $\\alpha\\!-\\!\\Omega$ dynamo cannot work in such a situation. Several alternative dynamo concepts have been developed for fully convective stars \\citep[e.g.][]{Giampapa96,Kueker99,Dobler06,Chabrier06,Browning07,Voegler07,Graham10}. Although the reliability of these theoretical models is not entirely clear, there is good evidence for the simultaneous presence of an $\\alpha\\!-\\!\\Omega$ dynamo \\textit{and} some kind of a small-scale turbulent dynamo in the convection zone of our Sun \\citep[e.g.,][]{Durney93,Bueno04}. Direct observations of the relation between rotation and magnetic activity for stellar samples spanning a wide range of ages can provide fundamental constraints on the theoretical models. Numerous large datasets are available for samples of main-sequence stars as well as for young stellar clusters with ages as young as $30$~Myr \\citep[e.g.][]{Prosser96,Stauffer97a,Stauffer97b,Randich00,Jeffries11}. For younger ages, however, there is still a clear lack of reasonably large stellar samples for which good activity \\textit{and} rotation data are available. The data obtained in the COUP and XEST observations do both indicate that there are very young stellar populations of only $\\la 1$~Myr old \\citep[see, e.g.,][]{Weights09,Dib10,Luhman10}. Since the stellar rotation, the magnetic activity levels, and other basic stellar parameters evolve strongly in the age range between 1~Myr and 30~Myr, a sample of stars with an age of a few Myr can provide valuable information. With an age of $\\approx 3$~Myr \\citep{Luhman03,Mayne07}, the young cluster IC~348 is very well-suited in this respect. This age is particularly interesting because it corresponds to the point in time when the structure of solar-mass stars changes from a fully convective interior to a radiative core plus convective envelope structure, and this should affect the dynamo processes that are the ultimate source of the magnetic activity. ", "conclusions": "Our analysis of the relation between X-ray activity (as traced by deep \\textit{Chandra} observations) and the rotation properties of the TTS in the young cluster IC~348 has yielded the following results. First, we have shown that there is no correlation between the fractional X-ray luminosity and the rotation periods of the TTS; even the rather slowly rotating stars (periods $\\ga 10$~days) display very strong X-ray activity. Second, according to the Rossby numbers (that we have determined based on detailed stellar models for pre-main sequence stars), all TTS are in the saturated regime of the rotation--activity relation defined by main-sequence stars. Third, we have found no significant evidence of super-saturation among the most rapid rotators, although our data suggest that stars with extremely low Rossby numbers ($Ro \\leq 0.006$) have slightly lower activity levels. This seems to agree with the results of \\citet{Jeffries11}, who claim that in their NGC~2547 sample supersaturation occurs only for Rossby numbers lower than 0.005. In all three of these aspects, the TTS in IC~348 behave similarly to the TTS in the Orion Nebula Cluster. A remarkable difference is however seen in the scatter in the X-ray activity levels for the TTS in the saturated rotation regime. The scatter seen in the $\\approx 3$~Myr old IC~348 sample is considerably smaller than that in the $\\approx 1$~Myr old Orion Nebula Cluster sample, but, at the same time, considerably larger than seen in the $\\approx 30$~Myr old stars in NGC~2547 as well as for main-sequence stars. This suggests that some process reduces the wide distribution of activity levels seen in the youngest stars towards the much narrower distribution in older pre-main sequence and (young) main sequence stars during the first $\\sim 30$~Myr period, whereas the absolute level of the X-ray activity remains roughly constant during that time. A possible explanation of this effect may be related to the stellar interior structure and the corresponding dynamo mechanisms that are the basis of the magnetic activity that produces the observed X-ray emission. At an age of $\\leq 1$~Myr (e.g., in the Orion Nebula Cluster), almost all low- and intermediate mass stars ($M \\leq 2\\,M_\\odot$) are fully convective. As already mentioned, the solar-like $\\alpha\\!-\\!\\Omega$ dynamo cannot operate in such a situation and some kind of small-scale convective dynamo must be the source of the magnetic activity. The more or less chaotic nature of such a dynamo may be responsible for the very wide scatter in activity levels seen for stars with similar rotation rates and stellar parameters. According to the pre-main sequence stellar evolution models of \\citet{Siess00}, a $1.5\\,M_\\odot$ star develops a radiative core at an age of about 1~Myr. As time proceeds, stars with increasingly lower masses develop radiative cores (at $\\approx 1.7$~Myr for $1.2\\,M_\\odot$ stars, at $\\approx 2.3$~Myr for $1.0\\,M_\\odot$ stars, and at $\\approx 4.2$~Myr for $0.8\\,M_\\odot$ stars). As soon as a star has a radiative core, the conditions for the operation of a solar-like $\\alpha\\!-\\!\\Omega$ dynamo are met, which then gradually replaces the convective dynamo. When the total magnetic activity is dominated by the $\\alpha\\!-\\!\\Omega$ dynamo, stars in the saturated rotation regime have a rather homogeneous level of X-ray activity, i.e.~a small scatter in their $L_{\\rm X}/L_{\\rm bol}$ values. At the 3~Myr age of IC~348, all stars with $M \\ge 0.9\\,M_\\odot$ (corresponding to a spectral type of K6) should have a radiative core; this concerns about 25\\% of our stellar sample in IC~348 and can explain why the dispersion in the X-ray activity levels in IC~348 is smaller than in the Orion Nebula Cluster. As time proceeds, an increasing fraction of the lower-mass stars make the transition from a fully convective to a core/envelope structure, where the operation of an $\\alpha\\!-\\!\\Omega$ dynamo becomes possible, and this should continuously decrease the scatter in their $L_{\\rm X}/L_{\\rm bol}$ values. By the age of $\\approx 30$~Myr, nearly all stars have attained their final stellar structure, and thus the scatter in the activity levels has settled to the rather small dispersion as typical of (rapidly rotating) main-sequence stars. This scenario could be a qualitative explanation of the observed decrease in the scatter in the X-ray activity level of stars in the saturated rotation regime." }, "1112/1112.6406_arXiv.txt": { "abstract": "One key goal of the {\\it Hubble Space Telescope} Cosmic Assembly Near-Infrared Deep Extragalactic Legacy Survey is to track galaxy evolution back to $z\\approx 8$. Its two-tiered ``wide and deep'' strategy bridges significant gaps in existing near-infrared surveys. Here we report on $z\\approx 8$ galaxy candidates selected as F105W-band dropouts in one of its deep fields, which covers 50.1~arcmin$^2$ to 4~ks depth in each of three near-infrared bands in the Great Observatories Origins Deep Survey southern field. Two of our candidates have $J<26.2$~mag, and are $> 1$~mag brighter than any previously known F105W-dropouts. We derive constraints on the bright-end of the rest-frame ultraviolet luminosity function of galaxies at $z\\approx 8$, and show that the number density of such very bright objects is higher than expected from the previous Schechter luminosity function estimates at this redshift. Another two candidates are securely detected in {\\it Spitzer} Infrared Array Camera images, which are the first such individual detections at $z\\approx 8$. Their derived stellar masses are on the order of a few $\\times 10^9$~$M_\\odot$, from which we obtain the first measurement of the high-mass end of the galaxy stellar mass function at $z\\approx 8$. The high number density of very luminous and very massive galaxies at $z\\approx 8$, if real, could imply a large stellar-to-halo mass ratio and an efficient conversion of baryons to stars at such an early time. ", "introduction": "In the recent years, deep near-infrared imaging surveys have begun to yield a significant number of candidate galaxies at very high redshifts. The wide-field surveys from the ground have produced a handful of bright candidates at $z\\approx 7$ (e.g., Ouchi et al. 2009; Hickey et al. 2010; Castellano et al. 2010; Capak et al. 2011; Hsieh et al. 2012; Hathi et al. 2012), while the pencil-beam survey by the {\\it Hubble Space Telescope} (HST) Wide Field Camera 3 (WFC3) within the historical Advanced Camera for Surveys (ACS) Hubble Ultra Deep Field (hereafter ACS-HUDF; Beckwith et al. 2006) and its two parallel fields (HUDF09; PI. Illingworth) have probed the fainter populations at $z\\approx 7$--8 (Oesch et al. 2010; Bouwens et al. 2010, B10; Bunker et al. 2010; McLure et al. 2010, M10; Yan et al. 2010, Y10; Finkelstein et al. 2010) and possibly out to $z\\approx 10$ (Yan et al. 2010; Wyithe et al. 2011; Bouwens et al. 2011a). The WFC3 Early Release Science program (ERS, PI. O'Connell; Windhorst et al. 2011) has played an important role in connecting the ``wide'' and the ``deep'' ends of the exploration, allowing construction of $z\\approx 7$--8 samples at intermediate brightness levels (Wilkins et al. 2011; B11; Lorenzoni et al. 2011, L11). Meanwhile, the Hubble Infrared Pure Parallel Imaging Extragalactic Survey (HIPPIES; Yan et al. 2011) and the Brightest of Reionizing Galaxies Survey (Trenti et al. 2011a) have been exploring the bright end of the population through a large number of random, discrete WFC3 pointings obtained during the {\\it HST} parallel orbits. Surveys in foreground cluster fields utilizing gravitational lensing magnification have also resulted in a handful of promising candidates at $z\\gtrsim 7$--9 (e.g., Richard et al. 2006; Bradley et al. 2008, 2012; Laporte et al. 2011; Hall et al. 2012), which complement the surveys in blank-sky fields. The Cosmic Assembly Near-Infrared Deep Extragalactic Legacy Survey (CANDELS; PIs: Faber \\& Ferguson; see Grogin et al. 2011 and Koekemoer et al. 2011) employs a two-tiered, ``wide and deep'' strategy, which makes it uniquely positioned in bridging the significant gaps among existing surveys. In particular, its ``Deep'' component will cover $\\sim 125$~arcmin$^2$ in two fields upon completion, and its data will be ideally suited for studying galaxies at very high redshifts. In this paper, we report our preliminary results from the Deep observations in the Great Observatories Origins Deep Survey (GOODS; Giavalisco et al. 2004) southern field, where we use the nearly complete data set to study the galaxy population at $z\\approx 8$. We use the following cosmological parameters throughout: $\\Omega_M=0.27$, $\\Omega_\\Lambda=0.73$ and $H_0=71$~km~s$^{-1}$~Mpc$^{-1}$. The quoted magnitudes are all in the AB system. ", "conclusions": "Our samples are unique in two aspects. (1) We have two very bright candidates at $J_{125}\\leq 26.0$~mag, which are at least $\\sim 1$~mag brighter than the brightest $Y_{105}$-dropouts previously found. They are comparable in brightness to the brightest $Y_{098}$-dropouts reported by Yan et al. (2011) and Trenti et al. (2011a,b) in their WFC3 parallel surveys, but could be at higher redshifts because the redshift window of this current $Y_{105}$-dropout selection is higher than that of the existing $Y_{098}$-dropouts. (2) Two other $Y_{105}$-dropouts, which are not among the brightest in the WFC3 bands, are securely detected in IRAC, which is in sharp contrast to previous results where only non-detection in IRAC have been reported for $Y_{105}$-dropouts. Here we discuss the implications of both. \\subsection{Stellar Population and Stellar Mass} We investigate the stellar populations of the IRAC-detected $Y_{105}$-dropouts, {\\tt AUTO\\_035} and {\\tt AUTO\\_293}, by analyzing their spectral energy distributions (SED) through template fitting. The approach here is similar to that of Finkelstein et al. (2011). We first use the EAZY code (Brammer et al. 2008) and the photometry of the GOODS ACS \\& CANDELS WFC3 data to estimate photometric redshifts ($z_{phot}$). The full SED (including the IRAC photometry) is then fitted to a suite of templates based on the updated models of BC03 (the so-called ``CB07'' models). A Salpeter initial mass function (Salpeter 1955) and the metallicities of 0.02--1$Z_\\odot$ are adopted. The templates have a range of exponentially decreasing and increasing SFHs (see also Papovich et al. 2011), and can include nebular lines based on the number of ionizing photons and metallicity of a given model (B. Salmon et al. in preparation). We assume the extinction law of Calzetti (2001) with $E(B-V)$$=$0--0.5~mag and the H I absorption as formulated in Madau (1995). The results are summarized in Figure 4. We note that fitting redshift and other properties simultaneously does not signficantly change these results, and the differences are captured in the errors that we quote here. The best-fit photometric redshifts are $8.5^{+0.2}_{-4.5}$ and $8.9^{+0.2}_{-0.5}$ for {\\tt AUTO\\_035} and {\\tt 293}, respectively, which are consistent with the redshift window of our color selection. Adding the formal $I_{814}$ limit to the fitting process does not change the results. We fit the SEDs with and without the contribution from nebular lines, and both results suggest high stellar mass ($\\mathcal M_*$) in the range of $10^{9}$ to $10^{10} M_\\odot$ for both objects. Including nebular emission lines, we obtain $2.5^{+4.7}_{-1.4}\\times 10^{9}$ and $0.9^{+3.0}_{-0.1}\\times 10^{9}$~$M_\\odot$ for {\\tt AUTO\\_035} and {\\tt 293} respectively, while without nebular emission lines these are $10.5^{+1.8}_{-8.1}\\times 10^{9}$ and $2.6^{+1.8}_{-1.4}\\times 10^{9}$~$M_\\odot$, respectively. The error bars reflect the 68\\% confidence level of the fit. Labb\\'e et al. (2010a) derived the stellar masses of galaxies at lower redshifts of $z\\approx 7$ using models without nebular emission, and the average value is $\\sim 10^9$~$M_\\odot$. The two objects in our sample have comparable or even larger values using similar models, which are surprisingly high at such an early epoch. Using the $\\mathcal M_*$ estimates, we obtain the first measurement of the mass function (MF) of galaxies at $z\\approx 8$ at the high-mass end. As we only have two objects whose $\\mathcal M_*$ values are reasonably similar, we opt to count them within one mass bin of a $\\pm 0.4$~dex bin size. The two different sets of models result in significantly different mass estimates, and therefore the bin center is different for each case. Assuming a top-hat selection function of our survey within $7.7\\leq z\\leq 8.7$, we get $\\phi(log\\frac{M_*}{M_\\odot})|_{9.2\\pm 0.4} = (1.9^{+2.4}_{-0.6})\\times 10^{-5}$~Mpc$^{-3}$~dex$^{-1}$ if we adopt the values derived using the models with nebular emission. If we apply the effective volume ($V_{eff}$) correction of the survey instead (see \\S 3.4), we obtain $(3.1^{+4.0}_{-1.0})\\times 10^{-5}$~Mpc$^{-3}$~dex$^{-1}$. Here we use $V_{eff}=\\int V_{eff}(m_J)dm_J=\\int\\int dm_JdzP(m_J,z)dV/dz$, where $dzdV/dz$ is the unit co-moving volume at redshift $z$, and $P(m_J,z)$ is the redshift selection function at different magnitudes $m_J$ as derived through simulations (see Figure 3). The error bars here reflect the 68\\% interval of the uncertainties caused by the Poisson noise in the sample. For the case of using the $\\mathcal M_*$ estimates based on the models without nebular lines, all the above values are applicable at $\\phi(log\\frac{M_*}{M_\\odot})|_{9.7\\pm 0.4}$. The contribution of these objects to the global stellar mass density, $\\rho_{*}$, at $z\\approx 8$ within our survey volume is $(3.3^{+7.4}_{-1.4}) \\times 10^{4}$ and $(12.6^{+3.4}_{-9.0}) \\times 10^{4}$~$M_\\odot$~Mpc$^{-3}$ based on the models with and without nebular emission lines, respectively, if assuming a top-hat selection function at $7.7\\leq z \\leq 8.7$. The errors reflect the 68\\% confidence level of the fitting results. If we apply the same correction for $V_{eff}$ as above, we obtain $\\rho_{*}=(5.5^{+12.6}_{-2.4}) \\times 10^{4}$ and $(21.2^{+5.8}_{-15.3}) \\times 10^{4}$~$M_\\odot$~Mpc$^{-3}$. We derive the mass-to-light ratio of these two objects using their 4.5~$\\mu$m flux as a proxy to $L_V$. Using the fit results with contribution from nebular lines, we obtain $\\mathcal M_*/L_V=0.07^{+0.14}_{-0.04}$ and $0.04^{+0.13}_{-0.00}$ for {\\tt AUTO\\_035} and {\\tt 293}, respectively. Without the contribution from nebular lines, we obtain $\\mathcal M_*/L_V=0.31^{+0.06}_{-0.24}$ and $0.12^{+0.08}_{-0.06}$. The latter values are in general agreement with those obtained by Labb\\'e et al. (2010a,b) through stacking analysis of the IRAC data of the $Y_{098}$-dropouts and the $Y_{105}$-dropouts in the WFC3 ERS field (see also Gonz\\'alez et al. 2011) and the HUDF using similar models without nebular lines. On the other hand, Labb\\'e et al. (2010b) estimated that nebular lines have a small impact on $\\mathcal M_*/L_V$, reducing it by $\\sim 0.2$~dex. Our results suggest a stronger effect, reducing $\\mathcal M_*/L_V$ by $\\sim 0.5$--0.6 dex. \\begin{figure}[tbp] \\centering \\includegraphics[width=\\linewidth]{fig4.eps} \\caption{Summary of SED fitting results for {\\tt AUTO\\_035} (left) and {\\tt 293} (right). The results obtained from the fit with and without the contributions from nebular lines are coded in blue and red, respectively. The top panels show the observed SEDs (grey circles) and the best-fit templates, and the insets display their 3.6 and 4.5~$\\mu$m image ($18\\farcs 6 \\times 18\\farcs 6$). The bottom panels show the likelihood functions of $z_{phot}$ and $M_*$, and the best-fit values are indicated by the vertical lines. } \\end{figure} \\subsection{Constraint on the LF} Here we use our sample to constrain the LF at $z\\approx 8$. We adopt the $J_{125}$ {\\tt MAG\\_AUTO} values of the candidates as their total magnitudes. After applying the correction for $V_{eff}$, we obtain the stepwise LF in the five 0.5-mag bins (25.45, 25.95, 26.45, 26.95, 27.45)~mag as $\\phi = (2.5^{+5.8}_{-0.6}, 2.5^{+5.8}_{-0.6}, 2.8^{+6.3}_{-0.6}, 3.6^{+8.3}_{-0.8}, 24.2^{+22.6}_{-5.6})\\times 10^{-5}$~Mpc$^{-3}$~mag$^{-1}$. These number densities are shown in Figure 5, and compared to the predictions from a number of Schechter LF estimates at $z\\approx 8$ (M10; Y10; B11; L11). The black squares are our observed densities, while the red squares are the densities after correcting for $V_{eff}$. The error bars represent a Bayesian 68\\% credible interval, indicating the central 68 percentile range for the posterior distribution of the true number density assuming Poisson statistics. These uncertainties account for the incompleteness, but do not account for any systematic uncertainties from the contamination due to low-z interlopers. For ease of direct comparison to observations, surface density and apparent magnitude scales are also provided on the same figures to present these results in terms of differential number densities versus apparent $J_{125}$ magnitudes. For comparision, we also plot the {\\it observed} densities (i.e., before corrections for their corresponding $V_{eff}$ values) extracted from the sample of Y10 in the HUDF09 proper and those of B11 in the HUDF09, HUDF09P1 and HUDF09P2 {\\it after} applying the additional criteria of $Y_{105}-J_{125}\\geq 0.8$~mag and $J_{125}-H_{160}\\leq 0.3$~mag. \\begin{figure}[tbp] \\centering \\includegraphics[width=\\linewidth]{fig5.eps} \\caption{Constraints on the very bright-end of the LF at $z\\approx 8$ based on the sample of $Y_{105}$-dropouts in CANDELS/Deep from this current work. The black squares are the differential surface densities inferred from the number counts in our sample, while the red squares are those after correction for $V_{eff}$ with respect to the volume within $7.7\\leq z\\leq 8.7$. The count predictions over $7.7\\leq z\\leq 8.7$ are from the Schechter LF estimates at $z\\approx 8$ from B11, M10, Y10, and L11 are shown as various black curves, together with a non-evolution one at $z\\approx 7$ from B11 as the blue curve. For comparison, different symbols at the faint-end show the raw (observed) surface densities of $Y_{105}$-dropouts based on the following samples: Y10 (HUDF -- stars), B11 (HUDF -- crosses; HUDFP1 -- pentagons; HUDFP2 -- triangles). } \\end{figure} An intriguing feature that our samples reveal is that there could be an excess of bright objects at $J_{125}\\lesssim 26.2$~mag with respect to any of the existing Schechter LF estimates at $z\\approx 8$ in the literature would predict. The inferred number density at this bright end is even higher than the prediction from a non-evolution $z\\approx 7$ LF (e.g., the one from B11). This excess is still evident with the observed counts {\\it before} applying the incompleteness correction. The observed counts at $J_{125}\\lesssim 26.2$~mag in Figure 5 comes from objects {\\tt AUTO\\_048} and {\\tt AUTO\\_100}. Based on their photometry in $J_{125}$ and assuming $z=8$, they have $M_{UV}$ of $-21.54$ and $-21.21$, respectively. While such high luminosities have been observed in the spectroscopically confirmed $z\\approx 7$ galaxies of Ono et al. (2012), they are 5--9$\\times$ more luminous than any previously known $Y_{105}$-dropouts. Such an excess is unlikely due to cosmic variance, whose effect is expected to be small compared to the Poisson errors for our observed number density if we use the formalism of Trenti \\& Stiavelli (2008). We note that, however, that O12 do not see such an excess, even though their sample also includes our brightest candidate {\\tt AUTO\\_048}. The main reason for this discrepancy could be that O12 use a larger volume in the calculation. They have included part of the CANDELS Wide field in the GOODS-S region, which is a major contributor to their larger volume. However, this area does not have sufficiently deep $Y_{105}$ data to apply our selection criterion of $Y_{105}-J_{125}\\geq 0.80$~mag down to $J_{125}\\approx 26.0$~mag. O12 have adopted a less stringent criterion of $Y_{105}-J_{125}\\geq 0.45$~mag, which allows them to construct a $Y_{105}$-dropout sample in this area at the cost of including candidates at lower redshifts ($z\\approx 7.2$). The advantage of their approach is that they can incorporate the WFC3 ERS field in the GOODS-S region as well, which further increases their survey volume. While the WFC3 ERS field has $Y_{098}$ but not $Y_{105}$ data, the redshift range probed by their $Y_{098}$-dropouts in this area overlaps that of their $Y_{105}$-dropouts elsewhere, and therefore O12 has combined the $Y_{098}$- and the $Y_{105}$-dropouts to address the LF averaged over a wide redshift range ($z\\approx 7$--8.7). In contrast, our color criteria select galaxies in a higher range of redshifts, $7.7\\lesssim z\\lesssim 8.7$, and the bright-end excess (with respect to a smooth Schechter function) that we find suggests that our understanding of the LF at $z\\approx 8$ is still far from being complete. The significance of this excess is still subject to small number statistics, and we cannot rule out the possibility that one or both of our brightest objects could be interlopers at low redshifts. However, we should point out that similar bright-end excess at $z\\sim 8$ has also been suggested in other studies, for example in the bright $Y_{098}$-dropout study in HIPPIES (Yan et al. 2011) at similar depths, in the ground-based survey of Laporte et al. (2011) at a brighter level, and in the much wider survey in Capak et al. (2011) at an even brighter limit \\footnote{While Bowler et al. (2012) derived different $z_{ph}$ ($\\sim 2$) for the candidates in Capak et al. (2011), the true nature of these objects still remains uncertain.}. \\subsection{Implications for Star Formation in the Early Universe } Our $z\\approx 8$ candidate galaxy samples have two very luminous objects and two high stellar mass objects. We have presented tentative evidence that the bright end of the galaxy LF at $z\\approx 8$ might not follow the exponential cut-off of the Schechter function. While the stellar mass function at this redshift has not yet been determined, we have provided the first, albeit still tentative measurement at the high-mass end and shown that some galaxies with stellar masses as high as a few $\\times 10^9$~$M_\\odot$ might already in place at $z\\approx 8$. The physical origin of the exponential cut-off at the bright end of the LF or the high mass end of the MF seen at lower redshifts remains a matter of debate. At $z\\sim 0$, star formation is ``quenched'' in halos with masses greater than $\\sim 10^{12}$~$M_\\odot$ (``quenching mass''), perhaps by AGN feedback (e.g., Croton et al. 2006; Somerville et al. 2008; Gabor et al. 2011). However, at higher redshift ($z\\lesssim 2$), the quenching mass may be higher (Dekel et al. 2009; Behroozi et al. 2010), and the existence of massive, rapidly star forming galaxies at $z>2$ is well established. At $z\\sim 2$, the observed exponential cut-off in the rest-frame UV LF appears to be due to dust (Reddy et al. 2010) rather than quenching. At very high redshifts, it has been noted before that the very blue rest-frame UV colors of $z\\approx 7$ candidate galaxies (some of them confirmed) suggest that they may contain little dust (e.g., Oesch et al. 2010; M10; Y10; Bunker et al. 2010). Thus, at $z\\approx 8$, when (1) the quenching mass is much higher than the typical halo mass, and (2) dust has little effect, perhaps we should not expect to see an exponential cutoff in the luminosity or mass functions of UV-bright galaxies. It is interesting that the estimated stellar masses and number densities could imply a rather high efficiency of conversion of baryons into stars. For the lower stellar mass estimates, which are obtained using the models with nebular emission (the average is $\\sim 2\\times 10^9$~$M_\\odot$), and assuming that $\\sim 20$\\% of the available baryons have been converted to stars, the implied halo masses are on the order of $\\sim 10^{11}$~$M_\\odot$, for which the expected number density of dark matter halos in the currently favored LCDM cosmology is comparable to the observed number density of objects. This is consistent with the predicted star formation efficiencies and host halo masses at $z\\approx 8$ from cosmological hydrodynamic simulations. However, if the true stellar masses are higher, or the star formation efficiency is lower, a rapidly growing tension arises between the number density of dark matter halos and the observed number density of galaxies (above halo mass $\\sim 10^{11}$~$M_\\odot$, the halo number density declines by about two orders of magnitude for a factor of $\\sim 3$ increase in mass). While converting 20\\% of the available baryons into stars may not sound excessive, this is in fact the \\emph{maximum} value that has been inferred at any epoch. Due to the presumably very low metallicity of the gas in these early objects, we might have expected much lower star formation efficiencies than are seen locally (e.g. Krumholz \\& Dekel 2010)." }, "1112/1112.3824_arXiv.txt": { "abstract": "{Recently, for the first time the abundance of P has been measured in disk stars. This provides the opportunity of comparing the observed abundances with predictions from theoretical models.} {We aim at predicting the chemical evolution of P in the Milky Way and compare our results with the observed P abundances in disk stars in order to put constraints on the P nucleosynthesis.} {To do that we adopt the two-infall model of galactic chemical evolution, which is a good model for the Milky Way, and compute the evolution of the abundances of P and Fe. We adopt stellar yields for these elements from different sources. The element P should have been formed mainly in Type II supernovae. Finally, Fe is mainly produced by Type Ia supernovae.} {Our results confirm that to reproduce the observed trend of [P/Fe] vs. [Fe/H] in disk stars, P is formed mainly in massive stars. However, none of the available yields for P can reproduce the solar abundance of this element. In other words, to reproduce the data one should assume that massive stars produce more P than predicted by a factor of $\\sim$ 3. } {We conclude that all the available yields of P from massive stars are largely underestimated and that nucleosynthesis calculations should be revised. We also predict the [P/Fe] expected in halo stars. } ", "introduction": "Recently, Caffau et al. (2011a) have measured the P abundance in a sample of 20 cool stars in the Galactic disk. They found that the [P/Fe] ratio behaves like the [S/Fe] ratio, namely increasing towards lower metallicity stars. This was the first time that P was observed in Galactic stars in spite of the fact that P is among the top 20 most abundant elements in the Universe. There is only one single stable isotope of P, $^{31}$P, which is thought to be formed by neutron capture on $^{29}$Si and $^{30}$Si in massive stars. By means of a detailed model for the chemical evolution of the Milky Way one can compute the P evolution and compare it to the observations. This can allow us to understand the origin of this element and impose constraints on its formation inside stars. In this paper, we adopt a good model for the chemical evolution on the Galaxy, already tested on a large number of chemical species (Fran\\c cois et al. 2004). In particular, the model takes into account detailed stellar nucleosynthesis and supernova progenitors (SNe Ia, II, Ib/c) as well as the stellar lifetimes. The stellar yields of P have been computed by several authors such as Woosley \\& Weaver (1995), Kobayashi et al. (2006) as functions of the initial stellar metallicity. In this paper we like to compare the predictions of our chemical evolution model for P, obtained by including different stellar yields, with the recent data. In Section 2 we briefly describe the chemical evolution model adopted, in Section 3 we describe the observational data. In Section 4 the results are compared to the data and finally, in Section 5, some conclusions are drawn. ", "conclusions": "In this paper we have compared model predictions adopting different sets of yields with recent data on the abundance of P in the Galactic disk. Our conclusions can be summarized as follows: \\begin{itemize} \\item The observed fall of the abundance of [P/Fe] in Galactic disk stars suggests that P is mainly produced by core-collapse SNe with a small contribution from SNe Ia. Otherwise, if SNe Ia were important the [P/Fe] ratio would remain fairly constant. \\item The metallicity dependent yields of P from massive stars of K06 together with the P yields from SNe Ia from Iwamoto et al. (1999) well reproduce the data if the yields from massive stars are increased by a factor of $\\sim$ 3. This suggests that the yields of P available in the literature are underestimated. Both the neutron rich isotopes of $^{29,30}$Si and $^{31}$P, which derives from neutron capture on the two Si isotopes, are produced in the oxygen and neon burning shells in massive stars (WW95). Therefore, in order to have a larger P production, the O and Ne shell- burnings as well as the neutron capture on the $^{29,30}$Si isotopes and the destruction of $^{31}$P by (p, $\\alpha$) reactions should be revised. \\item We predict also the behaviour of [P/Fe] in halo stars and we suggest that it should show a plateau between [Fe/H] =-1.0 and -3.0 dex corresponding to [P/Fe] $\\sim$ +0.5 dex if yields (corrected) from normal SNe II are adopted, and to $\\sim$+0.2dex if hypernova yields (corrected) are adopted. In order to distinguish between these two cases we await for observations of the P abundance in halo stars. \\end{itemize}" }, "1112/1112.5319_arXiv.txt": { "abstract": "{We used the new IRAM~30-m FTS backend to perform an unbiased $\\sim$15~GHz wide survey at 3~mm toward the \\pipe\\ young diffuse starless cores. We found an unexpectedly rich chemistry. We propose a new observational classification based on the 3~mm molecular line emission normalized by the core visual extinction (\\Av). Based on this classification, we report a clear differentiation in terms of chemical composition and of line emission properties, which served to define three molecular core groups. The ``diffuse'' cores, \\Av$\\la$15, show poor chemistry with mainly simple species (e.g. \\cs\\ and \\cdh). The ``oxo-sulfurated'' cores, \\Av$\\simeq$15--22, appear to be abundant in species like \\so\\ and \\sod, but also in HCO, which seem to disappear at higher densities. Finally, the ``deuterated'' cores, \\Av$\\gsim$22, show typical evolved chemistry prior to the onset of the star formation process, with nitrogenated and deuterated species, as well as carbon chain molecules. Based on these categories, one of the ``diffuse'' cores (Core~47) has the spectral line properties of the ``oxo-sulfurated'' ones, which suggests that it is a possible failed core.} ", "introduction": "A new generation of sensitive receivers and wideband backends allows to study in detail the chemistry of faint starless cores. Several surveys have been performed toward them reporting a rich but relatively simple chemistry: essentially carbon chemistry with significant sulfur and nitrogen bearing molecules, followed by later deuteration which can be used as a chemical clock (e.g., \\citealp{turner94,turner00,hirota06,tafalla06,bergin07}). Recently, from the theoretical side, several papers have tried to model the starless core chemistry self-consistently \\citep{aikawa01,garrod05,keto08}. The \\pipe\\ is a nearby (145~pc: \\citealp{alves07}) cloud that harbors more than one hundred of low mass ($\\sim$1~M$_{\\odot}$) starless cores, most of them gravitational unbound but confined by the thermal/magnetic pressure of the whole cloud \\citep{alves08,lada08,franco10}. The \\pipe\\ differs from the other nearest dark cloud complexes such as Taurus or $\\rho$~Ophiuchus because it has a very small star formation efficiency \\citep{onishi99,forbrich09,roman10,roman11}. Thus, the \\pipe\\ is an ideal target to study the physical and chemical conditions in a pristine environment prior to the onset of the star formation process, as the recent numerous studies have shown (e.g., \\citealp{brooke07,muench07,rathborne08}). \\citet{frau10} present the first results of an extensive continuum and molecular line study on a subset of a selected sample of cores distributed in the different regions of the \\pipe: {\\it bowl}, {\\it stem}, and B59. The cores are in general less dense and less chemically evolved than starless cores in other star forming regions studied (e.g. \\citealp{crapsi05}). We find very different morphologies and densities, and no clear correlation of the chemical evolutionary stage of the cores with their location in the cloud. The \\pipe\\ starless cores have shown to be more heterogeneous than expected. In this work, we present a wide ($\\sim$15~GHz) unbiased chemical survey at 3~mm toward a larger sample of \\pipe\\ starless cores, spanning a factor of 6 in their visual extinction (\\Av) peaks. This is a first step to characterize their varied chemistry in order to proceed to future modeling. ", "conclusions": "} The chemistry detected toward the sample of fourteen starless cores is unexpectedly rich taking into account their low temperatures (10--15 K: \\citealp{rathborne08}) and visual extinctions. The apparent correlation within the sample of the 3~mm molecular transition normalized intensities to visual extinction allow us to propose an observational classification (see Fig.~\\ref{fig_spec}). We define three groups of starless cores, which are probably related with their dynamical age: {\\it ``diffuse''}, {\\it ``oxo-sulfurated''}, and {\\it ``deuterated''} cores. This classification can be useful in future wide band 3~mm observations of molecular clouds. {\\it -- ``Diffuse'' cores}: a set of cores with small column densities (\\Av$\\lsim$15~mag $\\sim$ $N_{\\rm H_2}$$\\lsim$1.2$\\times$10$^{22}$~cm$^{-2}$) lies above the blue dot-dashed horizontal line in Fig.~\\ref{fig_spec}. Their spectra is rather poor, showing only significant normalized intensity in the transitions of the main isotopologues of abundant species like \\cdh, HCN (and likely HNC), \\hcom\\ and SO. Such a simple observational chemistry suggests that these are very young starless cores, or even transient clumps on which essentially the cloud chemistry is better detected due to density enhancements. Core~47 is a clear exception and it is discussed later in the text. {\\it -- ``Oxo-sulfurated'' cores}: a group of denser cores (\\Av$\\simeq$15--22~mag $\\sim N_{\\rm H_2} \\simeq$1.2$\\times$10$^{22}$--1.7$\\times$10$^{22}$~cm$^{-2}$) that show richer chemistry but not yet significant deuteration to be observed. In Fig.~\\ref{fig_spec} this group lie between the blue dot-dashed and red dashed horizontal lines. All the transitions detected in the {\\it ``diffuse''} cores are also present. The \\so~3$_2$--2$_1$ transition is the main signpost as it is very bright. Many other oxo-sulfurated molecules (\\tso, \\sod, and \\ocs), as well as \\hco, exhibit a same trend, but they are not detected at higher densities. This suggests an increase of these chemically related species in the gas-phase in this \\Av\\ range, followed by a later depletion/destruction as density increases. These cores might be in-the-making cores, which have developed a richer chemistry and piled up more material, probably in a stage near to the onset of collapse \\citep{ruffle99}. Core~102 is an exception in this group and is discussed later in the text. {\\it -- ``Deuterated'' cores}: the densest cores of the sample (\\Av$\\gsim$22~mag $\\sim$ $N_{\\rm H_2}$$\\gsim$1.7$\\times$10$^{22}$~cm$^{-2}$), shown below the red dashed horizontal line in Fig.~\\ref{fig_spec}. Core~12, the densest one, sets the upper limit at \\Av=67.2~mag ($N_{\\rm H_2}$$\\simeq$5.3$\\times$10$^{22}$~cm$^{-2}$). These cores are generally bright in the transitions typical of the other two groups. The oxo-sulfurated molecules are the exception, hardly present, probably depleted/destructed at the densities reached. The main signpost are the emission, only present in this group, in rare isotopologues of the nitrogenated ubiquitous lines (\\htcn, \\hcqn, \\hntc, and \\hqnc), deuterated forms of abundant species (\\cthdt, \\nhdd, and \\dctn), and carbon-chain molecules (\\cqh\\ and \\chtcdh). These cores might be stable starless cores with a life-time long enough to achieve the densities needed to synthesize efficiently carbon chains and deuterated species \\citep{roberts00,gwenlan00}. As already said, Core~47 does not match the chemical properties of the diffuse cores. It shows a chemistry matching the oxo-sulfurated group, which proved to be very sensitive to density. This suggests that it might be a failed core which developed a rich chemistry and is now merging back into the cloud. This scenario can increase the oxo-sulfurated chemistry detected \\citep{garrod05}. Core~47 is located close to Core~48 in the only \\pipe\\ region with superalfv\\'enic turbulence, as shown by optical polarization observations \\citep{franco10}. Therefore, it is possible that an external source of turbulence is disrupting the medium in this area and dispersing the cores. On the other hand, Core~102, in the oxo-sulfurated group, shows a chemistry similar to that of the diffuse cores. Similarly, Core~87, among the deuterated cores, shows features of the oxo-sulfurated group. This suggests that these cores might have piled up material so quickly that a more complex chemistry had no time to be synthesized. Both cores lie in the same N-S oriented high-density structure \\citep{roman10} where \\citet{franco10} report a N-S magnetic field. The fast evolution might have been driven by magnetic fields with the surrounding mass collapsing in this direction. The FTS chemical survey toward the starless cores of the \\pipe\\ showed a chemistry much more rich than expected for a cloud giving birth to low-mass stars at very low efficiency. A good interpretation of the results demands chemical modeling to investigate a possible evolutive track, which will be the purpose of a forthcoming study." }, "1112/1112.0016_arXiv.txt": { "abstract": "VFTS 682, a very massive and very hot Wolf-Rayet (WR) star recently discovered in the Large Magellanic Cloud near the famous star cluster R136, might be providing us with a glimpse of a missing link in our understanding of Long Gamma-Ray Bursts (LGRBs), including dark GRBs. It is likely its properties result from chemically homogeneous evolution (CHE), believed to be a key process for a massive star to become a GRB. It is also heavily obscured by dust extinction, which could make it a dark GRB upon explosion. Using {\\em Spitzer}\\/ data we investigate the properties of interstellar dust in the vicinity of R136, and argue that its high obscuration is not unusual for its environment and that it could indeed be a slow runaway (``walkaway'') from R136. Unfortunately, based on its current mass loss rate, VFTS 682 is unlikely to become a GRB, because it will lose too much angular momentum at its death. If it were to become a GRB, it probably would also not be dark, either escaping or destroying its surrounding dusty region. Nevertheless, it is a very interesting star, deserving further studies, and being one of only three presently identified WR stars (two others in the Small Magellanic Cloud) that seems to be undergoing CHE. ", "introduction": "Ideally, in order to fully understand the mapping between massive star progenitors and when and how they explode (or not, \\citealt{Kochanek08}), we would have extensive multi-wavelength data obtained for many such explosions both BEFORE and AFTER the event. The AFTER part has certainly undergone an explosive growth in the last decade or so, due to many successful supernova (SN) searches such as the galaxy-targeted {\\em Lick Observatory Supernova Search} (LOSS), the {\\em Catalina Real-time Transient Survey} (CRTS), the {\\em Robotic Optical Transient Search Experiment} (ROTSE), the {\\em Palomar Transient Factory} (PTF); and gamma-ray burst searches such as the {\\em High Energy Transient Explorer-2 (HETE-2)} and {\\em Swift}. The BEFORE part is naturally limited by the fact that massive stars, while relatively bright, are only observable in the local Universe ($d\\lesssim10\\;$Mpc). Here we have had to rely on proximity (SN 1987A, \\citealt{SN1987A}), luck based on archival data (e.g., \\citealt{Smartt09}), or the first systematic campaign to monitor future SN progenitors (\\citealt{Szczygiel11}), where many nearby galaxies are observed to sufficient depth so eventually they will provide progenitor data for a significant number of future SNe. We have by now compiled a fair amount of information on SN progenitors, such as the blue supergiant progenitor for SN 1987A (\\citealt{White87}) or dusty progenitors of SN 2008S and 2008 NGC 300 transient (\\citealt{Prieto08}; \\citealt{Prieto08b}). However, for very rare explosive events, such as long-duration gamma-ray bursts (LGRBs), the systematic approach that works for the normal core-collapse SNe is not yet possible because the events are so rare. Therefore the prospect of future systematic observational efforts that could identify nearby LGRB progenitors is very dim. Here we have to rely on a more extended chain of reasoning, with data being supplemented with reasonable, theoretical guesses. Discovery of the connection between LGRBs and broad-lined Type Ic SNe (e.g., \\citealt{Stanek03}; \\citealt{Hjorth03}), thought to result from core-collapse of hydrogen-free massive stars, favors two possible progenitor models: single massive Wolf-Rayet (WR) stars with rapidly rotating cores (\\citealt{FS85}), or lower mass helium stars stripped by a close binary companion (\\citealt{Podsiadlowski04}). The prompt emission of an initial LGRB, lasting seconds or minutes, is usually followed by a multi-wavelength afterglow which lasts days to even years. Although X-ray afterglows of LGRBs are nearly always detected by {\\em Swift} and {\\em Fermi}, detection of optical and infrared afterglows is less common. A dark GRB is defined by either an absent or faint optical afterglow relative to its X-ray emission. The rapid detection of X-ray afterglows with {\\em Swift} revealed that the dark fraction of LGRBs is about $30\\%$ of GRBs (\\citealt{AS07}; \\citealt{Perley09}). In general, besides the intrinsically optically faint GRBs, the optical attenuation of dark GRBs can be caused by dust extinction in GRB host galaxies, foreground extinction, or Lyman-$\\alpha$ absorption by neutral hydrogen at high redshifts (\\citealt{Perley09}). Study of massive stars in very nearby galaxies can supply the other missing links in our understanding of LGRBs. The recently uncovered very massive stars up to $300M_{\\odot}$ in the Large Magellanic Cloud (LMC) young cluster R136 extend our knowledge of massive star formation and evolution (\\citealt{Crowther10}). In particular, the newly discovered very massive WR star VFTS 682, located 30 pc away from R136, drew our attention, mainly because of its very high foreground dust extinction and possibility of being a GRB progenitor (\\citealt{VFTS682}). Its unusually high effective temperature can be understood as the consequence of chemically homogeneous evolution (CHE), which was proposed as the crucial part of the process of producing LGRBs (\\citealt{Yoon05}). If VFTS 682 will indeed make a GRB at the end of its evolution, the high foreground dust extinction could mean that it will become a dark GRB. However, its high mass and strong mass-loss seems to prevent it from being a GRB progenitor, and the potential afterglow might destroy the surrounding dust even if VFTS 682 eventually produces a GRB. Therefore, it is worthwhile to look into the details of VFTS 682's evolution and the fate of its circumstellar environment. Section \\ref{Data} describes the dust properties in the VFTS 682 vicinity region. Section \\ref{Progenitor} investigates the possibility of VFTS 682 producing a LGRB and Section \\ref{Dark} discusses the fate of dusty clouds around VFTS 682. Conclusions are presented in Section \\ref{Con}. ", "conclusions": "\\label{Con} We give a first order estimate whether VFTS 682 can be a LGRB progenitor in the sense of CHE, which is still a question mark in previous work. VFTS 682 will most likely lose more than 85\\% of its current angular momentum and cease its CHE evolution before it leaves MS with a remaining mass $\\sim50-70M_{\\odot}$. In general VFTS 682 will fail to produce a GRB due to its strong mass-loss and the resulting angular momentum loss, unless it has a currently extreme rapid rotation ($v_{0}>700$ km s$^{-1}$) which helps it maintain CHE and eventually produce a GRB associated with a hypernova. Wind anisotropies and shorter lifetime could leave a more massive faster rotating star, but it is doubtful whether a relativistic jet could travel through such a thick stellar envelope and break out of the stellar surface, unless it is a rare long-lived ($\\gtrsim100$ s) LGRBs. Similarly, it is unlikely that VFTS 682 will be heavily obscured at death and produce a dark GRB. Its proper motion will probably cause it out of the dusty region, and a GRB of an early optical afterglow $L_{\\rm opt}>10^{50}$ erg s$^{-1}$ would destroy the dust clouds within 30 pc, otherwise the dusty clouds can be heated and ionized up to a region of $\\sim$ 100 pc by X-ray and optical radiation given by the death of VFTS 682. CHE is believed to be the crucial process in the evolution path towards LGRBs. However, the observed sample of WRs likely undergoing CHE is quite small. The only observation of WR stars besides VFTS 682 which might be undergoing are two WNh stars in the SMC (i.e., SMC-WR1 and WR2 in \\citealt{Martins09}), which makes these three WR stars extremely important to understanding the evolution of WR stars undergoing CHE and the related problem of LGRB formation. The angular momentum losses of the two WR stars in the SMC will be much less significant than VFTS 682 in the LMC. Taking the stellar parameters in \\cite{Martins09}, we estimate that the nuclear timescale of SMC-WR1 (WR2) $\\tau_{\\rm nuc}\\sim4$ Myr (5 Myr) is shorter than the wind timescale $\\tau_{\\rm wind}\\sim10-20$ Myr, and the final stellar rotation velocity should be 90\\% of the current velocity. Therefore, the two WR stars in the SMC are more likely GRB progenitors in the scenario of CHE, although the metallicity threshold is still an issue (\\citealt{Martins09}). In any case, finding CHE WR stars in the LMC and SMC promotes a future work on theory models. VFTS 682 is a very interesting star. As mentioned in Section 2, because there is no observation evidence to show that VFTS 682 is unusual compared to other stars in the 30 Dor region, the interesting possibility of existence of others massive stars like VFTS 682 in the vicinity of R136 is not excluded. Note that R136 is sufficient young and massive ($\\leq5.5\\times10^{4}M_{\\odot}$) to generate runaway stars beyond $150M_{\\odot}$. Recently \\cite{Banerjee11} gives a theoretical model study on the dynamical ejection of runaway massive stars from R136. We suggest that there might be other massive stars that ``walked away'' from R136, but are currently hidden behind even more dust than VFTS 682. Since mid-IR date cannot be used to flag such stars, spectroscopic observations, which are beyond the scope of this Letter, should be further investigated to show the possibility of their existence." }, "1112/1112.1589_arXiv.txt": { "abstract": "{Generation and dissipation of magnetic fields is a fundamental physical process on the Sun. In comparison to flux emergence and the initial stages of sunspot formation, the demise of sunspots still lacks a comprehensive description. } {The evolution of sunspots is most commonly discussed in terms of their intensity and magnetic field. Here, we present additional information regarding the three-dimensional flow field in the vicinity of sunspots towards the end of their existence.} {We present a subset of multi-wavelengths observations obtained with the Japanese \\textit{Hinode} mission, the \\textit{Solar Dynamics Observatory} (SDO), and the \\textit{Vacuum Tower Telescope} (VTT) at \\textit{Observatorio del Teide}, Tenerife, Spain during the time period from 2010 November 18--23. Horizontal proper motions were derived from G-band and Ca\\,\\textsc{ii}\\,H images, whereas line-of-sight velocities were extracted from VTT Echelle H$\\alpha$~$\\lambda656.28$~nm spectra and Fe\\,\\textsc{i} $\\lambda630.25$~nm spectral data of the \\textit{Hinode/Spectro-Polarimeter}, which also provided three-dimensional magnetic field information. The \\textit{Helioseismic and Magnetic Imager} on board SDO provided continuum images and line-of-sight magnetograms as context for the high-resolution observations for the entire disk passage of the active region.} {We have performed a quantitative study of photospheric and chromospheric flow fields in and around decaying sunspots. In one of the trailing sunspots of active region NOAA~11126, we observed moat flow and moving magnetic features (MMFs), even after its penumbra had decayed. We also noticed a superpenumbral structure around this pore. MMFs follow well-defined, radial paths from the spot all the way to the border of a supergranular cell surrounding the spot. In contrast, flux emergence near the other sunspot prevented it from establishing such well ordered flow patterns, which could even be observed around a tiny pore of just 2~Mm diameter. After the disappearance of the sunspots/pores a coherent patch of abnormal granulation remained at their location, which was characterized by more uniform horizontal proper motions, low divergence values, and diminished photospheric Doppler velocities. This region, thus, differs significantly from granulation and other areas covered by G-band bright points. We conclude that this peculiar flow pattern is a signature of sunspot decay and the dispersal of magnetic flux.} {} ", "introduction": "Sunspots are a thought-provoking aspect of solar activity because of the close interaction between plasma motions and magnetic fields. Recent progress in MHD simulations \\citep[e.g.,][]{Rempel2011} provide a comprehensive framework for the interpretation of high-resolution sunspot observations. The formation of a penumbra around a sunspot is a rapid phenomenon, i.e., within a few hours a sunspot can develop a penumbra \\citep{Leka1998, Yang2003a}, which is intimately linked to more inclined magnetic field lines and the onset of the Evershed flow. \\citet{Schlichenmaier2010b} observed the growth of a penumbra where only the newly formed penumbra contributed to the increase in spot size while the umbra remained stable. The formation of a penumbra, which would surround the entire spot, was hindered by continuous flux emergence between the spots of the bipolar region. Quite the opposite, the decay of a sunspot is a slow process. Decay rates for stable leading sunspots and irregular follower spots are different \\citep{MartinezPillet2002}. A number of decay laws were proposed such as a linear decay law described by \\citet{Bumba1963} and a parabolic decay law proposed by \\citet{Petrovay1997}. \\citet{MartinezPillet2002} critically reviews various diffusion models, concludes that they explain well how flux is spread over larger areas while the spot is decaying, but they fail to satisfactorily describe the flux removal process. The initial stages of sunspot decay, i.e., when the spot looses its penumbra, are exemplary described in \\citet{BellotRubio2008}, who discovered finger-like structures, which are neither related to penumbral filaments nor the Evershed flow. These features might be penumbral field lines rising to the chromosphere, thus contributing to the decay of the sunspot penumbra. When a sunspot looses its penumbra, its decay reaches a critical point. Magnetic field lines become more vertical and convective motions in its vicinity begin to change. These ideas of a critical inclination angle and convective motions were put together by \\citet{Rucklidge1995}, who explains in a simple model why small sunspots can have a penumbra while larger pores do not possess one. The moat flow is a large-scale flow pattern commonly observed around sunspots \\citep{Meyer1974}. However, flux removal and dispersal can only be understood in the context of the moat flow's fine structure. Moving magnetic features (MMFs) play a major role in the flux dispersal process and they are only associated with decaying sunspots \\citep{Harvey1973}. The total flux carried by MMFs is several times larger than the flux contained within the sunspot itself. Thus, the polarity of MMFs has to be considered for a balance of the net flux. MMFs move radially outward with a velocity of 1~km~s$^{-1}$ before they reach and dissolve within the network, i.e., at the boundaries of the supergranular cell containing the sunspot. \\citet{Zuccarello2009} showed evidence that MMFs and moat flow are present even in the vicinity of pores, i.e., in the absence of penumbral filaments and Evershed flow \\citep[cf.,][]{CabreraSolana2006}. \\citet{Deng2007} also detected a persistent moat flow after the penumbra around spot disappeared leaving only a pore. Even though the moat flow might not be closely tied to the Evershed flow, the sub-photospheric interaction of magnetic field lines and flows can still produce the observed flow patterns. \\citet{Verma2011} described a local correlation tracking (LCT) method to measure horizontal flows based on \\textit{Hinode} G-band images. In this study, we perform a case study, where we put such horizontal flow fields in the context of other photospheric and chromospheric data. In particular, we are interested in the final stages of sunspot decay. In Sect.~\\ref{SEC02}, we present a subset of multi-wavelengths observations, which were obtained within the scope of \\textit{Hinode Operation Plan} (HOP) 0176. The temporal evolution of active region NOAA~11126 in terms of intensity, morphology, and flow as well as magnetic fields is described in Sect.~\\ref{SEC03} and discussed in Sect.~\\ref{SEC04}. \\begin{table*}[t] \\caption{Observing characteristics and physical parameters}\\label{TAB01} \\tiny \\begin{tabular}{llcccccp{65mm}} \\cline{1-7}\\vspace*{-2.5mm}\\\\\\cline{1-7} \\multicolumn{2}{l}{November} & 18 & 19 & 20 & 21 & 22 & \\rule[-3pt]{0pt}{12pt}\\\\ \\cline{1-7} $B$ & & S32.6$^{\\circ}$ & S32.6$^{\\circ}$ & S32.6$^{\\circ}$ & S32.6$^{\\circ}$ & S32.6$^{\\circ}$ & \\rule{0pt}{9pt}\\\\ $L$ & & E5.5$^{\\circ}$ & W6.3$^{\\circ}$ & W20.5$^{\\circ}$ & W31.5$^{\\circ}$ & W44.1$^{\\circ}$ & \\\\ $\\mu$ & & $0.81$ & $0.81$ & $0.77$ & $0.70$ & $0.58$ & \\\\ $t_{0,\\mathrm{GB}}$ & UT & 10:23 & 09:05 & 09:07 & 09:05 & 09:05 & \\\\ $t_{0,\\mathrm{SP}}$ & UT & 11:46 & 09:05 & 09:07 & 09:04 & 09:04 & \\\\ $t_{0,\\mathrm{ES}}$ & UT & 10:22 & 08:59 & 14:16 & 09:31 & 08:47 & \\rule[-4pt]{0pt}{6pt}\\\\ \\cline{1-7} $A_\\mathrm{Spot\\ A}$ & Mm$^{2}$ & 33.8 & 21.5 & & & & \\rule{0pt}{9pt}\\\\ $v_\\mathrm{Spot\\ A}$ & km s$^{-1}$ & $0.22 \\pm 0.19$ & $0.15 \\pm 0.15$ & & & & \\\\ & & $0.20 \\pm 0.25$ & $0.15 \\pm 0.14$ & & & & \\\\ $v_\\mathrm{Ring\\ A}$ & km s$^{-1}$ & $0.44 \\pm 0.23$ & $0.37 \\pm 0.19$ & & & & \\\\ & km s$^{-1}$ & $0.44 \\pm 0.23$ & $0.38 \\pm 0.19$ & & & & \\\\ $\\left|\\nabla v\\right|_\\mathrm{Ring\\ A}$ & $10^{-3}$ s$^{-1}$ & $0.83 \\pm 0.78$ & $0.65 \\pm 0.54$ & & & & \\\\ & & $1.28 \\pm 1.33$ & $1.13 \\pm 1.06$ & & & & \\rule[-4pt]{0pt}{6pt}\\\\ \\cline{1-7} $A_\\mathrm{Spot\\ B}$ & Mm$^{2}$ & 54.7 & 21.6 & 54.2 & 17.5 & 2.7 & \\rule{0pt}{9pt}\\\\ $v_\\mathrm{Spot\\ B}$ & km s$^{-1}$ & $0.22 \\pm 0.15$ & $0.15 \\pm 0.15$ & $0.16 \\pm 0.07$ & $0.18 \\pm 0.05$ & $0.28 \\pm 0.06$ & \\\\ & & $0.20 \\pm 0.14$ & $0.11 \\pm 0.05$ & $0.15 \\pm 0.08$ & $0.25 \\pm 0.06$ & $0.37 \\pm 0.06$ & \\\\ $v_\\mathrm{Ring\\ B}$ & km s$^{-1}$ & $0.31 \\pm 0.19$ & $0.20 \\pm 0.12$ & $0.28 \\pm 0.14$ & $0.25 \\pm 0.14$ & $0.30 \\pm 0.09$ & \\\\ & & $0.32 \\pm 0.20$ & $0.20 \\pm 0.12$ & $0.29 \\pm 0.15$ & $0.28 \\pm 0.13$ & $0.40 \\pm 0.09$ & \\\\ $\\left|\\nabla v\\right|_\\mathrm{Ring\\ B}$ & $10^{-3}$ s$^{-1}$ & $0.63 \\pm 0.68$ & $0.37 \\pm 0.36$ & $0.46 \\pm 0.49$ & $0.47 \\pm 0.51$ & $0.28 \\pm 0.33$ & \\\\ & & $0.91 \\pm 1.20$ & $0.47 \\pm 0.59$ & $0.76 \\pm 1.04$ & $0.63 \\pm 0.81$ & $0.44 \\pm 0.44$ & \\rule[-4pt]{0pt}{6pt}\\\\ \\cline{1-7} $A_\\mathrm{mag}$ & Mm$^{2}$ & $136.9$ & $83.5$& $98.9$ & $44.7$ & $22.2$ & \\rule{0pt}{9pt}\\\\ $N_\\mathrm{mag}$ & & 10 & 5 & 9 & 6 & 2 & \\\\ $v_\\mathrm{mag}$ & km s$^{-1}$ & $0.22 \\pm 0.12$ & $0.15 \\pm 0.09$ & $0.16 \\pm 0.07$ & $0.18 \\pm 0.05$ & $0.28 \\pm 0.06$ & \\\\ $v_\\mathrm{bp}$ & km s$^{-1}$ & $0.27 \\pm 0.19$ & $0.27 \\pm 0.16$ & $0.28 \\pm 0.16$ & $0.23 \\pm 0.12$ & $0.26 \\pm 0.10$ & \\\\ $v_\\mathrm{gran}$ & km s$^{-1}$ & $0.40 \\pm 0.24$ & $0.35 \\pm 0.20$ & $0.37 \\pm 0.20$ & $0.34 \\pm 0.19$ & $0.34 \\pm 0.18$ & \\\\ $\\left|\\nabla v\\right|_\\mathrm{gran}$ & $10^{-3}$ s$^{-1}$ & $0.88 \\pm 0.89$ & $0.72 \\pm 0.67$ & $0.85 \\pm 0.79 $ & $0.95 \\pm 0.88$ & $1.04\\pm 0.98$ & \\\\ $\\left|v\\right|_\\mathrm{mag, LOS}$ & km s$^{-1}$ & $0.03 \\pm 0.17$ & $0.02 \\pm 0.12$ & $0.04 \\pm 0.20$ & $0.02 \\pm 0.12$ & $0.00 \\pm 0.02$ &\\\\ $\\left|v\\right|_\\mathrm{bp, LOS}$ & km s$^{-1}$ & $0.43 \\pm 0.37$ & $0.44 \\pm 0.37$ & $0.44 \\pm 0.38$ & $0.41 \\pm 0.38$ & $0.34 \\pm 0.30$ &\\\\ $\\left|v\\right|_\\mathrm{gran, LOS}$ & km s$^{-1}$ & $0.51 \\pm 0.41$ & $0.50 \\pm 0.41$ & $0.55 \\pm 0.45$ & $0.57 \\pm 0.47$ & $0.55 \\pm 0.45$ &\\rule[-4pt]{0pt}{6pt} \\raisebox{23mm}[-23mm]{\\parbox{65mm}{ The parameters in the first column of the table refer to heliographic latitude $B$, heliographic longitude $L$, cosine of the heliocentric angle $\\mu$, start of the G-band (GB), spectro-polarimeter (SP), and Echelle spectrograph (ES) observing sequences $t_{0}$, spot area $A$, horizontal velocity $v$, mean divergence $\\left|\\nabla v\\right|$, number of magnetic elements $N_\\mathrm{mag}$. The indices refer to G-band bright points (bp), granulation (gran), magnetic elements (mag), the two sunspots (Spot~A and Spot~B), and the four-megameter wide annuli around both spots (Ring~A and Ring~B). If two rows are given for a physical parameter, then the top row refers to G-band data, whereas the bottom row was derived from Ca\\,\\textsc{ii}\\,H data. If present, the standard deviation refers to the variation of the physical parameters within the specified regions rather than to any formal error.}}\\\\ \\cline{1-7} \\end{tabular} \\end{table*} ", "conclusions": "} We have presented a detailed account of the final stages in the decay of the active region NOAA~11126, which did not obey the Hale-Nicholson polarity law \\citep[e.g.,][]{Zirin1988a}. Since only one out of ten active regions shows such a behavior \\citep{Howard1990} and we only present a case study, our results might not be representative for sunspot decay in general. However, space missions such as \\textit{Hinode} and SDO provide nowadays data of sufficient coverage, resolution, and cadence that statistical properties of sunspot decay become accessible. Furthermore, previous studies of non-Hale regions \\citep[e.g.,][and reference therein]{LopezFuentes2000} were centered on flux emergence, $\\delta$-spots, and strong solar flares. The present study can consequently be considered as an extension of these studies with a focus on a much quieter magnetic field topology, which might be representative for the lower solar activity during cycle No. 24 \\citep{Petrovay2010, Nielsen2011}. The major findings of our study can be summarized as follows: (1) MMFs were observed in the vicinity of spot~\\textsf{A} until it decayed. Mostly type~II and a few interspersed \\textsf{U}-shaped type~I MMFs contributed to the observed moat flow, which also left a clear signature in the time-averaged Ca\\,\\textsc{ii}\\,H images \\citep{MartinezPillet2002}. (2) Even though penumbral filaments had almost completely disappeared in photospheric G-band images of spot~\\textsf{A} on November~18, H$\\alpha$ line core images clearly exhibited a structure reminiscent of a superpenumbra. Thus, filamentary structures including the inverse Evershed flow \\citep{Maltby1975, Georgakilas2003} might be more persistent in the chromosphere. (3) We have also observed MMFs in the vicinity of a tiny pore with a diameter of about 2~Mm, which did not show any indication of penumbral filaments. Such an observation argues strongly against a close tie between Evershed flows and MMFs \\citep[cf., ][]{VargasDominguez2008, VargasDominguez2010, CabreraSolana2006}. \\citet{Rempel2011} argues based on MHD simulations of sunspots that penumbral flows can be separated in two components, where the shallow one corresponds to the Evershed flow and the deeper one is related to moat flow. (4) The strong rotation and twist seen in spot~\\textsf{B} might explain, why this trailing spot never advected sufficient magnetic flux to establish more than a rudimentary penumbra and remained highly fragmented during its entire life cycle. (5) The photospheric and chromospheric maps of horizontal flows show a peculiar pattern, once the last dark feature of the active region has disappeared. In general, the flow field in this region is less structured than regions covered by granulation, i.e., the dispersed magnetic field still significantly affects the convective pattern. Chromospheric flows have increased notably compared to times when spots and pores were still present. Most prominently, a contiguous area of low divergence appears towards the end of sunspot decay." }, "1112/1112.4479_arXiv.txt": { "abstract": "We study the buildup of magnetic fields during the formation of Population III star-forming regions, by conducting cosmological simulations from realistic initial conditions and varying the Jeans resolution. To investigate this in detail, we start simulations from identical initial conditions, mandating 16, 32 and 64 zones per Jeans length, and studied the variation in their magnetic field amplification. We find that, while compression results in some amplification, turbulent velocity fluctuations driven by the collapse can further amplify an initially weak seed field via dynamo action, provided there is sufficient numerical resolution to capture vortical motions (we find this requirement to be 64 zones per Jeans length, slightly larger than, but consistent with previous work run with more idealized collapse scenarios). We explore saturation of amplification of the magnetic field, which could potentially become dynamically important in subsequent, fully-resolved calculations. We have also identified a relatively surprising phenomena that is purely hydrodynamic: the higher-resolved simulations possess substantially different characteristics, including higher infall-velocity, increased temperatures inside 1000 AU, and decreased molecular hydrogen content in the innermost region. Furthermore, we find that disk formation is suppressed in higher-resolution calculations, at least at the times that we can follow the calculation. We discuss the effect this may have on the buildup of disks over the accretion history of the first clump to form as well as the potential for gravitational instabilities to develop and induce fragmentation. ", "introduction": "The formation of the first stars in the Universe is a process well-suited to numerical computation. While direct observations of these stars are still an optimistic, yet-unrealized prospect \\citep{2005ApJ...629..615W, 2009ApJ...698L..68F, 2005Natur.434..871F, 2011Natur.477...67C}, simulations are able to begin with well-posed initial conditions and directly simulate their formation. These calculations, typically spanning many orders of magnitude in both spatial and density scales \\citep{2008AIPC..990...16T, 2008Sci...321..669Y}, include the effects of dark matter, hydrodynamics, radiative cooling, chemical heating and cooling, and the non-equilibrium, self-consistent evolution of multiple ionization and molecular states of the gas \\citep{abel97, anninos97, RA04, 2008AIPC..990...25G, 2008MNRAS.388.1627G}. For nearly a decade, the consensus viewpoint had been that the first stars formed in isolation, between $30-300~\\Msun$ in mass, and only one per $\\sim 10^{6}~\\Msun$ halo. This had been confirmed by divergent methods: both Smoothed Particle Hydrodynamics (SPH) simulations and Adaptive Mesh Refinement (AMR) calculations showed similar results \\citep{ABN02, 2003ApJ...592..645Y, oshea07a}. However, recently aspects of this picture are being challenged. Advancements in the current generation of simulations have taken place in three primary areas. The first is that the effect of heating from the formation of molecular hydrogen via three-body reactions \\citep{2011ApJ...726...55T} is now being taken into account; this will heat the gas during later stages of collapse and through thermal regulation of the accretion rate produce higher infall velocities onto the central core. The second aspect is related to the so-called ``Courant myopia'' of calculations proceeding to high densities. Because the Courant timescale becomes extremely short at high densities, out of necessity previous simulations stopped after the formation of the first high-density core. Recent simulations have followed accretion for up to several thousand years, using the numerical technique of sink particles, accreting Lagrangian particles representing protostars below the resolution limit of the simulation \\citep{2011ApJ...737...75G, 2011Sci...331.1040C, 2010MNRAS.403...45S}. While this approach allows to follow the calculations further in time it does lack the mathematical rigor that may allow one to prove the result to be correct. The final effect now being added is simply that of sampling: in the published literature, very few calculations had been performed; compared to the wealth of calculations of galaxy mergers, cluster formation, and so on, the sampling of Population III star formation was dramatically underserved, with only a handful of papers discussing even multiple simulations \\citep[see, e.g.,][]{oshea07a}. These recent shifts have suggested that these halos have fragmented into either a small-number multiplicity of protostars \\citep{2009Sci...325..601T, 2010MNRAS.403...45S} or many small-mass pre-stellar objects \\citep{2011Sci...331.1040C, 2011ApJ...737...75G}. However, none of these simulations have included the effects of magnetic fields in their calculations, nor have they self-consistently followed the growth of seed magnetic fields over the collapse and virialization of these first minihalos. Similarly, the fragmentation has been only studied for times of less than 1\\% of the accretion time scale of the massive stars being studied. It has been difficult to show whether possible early fragments would survive and not grow substantially in mass and simply merge with the central proto-star. The influence of magnetic fields on the collapse of the first stars has received a renewed interest in recent years \\citep{2008ApJ...685..690M}. Early numerical simulations \\citep{2008ApJ...688L..57X} including the Biermann battery effect suggested that the dynamical effect of magnetic fields on primordial gas is very small. Though their resolution was rather low, these simulations found that fields were primarily built up by collapse, leading to $B \\propto \\rho^{2/3}$. The thermodynamics of primordial gas are quite sensitive to the $\\mathbf{B} - \\rho$ relationship via heating from ambipolar diffusion \\citep{2010ApJ...721..615S}. Although magnetic fields are believed to be unlikely to affect the characteristic fragmentation scale for Population III stars, they may increase the temperature in very high density gas, leading to higher accretion rates onto the protostars \\citep{2009ApJ...703.1096S}. The ability of turbulent fluid motions to drive dynamo action during primordial protostellar collapse has been predicted analytically by \\citet{2010A&A...522A.115S} and confirmed numerically by \\citet{2010ApJ...721L.134S} and \\citet{2011ApJ...731...62F} for idealized collapse calculations. The latter work also suggested that a minimum resolution of 32 elements per Jeans length is required to capture dynamo action, giving a very strong motivation for higher resolution cosmological MHD calculations. Properly resolved, dynamo amplification leads to exponential growth of the field over a turbulent eddy turnover time, and can radically change the strength of the magnetic field at any given density during collapse. The \\citet{2011ApJ...731...62F} simulations considered the collapse of a nearly-isothermal Bonnor-Ebert sphere with turbulence seeded by velocity perturbations. While these conditions were idealized from earlier hydrodynamic cosmological computations, we seek a more complete picture by following the full magnetohydrodynamic (MHD) evolution of primordial gas from cosmological initial conditions. This allows us to self-consistently study the interaction of primordial gas dynamics, turbulence, and dynamo action. In this paper, we present the first highly-resolved calculations of the formation of the first stars in the Universe from cosmological initial conditions, taking into account a full suite of chemical reaction rates, chemical heating due to three-body reactions, magnetic fields from primordial seed fields, and a resolution of 64 zones per Jeans length. Furthermore, it has been conducted with an open source, community-built simulation code available for inspection and contribution. In this paper, we examine both the large-scale and small-scale amplification of the magnetic field, as well as the manner in which resolution affects the chemo-kinetic state of the inner molecular cloud. A forthcoming paper will study the growth of turbulence in more detail. ", "conclusions": "We have presented on the first fully-cosmological calculations of Population III star formation that include all relevant chemical processes, as well as magnetic fields. We see, in agreement with \\cite{2011ApJ...731...62F}, that a critical resolution exists above which we are able to resolve small-scale dynamo action resulting in increased magnetic-field field growth. While we adequately resolve the \\textit{action} of this small-scale dynamo, we have not yet resolved the amplification caused by that dynamo, as demonstrated by the nascent J128 simulation. In fact, while these calculations have yet to show a dynamically-important magnetic field, estimates of the possible saturation level of the magnetic field indicate that with increased resolution or stronger seed fields, magnetic fields may in fact become dynamically important. A side effect of conducting these resolution studies has been that we see substantial variation in the chemical, kinetic, and velocity structure of collapsing metal-free, star-forming clouds. In particular, the substantial difference in the J64 run suggests that under-resolving the hydrodynamics can result in incorrect solutions: the fraction of hydrogen gas that is in molecules, the speed of sound, the infall velocity and the turbulent support all depend strongly on hydrodynamic resolution. Our calculations, while falling short of following potential fragmentation in detail, suggest that resolution effects may be an additional complication in determining the initial mass function of the first stars; in fact, they suggest that previous calculations in the literature have not yet converged on a solution. Future calculations, of much higher resolution and attaining a greater saturation level of magnetic fields, will help to resolve these outstanding questions." }, "1112/1112.1956.txt": { "abstract": "We present single-S\u00e9rsic two-dimensional model fits to $167,600$ galaxies modelled independently in the $ugrizYJHK$ bandpasses using reprocessed Sloan Digital Sky Survey Data Release Seven (SDSS DR7) and UKIRT Infrared Deep Sky Survey Large Area Survey (UKIDSS-LAS) imaging data available from the GAMA database. In order to facilitate this study we developed SIGMA, an R wrapper around several contemporary astronomy software packages including Source Extractor, PSF Extractor and GALFIT 3. SIGMA produces realistic 2D model fits to galaxies, employing automatic adaptive background subtraction and empirical PSF measurements on the fly for each galaxy in GAMA. Using these results, we define a common coverage area across the three GAMA regions containing $138,269$ galaxies. We provide S\u00e9rsic magnitudes truncated at $10$ $r_{e}$ which show good agreement with SDSS Petrosian and GAMA photometry for low S\u00e9rsic index systems ($n<4$), and much improved photometry for high S\u00e9rsic index systems ($n>4$), recovering as much as $\\Delta m=0.5$ magnitudes in the $r$ band. We employ a $K$ band S\u00e9rsic index/$u-r$ colour relation to delineate the massive ($n>\\sim2$) early-type galaxies (ETGs) from the late-type galaxies (LTGs). The mean S\u00e9rsic index of these ETGs shows a smooth variation with wavelength, increasing by $30\\%$ from $g$ through $K$. LTGs exhibit a more extreme change in S\u00e9rsic index, increasing by $52\\%$ across the same range. In addition, ETGs and LTGs exhibit a $38\\%$ and $25\\%$ decrease respectively in half-light radius from $g$ through $K$. These trends are shown to arise due to the effects of dust attenuation and stellar population/metallicity gradients within galaxy populations. ", "introduction": "\\label{sec:intro}The shapes and sizes of galaxies are not random but are defined by the orbital motions of their constituent stellar populations, arranging themselves into elliptical, bulge, disk, and bar like structures. Exactly why and how these structures come about is somewhat a mystery which no doubt relates to a complex formation history involving collapse, merging, infall, secular evolution, and feedback processes as well as the precise nature of the coupling between the dark matter, gas, dust and stars and the influence of the larger halo in which the galaxy might reside (group, cluster, etc.), and the broader environment (filament, void, nexus, etc.). The combination of variations in, for example, galaxy structure, formation history, evolution and relative environment lead to distinct measurable effects on global galaxy parameters such as colour, concentration and size. The ultimate goal of structural analysis is to inform this discussion by robustly isolating and quantifying these parameters and exploring correlations between these properties and those obtained by other means, such as dynamical information. Once the underlying structure of a galaxy is understood the overarching morphological class may be determined, and from this we can explore correlations with, for example, environment through the well known morphology-density relation \\citep{Dressler1980}, i.e., the apparent preference for red, passive galaxies in the dense cores of galaxy groups and clusters. Several mechanisms have been suggested to explain this feature, most notably the combined effects of strangulation \\citep{Larson1980,Kauffmann1993,Diaferio2001}, ram pressure stripping \\citep{Gunn1972}, harassment \\citep{Moore1996} and tidal interactions and merging \\citep{Park2008}. Recent studies by, e.g., \\citet{VanderWel2008,Welikala2008,Welikala2009,Bamford2009} and others confirm this morphology-environment connection; however, they suggest that the relation between structure and morphology is less apparent. Indeed, it appears that the mass of a galaxy rather than the environment in which it resides is more influential in determining its structure, highlighting the importance of stellar mass estimates. As an example of the connection between galaxy structure and the physical processes of galaxy formation \\citet*{Dalcanton1997} and independently \\citet*{Mo1998}, both following on from \\citet*{Fall1980}, relate the scale length of the disk to the angular momentum of a galaxy's dark matter halo. In addition, numerous properties of the bulge component are now known to relate to the mass of the super-massive black hole \\citep*[e.g.,][]{Haring2004,Novak2006,Graham2007}. Variations in structural properties as a function of wavelength \\citep[e.g.,][]{LaBarbera2010a} enable the extraction of colour gradients, potentially implying the direction of disk growth (e.g., inside out, \\citealt{Barden2005,Bakos2008,Trujillo2009,Wang2010}), or arguing for the redistribution of populations from the inner to outer regions \\citep{Roskar2008}, possibly coupled with bar formation \\citep{Debattista2006}. The physics underpinning galaxy structure is relatively immature, despite the very long history dating back to Knox-Shaw, Reynolds and Hubble and essentially consists of spot-check simulations which focus on a particular phenomena in a non-cosmological context (for recent developments see \\citealt{Roskar2010} and \\citealt{Agertz2010}). For example, numerical models can readily produce bar, pseudo-bulge, spiral patterns, and spheroidal structures through coupled rotation, secular evolution, shock-wave propagation, and merging history. Until recently the very thin nature of the spiral disks has presented a particular challenge for numerical models, with numerical simulations in particular forming small thick disks, mainly because of the high level of merging \\citep{Navarro1997}. This is in stark contrast to a number of independent empirical studies \\citep[e.g.,][]{Driver2007,Gadotti2009,Tasca2011} which estimate that approximately $60\\%$ of the stellar mass in the universe today lies within disk systems, suggesting a more quiescent merger history (however, see \\citealp{Hopkins2010} on the stability of gas rich disks). In addition, studies by \\citet{MenendezDelmestre2007} suggest that up to $67\\%$ of spiral galaxies contain a barred structure, further complicating simulation efforts. However, numerical simulations are now starting to produce realistic disk systems \\citep{Governato2007,Governato2009,Agertz2009,Agertz2010} albeit with heavily controlled initial conditions, more quiescent merger histories and a greater degree of gas infall. Beyond a distance of $\\sim100$ Mpc detailed structural studies have been relatively rare and mostly confined to the deep yet very narrow pencil beam surveys from the Hubble Space Telescope (HST). It was only following the refurbishment of HST that structural analysis once again became a mainstream study \\citep{Driver1995a,Driver1995b,Driver1998}. HST provides kpc resolution across the full path length of the Universe, which is now also becoming possible with AO ground-based systems \\citep{Huertas-Company2007}. The conjunction of development in numerical models and this new ability to resolve the shapes and sizes of galaxies at any distance has led to a dramatic renewed interest in structural analysis. One interesting claim is the apparent remarkable growth of galaxy sizes since intermediate redshifts (e.g., \\citealp{Trujillo2005,Barden2005,McIntosh2005,Trujillo2006,Trujillo2007,Weinzirl2011}), potentially supporting the notion of recent growth in disk systems following an earlier aggressive merger phase at $z\\sim2$ \\citep[see][]{Driver1996,Driver2005}. An alternative suggestion which does not require galaxy growth through mergers is the transformation of some of these so-called {}``red nuggets'' into the bulges of disk galaxies via the accretion of a cold gas disk \\citep{Graham2011}. However, structural analysis is not trivial to implement and interpret correctly, and is plagued by a number of key issues. In particular: \\begin{enumerate} \\item Wavelength bias. At different wavelengths, light traces varying stellar populations (\\citealp{Block1999}). Typically this is a young stellar population at shorter wavelengths and an older stellar population at longer wavelengths. For this reason, it is vital when comparing structural properties to compare properties measured at the same rest wavelength. \\item Dust attenuation. Dust is predicted to modify not only the recovered total flux as a function of wavelength (e.g., \\citealp{Tuffs2004,Pierini2004}) but also galaxy sizes, shapes and concentrations (see for example \\citealp{Mollenhoff2006,Graham2008b}). Dust can vary enormously from system to system with significant environmental dependencies and strong evolution with redshift. Each individual galaxy ultimately requires either a dust correction or analysis at rest-NIR wavelengths where dust will have a smaller impact (see photon escape fraction curve in \\citealt{Driver2008}). It is obvious that any attempt to model the dust in galaxies raises the larger problem of degeneracies appearing between additive and subtractive flux components. \\item Local minima during the minimisation process. For a single profile fit there are often 7 free parameters, with that number rising for multi-component fits. The surface within this parameter space is known to be complex, containing multiple local minima representing potentially non-physical results, e.g., a bulge which contributes significantly more flux to the outer regions of a galaxy than the disk (\\citealt{Graham2001}). Other than manual checks of the output, various methods may be employed to reduce the risk of divergence on an incorrect result including; constraints applied during the minimisation routine and employing an automated logical filter (e.g., \\citealt{Allen2006}). \\item What lies below the limiting isophote. Whilst the surface brightness profile of some galaxies behaves as expected out to very faint magnitudes (e.g., NGC 300:\\citealp{Bland-Hawthorn2005,Vlajic2009}, NGC 7793: \\citealp{Vlajic2011}), the potential myriad of phenomena present in the outer wings of many systems may cause deviations away from a typical light profile. These include truncated and anti-truncated disks \\citep{Erwin2005,Pohlen2006}, UV excesses \\citep{Bush2010}, tidal debris, halos \\citep{Barker2009,McConnachie2009} and minor merger fossil records \\citep{Martinez-Delgado2010}. In fact, the outer regions of galaxies may defy any systematic profile fitting into a restricted number of structures. The accuracy of any estimation of the background sky and gradients therein will also no doubt affect analyses of these outer structures. \\item The number of components required. When considering the structure of very nearby galaxies, the deeper one looks the more one finds. Some galaxies, even in the dust free $3.8$ micron bands, require up to six components \\citep{Buta2010} before a satisfactory fit can be obtained. In many cases there is uncertainty as to how many components are required, how to quantitatively decide this in an automated and repeatable fashion, and which components are fundamental and which perhaps secondary. For example, should bar and pseudo-bulge flux be incorporated into a single disk model or kept distinct. \\item Sky estimation. Understanding the background sky level at the position of your primary object of interest is crucial in producing meaningful measurements of that galaxy. Considerations must be made in regards accuracy and speed of estimating the background. \\item Systematic selection bias. Sample bias will be introduced due to size, resolution, orientation, profile shape and smoothing scale limitations \\citep{Phillipps1986,Driver1999}. Samples of galaxies are usually selected based on global criteria, such as magnitude. However, it becomes non-trivial to transcribe these global limits into appropriate limits for galaxy sub-components, e.g., a certain type of disk may only have been detected because it also contains a prominent bulge. \\end{enumerate} There are several publicly-available galaxy modelling codes in common usage including GIM2D \\citep{Simard2002}, BUDDA \\citep{DeSouza2004}, GASPHOT \\citep{Pignatelli2006} and GALFIT 3 \\citep{Peng2010a}. In addition, there are a number of software pipelines, wrappers around contemporary astronomy software, that aim to automate the process of galaxy modelling including GALAPAGOS (Barden et al., 2011, submitted) and \\noun{PyMorph} \\citep{Vikram2010}. These packages all have their advantages and disadvantages and have been compared in a number of external studies (e.g. \\citealp{Haussler2007,Hoyos2011}) in addition to their own internal comparisons, and so we refer the reader to these papers for discussions of the pros and cons between 1D v 2D fitting and the actual minimisation algorithms employed. For this body of work GALFIT was chosen for its ease of use and high-quality realistic model outputs, plus the ability to perform simultaneous modelling of nearby neighbours to the primary galaxy. In this series of papers we introduce and utilise SIGMA, an automated code designed to produce single-S\u00e9rsic and multi-component profile fits for galaxies in the GAMA dataset. Using SIGMA, this paper presents one of the largest catalogues of multi-wavelength single-S\u00e9rsic model fits; $167,600$ galaxies modelled independently across $9$ bandpasses. This catalogue is currently in use to aid in measurement of the evolution in the size-(stellar mass) distribution of galaxies (Baldry et al., 2011); explore star formation trends as a function of morphology (Bauer et al., in prep.); to further understand the cosmic SED from 0.1 micron to 1 mm (Driver et al., 2011); to apply dust corrections to galaxies observed at multiple inclinations (Grootes et al., 2012); to explore the dust properties and star-formation histories of local submillimetre selected galaxies \\citet{Rowlands2011}; better constrain stellar mass measurements by providing total flux corrections \\citep{Taylor2011}; comment on the quenching of star formation in the local universe (Taylor et al., in prep.); explore the relation between galaxy environments and their star formation rate variations (Wijesinghe et al., in prep.); provide a new method for automatic morphological classification (Kelvin et al., in prep.); and further explore the relation between environment (i.e., halo mass; \\citealp{Robotham2011}), morphology and structure (Kelvin et al., in prep.). Future studies building on SIGMA will incorporate advanced logical filtering and profile management to produce multi-component fits for a low redshift sample, allowing full structural decomposition into bulge-disk-bar, etc. (Kelvin et al., in prep). This paper is organised as follows. We first outline the GAMA data in Section \\ref{sec:data}. We describe the SIGMA (Structural Investigation of Galaxies via Model Analysis) pipeline developed to process this data and produce robust 2D galaxy models in Section \\ref{sec:sigma} and present SIGMA single-S\u00e9rsic results for $167,600$ objects modelled independently in $ugrizYJHK$ from the GAMA Phase I study in Section \\ref{sec:output}. From this large catalogue we establish a common coverage sample of $138,269$ galaxies. Finally, we further explore the wavelength dependence on recovered structural parameters in Section \\ref{sec:wavelength}. A standard cosmology of $H_{0}=70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{m}=0.3$, $\\Omega_{\\Lambda}=0.7$ is assumed throughout. ", "conclusions": "\\label{sec:conclusion}We have produced high-fidelity automated two-dimensional single-S\u00e9rsic model fits to $167,600$ galaxies selected from the GAMA input catalogue. These have been modelled independently across $ugrizYJHK$ using reprocessed SDSS and UKIDSS-LAS imaging data. These data have subsequently been delivered to the GAMA database in the form of the catalogue \\emph{SersicCatv07}. In order to facilitate the construction of this dataset, SIGMA, an extensive multi-processor enabled galaxy modelling pipeline, was developed. SIGMA is a wrapper and handler of several contemporary astronomy software packages, employing adaptive background subtraction routines and empirical PSF generation on a per-galaxy per-band basis to tailor input data into the galaxy modelling software GALFIT 3. Output results from GALFIT are analysed for pre-determined modelling errors such as positional migration, extreme model shape and/or size parameters and adverse nearby neighbour flux. Nearby object masking is employed as a last resort, with secondary neighbours being preferentially modelled simultaneously with the primary galaxy in the first instance. Using this dataset, we have defined a common coverage area across the three GAMA regions that encompasses $138,269$ galaxies, $82.5\\%$ of the full sample. This common area contains only those galaxies which have been observed in all nine bands, providing a useful basis upon which to further explore wavelength trends. We define a S\u00e9rsic magnitude system that truncates S\u00e9rsic magnitudes at $10$ $r_{e}$. This ensures that flux is not extrapolated below the typical limiting isophote into regions where data quality and quantity is not sufficient to constrain the form of the galaxy light profile. Truncated S\u00e9rsic magnitudes appear to be a good descriptor of global galaxy colours and total galaxy flux. For well-resolved disk-like galaxies ($n<2$), traditional aperture-based methods are in good agreement with truncated S\u00e9rsic magnitudes. For high centrally-concentrated systems however ($n>4$), it appears that traditional aperture-based, such as Petrosian magnitudes, may miss as much as $\\Delta m_{r}=0.5$ magnitudes from the total flux budget which is only recovered through S\u00e9rsic modelling. When considering the dataset in n--colour space we find galaxies appear to exist in two distinct groups. For the most massive systems, we associate these two groups with the spheroid-dominated early-type galaxy (ETG) and disk-dominated late-type galaxy (LTG) populations. Owing to the nature of our input sample selection, these definitions do not extend down to the fainter dwarf population, and so subsequent trends will not represent those systems. We use the longest wavelength $K$ band S\u00e9rsic index measurements in conjunction with rest-frame $u-r$ colour to define these two populations. Using these definitions, we are able to further probe the variations in recovered structural parameters with wavelength for each population. We find that the S\u00e9rsic indices of ETGs remain reasonably stable at all wavelengths, increasing by $0.11$ dex ($+30\\%$) from $g$ to $K$ and becoming very stable beyond the $z$/$Y$ interface. In contrast to this, we find that LTGs exhibit larger variations in S\u00e9rsic index with wavelength, increasing by $0.18$ dex ($+52\\%$) across the same wavelength range. Recovered sizes for both the spheroid and disk systems show a significant variation with wavelength, showing a reduction in half-light radii of $0.20$ dex ($-38\\%$) in ETGs and $0.13$ dex ($-25\\%$) in LTGs from $g$ to $K$. Size variation of this scale for disk systems has been well predicted by dust models, highlighting the important role dust attenuation plays when considering structural variations across a broad wavelength range. We note that spheroidal systems exhibit a larger size variation with wavelength than that found in disk systems. Possible physical explanations for this behaviour include low levels of unresolved dust or the effects of AGN feedback in the core of the galaxy, both of which would affect S\u00e9rsic profiling. Significant amounts of dust, such as an increased dust attenuation optical depth parameter $\\tau_{\\mathrm{B}}^{f}$, may allow current dust models to accurately describe the variation in half-light radii we find. It is unlikely however that a significant fraction of our spheroid-dominated population contain sufficient amounts of dust for this to be the case. Large stellar population/metallicity gradients present within individual structures of the galaxy would cause galaxies to look markedly different in different wavelengths, contributing to any concentration-wavelength/size-wavelength variation. In addition to these factors, uncertainties on the measured PSF and background sky must be considered. However, when considering variations in half-light radius and S\u00e9rsic index together with wavelength we find that the large fluctuations in spheroidal parameters amount to a relatively modest impact on the recovered light profile. A comparatively larger effect is noted for the disk systems, particularly in the core region, supporting the presence and effect of dust attenuation in addition to stellar population/metallicity gradients. At a distance of $1$ pixel from the central region, spheroid systems display a variation in surface brightness of $0.49$ magnitudes from $u$ through to $K$. In disk systems, the comparative figure is $0.86$ magnitudes, an increase of $75\\%$. This highlights the importance of not considering recovered parameters in isolation, as the interplay between them has the possibility of masking underlying trends. The effects of dust attenuation appear to be the dominant factor constraining the variations in structural parameters with wavelength, notably so for the disk-dominated population. In contrast with this, apparent large structural variations in the spheroid-dominated population appear to have a relatively minor effect on the underlying surface-brightness profile than might have been expected. Future studies in Kelvin et al. (2011; in prep.), focussing on a limited sub-sample of this dataset, will provide a deeper understanding of these structural variations with wavelength, enabling us to comment further on the key mechanisms involved in varying structural parameters with wavelength for a host of different morphologies." }, "1112/1112.4223_arXiv.txt": { "abstract": "Any symmetry reduces a second-order differential equation to a first integral: variational symmetries of the action (exemplified by central field dynamics) lead to conservation laws, but symmetries of only the equations of motion (exemplified by scale-invariant hydrostatics) yield first-order {\\em non-conservation laws} between invariants. We obtain these non-conservation laws by extending Noether's Theorem to non-variational symmetries and present an innovative variational formulation of spherical adiabatic hydrostatics. For the scale-invariant case, this novel synthesis of group theory, hydrostatics, and astrophysics allows us to recover all the known properties of polytropes and define a {\\em core radius}, inside which polytropes of index $n$ share a common core mass density structure, and outside of which their envelopes differ. The Emden solutions (regular solutions of the Lane-Emden equation) are obtained, along with useful approximations. An appendix discusses the $n=3$ polytrope in order to emphasize how the same mechanical structure allows different thermal structures in relativistic degenerate white dwarfs and zero age main sequence stars. ", "introduction": "Noether's Theorem relates every {\\em variational symmetry,} a symmetry of an action or similar integral, to a conservation law, a first integral of the equations of motion~\\cite{Bluman}. By an extension of Noether's Theorem, {\\em non-variational symmetries,} --- symmetries of the equations of motion which are not in general variational symmetries --- also lead to first integrals, which are not conservation laws of the usual divergence form, as discussed in a previous article~\\cite{BludKenI}. There it was shown that a Lagrangian $\\mathcal{L}(t,q_i,\\dot{q_i})$ and action $S=\\int{\\mathcal{L}(t,q_i,\\dot{q_i})} dt$, with degrees of freedom $q_i$, can be transformed under an infinitesimal point transformation $\\delta (t,q_i), \\delta q_j (t,q_i)$: \\bea \\delta \\mathcal{L}=\\dot{\\mathcal{L}}\\delta t+(\\partial\\mathcal{L}/\\partial q_i)\\delta q_i+(\\partial\\mathcal{L}/\\partial\\dot{q_i}) \\Bigl[\\frac{d \\delta q_i}{d t}-\\dot{q_i}\\frac{d \\delta t}{d t}\\Bigr]=\\Bigl [\\frac{dG}{dt}- \\mathcal{L}\\cdot\\frac{d(\\delta t)}{dt}+\\mathcal{D}_i\\cdot(\\delta q_i-\\dot{q_i}\\delta t)\\Bigr] \\quad, \\eea in terms of the total derivative of the {\\em Noether charge,} $G:=\\mathcal{L}\\cdot\\delta t+p_i\\cdot(\\delta q_i-\\dot{q_i}\\delta t),$ and the variational derivative $\\mathcal{D}_i:= \\partial\\mathcal{L}/\\partial q_i-d(\\partial\\mathcal{L}/\\partial\\dot{q_i})/dt$. For transformations that leave initial and final states unchanged, the variation in action is \\be \\delta S_{if}= G(f)-G(i)+\\int_i^f dt\\ \\Bigl[\\delta q_i\\cdot\\mathcal{D}_i+\\delta t\\cdot \\Bigl(\\frac{d \\mathcal{H}}{dt}+ \\frac{\\partial\\mathcal{L}}{\\partial t}\\Bigr) \\Bigr]\\quad , \\ee if the term in $d(\\delta t)/dt$ is integrated by parts. If the system evolution obeys an action principle, that this variation vanish for independent variations $\\delta q_i, \\delta t$ that vanish at initial and final times, the system obeys the Euler-Lagrange equations $\\mathcal{D}_i=0$ and $d \\mathcal{H}/dt=-\\partial\\mathcal{L}/\\partial t$, the rate of change of the Hamiltonian in non-conservative systems. On-shell, where $\\mathcal{D}_i=0$, \\bea \\delta S_{if}=\\int_i^f{ \\bar{\\delta}\\mathcal{L}}\\ dt=G(f)-G(i) \\\\ \\frac{dG}{dt} =\\bar{\\delta}\\mathcal{L}:=\\delta\\mathcal{L}+\\mathcal{L}\\cdot (d\\delta t/dt) . \\eea This is {\\em Noether's equation}, giving the evolution of a symmetry generator or Noether charge, in terms of the Lagrangian transformation that it generates. It expresses the Euler-Lagrange equations of motion as the divergence of the Noether charge. This divergence vanishes for a variational symmetry, but not for any other symmetry transformation. Noether's equation (4) could have been derived directly from the definition of the Noether charge. But using the action principle make manifest the connection between Noether's equation and the Euler-Lagrange equations. We use the action principle and this connection to reformulate the theory of hydrostatic barotropic spheres, which is integrable if they are scale symmetric, even where this scale symmetry is not a symmetry of the action (Section II). The first integrals implied by any symmetry of the equations of motion, while generally not vanishing-divergence conservation laws, are still useful dynamical or structural first-order relationships. Because it neglects all other structural features, scaling symmetry is the most general simplification that one can make for any dynamical system. For the radial scaling transformations we consider, $\\delta r=r$, the Lagrangian scales as some scalar density $\\delta \\mathcal{L}=-2\\tilde{\\omega}\\mathcal{L};$ and the action scales as $\\delta S=(1-2\\tilde{\\omega}) S$. The Noether charge generating the scale transformation evolves according to a {\\em non-conservation law} $dG/dt =(1-2\\tilde{\\omega})\\mathcal{L}$, a first-order equation encapsulating all of the consequences of scaling symmetry~\\cite{BludKenI}. From this first-order equation follow directly all the properties of index-$n$ polytropes, as established in classical works~\\cite{Chandra,Schwarzschild}, modern textbooks~\\cite{Kippen,Hansen}, and the recent, excellent treatments of Horedt and Liu~\\cite{Horedt,Liu}. Our secondary purpose is to present an original variational formulation of spherical hydrostatics and to extend Noether's Theorem to non-variational scaling symmetry, which yields a {\\em scaling non-conservation law} (Section II). For spherical hydrostasis, we define a {\\em core radius}, inside which all stars exhibit a common mass density structure. Outside this core, polytropes of different index $n$ show different density structures as the outer boundary is felt (Section III). Section IV completes the integration of the Lane-Emden equation by quadratures and obtains useful approximations to the Emden function $\\theta_n (\\xi)$. An appendix reviews the thermodynamic properties of the physically important polytropes of index $n=3$~\\cite{Hansen,Kippen,BludKenI}. What is original here is the explanation of the the differences between relativistic degenerate white dwarf stars and ideal gas stars on the zero-age main sequence (ZAMS), following from their different entropy structures. Our original approximations to $\\theta_3 (\\xi)$ should prove useful in such stars. ", "conclusions": "% We have explored how a symmetry of the equations of motion, but not of the action, reduces a second-order differential equation to first-order, which can be integrated by quadrature. In scale-invariant hydrostatics, the symmetry of the equations yields a first integral, which is a first-order equation between scale invariants, and yields {\\em directly} all the familiar properties of polytropes. We observe that, like all stars, polytropes of index $n$ share a common core density profile and defined a {\\em core radius} outside of which their envelopes differ. The Emden functions $\\theta_n(\\xi)$, solutions of the Lane-Emden equation that are regular at the origin, are finally obtained, along with useful approximations. The Appendix reviews the astrophysically most important $n=3$ polytrope, describing relativistic white dwarf stars and zero age main sequence stars. While reviewing these well-known applications~\\cite{Hansen,Kippen}, we stress how these {\\em same} mechanical structures differ {\\em thermodynamically} and the usefulness of our original (Section IV) approximations to these Emden functions. \\appendix*" }, "1112/1112.6226_arXiv.txt": { "abstract": "The ejection of matter in the close vicinity of a young stellar object is investigated, treating the accretion disk as a gravitationally bound reservoir of matter. By solving the resistive MHD equations in 2D axisymmetry using our version of the Zeus-3D code with newly implemented resistivity, we study the effect of magnetic diffusivity in the magnetospheric accretion-ejection mechanism. Physical resistivity was included in the whole computational domain so that reconnection is enabled by the physical as well as the numerical resistivity. We show, for the first time, that quasi-stationary fast ejecta of matter, which we call {\\em micro-ejections}, of small mass and angular momentum fluxes, can be launched from a purely resistive magnetosphere. They are produced by a combination of pressure gradient and magnetic forces, in presence of ongoing magnetic reconnection along the boundary layer between the star and the disk, where a current sheet is formed. Mass flux of micro-ejection increases with increasing magnetic field strength and stellar rotation rate, and is not dependent on the disk to corona density ratio and amount of resistivity. ", "introduction": "Highly collimated outflows have been observed from AGNs to Young Stellar Objects (YSOs) and young brown dwarfs \\citep{wh1,wh2}. Accreting compact stars, like accreting white dwarfs in symbiotic binaries \\citep{sk03} and neutron stars like Cir X-1 \\citep{hei07}, also show similar outflowing phenomena. An outflow is characterized as a jet if it is super-magnetosonic, collimated into an apparent narrow opening, and reaches a stationary or quasi-stationary state. Such high-velocity outflowing fluxes of matter are an integral part of stellar evolution. Observations in multiple wavelengths are reaching closer and closer to the objects that drive them. Among all the systems, models of launching outflows in YSOs are closest to scrutiny by observations due to available data from star--forming regions. An accretion disk, through which matter accretes onto the young star with velocities close to a free--fall, is often associated with a jet--driving YSO \\citep{e94,e06}. A correlation between the accretion rate and the high-velocity jet power was found in many Classical T-Tauri Stars (CTTSs) \\citep{Cab90}. The ratio of mass loss in the outflow to disk accretion rate, $\\dot{M}_{\\rm w}/\\dot{M}_{\\rm a}$, extracted from observations is hard to constrain. It is best estimated to be approximately 0.1 through measurements of optical forbidden lines and veiling --- see e.g.\\ \\citet{hart95} and \\citet{edw08}. Recently, He~I~$\\lambda$10830 line has been used as a probe into the high-velocity winds originating from the inner region where the star interacts with the disk \\citep{e03,kw07}. Despite the potential diagnostic power of such emission lines, the actual structure and physical conditions of outflows can be more complex. Star-disk interaction is also investigated by means of X-rays, which offer deepest probe into the launching region. During flare events from the protostellar objects X-rays are emitted and have been observed (\\citet{fav05}, \\citet{gia06}, \\citet{aar10}, \\citet{mclw11}). Production of such X-rays is possible by either high temperature or high flow velocities, which can be related to the stellar surface, shocks in the stellar magnetosphere or, as we will suggest, to the reconnection events in the star-disk magnetosphere. Such events will leave imprint in the chemical and physical properties of the object \\citep{shuetal07}. To further interpret the observed line profiles, and the origins of outflows from the close vicinity of CTTSs, predictions from both theoretical and numerical models are required. \\begin{table*} \\caption{In many resistive-MHD simulations performed up to date, assumptions for initial conditions vary much and duration of simulations is very different. We list some of important works to position our work in the context. In the second column, ``fast'' stellar rotation means that corotation radius is smaller than the disk truncation radius, $R_{cor}R_t$. Disk is rotating in all the cases. Columns about physical resistivity and viscosity specify only if the option is at all included in the code; for where it is really effective, reader should check the related publication. When present, viscosity is important only in the disk. } \\begin{tabular}{|c||c|c|c|c|c|c|} \\hline & & & & & & \\\\ \\ \\ & $\\Omega_\\ast$ & corona & $\\kappa=\\rho_d/\\rho_c$ & $T_{\\mathrm max}/T_0$ (days) & resistive & viscous\\\\ & & & & & & \\\\ \\hline & & & & & & \\\\ \\ \\ \\citet{hay96} & non-rotating & non-rotating & 10$^3$ & 5 & yes & no\\\\ \\ \\ \\citet{hir} & non-rotating & rotating & 10$^4$ & 16.5 & yes & no\\\\ & & rotating different than disk & & & & \\\\ \\ \\ \\citet{ms97} & slow & solid body rotation & 10$^2$ & 0.3 & yes & no \\\\ & & corotating with star at $R_{\\rm cor}$ & & & & \\\\ \\ \\ \\citet{gods99a} & fast & rotating & $10^4$ & 100 & yes & no \\\\ \\ \\ \\citet{rukl02} & slow & corotating with star & 10$^2$ & 100 & no & yes \\\\ & & for R$\\leq R_{\\rm cor}$, else with disk & & & & \\\\ \\ \\ \\citet{kuk03} & slow & not in hydrostatic balance, & 10$^3$ & 1000 & yes & yes \\\\ & & non-rotating & & & & \\\\ \\ \\ \\citet{u06} & fast & corotating with star & 10$^3$ & 2000 & yes & yes \\\\ & & for R$\\leq R_{\\rm cor}$, else with disk & & & & \\\\ \\ \\ \\citet{R09} & fast & corotating with star & 10$^4$ & 2000 & yes & yes\\\\ & and slow & for R$\\leq R_{\\rm cor}$, else with disk & & & & \\\\ \\hline & & & & & & \\\\ \\ \\ M\\v{C} et al. (present paper) & slow & corotating with star & 10$^4$ & 1500 & yes & no \\\\ & & for R$\\leq R_{\\rm cor}$, else with disk & & & & \\\\ \\hline \\end{tabular} \\label{tabla1} \\end{table*} Outflows driven by energy derived from accretion are particularly appealing in the scenarios of jet launching. Many models have been proposed based on the concept of magnetocentrifugal wind mechanisms \\citep{bp82}, differing in the origins of the underlying magnetic fields and locations of matter launching. An outflow could be a disk wind driven by magnetic fields dragged in from the envelope or generated by the disk dynamo, or an inner disk wind anchored to the narrow innermost region as in the X-wind model powered by an enhanced dynamo from the star-disk interaction \\citep{shuetal94, shuetal97}, simultaneously with an accretion funnel \\citep{osshu95}. It might also be a stellar wind driven along the open field lines from the stellar surface by thermal or magnetic pressure \\citep{mattpu05,mattpu08}, or some combination of the different possibilities. Related to the launch of winds, magnetospheric accretion has been described in works by \\citet{koe91}, \\citet{osshu95} and \\citet{kold02} in the context of a magnetosphere interacting with the surrounding disk, sharing some similarities with the compact objects like neutron stars \\citep{ghla79a,ghla79b}. Except for the pure disk wind models, a magnetically connected star-disk system plays an important role in the making of the young stellar system and the evolution of angular momentum through the generation of outflows during the main phase of accretion. Numerical investigations have been followed up on the time-dependent evolution of a system where the central star is magnetically connected to its accretion disk and their connection to jet formation and accretion. In one of the earliest attempts by \\citet{hay96}, where simulations of only a few rotation periods were obtained, a dipole magnetosphere corotating with the central star threaded the accretion disk that was in Keplerian rotation. Magnetic field lines connecting both the disk and the star inflate outwards due to shear, and reconnection blows out the matter along with the field, partially opening up the originally closed dipole loops. Gas can outflow from those opened field lines and might form part of the X-emission that is often associated with flares. Reconnection as a possible origin of X-rays from such systems has also been indicated in \\citet{EdP10}. \\citet{hir} investigated a magnetized star interacting with a truncated disk that was threaded with an initially uniform field dragged in from the outer core, in the same direction of the magnetosphere, but separated by a neutral current sheet in the equatorial plane as a result of interaction between the fields brought together. For simplicity, the star was not rotating, but the differentially rotating disk could anyway provide enough shear to make the field inflate outwards, followed by a reconnection event and mass transfer onto the magnetosphere. The transferred mass diverted into two directions: one that falls onto the star and the other that flows out along the opened stellar field lines. Longer simulations by \\citet{gods97} with an aligned dipole and a conducting accretion disk showed that differential rotation of the disk can drive episodes of loop expansion. Such expansion can drive two outflow components of gas: one hot convergent flow along the rotation axis, and another, slower cold flow on the disk side of the expanding loop. \\citet{ms97}, on the other hand, investigated interactions of magnetospheres with accretion disks under three different magnetic configurations and their respective dynamical evolution. \\citet{kuk03} solved the disk in 1D with a radiative hydrodynamic code by \\citet{kley89}, and then extrapolated the solution to 2D as their initial condition. For the full 2D axisymmetric MHD problem, the induction equation, Lorentz force and Ohmic dissipation were now included into Kley's code, with the assumption of equal viscous and resistive dissipations. With $Pr\\sim 1$ magnetic field lines would not bend towards the axis. However, because of non-equilibrium initial conditions, they bunched close to the star. The main result was that, with the assumed mass accretion rate of $\\dot{M}_0=10^{-7}M_\\odot$~yr$^{-1}$, for a smaller magnetic field than 1~kG the disk was not disrupted; but for a larger field of the order of 1-10 kG, an outflow could be launched from the disk. In \\citet{rukl02} and \\citet{lrl05}, much better initial equilibrium has been set than in any previous simulations, with matter continuing to inflow through the disk because of viscosity. It was an improvement, as it was proceeding with the viscous time-scale, because of a slow accretion of matter. Without such initial equilibrium, the non-stationary initial conditions determine the flow in the disk, and influence the simulation. A star and part of the magnetosphere corotated up to the corotation radius, and the magnetosphere corotated with the disk farther out. The simulations were performed in the ideal MHD regime, with effective numerical resistivity diffusing magnetic field in the radial direction. They found funnel flows onto the central object, spinning up or down the star, depending on the ratio of rotation rate of the star to the rotation rate of the disk inner rim. In \\citet{u06} and \\citet{R09} the effects of physical viscosity and resistivity on the outflow were investigated. In \\citet{R09} one of studied cases is with the magnetic Prandtl number, the ratio between the viscosity and the resistivity $\\mathrm{Pr}=\\nu/\\eta$, greater than one. Magnetic field lines are bent towards the magnetosphere in the gap and magnetic energy increases, enabling the outflow. It has been found that, in addition to the fast and light jet, there is another, new conical wind flowing up to 30 percent of the matter from the innermost portion of the disk. Two different cases were considered in their simulations, one for a fast (with the setup as in \\citet{u06} and a slightly different parameters), and the other for a slowly rotating star. In both cases, to enable the smooth start of the simulation, initially slow rotation of the star was gradually speeded up to its maximum value, with matter slowly inflowing from the outer boundary, to obtain stellar magnetic field compressed towards the magnetosphere in the gap. In the latter case, disk was not initially present in the computational box, but was formed from the matter inflowing from the outer boundary. Such setup was different from most other simulations in the literature. With non-stationary initial conditions in the disk, accretion tends to be too fast. Most of the mentioned works involve resistive MHD models of accretion disks. Introduction of viscosity in the disk helps to obtain slow accretion, with a viscous time-scale. We do not include physical viscosity in our simulations, but we set the resistivity in the whole computational box, not only in the disk. Resistivity, which controls the onset of magnetic reconnection, triggers the necessary change in the magnetic field geometry needed for any launching. In our previous work, in \\citet{cf04}, we reported on the result (with the same code) when only a disk is present, without the magnetosphere of the central object taken into account. Propagation of the outflow through the resistive corona (with the disk set as a boundary condition) we investigated in \\citet{FC02}. One important parameter to distinguish the investigated MHD regime is the already mentioned magnetic Prandtl number. As there is no physical viscosity included, our resistive simulations are in the regime of $\\mathrm{Pr}\\lesssim 1$. Violent initial conditions now helped to bring magnetic flux closer towards the star, helping the launching of matter outwards. Another parameter whose effect we study is the density ratio between the disk and the corona. It is usually included as a free parameter of the order of $10^2$ or $10^3$, at best $10^5$. We investigate the influence of this ratio on the mass and angular momentum flux in the launching of outflows. There are other possibilities in the setup, which we did not investigate here, e.g.\\ inclusion of the stellar wind, which would probably affect the open stellar field. Table \\ref{tabla1} lists the kinematic and thermodynamic assumptions adopted in earlier works, each of them being usually repeated with a variety of parameters or methods, with or without physical resistivity and viscosity included in the code. We put our work in context of setups and assumptions of those works, as our results differ in the mass fluxes carried in the ejected gas. The organization of the paper is as follows. We first describe our implementation of the boundary and initial conditions. In \\S\\ref{props} we report regimes we found under a broad range of parameters. We investigated the influence of corona to disk density ratio, strength of magnetic field and the physical resistivity. In \\S\\ref{recon} we address the role of reconnection in the launching, in \\S\\ref{elsase} we check a criterion for the site of launching, and in \\S\\ref{trunc} we compare position of the disk truncation radius in our simulations with some theoretical predictions. Then we discuss investigated parameters and the resulting ejections. ", "conclusions": "In this work, we for the first time demonstrate launching fast, light quasi-stationary flows of matter, which we call micro-ejections, from the magnetosphere above the gap in the star-disk system. In our simulations we included physical resistivity not only in the disk, but in the whole computational box. We investigate effects of anomalously large resistivity (when compared to microscopic resistivity), modeled as a function of matter density, to the launching of such micro-ejections in the case of slowly rotating star. The physical resistivity, when included in previous simulations in the literature, was limited to the disk, decreasing effectively to zero out of it. The reconnection above the disk, which is necessarily happening during such simulations, was at the mercy of numerical resistivity, whose effects are different from physical resistivity In order to better understand differences between a variety of setups, we compare resistive simulations performed to date --- some of them are shown in Table \\ref{tabla1}. Resistivity can only moderately modify the flow shape, on a slower time scale than the formation of a magnetic wall \\citep{lyb96,lyb03}, so that if the resistivity is not large enough for reconnection to occur, the geometry of the flow will be different. As sketched in Figure \\ref{geommag}, common phenomena have been identified in our simulations, which occur, in those simulations of interacting star-disk systems from Table \\ref{tabla1} which have violent initial relaxation and/or weak magnetic field. After the relaxation of the system from unrealistic initial condition to a more evolved configuration, magnetic reconnection adjusts the topology of compressed field lines. Originally closed loops partially open. In this picture, the resistivity plays an important role, as it enables the reconnection. We model physical resistivity as an anomalous turbulent resistivity dependent on the matter density $\\rho$, following \\cite{FC02}, with $\\eta=\\eta_0 \\rho^{1/3}$. Previously, resistivity has been included in the whole computational domain in \\citep{fen09}, but with the disk only as a boundary condition. Those simulations followed the propagation of outflow during few thousands of rotations of the inner disk radius. With the disk included in the computation, relaxation from initial conditions is much more complicated, and it is not easy to follow the evolution of the system for hundreds of rotations. We set up our resistive version of Zeus-3D code without employing any special procedure to prevent the violent relaxation from non-equilibrium initial conditions\\footnote{Such procedure could consist of the preparation of the particularly suitable disk model, as in \\citet{kuk03} or, as in the seminal paper of \\citet{R09}, slow introduction of the disk matter into the computational box, with gradual increase of the stellar rotation rate.}. Because of numerical difficulties with resistive MHD simulations in our setup with violent relaxation, our simulations tend to cease during relaxation, or last for too short after it for concluding about the results with a stable disk gap at a longer timescale. To perform the parameter study, we modified boundary conditions in the disk equatorial plane, to enforce a stable disk gap. For a sufficiently large stellar magnetic field, a quasi-stationary micro-ejection forms. The magnetic field lines are opened above the star, with the more or less episodic, very light axial flow nearby the axis, which is artificial, of numerical origin. The magnetosphere between the star and the disk, along the boundary layer between the stellar and the disk field, is the site of launching of slower and more massive micro-ejection under a wider angle. The mechanism is similar to the launching of solar micro-flares, only that in our case magnetosphere of the star-disk system is the site of launching. Such events would leave trace in the chemical and physical properties of the object, as indicated in \\citet{shuetal07}. The role of reconnection in accretion disks with jets has been discussed also in \\citet{EdP10}. We will address this question in future work, together with the question of reconnection with different models of resistivity. In our parameter study we investigated influence of the density contrast between the disk and corona, $\\kappa$. There is no difference in results for $\\kappa$ from $10^{3}$ to $10^{6}$. For $\\kappa=10^{2}$, the mass flux is always larger for a factor of few, and we conclude that results of simulations with $\\kappa=10^{2}$ could be unrealistic. Another parameter we checked was dependence of mass flux on stellar rotation rate. We found that mass flux increases with stellar rotation rate. Both fluxes also increase with increasing magnetic field--we vary the strength of the stellar dipole in the range of (0.1--200)\\,G for the disk accretion rate $10^{-6}M_\\odot/$yr. Both fluxes increase with increasing magnetic field. There is a limit to increase of magnetic field in simulation. A very strong initial field may cause strong shear motion at the beginning, and can largely disturb the relaxation process, especially in the case of a non-resistive corona, because pinching and reconnection of the magnetic field depend critically on the conditions in the magnetosphere \\citep{lov95,gods97}. Because resistivity enables the reconnection and reshaping of magnetic field, we also investigated if fluxes in micro-ejection depend on resistivity. We found that there is no such dependence. Without resistivity, numerical or physical, there is no reconnection, but once it occurs, the level of resistivity does not change the outcome of our simulations. Further study of this dependence is needed, with different resolutions. Our simulations do not scale well, and in the present setup we could not perform such study. In ideal MHD simulations performed with quasi-equilibrium initial conditions and slightly higher resolution than in our simulations --- see e.g.\\ \\citet{rukl02} --- numerical resistivity can mimic resistive effects, and such results could be more realistic than those in the high resolution ideal-MHD simulations \\citep{Y86}. We note that in high resolution simulations, physical resistivity should be included, or results could depend on numerical resistivity, which is not related to physical quantities of the setup. Any study of results is then necessarily flawed by purely numerical effects. Different models for resistivity should be investigated, too. A model for physical resistivity itself can affect the results, as was pointed out in \\cite{gods97}. To identify locations in which the launching is possible in our computational box, we compare different indicators for a successful launching. The best such indicator in general is the line where the magnetic and matter pressure are equal ($\\beta=1$). We also find that in cases with weak magnetic field, the Elsasser number $\\Lambda$, could serve as an indicator for successful launching. There is no disk component of flow in our results, and this is why they are very light. Matter in micro-ejections comes from the magnetosphere above the disk gap, which is just a small part of the matter from the disk inner radius, which slipped through the magnetic field lines." }, "1112/1112.5539_arXiv.txt": { "abstract": " ", "introduction": "} The Stellar Observations Network Group (SONG) is an initiative to design and build a global network of small telescopes. The goal of SONG is to become a key facility in both asteroseismology and planet search research programmes, a facility that provides state-of-the-art data of a quality that could not be achieved by use of any other space-based or ground-based facility \\cite{jcd06,gru06,gru08a,gru08b,gru09}. SONG is proposed to have a total of eight nodes, four located in the northern hemisphere and four in the south. By placing the telescopes at roughly equally-spaced longitudes, long-term nearly continuous observations can be obtained. Each telescope has an aperture of 1m and will be equipped with a high resolution spectrograph for measuring very precise doppler velocities and dual--color lucky-imaging cameras for photometry of faint stars in crowded fields. The prototype of SONG has been fully financed and a site on the Observatorio del Teide is being prepared for its installation. ", "conclusions": "" }, "1112/1112.5363_arXiv.txt": { "abstract": "We study the stability of the (87) Sylvia system and of the neighborhood of its two satellites. We use numerical integrations considering the non-sphericity of Sylvia, as well as the mutual perturbation of the satellites and the solar perturbation. Two numerical models have been used, which describe respectively the short and long-term evolution of the system. We show that the actual system is in a deeply stable zone, but surrounded by both fast and secular chaotic regions due to resonances. We then investigate how tidal and BYORP effects modify the location of the system over time with respect to the instability zones. Finally, we briefly generalize this study to other known triple systems and to satellites of asteroids in general, and discuss about their distance from mean-motion and evection resonances. ", "introduction": "A large number of satellites of asteroids have been discovered since the discovery of the satellite Dactyl, thanks to the Galileo flyby of (243) Ida (\\citealt{belton1996}). As of today, there is 206 known systems (binary, triple and quintuple), following the Johnston's archive online database\\footnote{\\url{http://www.johnstonsarchive.net/astro/asteroidmoons.html}} (see also the online database\\footnote{\\url{http://www.asu.cas.cz/~asteroid/binastdata.htm}} described by \\cite{Pravec2007} and \\cite{Pravec2011}) and it is believed that small binaries could represent a fraction of 15\\% of the NEA population (\\citealt{Margot2002}; \\citealt{Pravec2006}). Triple systems are rare and only nine known systems have been reported up to now in the entire Solar System. The dynamical evolution and formation mechanisms of these systems are highly dependent on the size ratio between the secondaries and the primary. If this ratio is very small, as in the case of the Main-Belt asteroid (243) Ida, the systems are similar to the classical dynamical problem of a massless satellite orbiting a planet (see for example \\citealt{Kozai1959,Kozai1962}), this one being replaced by a possibly highly elongated ellipsoid (\\citealt{Chauvineau1993}; \\citealt{Scheeres1994}; \\citealt{Scheeres1996}; \\citealt{Compere2012a}). On the other hand, systems with similar size components, as the Near-Earth asteroid (66391) 1999 KW4, have to be described taking into account both their shapes and their rotations. A lot of studies have been realized on the expression of the full two-body problem and the study of its characteristics (\\citealt{Maciejewski1995}; \\citealt{Scheeres2002}; \\citealt{Fahnestock2008}; \\citealt{Boue2009}). Similarly, emphasis has been given during the past decade on the description of dissipative effects on binary systems, like tidal effects (\\citealt{Mathis2009}; \\citealt{Goldreich2009}; \\citealt{Taylor2010,Taylor2011}) or BYORP (\\citealt{Cuk2005}; \\citealt{Cuk2010}; \\citealt{Mcmahon2010}; \\citealt{Steinberg2011}). We studied in this paper the dynamics and stability of the system (87) Sylvia, which was the first triple asteroid system discovered (\\citealt{Marchis2005a}). The specificities of this system place it in the first class described above. Sylvia, discovered in 1866, is a low-eccentric and midly-inclined asteroid located in the outer Main Belt. Its long-term evolution has been investigated through the AstDys project (\\citealt{Milani1998}; \\citealt{Knezevic2003}) giving its proper orbital elements ($\\overline{a}$ = 3.486 AU, $\\overline{e}$ = 0.0537, $\\overline{i}$ = $9.85^{o}$) and its secular fondamental frequencies ($n=55.297^{o}$/yr, $g = 134.798$\"/yr, $s = -130.782$\"/yr). Its orbit has been found to be slightly chaotic, exhibiting a Lyapunov time of $\\sim$ 1.4 Myr. The two satellites of Sylvia present near-circular and near-equatorial orbits, and have a mass ratio of about $10^{-4}$ and $10^{-5}$ with Sylvia. The outermost satellite, Romulus, is approximately ten times more massive than the innermost one, Remus, for a semi-major axis twice as important. \\cite{Winter2009} studied the system and found that the satellites could be highly unstable when the oblateness of Sylvia (even a small fraction) is not taken into account. Indeed, the oblateness of the asteroid, as well as the short distance of the satellites from its surface ($\\sim$ 5 and 10 radius of Sylvia), critically increase the precession frequencies and prevent them from commensurabilities with frequencies arising from other gravitational perturbations. Our aim is the understanding of the dynamical mechanisms present in the system and in its neighborhood. We then generalize some of the results to the other triple systems, and, in a general way, to the systems similar to (87) Sylvia, e.g. with a small size ratio and a primary diameter of the order of $\\sim$ 100 km. ", "conclusions": "We studied the dynamics and stability of the (87) Sylvia triple system by using numerical integrations of the complete and averaged equations of motion. We used a shape of Sylvia derived from light-curves observations and up to $C_{4,4}, S_{4,4}$ for the complete integrations. The position of the actual system lies in a very stable zone. We showed the possible evolutions of the system trough tidal and BYORP effects, and showed that the system, currently lying between the mean-motion resonances 2:1 and 3:1, will likely evolve through the evection resonance before the MMR 2:1 in the future. The other known triple system considered here, except (216) Kleopatra, also lie between the MMR 2:1 and 3:1. Finally, we show that the evection resonance could limit the outward evolution of the satellites. \\paragraph*{Aknowledgments} The authors warmly thank A. Lemaitre for her constructive comments and her careful reading of the manuscript. Numerical simulations were made thanks to the local computing resources (Clusters ISCF and URBM-SYSDYN) at the University of Namur (FUNDP, Belgium). This work was supported by FAPESP (process n\\degre 2010/52715-5)." }, "1112/1112.2775_arXiv.txt": { "abstract": "Using data mining techniques applied on emission line characteristics of Be stars spectra we attempted to find new Be stars candidates in SDSS SEGUE survey. The mid-resolution spectra of confirmed Be stars obtained from VO-compatible archive of Ond\\v{r}ejov observatory 2m telescope were transformed to the spectral resolution of SDSS and important characteristics of emission line profiles were estimated, to be used as a training base of supervised learning methods. The obtained knowledge base of the characteristic shapes and sizes of Be emission lines was finally used to identify new potential candidates in SDSS spectral survey. The several newly found Be stars candidates justify our approach and approve Astroinformatics as a viable research methodology. ", "introduction": "Current data deluge in astronomy requires applying data mining techniques to extract new information about the physical nature of celestial objects. The possibility of cross-matching several surveys via Virtual Observatory protocols may play a key role in future discoveries. Data mining of large collections of spectra seems to be one of promising as well as challenging topics. We have focused on obtaining new candidates of H$\\alpha$ emission stars using supervised data mining method of Decision Trees on almost 200,000 spectra in SDSS SEGUE \\citep{yanny2009segue} spectral survey. ", "conclusions": "\\begin{figure}[!htb] \\plotfiddle{P141_f4.eps}{6.2cm}{0}{70}{50}{-200}{20} \\caption{The example of candidate Be star found in SDSS SEGUE survey} \\end{figure} \\noindent The classifier has identified 1110 Be stars candidates in SEGUE, however most of them are probably of different nature (e.g. AGNs, young stellar objects or reduction artifacts). Nevertheless, there are as well several highly probable Be stars like the one on Fig.~3." }, "1112/1112.0066_arXiv.txt": { "abstract": "In this work we propose a new orbital architecture for the two proposed circumbinary planets around the polar eclipsing binary HU Aquarii. We base the new two-planet, light-travel time model on the result of a Monte Carlo simulation driving a least-squares Levenberg-Marquardt minimisation algorithm on the observed eclipse egress times. Our best-fitting model with $\\chi_{r}^2=1.43$ resulted in high final eccentricities for the two companions leading to an unstable orbital configuration. From a large ensemble of initial guesses we examined the distribution of final eccentricities and semi-major axes for different $\\chi_{r}^2$ parameter intervals and encountered qualitatively a second population of best-fitting parameters. The main characteristic of this population is described by low-eccentric orbits favouring long-term orbital stability of the system. We present our best-fitting model candidate for the proposed two-planet system and demonstrate orbital stability over one million years using numerical integrations. ", "introduction": "In the past two decades, astrophysical timing measurements has been used to infer the existence of multiple low-mass planetary objects. \\cite{WolszczanFrail1992} announced the first detection of a planetary system around the pulsar PSR1257+12. A two-planet system orbiting the short-period subdwarf B of the eclipsing binary HW Virginis was first presented in the works by \\cite{Lee2009}, and recently \\cite{Beuermann2010}, \\cite{Potter2011}, and \\cite{Doyle2011} announced the existence of two circumbinary planets in possible mean motion resonances around NN Serpentis, two giant planets orbiting the eclipsing binary UZ Fornacis, and a single circumbinary planet around the stars of the binary system Kepler 16, respectively. The studies by \\cite{Beuermann2010} and \\cite{Potter2011} inferred the presence of additional massive objects by explaining the observed timing anomalies with the light-travel time (LTT, hereafter) effect. In the ideal case, the stellar components of an eclipsing binary system are orbiting their common centre of mass with a constant period. Timing irregularities can occur if the binary system is accompanied by an additional massive object. In this case, the binary centre of mass is orbiting the system's centre of mass. At some times the binary will be closer to the observer while at other times, it will be farther away, giving rise to the LTT effect on measured eclipse egress times. In a recent study, the measurements of eclipse egress times of the eclipsing polar binary HU Aquarii (HU Aqr, hereafter) was used to infer the presence of a circumbinary planet around this system \\citep{Qian2011}. These authors modeled the complete timing data set by adding the LTT effects from two circumbinary planets and found that the pericentre of the outer planetary companion is inside the orbit of the inner planet. This orbital architecture implies a crossing orbit configuration which points to strong mutual perturbations between the two planets. \\cite{Horner2011} subsequently carried out a detailed stability analysis of these bodies and found that almost all their initial conditions lead to unstable orbits on short-time scales. These authors concluded that the HU Aqr planetary system has most likely a different orbital architecture than proposed by \\cite{Qian2011}. Three possibilities exist to explain the instability of the proposed planetary system. Either i) the LTT parameter space was not explored thoroughly while orbital stability constrains were imposed [omitted in \\cite{Qian2011}], or, ii) the applied LTT model was not complete and was unable to explain the timing data set properly, or iii) the data set is not yet large enough to draw reliable conclusions on the existence of additional low-mass companions. In this work we aim to find a more plausible orbital architecture that best describes the timing data while also being conform with orbital stability requirements. We regard the stability condition as an observable that places additional constraints on the fitting process \\citep{Gozdziewski2005}. We carried out a large-scale Monte Carlo, least-squared parameter survey by fitting various LTT models to the complete data set. In parallel to the fitting process, we performed a stability analysis of various LTT model orbits which best describe the data set. By parameterising planet's Hill radii, we imposed orbital stability constraints to our parameter survey in order to explore orbital architectures that result in stable orbits. We extended the models by \\citet{Qian2011} to systems that also allow the inner planet to attain an eccentric orbit. Our results suggest that two-planet LTT models in general tend to produce orbits with higher eccentricities which are likely unstable. The outline of this paper is as follows. In section 2 we present details of the adopted LTT model and provide a brief description of the least-squared minimisation algorithm. In section 3 we introduce our orbital stability constrains and in section 4 we describe results of numerical experiments. Finally, we give a summary and discussion in section 5. ", "conclusions": "In a recent study \\cite{Qian2011} announced the discovery of a two-planet system around the eclipsing polar binary HU Aquarii. In a subsequent paper, the stability of their best-fitting orbital parameters was examined by \\citet{Horner2011} who showed that the system is highly unstable. The most likely reason for this instability is the high eccentricity of the outer companion. In an attempt to find a system with stable orbits, we calculated a large number of two-planet LTT models using a Monte Carlo approach. This approach provides statistics on the distribution of the final goodness-of-fit $\\chi_{r}^2$ parameter in combination with constrains on requiring initial orbital stability. In this work, we analyzed the complete available timing data of HU Aquarii. We made the following assumptions in our analysis. First we allowed the inner planet to attain an eccentric orbit in response to possible eccentricity excitation by an outside perturber. Compare to the study by \\cite{Qian2011}, this assumption resulted in introducing three extra free parameters. We also did not include a secular term which would account for timing variations due to mass-/angular momentum transfer and/or magnetic activity of the binary component. \\citet{Qian2011} argue that the secular term in their work is most likely caused by the presence of a third circumbinary massive object. However, the existence of this body in their data is far from obvious as the observational time span is much smaller than the orbital period of the proposed third circumbinary (planetary) companion. The confirmation of the third planet would require a few decades of future photometric follow-up monitoring. Finally, we did not consider mutual gravitational interactions between the two planets or gravitational perturbations on the binary stars. Our best fit model had a goodness-of-fit parameter of $\\chi_{r}^2=1.43$ occurring fourteen times out of 111.844 initial guesses. From studying the histogram distribution of final $\\chi_{r}^2$ parameters, we were faced with a dichotomy. On the one hand, our best-fitting model of $\\chi_{r}^2=1.43$ had a low occurrence frequency (Fig.~\\ref{chi2histogram}) and resulted in high eccentricities (Fig.~\\ref{e-distribution}a). On the other hand, we encountered a significant higher occurrence frequency of larger $\\chi_{r}^2$ values. At the moment we have no explanation for this trend. One possibility is the inability of the LM minimisation algorithm to escape a local minimum in the non-linear parameter space. To test this possibility, one can use a least-squared minimisation procedure based on a genetic algorithm (GA) or a Bayesian Markov Chain Monte Carlo method. The results of our study suggest that we should be able to trust the nominal $\\chi_{r}^2=1.43$ best-fitting parameters as this fit represents, most likely, the global minimum in $\\chi_{r}^2$ parameter-space, and provides the best possible description of the observed timing data set. However, as we have shown, the resulting best-fitting orbit with $\\chi_{r}^2=1.43$ was unstable. Models in the neighbourhood of this system were also unstable with a collision or ejection taking place on very small time scales. This instability can be attributed to high values of the final fitted eccentricities for one or both planetary companions. It is worth mentioning that models with goodness-of-fit close to the nominal best-fitting model did not allow low-eccentricity/circular orbits which would have rendered them most likely to be stable. We would like to mention that other authors have encountered similar (in)stability problem for the two planets in the systems of NN Serpentis and UZ Fornacis as well \\citep{Potter2011, Beuermann2010}. The two-planet system HW Virginis \\citep{Lee2009} also appears to be unstable (B. Funk, private communication). In particular, it is interesting to note that the best-fitting models of the proposed UZ Fornax circumbinary planetary system also attained high final eccentricities \\citep{Potter2011}. Examining the $\\chi_{r}^2$ frequency distribution in detail resulted in the emergence of a second population of models with fitted parameters that fulfill our orbital stability constraints. Orbital stability was examined for models with $\\chi_{r}^{2}=1.89$. Although this population had a larger $\\chi_{r}^2$ goodness-of-fit than the nominal case, we were able to find stable orbits due to significantly lower orbital eccentricities. Combining/pairing a least-squared minimisation algorithm with a stability analysis enabled us to determine a new set of orbital elements for the two planets which are different from those given by \\citet{Qian2011}. The latter is another implication of our study which points to the existence of a variety of stable configurations when considering models with increasingly larger $\\chi_{r}^2$. Assuming that no other physical effects are capable of explaining the observed timing variations in the light curve of HU Aquarii, we are left with the two-planet model as the most probable cause for the observed timing variation. Our study indicates that single one-planet LTT sinusoidal timing variations are inadequate to describe the observational data. However, it is necessary to mention that timing anomalies could also be caused by direct perturbations of a circumbinary planet on the binary orbit, though these perturbations would introduce a high-frequency (short-period) component to the eclipse timing variations due to the short orbital period of the binary system. A second possibility worthy to consider in future work would be to include mutual perturbations between the two proposed circumbinary planets. Explaining eclipse timing variations caused by the latter effect has not been explored for circumbinary planetary systems. Finally, a similar study by \\cite{Wittenmyer2011} also addresses the instability of the two proposed companions around HU Aqr from revised LTT models. In their work the authors point out the need to obtain a better understanding of the mutual interaction between the two binary components. Binary orbital period modulation caused by magnetic interaction and/or mass-transport between the two binary components could also possibly explain the observed timing anomaly for HU Aqr. However, a combination of the above effects might also provide a satisfactory description of the observed timing data. At this point, we recommend future discoveries of multi-planet circumbinary systems also include a preliminary orbital stability study as a necessary condition to increase the likelihood of the existence of such systems. At the moment it seems that best-fitting single LTT models in superposition ($\\tau_1 + \\tau_2$) are inadequate to result in stable multi-planetary circumbinary systems. Though such a model provide a good description to the observational timing data. Furthermore, we also recommend future studies to obtain and publish timing data of the secondary eclipse (if available). Any timing variations introduced from gravitational interactions and/or LTT should also provide an adequate description of timing measurements obtained from the secondary eclipse. Further constraining of our model requires obtaining additional timing data of the HU Aqr system. Using the results of our study involving stable planetary orbits, we predict that future mid-egress timing measurements would result in $O-C$ timing differences of -10.0 to 0.0 seconds (predicted at epoch $E = 40000$ in Fig.~\\ref{BestFitOmC_2}). We emphasise that high-accuracy timing measurements are crucial to unveil the true nature of the observed timing anomalies as well as the development of improved models describing eclipse (egress) timing variations of a binary star system." }, "1112/1112.2580_arXiv.txt": { "abstract": "LOFAR is a groundbreaking low-frequency radio telescope currently nearing completion across northern europe. As a software telescope with no moving parts, enormous fields of view and multi-beaming, it has fantastic potential for the exploration of the time-variable universe. In this brief paper I outline LOFAR's capabilities, as well as our plans to use it for a range of transient searches and some crude estimated rates of transient detections. ", "introduction": "LOFAR, the Low Frequency Array, is a large, low-frequency radio telescope in northern Europe, led by ASTRON. Construction of the array, which has its core collecting area in The Netherlands, with international stations in France, Germany, Sweden and The UK, is nearly complete, and astronomically interesting data are now being taken. LOFAR operates in the 30--80 and 120--240 MHz frequency ranges. The 80--120 MHz frequency gap corresponds to the FM radio bands at which frequencies astronomical observations would be impossible\\footnote{Unless northern europe could be persuaded to stop night-time FM radio broadcasts for a few weeks in the interests of finding the {\\em Epoch of Reionisation} signal..}. Construction of the array is almost complete -- See Fig \\ref{europe} for the distribution of operating LOFAR stations across Europe. In addition, observations are occasionally possible to frequencies as low as 15 MHz. For a full reference paper on LOFAR see van Haarlem et al. (2012). \\begin{figure}[t] \\begin{center} \\includegraphics[width=11.0cm]{europe.eps} \\caption{The distribution of complete LOFAR stations across Europe.} \\label{europe} \\end{center} \\end{figure} LOFAR has six {\\em key science projects} (KSPs), one of which is {\\em Transients} (principal investigators Fender, Stappers, Wijers). The remit of the TKSP covers essentially all transient and variable astrophysics, including commensal searches of all data (ultimately in near-real-time, although this functionality is not yet implemented). The TKSP covers both time-series and image-plane searches for transients and variables, including pulsars (Stappers et al. 2011). The adoption of transients and variables as key science drivers for the project is a theme for most of the large SKA pathfinders and precursors, in general unlike older radio facilities. However, time-series and image-plane transients have been separated for both ASKAP (which has {\\em CRAFT} and {\\em VAST} respectively) and MeerKAT ({\\em TRAPUM} and {\\em ThunderKAT}). This makes some sense from a techniques point of view, although there is some overlap in the science. ", "conclusions": "LOFAR has a wide range of diverse capabilities (multi-beaming, simultaneous timing and imaging modes, splitting of array in large numbers of individual stations, re-imaging in the past with the TBBs) which are ideally suited for exploring transient parameter space. Just as importantly, there is the will to support this science, with a Key Science Project dedicated to precisely this exploration. Based on current estimates, it should find many 1000s of interesting radio transients per year, providing a huge target list for multiwavelength follow-up, and providing new tests of our ability to automatically detect, classify and report such events efficiently." }, "1112/1112.2549_arXiv.txt": { "abstract": "One of the main debated astrophysical problems is the role of the AGN feedback in galaxy formation. It is known that massive black holes have a profound effect on the formation and evolution of galaxies, but how black holes and galaxies communicate is still an unsolved problem. For Radio Galaxies, feedback studies have mainly focused on jet/cavity systems in the most massive and X--ray luminous galaxy clusters. The recent high--resolution detection of warm absorbers in some Broad Line Radio Galaxies allow us to investigate the interplay between the nuclear engine and the surrounding medium from a different perspective. We report on the detection of warm absorbers in two Broad Line Radio Galaxies, 3C~382 and 3C~390.3, and discuss the physical and energetic properties of the absorbing gas. Finally, we attempt a comparison between radio--loud and radio--quiet outflows. ", "introduction": "In the last few years, high--resolution X--ray spectroscopy has made progress in the exploration of the circumnuclear environment of radio--loud (RL) AGNs. While the presence of X--ray emitting and absorbing gas is well established in Seyfert galaxies, for RL sources the investigation of the nuclear environment through this technique is very recent. Specifically, in Broad Line Radio Galaxies (BLRG), the RL counterpart of Seyfert 1s, the detection of warm absorbers (WA) \\footnote{With the term ``warm absorber'' we intend ionized outflowing gas in our line--of--sight that produces narrow absorption lines in the soft X--ray spectrum.} was expected to be more difficult because of their small number in the local Universe and because the Doppler amplification of the jet emission could mask the absorption features. However, steps forward have been recently made thanks to high--resolution X--ray spectroscopy. Here we summarize the results concerning the discovery of WAs in two BLRGs, 3C~382 and 3C~390.3. ", "conclusions": "We report on the detection of WAs in two BLRGs, 3C~382 and 3C~390.3, and discuss the physical and energetic properties of the absorbing gas: {\\bf (i)} the outflows are highly ionized (log$\\xi >$2~erg~cm~s$^{-1}$) and slow, with velocities ranging between 10$^{2}$--10$^{3}$~km~s$^{-1}$; {\\bf (ii)} the mass outflow rates are higher than the mass accretion rates if a volume filling factor (C$_{\\rm v}$) equal to 1 is assumed. Therefore a gas clumpy configuration (C$_{\\rm v} <$1) is expected; {\\bf (iii)} the kinetic luminosity associated to these slow outflows is always lower than the accretion luminosity and the jet kinetic power; {\\bf (iv)} although RL and RQ WA physical properties appear very similar, at least at zeroth order, the mass outflow rate and the radio--loudness parameter (R) seem to be correlated. This correlation could indicate a different gas distribution or alternatively, if the gas distribution is the same, powerful jets could favor the escape of more massive winds. \\begin{figure}[pbh] \\begin{center} \\psfig{file=Torresi_1_f2.eps,width=3.7 cm, angle=0} \\vspace*{8pt} \\caption{Mass outflow rate plotted against the radio--loudness parameter (R). {\\it Red circles}: BLRGs considered in this work; {\\it black squares}: type 1 RQ AGNs belonging to our reference sample. \\label{f2}} \\end{center} \\end{figure}" }, "1112/1112.0299_arXiv.txt": { "abstract": "Reports of the death of the precursor of Supernova (SN) 1961V in NGC 1058 are exaggerated. Consideration of the best astrometric data shows that the star, known as ``Object 7,'' lies at the greatest proximity to SN 1961V and is the likely survivor of the ``SN impostor'' super-outburst. SN 1961V does not coincide with a neighboring radio source and is therefore not a radio SN. Additionally, the current properties of Object 7, based on data obtained with the {\\sl Hubble Space Telescope}, are consistent with it being a quiescent Luminous Blue Variable (LBV). Furthermore, post-explosion non-detections by the {\\sl Spitzer Space Telescope\\/} do not necessarily and sufficiently rule out a surviving LBV. We therefore consider, based on the available evidence, that it is yet a bit premature to reclassify SN 1961V as a {\\it bona fide\\/} SN. The inevitable demise of this star, though, may not be too far off. ", "introduction": "The evolution of the most massive stars is not well known. It is thought that main sequence stars with $M_{\\rm ZAMS} \\gtrsim 30\\ M_{\\sun}$ proceed through a blue supergiant phase, possibly to a red supergiant phase, or straight to a luminous blue variable (LBV) phase, and to the Wolf-Rayet (WR) phase prior to explosion as supernovae (SNe). That WR stars typically possess $M\\lesssim 20\\ M_{\\sun}$ \\citep{crowther07} requires their precursor stars to shed most of their mass, presumably through eruptive mass ejections as a LBV \\citep{humphreys94,smith06}. The term ``supernova impostor'' has been coined \\citep{vandyk00} to describe these eruptive events for massive stars, such as $\\eta$~Carinae \\citep[e.g.,][]{davidson97}. This is because various observational aspects of the eruptions can mimic the properties of true SNe. During $\\eta$~Car's Great Eruption in the 1800s, the star greatly exceeded the Eddington limit, with the bolometric luminosity increasing by $\\sim$ 2 mag. The total luminous output of such an eruption ($\\sim 10^{49}$--$10^{50}$ erg) can rival that of a SN. However, a SN, by strict definition, is an explosive event that ends the life of a star (although a compact object may form in the process). If an optical transient is observed with energetics comparable to that of a true SN, and, after a sufficient period of time has elapsed, a star still exists at the exact position of the transient, then that transient is a SN impostor. The most studied example of a SN impostor, SN~1961V in NGC 1058, remains controversial to this day. A debate has continued as to whether or not this event was a true SN. In fact, it is truly impressive how many journal pages have been spent on this one object. This includes two recent studies, both of which present arguments that claim SN 1961V as a genuine SN. We argue here that SN 1961V is an impostor and that its precursor has survived what was a powerful eruption. Many of our points below were first presented in our review of SN impostors \\citep{vandyk11}. ", "conclusions": "Based on our analysis of the available data, we consider SN 1961V (still) to be a SN impostor and that Object 7 is the survivor of this event. All of the arguments we have made above, as part of this analysis, support this conclusion. We have also attempted to dispel several erroneous suppositions about the nature of this event. The survivor is clearly observed in very recent years to exist. The star quite plausibly has the properties of a quiescent, massive LBV. The star has not yet exploded and is not a radio SN. The star is surrounded by a circumstellar shell or nebula, however, this shell is not nearly as dusty as required by \\citet{kochanek11} and previous investigators. We find that the visual extinction to Object 7 is in the range $A_V=1.8$--2.3 mag, and we suggest that most of this extinction is arising from the interstellar medium along our line-of-sight within the host galaxy, rather than from the shell itself. As a result, therefore, the shell need not be nearly as luminous an IR emitter as $\\eta$ Car. This conclusion is further supported by the mid-IR observations of other SN impostors. The positional proximity of Object 7 to SN 1961V, and the positional offset between SN 1961V and the radio source centroid, are both inescapable facts. The only conceivable way to counter the former fact is to conclude that Object 7 is either a physical or an optical double to the progenitor of an actual SN. This, of course, is possible. The uncertainty alone in the absolute position, $0{\\farcs}1$, corresponds to $\\sim 5$ pc, certainly allowing for the possibility, in the relatively crowded cluster environment of SN 1961V, that Object 7 is merely a neighbor to the progenitor. Also, the probability of core-collapse SNe for high-mass stars in binaries is relatively high \\citep[e.g.,][]{kochanek09}. Although, if Object 7 {\\em were\\/} the binary companion to the star that exploded, we might expect Object 7 to have been stripped or otherwise affected by what would have been a very powerful explosion, potentially leading to an unusual brightness or color. This is not supported by the available photometry or inferred luminosity of Object 7. Instead, it appears to have the properties of a ``run-of-the-mill,'' evolved, high-mass star. Nonetheless, the positional offset between SN 1961V and the radio centroid is insurmountable --- these two are not one and the same, which is essentially at the heart of the case that has been made for SN 1961V being a true SN. No need exists for the supposed core-collapse explosion ``hybrid'' of a SN II-P and SN IIn. Furthermore, we disagree with the statement made by \\citet{smith11} that SN 1961V is somehow unique among the impostors. It has direct analogs in both SN 2000ch \\citep{wagner04,pastorello10} and SN 2009ip in NGC 7259 \\citep{smith10,foley11}, specifically in terms of the inferred expansion velocities at eruption and of the light curve behavior. The recent, more powerful revival of SN 2000ch displays ``unusually'' high velocities (FWHM) of 1500--2800 km s$^{-1}$ \\citep{pastorello10}. SN 2009ip shows blueshifted He\\,{\\sc i} absorption at 3000--5000 km s$^{-1}$, which \\citet{smith10} speculate for SN 2009ip could be due to a fast blast wave, which occurred quasi-contemporaneously with the origin of the slower ejecta. SN 2000ch exhibits multiple P Cygni absorption components, with profile edges up to 3000--3500 km s$^{-1}$. A sustained high-luminosity pre-eruption state was seen for both SNe 2000ch and 2009ip; SN 2000ch was at $M_R\\simeq -10.7$ mag prior to the 2000 eruption \\citep{wagner04}, and SN 2009ip had a pre-eruption luminosity of $M_V\\simeq -10$ mag \\citep{smith10}. Although not nearly the $M_{\\rm pg}\\approx -12$ mag brightness of SN 1961V, both SNe 2000ch and 2009ip, therefore, both may also have been in a super-Eddington phase prior to their giant eruptions. Even though the line widths were different, it is interesting to note that the H$\\alpha$ line profiles, in particular, of SN 1961V \\citep[][see his Figure 2]{zwicky64} and SN 2000ch \\citep[][see their Figure 7]{wagner04}, near maximum of the 2000 eruption, were very similar --- a sharp drop-off to the blue wing and extended wing to the red. The H$\\alpha$ emission line in 2002 for Object 7/SN 1961V also showed a similar asymmetric profile (and, we emphasize that this profile had a very similar shape as the much broader line profile for SN 1961V in 1962, as shown by \\citeauthor{zwicky64}). A possible explanation could be the effects of dust extinction in the expanding ejecta, although we have shown that the extinction from the ejecta is likely relatively modest for SN 1961V. A detailed analysis of this effect should be explored, although we consider this to be beyond the scope of this paper. We do agree with \\citet{smith11} and \\citet{kochanek11}, however, that the ``undulations'' in SN 1961V's post-maximum light curve may well have arisen from the blast wave overtaking previously-ejected shells of matter ahead of the shock. We speculate here that the star was already in a sustained, eruptive outburst prior to 1960 and that the onset of the super-outburst, which peaked in luminosity in 1961 December, could possibly have been due to the interaction of the fast-moving ($\\sim$2000 km s$^{-1}$), dense, massive shell with the pre-existing, slower, less dense mass loss. A potential analog is the behavior of the SN IIn 1994W in NGC 4041 --- \\citet{dessart09} modeled the high luminosity, sustained plateau, and sudden drop-off of this SN as the interaction of a fast ($\\gtrsim 1500$ km s$^{-1}$), dense shell interacting with a slower, less dense shell. Furthermore, both the brief and more sustained plateaus in the SN 1961V light curve could have arisen from the interaction of the fast blast wave with various regimes of previously ejected matter, or to periods of partial recombination in the expanding ejecta, analogous to the recombination wave in SN II-P ejecta. Alternatively, these plateaus could also have been due to subsequent lesser eruptions by the star (Humphreys et al.~\\citeyear{humphreys99}). The fast-moving massive shell radiated for many years after the super-outburst, albeit more faintly. \\citet{goodrich89} detected the broad emission-line component at $F_{{\\rm H}\\alpha} \\simeq 2.2 \\times 10^{-16}$ erg cm$^{-2}$ s$^{-1}$ in 1986. We can set a limit on detection of this component in the {\\sl HST}/STIS spectrum from 2002 at $F_{{\\rm H}\\alpha} < 3 \\times 10^{-16}$ erg cm$^{-2}$ s$^{-1}$ (see Figure~\\ref{figspec}). \\bibpunct[; ]{(}{)}{,}{a}{}{;} It is correct to point out the similarities in the variety of SN impostor and SN IIn properties. The two are very likely intimately intertwined, as we now know that at least one SN IIn arose from a SN impostor \\citep[SN 2005gl;][]{galyam07,galyam09}. The SN IIn are also a diverse class of objects (if the term ``class'' is truly applicable in the case of such heterogeneity), and it is still not clear how many SNe IIn have, in fact, really been impostors. For instance, SN 1994W itself may be an impostor \\citep{dessart09}. Obviously, we have extrapolated fairly extravagantly from two photometric measurements of Object 7 from recent {\\sl HST\\/} data. We have had to assume that the colors for the star from 1991 are applicable to 2008 as well. What is ultimately required is a set of deep, multiband optical and near-IR observations of the SN 1961V environment using the modern {\\sl HST} with the Wide Field Camera 3, which would vastly improve upon the pre-furbishment WF/PC-1 images and the motley assortment of WFPC2 and unfiltered STIS image data analyzed to date. Such future observations would only require a rather modest investment of {\\sl HST\\/} time --- to image in $BVRIJH$ at $S/N\\gtrsim 10$ would require only 3 orbits. Looking beyond {\\sl HST}, observations with the {\\sl James Webb Space Telescope\\/}, particularly with the Mid-IR Instrument (MIRI), would reveal the actual dust emission from the surviving star. This object has clearly garnered much attention and speculation over the decades and continues to provide us with invaluable insights into the evolution of the most massive stars. With these space-based data, the definitive nature of the SN 1961V precursor may well be established once and for all. \\bibpunct[; ]{(}{)}{;}{a}{}{;} Finally, NGC 1058 should continue to be regularly monitored by SN searches in nearby galaxies. For Object 7 as the SN 1961V survivor, we might expect, much as was the case for SN 2006jc \\citep{foley07,pastorello07}, that the star will {\\em truly\\/} explode as the SN that we believe other authors have erroneously concluded has already occurred. The super-outburst from the 1960s could well be the forerunner of a core-collapse event. It is possible that the star will explode as a high-luminosity, hydrogen-rich SN IIn, such as SN 2006gy \\citep{ofek07,smith07b}, or as a more helium-rich SN Ibn, such as SN 2006jc \\citep[as suggested by][]{kochanek11}." }, "1112/1112.2255_arXiv.txt": { "abstract": "{}{We analyze coronagraph observations of a polar jet observed by the Sun Earth Connection Coronal and Heliospheric Investigation (SECCHI) instrument suite onboard the Solar TErrestrial RElations Observatory (STEREO) spacecraft.} {In our analysis we compare the brightness distribution of the jet in white-light coronagraph images with a dedicated kinetic particle model. We obtain a consistent estimate of the time that the jet was launched from the solar surface and an approximate initial velocity distribution in the jet source. The method also allows us to check the consistency of the kinetic model. In this first application, we consider only gravity as the dominant force on the jet particles along the magnetic field.} {We find that the kinetic model explains the observed brightness evolution well. The derived initiation time is consistent with the jet observations by the EUVI telescope at various wavelengths. The initial particle velocity distribution is fitted by Maxwellian distributions and we find deviations of the high energy tail from the Maxwellian distributions. We estimate the jet's total electron content to have a mass between $3.2\\times10^{14}$ and $1.8\\times10^{15}$ g. Mapping the integrated particle number along the jet trajectory to its source region and assuming a typical source region size, we obtain an initial electron density between $8\\times10^9$ and $5\\times10^{10}$ cm$^{-3}$ that is characteristic for the lower corona or the upper chromosphere. The total kinetic energy of all particles in the jet source region amounts from $2.1\\times10^{28}$ to $2.4\\times10^{29}$ erg.}{} ", "introduction": "Polar coronal jets were originally observed by the Extreme-Ultraviolet Imaging Telescope (EIT) onboard the Solar and Heliospheric Observatory (SOHO)~ \\citep{Domingo:etal:1995}. In white light, the polar jets discovered by LASCO/SOHO \\citep{StCyr:etal:1997} appear narrow and collimated, and expand rapidly as they travel through polar regions. They are often associated with an Extreme-Ultraviolet (EUV) jet seen near the solar surface \\citep{Wang:etal:1998}. These jets are often rooted in bright, low-lying loop features and are similar in appearance to Soft X-ray (SXR) jets \\citep{Shibata:etal:1992,Moses:etal:1997}. The essential acceleration mechanism for all these jets is very likely provided by magnetic reconnection. The difference between the above and other jet-like features, e.g. chromospheric jets, is the altitude where the magnetic reconnection is assumed to occur. The higher energy jets tend to be accelerated at a higher altitude than the lower energy jets \\citep{Shibata:etal:2007}. The EUV and SXR jets are often caused by the reconnection in the upper chromosphere or the lower corona. \\citet{Wang:etal:1998} analyzed 27 jets in EIT and LASCO data and characterized their motion by three different velocities: the leading-edge velocity $v_{lead}$, the centroid velocity $v_{cen}$ and the initial velocity $v_{init}$ of the centroid. In all cases $v_{cen}$ was much less than $v_{lead}$, indicating that the jets stretched out rapidly as they propagated through the corona. The authors also found that the bulk of the jet material decelerated as it propagated from the limb to the C2 field of view (FOV). This deceleration was attributed to solar gravity. However, the combined results of $v_{init} < v_{escape}$ and $v_{lead} > v_{escape}$, and the lack of evidence for downflow in EIT and C2 led Wang et al. to propose some in situ acceleration that prevents the bulk of the jet material from falling back onto the Sun after it was ejected. \\citet{Wood:etal:1999} improved the jet velocity estimates by \\citet{Wang:etal:1998} using on height-time plots. The authors determined the trace of the jet centroid from the EIT to the C2 FOV and found that the observed kinematic trajectories could be fitted with some success by ballistic orbits. They concluded, however, that gravity alone was not the only force controlling the jet propagation. Because of the similar behavior of the jets studied, both \\citet{Wood:etal:1999} and \\citet{Wang:etal:1998} suggested that by the time jets reached the C2 FOV, they were incorporated into the ambient solar wind. More recently, \\citet{Ko:etal:2005} studied a jet observed jointly by several instruments above the limb. These authors found that a ballistic model could explain most of the dynamical properties of this jet. In their model it was assumed that the gas was ejected upward from the surface with a range of initial speeds. The smooth change of the upflow-to-downflow speed at a certain altitude derived from the ballistic model was found to be consistent with the change of the line intensities from Doppler dimming observed by UVCS/SOHO. Owing to a lack of high cadence coronagraph observations Ko et al.'s study was essentially confined to heights below 1.64 $r_\\odot$. On June 7, 2007 a big eruptive jet was observed by EUVI, COR1, and COR2 on board the STEREO mission with higher spatial and temporal resolution compared to the data from EIT and LASCO C2. It extended from the solar surface to 5~$r_{\\odot}$. The event was also studied by \\citet{Patsourakos:etal:2008} from the stereoscopic viewpoint. They estimated the jet positions and the speed of the leading front at different times in the EUVI FOV. In this paper we will attempt to analyze the same jet based on white-light coronagraph observations at heights beyond about 1.5 $r_\\odot$. We extend the ballistic approach by \\citet{Ko:etal:2005} by quantitatively comparing the density variation from Thomson scattered white-light brightness at different heights to the variation expected from a ballistic model of the jet particles. This method has the advantage that it avoids the estimates of jet centroids and fronts. These fronts are not well-defined for a diffusively spreading plasma cloud, the jet centroids are often difficult to determine because a substantial part of jet material is hidden behind the occulter. Another advantage of our particle kinetic analysis lies is that it also provides a test of the validity of the ballistic model. We therefore receive much more information than the conventional leading edge, centroid velocity measurements. After the description of the observations in \\S2, we will introduce the ballistic model in \\S3. In \\S4 we present the results and try to extrapolate our findings to the jet source region and also discuss the limits of our model. Finally, we summarize our conclusions in the last section. ", "conclusions": "We have followed the evolution of a big eruptive jet event observed by SECCHI in both EUVI images and the COR1 and COR2 white-light coronagraphs. Based on the distance-time brightness analysis, we found that a ballistic model for the jet particles in general can explain quite well the brightness variation beyond 1.5 $r_\\odot$ in the COR1 and COR2 fields of view with gravity as the dominant acceleration of the jet particles. Additional parameters were derived or at least estimated, such as the initiation time, the initial velocity distribution, and the number of the jet particles. The derived initiation time is consistent with the EUVI observations at lower altitudes. The initial velocity distribution was fitted by two Maxwellian distributions with different mean kinetic energies. The good agreement with a Maxwellian for lower initial velocities may be due to collisions at heights below 1.4 $r_\\odot$. At high initial velocities, the distribution deviates from the Maxwellian toward a power law tail that may be a result of the jet acceleration process. The total jet particle number and kinetic energy sum up to about 1.6 to 8.9$\\times10^{38}$ and 2.1 to 24$\\times10^{28}$ erg, respectively. We neglected the effect of the magnetic mirror force and of Coulomb collision. As discussed, they might have some effect on the kinetics of the jet particles at lower altitudes. Note that these two forces counteract each other: while the collisions with the background plasma will decelerate the particles, the mirror force accelerates them away from the Sun. Especially the correct modeling of the Coulomb collisions below 1.4 $r_\\odot$ requires additional assumptions, e.g., about the coronal background density and its velocity. We outlined the basic idea of a new kinetic jet analysis. In the future, a more sophisticated kinetic model of the jet may be compared to white-light observations. We have shown that this comparison allows one to constrain details of the jet which could not be derived in previous studies. More work needs to be done in these directions. Moreover, more jet samples are required to find out to which extent a jet is embedded in the ambient solar wind and how the jet interacts with it. PROBA-3, which will be launched in a few years, will have a coronagraph with a FOV of 1.04 to 3 $r_\\odot$. It will provide us with a broader initial velocity coverage because the lower velocity limit depends on the occulter's size . Therefore we will have less uncertainty in the initial velocity distribution, the electron density in the jet source region, etc. The higher temporal observations with more wavelength coverage from AIA/SDO will help determine the jet initiation time more precisely. Recently, \\citet{Raouafi:etal:2008} found that a jet was very often succeeded by a plume above the jet launch site. Interestingly, in our jet study a plume was visible in both EUVI and COR1 before the jet. This phenomenon was also observed by \\citet{Lites:etal:1999} and other white-light observations in LASCO C2 \\citep{Llebaria:etal:2002}. However, no definite conclusion is given concerning the relation between plume and jet. A time series of 3D reconstruction of both plume and jet needs to be made to find the answer to this question." }, "1112/1112.4229_arXiv.txt": { "abstract": "Individual dark matter halos in cosmological simulations vary widely in their detailed structural properties such as shape, rotation, substructure and degree of internal relaxation. Recent non-parametric (principal component) analyses suggest that a few principal components explain a large fraction of the scatter in halo properties. The main principal component is closely linked with concentration, which in turn is known to be related to the mass accretion history of the halo. Here we examine more generally the connection between mass accretion history and structural parameters. The space of mass accretion histories has principal components of its own. We find that the strongest two can be interpreted as the overall age of the halo and the acceleration or deceleration of growth at late times. These two components only account for $\\sim70$\\%\\ of the scatter in mass accretions histories however, due to the stochastic effect of major mergers. Relating structural parameters to formation history, we find that concentration correlates strongly with the early history of the halo, while relaxation correlates with the late history. We examine the inferences about formation history that can be drawn by splitting haloes into subsamples, based on observable properties such as concentration and shape at some final time. This approach suggests interesting possibilities, such as the possibility of defining young and old samples of galaxy clusters in a rigorous, quantitative way, or testing the dynamical assumptions of galaxy formation models empirically. ", "introduction": "\\label{sec:introduction} Dark matter halos provide the framework for visible structure in the universe over a span of eight decades in mass, from the scale of rich galaxy clusters down to the scale of individual dwarf galaxies. Mass is normally assumed to be the main determinant of a halo's baryonic contents, and analytic models such as the halo occupation distribution \\citep[HOD --][]{Peacock, Seljak, Ma, White, Berlind, Cooray2002-halomodel} make this assumption explicitly. Yet the halos that form in cosmological simulations vary greatly in shape, concentration, spin, substructure and other structural properties. As simulations of increasing size and resolution provide a more and more detailed picture of halo properties, and as observational techniques including weak and/or strong gravitational lensing, X-ray and Sunyaev--Zel'dovich measurements \\citep[e.g.][]{Corless2009-WeaklensingShape,Oguri2010-WeaklensingShape25,Sereno2011-StronglensingShape,Morandi2011-xraylensing} reach a precision where they can determine structural properties such as shape and concentration reliably, it is important to understand how the structural features of a halo are interrelated, and what they can tell us specifically about its formation and evolution. Recently two groups, \\citet{Skibba2011-PCA} and \\citet{JeesonDaniel2011-PCAcorrelations} (S11 and J11 hereafter), have taken the important step of performing non-parametric principal component analyses of halo properties. Principal component analysis (PCA) searches for simplifying trends in a complex data set by finding the axes in a multi-dimensional parameter space that account for the largest fraction of the scatter. In the simplest case, it can uncover linear correlations in the data (e.g. fundamental lines or planes) and reveal hidden patterns or simplifications. The results of S11 and J11 agree on some basic aspects of halo structural properties. Overall, the scatter in halo properties spans a fairly high-dimensional space, with 4 principal components required to explain about 70\\% of the scatter. Nonetheless, a few strong components emerge. In terms of structural parameters, the first (strongest) principal component is best traced by concentration. The strong correlation with concentration suggests that this first principal component is linked to the overall age of the halo, and the strength of the correlation with $z_{0.5}$ confirms this. The origin of halo concentration has been studied extensively since it is a crucially important factor in many calculations, including strong lensing \\citep[e.g.][]{Broadhurst2005-stronglensing,Broadhurst2008-stronglensing2} and dark matter annihilation \\cite[e.g.][]{Taylor2003-darkmatterannihilation}. Several analytic models have been developed over the years to explain concentration in terms of formation history \\citep[e.g.][]{Bullock2001-concentrationMAH, Wechsler2002-MAH, Zhao09}, so the link between formation history and this particular structural property is fairly well understood. Older systems, that is systems that had already assembled most of their mass into one or a few progenitors at early times, are more concentrated on average, although the exact details of the connection between age and concentration vary from one analytic model to another. If the first principal component of halo structural properties is thus linked to age, we might naturally ask what the others correspond to. Are shape, spin or relaxation also related to the formation history, and if so how? To put this question in a quantitative framework, we first have to decide how to describe the ``formation history\" itself -- what should we take this to mean exactly, given the complex set of merger and accretion events through which halos form? We can take the mass accretion history (MAH) as a starting point, defining this as the function ${\\mathcal{M}}(z) \\equiv M(z)/M(0)$ which describes the mass of the main progenitor of a halo as a function of redshift, normalized to the value at $z = 0$ \\citep{vandenBosch2002-MAH}. Since ${\\mathcal{M}}(z)$ is a continuous function of a real variable, it contains an arbitrarily large amount of information about the history of a halo; equivalently, describing the MAH fully means specifying values of ${\\mathcal{M}}(z)$ at an infinite set of redshifts. To characterize the MAH in simpler terms, we can turn once again to principal component analysis, approximating each MAH as a vector of values ${\\mathcal{M}}(z_i)$ corresponding to the MAH evaluated at a finite, fixed set of redshifts $z_i$. PCA of these vectors can then tell us whether mass accretion histories are well described by a set of basis functions characterzed by a single parameter, as suggested by \\citet{vandenBosch2002-MAH} and \\citet{Wechsler2002-MAH}, or whether they require two or more variables to explain their diversity, as suggested by \\citet{Tasitsiomi} or \\citet{McBride2009-MAH}. Assuming a few principal components capture the main features of a halo's MAH, this will allow us to study correlations between structure and history in a well-defined and quantitative way. We note that this is only a first step towards understanding structure in terms of growth history; the MAH does not contain all the information about a halo's past by any means, and an alternate approach is to study the origin of a particular property such as shape or relaxation in detail, considering the physical processes involved \\citep[e.g.][]{VeraCiro2011-ShapeAquarius,Power2011-CorrelationsMAH}. In this paper, we generate halos in a set of cold dark matter (CDM) simulations covering three different mass scales. We record the MAH of each well-resolved halo and decompose the space of MAHs, taken as vectors of values ${\\mathcal{M}}(z_i)$, into principal components. These components provide a non-parametric description of the MAH and clarify its basic properties. We also record the structural properties of each halo at the present-day. We study the correlations between these properties themselves (reproducing the trends found by S11 and J11), and between structural properties and the principal components of the merger history. Finally, we discuss an application of our results, showing how samples of halos selected by concentration or shape will have systematically different formation histories. In this way, the observable properties of groups or clusters of galaxies can be used to infer their (unobservable) formation history. The outline of the paper is as follows. Section \\ref{sec:measuring_halo_properties} describes our n-body simulations and group finding, and how the structural properties of halos are defined and measured. Section \\ref{sec:MAHs} analyses the vector space of MAHs, decomposing them into principal components. In Section \\ref{sec:structural_cor}, we then analyze the correlations between structural components, and derive principal components in this second vector space. In Section \\ref{sec:application} we discuss applications of this work. We summarize our results in Section \\ref{sec:summary}. Throughout the paper, we consider a WMAP7 cosmology with parameters $\\Omega_M = 0.27$, $\\Omega_\\Lambda=0.73$, $\\Omega_b=0$ (in our CDM-only simulations), $H_{0} = 72$ $\\t{km} \\t{s}^{-1} \\t{Mpc}^{-1}$, $n_s = 0.96$, and $\\sigma_8 = 0.80$. ", "conclusions": "" }, "1112/1112.3123_arXiv.txt": { "abstract": "{ Recent LHC data showed excesses of Higgs-like signals at the Higgs mass of around 125\\,GeV. This may indicate supersymmetric models with relatively heavy scalar fermions to enhance the Higgs mass. The desired mass spectrum is realized in the anomaly-mediated supersymmetry breaking model, in which the Wino can naturally be the lightest superparticle (LSP). We discuss possibilities for confirming such a scenario, particularly detecting signals from Wino LSP at direct detection experiments, indirect searches at neutrino telescopes and at the LHC. } \\end{center} \\end{titlepage} \\setcounter{page}{2} Higgs mass contains very important information about low-energy supersymmetry (SUSY) models, which is well motivated because it provides a viable candidate of dark matter (DM) and also because it realizes the gauge coupling unification. In particular, in the minimal SUSY standard model (MSSM), the lightest Higgs boson cannot be heavier than the $Z$-boson at the tree level, while a sizable radiative correction may enhance the Higgs mass ~\\cite{Okada:1990vk}. The size of the radiative correction depends on the masses (and other parameters) of superparticles. The lightest Higgs mass becomes larger as superparticles (in particular, stops) become heavier. Thus, once the lightest Higgs mass is known, mass scale of superparticles is constrained. Recently, the ATLAS collaboration reported $3.6\\sigma$ local excess of the standard model (SM) Higgs-like event at $m_h\\simeq 126\\ {\\rm GeV}$ \\cite{ATLAS-CONF-2011-163}. In addition, the CMS collaboration also showed more than $2\\sigma$ local excess at $m_h\\simeq 124\\ {\\rm GeV}$ \\cite{HIG-11-032}.\\footnote{The excesses based on global probabilities, which take account of the look-elsewhere effect, are $2.3\\sigma$ (ATLAS) and $1.9\\sigma$ (CMS).} In order to achieve such a value of the lightest Higgs mass in the MSSM, relatively large values of the superparticle masses are required; the typical scale of the sfermion masses to realize $m_h\\simeq 125$\\,GeV is $10$\\,TeV--$10^3$\\,TeV~\\cite{hep-ph/0408240,arXiv:1108.6077}. Then, if the masses of all the superparticles are of the same order, it is difficult to find experimental signals of low-energy SUSY and the existence of SUSY is hardly confirmed. Although the sfermion masses are much larger than the electroweak scale, gauginos may be much lighter than sfermions and within the reach of collier and other experiments. One interesting possibility is the model in which the SUSY breaking scalar masses are from direct coupling to the SUSY breaking field while the gaugino masses are generated by the anomaly-mediation mechanism~\\cite{hep-th/9810155,hep-ph/9810442}; in this letter, we call such a model as anomaly-mediated SUSY breaking (AMSB) model. Even in the AMSB model, however, if the pure anomaly-mediation relation holds among the gaugino masses, gluino mass is about 8 times larger than the mass of Wino. Thus, if the Wino mass is a few hundred GeV, which is the lower bound on it from astrophysical and cosmological considerations as will be reviewed later, the gluino mass becomes multi-TeV; with such a heavy gluino, the discovery of the SUSY signal at the LHC becomes challenging because we consider the case that all the squarks are extremely heavy. Even so, there still exist possibilities of discovering signals of the AMSB scenario. In particular, in the present framework, the neutral Wino is the lightest superparticle (LSP) and may be DM. In such a case, pair annihilation cross section of the LSP and the scattering cross section of the LSP off the nuclei are both enhanced compared to the Bino LSP case, which has significant implications to direct and indirect detection of DM. Because the search of the superparticles at the LHC may be difficult, it is important to pursue these possibilities and explore how well we can study the AMSB scenario with these procedures.\\footnote{ The heavy SUSY particle spectrum and their detectability were discussed in a different context in Ref.~\\cite{Hall:2011jd}. } In this letter, motivated by the recent Higgs searches at the LHC, we discuss the detectability of the signals of AMSB scenario. We pay particular attention to the case of the Wino LSP. We focus on direct/indirect detection of the Wino DM at underground laboratories and neutrino telescopes. We also comment on the LHC reach for the direct Wino production. Since superparticles except for gauginos are heavy, standard methods for SUSY searches may not work. Even in this case, we will show that there are some windows for the confirmation of the SUSY. Let us first briefly discuss important properties of the AMSB scenario. We assume that the soft SUSY breaking scalar masses are generated by the direct coupling between the scalars and the SUSY breaking hidden sector field, while the gaugino masses are generated by the anomaly mediation mechanism. Adopting the pure AMSB relation, the gaugino masses are given by \\cite{hep-th/9810155, hep-ph/9810442} \\begin{eqnarray} M_a^{\\rm (AMSB)} = \\frac{b_a}{16\\pi^2} g_a^2 m_{3/2}, \\end{eqnarray} where $g_a$ ($a=1$--$3$) are gauge coupling constants of the SM gauge groups, $m_{3/2}$ is the gravitino mass, and $(b_1,b_2,b_3)=(11,1,-3)$. Then, the Wino becomes the lightest among the gauginos, and gaugino masses largely separate: $m_{\\tilde{B}}: m_{\\tilde{W}}: m_{\\tilde{g}}\\simeq 3:1:8$. Although the AMSB relation may be affected by Higgs and Higgsino loop diagrams \\cite{hep-ph/9810442, Gherghetta:1999sw}, we adopt the pure AMSB mass relation. With the gaugino masses being of $O(100)\\ {\\rm GeV}$--$O(1)\\ {\\rm TeV}$, the gravitino mass becomes of $O(10)\\ {\\rm TeV}$--$O(100)\\ {\\rm TeV}$. The sfermion masses are expected to be of the same order of the gravitino mass, which is preferred from the point of view of realizing $m_h\\simeq 125\\ {\\rm GeV}$. In particular, if the scalar masses are (almost) equal to the gravitino mass, $m_h\\simeq 125\\ {\\rm GeV}$ requires relatively small value of $\\tan\\beta\\sim$ a few (where $\\tan\\beta$ is the ratio of the vacuum expectation values of up- and down-type Higgs bosons) \\cite{arXiv:1108.6077}. Before discussing the detectability of the signals of AMSB model, we comment on the supersymmetric Higgs mass parameter (so-called $\\mu$-parameter). In the present setup, the soft SUSY breaking scalar mass parameters of up- and down-type Higgs bosons are expected to be of $O(10)\\ {\\rm TeV}$--$O(100)\\ {\\rm TeV}$. In order to have viable electroweak symmetry breaking, the $\\mu$-parameter (as well as heavy Higgs boson masses) is also expected to be of the same order; then, the Higgsinos become extremely heavy and the Wino becomes the LSP. Thus, we pay particular attention to the case of Wino LSP in the following. In some of our following analysis, however, we consider the case with $\\mu\\sim O(100)\\ {\\rm GeV}$--$O(1)\\ {\\rm TeV}$ taking account of the possibility of an accidental tuning of the parameters. This is because detection rates of some of signals (in particular, the direct detection rates) strongly depend on the value of $\\mu$. Taking account of the radiative correction due to the gauge boson loops, the neutral Wino becomes lighter than the charged one. Therefore, we focus on the case of neutral Wino LSP. In addition, we assume that the LSP (i.e., the neutral Wino) is the dominant component of DM. The Wino LSP accounts for the present DM density for $m_{\\tilde W} \\simeq 3$\\,TeV if it is produced only from thermal bath~\\cite{hep-ph/0610249}. In the AMSB scenario, however, the Wino LSP can be non-thermally produced from the gravitino or moduli decay \\cite{hep-ph/9810442, hep-ph/9906527}. If the reheating temperature takes an appropriate value, for example, the decay of gravitino produces the Wino LSP with correct relic density \\cite{hep-ph/0012052}, while thermal leptogenesis~\\cite{Fukugita:1986hr} works successfully~\\cite{Ibe:2011aa}. Thus the Wino is a good DM candidate in the present setup. Hereafter, we assume that the right amount of Wino is somehow produced in the early universe to be DM. We start with discussing direct detection experiments of DM. The scattering cross section of the Wino LSP off the nucleon significantly depends on $\\mu$. Since all scalars except for the lightest Higgs boson are expected to be heavy enough, it is only the lightest Higgs boson that mediates the spin-independent (SI) scattering. The DM-proton scattering cross section is given by~\\cite{Jungman:1995df} \\begin{equation} \\sigma = \\frac{4}{\\pi} \\left( \\frac{m_{\\tilde \\chi^0} m_N}{m_{\\tilde \\chi^0} + m_N} \\right)^2 \\left[ \\left( n_p f_p + n_n f_n \\right)^2 + 4 \\frac{J+1}{J} \\left( a_p \\langle s_p \\rangle + a_n \\langle s_n \\rangle \\right)^2 \\right], \\end{equation} where the first and the second term in the bracket are the contributions of SI and spin-dependent (SD) interaction, respectively. Here $m_{\\tilde \\chi_0}$ is the LSP mass, $m_N$ is the mass of the target nucleus, $n_p (n_n)$ is the number of proton (neutron) in the target nucleus, $J$ is the total nuclear spin, $a_p$ and $a_n$ are the effective DM-nucleon SD couplings, and $\\langle s_{p(n)} \\rangle$ are the expectation values of the spin content of the proton and neutron groups within the nucleus. The effective DM-proton coupling, $f_p$, is given by \\begin{equation} f_{p} = \\sum_{q=u, d, s} \\frac{f_q^{H}}{m_q} m_p f_{T_q}^{(p)} + \\frac{2}{27}f_{T_G} \\sum_{q=c, b, t} \\frac{f_q^{H}}{m_q} m_p, \\end{equation} where $f_{T_G}=1-\\sum_{u,d,s}f_{T_q}^{(p)}$, $m_p$ and $m_q$ denote the proton and quark masses, respectively, and $f_q^H$ is the effective DM-quark coupling obtained by the exchange of the Higgs boson. Since the DM-Higgs coupling is proportional to the magnitude of Wino-Higgsino mixing,\\footnote{ In the limit of pure Wino DM, the Wino-nucleon scattering cross section is too small to be detected~\\cite{arXiv:1004.4090}. } the cross section is enhanced if the Wino-Higgsino mixing is large. In Fig.~\\ref{fig:SI} we plot the Wino-proton SI and SD scattering cross section. In this plot we have used following values for the quark contents in the proton~\\cite{Ohki:2008ff} : $f_{T_u}^{(p)}=0.023, f_{T_d}^{(p)}=0.034, f_{T_s}^{(p)}=0.025$ and taken $\\tan \\beta = 3$ and $\\tan\\beta=20$. The XENON100 experiment~\\cite{arXiv:1104.2549} most severely constrains the SI cross section. The sensitivity is improved by a few orders of magnitude for the next generation 1 ton scale detectors, and then broad parameter regions up to $m_{\\tilde W}\\sim \\mu \\sim 1$\\,TeV will be explored. The IceCube searches for neutrino events arising from the DM annihilation in the Sun. Since the efficiency for the DM trapping into the Sun depends on the DM-proton scattering cross section, the high-energy neutrino observations give limits on it. For the SD cross section, the IceCube gives the most stringent limit, and it will be further improved by about one order of magnitude with the DeepCore instrument~\\cite{arXiv:1111.2738}. \\begin{figure}[tbp] \\begin{center} \\includegraphics[scale=0.5]{SI.eps} \\vskip 1cm \\includegraphics[scale=0.5]{SD.eps} \\caption{ Contours of spin-independent (SI) and spin-dependent (SD) Wino-proton scattering cross sections are plotted on the plane of $m_{\\rm wino}$ and $\\mu$. Shaded regions are excluded by the XENON100 experiment for SI, and IceCube experiment for SD. } \\label{fig:SI} \\end{center} \\end{figure} We have also calculated the detection rate at the IceCube DeepCore, arising from high-energy neutrinos produced by the Wino annihilation at the Galactic Center (GC). We distinguish two event classes following Refs.~\\cite{arXiv:0911.5188,arXiv:0912.3521,arXiv:1009.2068} : contained muon events and shower events. The contained muons correspond to those emerge inside the instrumental volume through the high-energy neutrino interactions with nucleons. The shower events are caused by charged current interactions of electron and tau neutrinos, and neutral current interactions of all neutrino species. They leave electromagnetic/hadronic shower inside the instrumental volume. The event rate of the contained muons is given by \\begin{equation} N_{\\mu^+\\mu^-} = \\int dE_{\\nu_\\mu}\\int_{E_{\\rm th}}^{E_{\\nu_\\mu}} dE_\\mu \\left[ \\frac{d\\Phi_{\\nu_\\mu}}{dE_{\\nu_\\mu}} \\left( \\frac{d\\sigma_{\\nu_\\mu p}^{\\rm (CC)}}{dE_\\mu}n_p +\\frac{d\\sigma_{\\nu_\\mu n}^{\\rm (CC)}}{dE_\\mu}n_n \\right) +(\\nu_\\mu \\leftrightarrow \\bar \\nu_\\mu) \\right] V_{\\rm eff}(E_\\mu), \\label{cont} \\end{equation} where $E_{\\nu_\\mu}$ is the incident neutrino energy, $E_\\mu$ is the muon energy resulting from the neutrino-proton (neutron) interactions, $E_{\\rm th}$ is the threshold energy above which the muon can be detected, $d\\Phi_{\\nu_\\mu}/dE_{\\nu_\\mu}$ is the neutrino flux at the Earth, $d\\sigma_{\\nu_\\mu p(n)}^{\\rm (CC)}/dE_\\mu$ denotes the neutrino-proton (neutron) charged current cross section for producing the muon energy with $E_\\mu$, $n_p (n_n)$ is the proton (neutron) number density in the detector material, and $V_{\\rm eff}$ is the effective volume for the muon detection. The incident neutrino flux generated by DM annihilation from the GC within cone half angle of $\\theta$ is given by \\begin{equation} \\frac{d\\Phi_{\\nu_i}}{dE_{\\nu_i}} = \\frac{R_{\\odot}\\rho_{\\odot}^2}{8\\pi m_{\\tilde W}^2} \\left( \\sum_{j=e,\\mu,\\tau} \\langle \\sigma v\\rangle \\frac{dN_{\\nu_j}}{dE_{\\nu_j}}P_{j\\to i} \\right) \\langle J_2 \\rangle_\\Omega \\Delta \\Omega. \\end{equation} Here $R_\\odot = 8.5$\\,kpc and $\\rho_\\odot=0.3\\,{\\rm GeV}\\,{\\rm cm^{-3}}$, $\\langle \\sigma v\\rangle$ is the Wino self-annihilation cross section including the non-perturbative effect~\\cite{hep-ph/0307216}, and $dN_{\\nu_j}/dE_{\\nu_j}$ is the energy spectrum of the neutrino produced by DM annihilation, which is calculated by the PYTHIA package for the $WW$ final state~\\cite{hep-ph/0603175}, $P_{j\\to i}$ is the probability that the $\\nu_j$ at the production is converted to $\\nu_i$ because of the neutrino oscillation effect, $\\Delta\\Omega=2\\pi(1-\\cos\\theta)$, and $\\langle J_2 \\rangle_\\Omega$ includes the information about the DM density profile in the Galaxy~\\cite{arXiv:0812.0219}. The shower event is evaluated in a similar way to the contained muon events (\\ref{cont}), except that the charged current interactions from $\\nu_e$ and $\\nu_\\tau$ as well as the neutral current interactions for all neutrino flavors are included. The background event is evaluated by inserting the atmospheric neutrino flux into the expression (\\ref{cont}). Fig.~\\ref{fig:ICDC} shows the signal-to-noise ratio at the IceCube DeepCore as a function of the Wino mass. Sensitivities for contained muon events (upper panel) and shower events (lower panel) with 1 year and 10 year observations are shown. We have adopted the NFW density profile and considered the neutrino flux from the cone half angle $\\theta = 10^{\\circ}$ and $\\theta = 25^{\\circ}$ around the GC. As noted in Ref.~\\cite{arXiv:1009.2068}, the sensitivity is maximized for $\\theta \\simeq 10^{\\circ}$. For this cone half angle, the flux dependence on the DM density profile is not large~\\cite{arXiv:0812.0219}. The effective volume for the contained and shower events are set to be $0.04\\,{\\rm km^3}$ and $0.02\\,{\\rm km^3}$, respectively~\\cite{arXiv:0911.5188}. The atmospheric background is taken from Ref.~\\cite{astro-ph/0611418}. It is seen that the signal-to-noise ratio is at most order one for the Wino mass of a few hundred GeV. We have also checked that the upward muon events expected at the KM3NeT detector~\\cite{848444}, assuming the effective area of $1\\,{\\rm km^2}$ and taking account of the energy loss of muons~\\cite{hep-ph/0012350}, provide similar sensitivities to the DeepCore. \\begin{figure}[tbp] \\begin{center} \\includegraphics[scale=1.7]{ICDC_cont.eps} \\vskip 1cm \\includegraphics[scale=1.7]{ICDC_show.eps} \\caption{ Signal-to-noise ratio at the IceCube DeepCore as a function of the Wino mass. Sensitivities for contained muon events (upper panel) and shower events (lower panel) with 1 year and 10 year observations are shown. We have considered the neutrino flux from the cone half angle $\\theta = 10^{\\circ}$ and $\\theta = 25^{\\circ}$ around the GC. } \\label{fig:ICDC} \\end{center} \\end{figure} The Wino DM annihilation may leave characteristic signatures on astrophysical observations. Gamma-ray observations by Fermi-LAT and HESS severely restrict the DM annihilation cross section (see, e.g., Refs.~\\cite{arXiv:1108.3546,arXiv:1110.6151} for recent works). The non-observations of DM-induced gamma-rays from dwarf galaxies excludes the Wino mass below $\\sim 400$\\,GeV~\\cite{arXiv:1108.3546}, although there are astrophysical uncertainties. On the other hand, the cosmic-ray positron excess observed by PAMELA satellite~\\cite{arXiv:0810.4995} may be explained by the Wino DM annihilation with mass of 200\\,GeV~\\cite{arXiv:0810.1892,arXiv:0812.4555} although it may confront the constraints from gamma-rays and anti-protons. The observations of light element abundances also give stringent bound on the DM annihilation cross section so as not to destroy light elements during Big-Bang Nucleosynthesis (BBN). It gives a lower bound on the Wino mass as $m_{\\tilde W} \\gtrsim 200$\\,GeV~\\cite{astro-ph/0405583,arXiv:0810.1892}. It may be encouraging that the cosmic lithium problem may be solved for the Wino mass of around this bound, which simultaneously may explain the PAMELA anomaly. DM annihilation also affects the recombination history of the Universe, which results in the modification on the cosmic microwave background (CMB) anisotropy~\\cite{arXiv:0905.0003,arXiv:0906.1197,arXiv:0907.3985}. The constraint is comparable to that from BBN. Taking these constraints into account, we conservatively consider that the Wino must be heavier than $\\sim 200$\\,GeV if it is the dominant component of DM. Finally, we comment on a possibility of discovering a signal of AMSB model at the LHC. If we adopt the AMSB mass relation among gauginos, gluino becomes relatively heavy. Then, colored superparticles are hardly produced at the LHC. Thus, we focus on the detection of a Wino signal. If the neutral Wino $\\tilde{W}^0$ is the LSP, we have a chance to observe the track of charged Wino $\\tilde{W}^\\pm$ \\cite{Feng:1999fu, hep-ph/0610277}. This is because the mass difference between charged and neutral Winos is so small ($\\sim 160\\ {\\rm MeV}$) that the decay length of $\\tilde{W}^\\pm$ becomes macroscopic ($c\\tau_{\\tilde{W}^\\pm}\\simeq 5\\ {\\rm cm}$). Some of the produced charged Winos may travel through several layers of inner trackers and their track may be reconstructed. In the ATLAS experiment, for example, the charged Wino track can be reconstructed with almost 100 \\% efficiency if $\\tilde{W}^\\pm$ hits the 3rd layer of the semiconducter tracker (SCT) before it decays \\cite{arXiv:0807.4987}. Then, because of the smallness of $c\\tau_{\\tilde{W}^\\pm}$ compared to the detector size, $\\tilde{W}^\\pm$ decays before going through the whole detector. Such a charged Wino is identified as a high $p_T$ track which disappears in the middle of the detector. Such a signal does not exist in the SM, and hence is a smoking gun evidence of the production of $\\tilde{W}^\\pm$. The Wino pair can be produced by the Drell-Yan process at the LHC. However, there is no high $p_T$ jet nor track in the final state in such an event, and hence the event cannot be recorded. In order to trigger on the Wino production events, one can use the event with high $p_T$ jet; such a jet can be from the initial state radiation. Then, at the parton level, the Wino production processes relevant for the present study are the following: \\begin{eqnarray*} && q\\bar{q} \\rightarrow \\tilde{W}^+ \\tilde{W}^- g,~~~ gq \\rightarrow \\tilde{W}^+ \\tilde{W}^- q,~~~ g\\bar{q} \\rightarrow \\tilde{W}^+ \\tilde{W}^- \\bar{q}, \\\\ && q\\bar{q}' \\rightarrow \\tilde{W}^\\pm \\tilde{W}^0 g,~~~ gq \\rightarrow \\tilde{W}^\\pm \\tilde{W}^0 q',~~~ g\\bar{q} \\rightarrow \\tilde{W}^\\pm \\tilde{W}^0 \\bar{q}'. \\end{eqnarray*} We calculate the cross section of the process $pp\\rightarrow\\tilde{W}\\tilde{W}j$; we perform the parton level calculation, and we approximate the $p_T$ of jet by that of final-state quark or gluon. In the calculation of the cross section, the helicity amplitude package HELAS \\cite{Murayama:1992gi} and the CT10 parton distribution functions \\cite{Lai:2010vv} are used. For the phase space integration, we use the BASES package \\cite{BASES}. In the calculation, we require that the transverse momentum of the jet be larger than 170, 270, and 370 GeV, and that at least one charged Wino travels more than 44.3 cm which is the distance to the 3rd layer of SCT from the beam pipe in the ATLAS detector \\cite{Aad:2008zzm}. In Fig.\\ \\ref{fig:cs}, the cross section is plotted as a function of the Wino mass. In the high luminosity run with ${\\cal L}=2\\times 10^{33}\\ {\\rm cm^{-2}s^{-1}}$, for example, the so-called j370 trigger is planned to be available, which requires a jet with $p_T>370\\ {\\rm GeV}$ \\cite{Aad:2008zzm}. Then, requiring 10 events with $p_T>370\\ {\\rm GeV}$ for the discovery, for example, Wino mass smaller than $270\\ {\\rm GeV}$ ($330\\ {\\rm GeV}$) is covered by the LHC with the luminosity of ${\\cal L}=100\\ {\\rm fb}^{-1}$ ($300\\ {\\rm fb}^{-1}$), where we have assumed that the background is negligible. Thus, in particular when $\\mu$ is large, the LHC experiment still have a chance to cover the parameter region which has not been excluded yet by the current direct and indirect DM searches. If the $p_T$ of the jet for the trigger can be reduced, the LHC can cover the region with larger Wino mass. So far, we have assumed the pure AMSB relation among gaugino masses. However, as we have mentioned, such a relation may be largely affected by the Higgs and Higgsino loop diagrams. With such an effect, the gluino mass may become $\\sim 1\\ {\\rm TeV}$ even when the Wino mass is a few GeV. In such a case, the conventional procedures of the SUSY search using the missing energy distribution may work. \\begin{figure}[tbp] \\begin{center} \\includegraphics[scale=1.7]{csCTEQ10.eps} \\caption{ Cross section for the process $pp\\rightarrow\\tilde{W}\\tilde{W}j$ (with $j=q$ or $g$), for $\\sqrt{s}=14\\ {\\rm TeV}$. The transverse momentum of $j$ is required to be larger than 170, 270, and 370 GeV from above. } \\label{fig:cs} \\end{center} \\end{figure} In summary, motivated by the recent report on the Higgs searches at the LHC, which indicated excesses of Higgs-like events at around $m_h\\simeq 125\\ {\\rm GeV}$, we have investigated prospects for confirmation of the AMSB scenario, particularly the detection of Wino LSP. We have considered the situation that the scalars except for gauginos and Higgsinos are heavy enough so that they cannot be produced at colliders. Even in this unfortunate case, the Wino DM may be detected through direct/indirect detection experiments. Direct detection efficiency crucially depends on the Higgsino mass, and if the Wino and Higgsino masses happen to be close, future experiments may find their signals. The neutrino telescopes such as IceCube DeepCore and KM3NeT also have a potential to discover the Wino LSP through the observation muon and/or shower events induced by high-energy neutrinos from DM annihilation at GC. ", "introduction": " ", "conclusions": "" }, "1112/1112.3409_arXiv.txt": { "abstract": "We consider the acceleration of very small dust grains including Polycyclic Aromatic Hydrocarbons (PAHs) arising from the electrostatic interactions of dust grains that have charge fluctuations in time due to charging events. We simulate the charge fluctuations of very small grains due to their sticking collisions with electrons and ions in plasma, and the emission of photoelectrons by UV photons using Monte Carlo method. We identify the acceleration induced by the charge fluctuations as the dominant acceleration mechanism of very small grains in the diffuse interstellar medium (ISM). We show that this acceleration mechanism is efficient for environments with low degree of ionization (i.e., large Debye length), where the charge fluctuations are slow but have a large amplitude. We also discuss the implications of the present mechanism for grain coagulation and shattering in the diffuse ISM, molecular clouds, and protoplanetary disks. ", "introduction": "Dust is an important constituent of the interstellar medium (ISM), molecular clouds, and accretion disks (see Whittet 2003; Draine 2009, 2011). It gets involved in many key processes, for instance, it controls heating and cooling of the ISM (see Draine 2003; Tielens 2005), reveals magnetic fields through grain alignment (see Lazarian 2007 for a review), and interferes with attempts to measure the temperature anisotropy and polarization of the cosmic microwave background (CMB) radiation (see Lazarian \\& Finkbeiner 2003; Fraisse et al. 2009; Dunkley et al. 2009). Very small dust grains of size $a\\le 100\\Angstrom$ with a notable fraction of polycyclic aromatic hydrocarbons (PAHs, hereafter very small grains and PAHs are used interchangeably) radiate electric dipole emission, which is an important component of CMB foregrounds (Draine \\& Lazarian 1998; Hoang et al. 2010, 2011). PAHs in protoplanetary disks (PPDs) around young intermediate mass (Herbig Ae/Be) stars (Acke \\& van den Ancker 2004) and in outer layers of low mass (T-Tauris) stars (Geers et al. 2006; Oliveira et al. 2010) may affect the magnetorotational instability (MRI) due to their influence on the environment ionization (Bai 2011; Perez-Becker \\& Chiang 2011). Most properties of dust, including light extinction, electron photoemission, and chemical activity depend not only on grain chemical composition, but also on their sizes. In astrophysical environments, grain size is affected by grain-grain collisions, which depends on their relative velocities. The minimal velocity of grains corresponds to their thermal velocity at the ambient gas temperature arising from Brownian motions. These velocities are usually assumed in the models of dust coagulation (Ossenkopf 1993; Dominik \\& Dullemond 2008). It has been long known that large scale hydrodynamic motions associated with turbulence can make grains move faster (see Draine 1985), but detailed calculations applicable to astrophysical environments started to appear only a few years ago. Since most astrophysical media are magnetized and grains are charged, the hydrodynamic treatment of acceleration is not adequate. A proper treatment of the grain acceleration through the resonant interactions of charged grains with magnetohydrodynamic (MHD) turbulence has been developed relatively recently (Lazarian \\& Yan 2002; Yan \\& Lazarian 2003; Yan et al. 2004; Yan 2009; Hoang et al. 2011b). This treatment makes an extensive use of the advances in the theory/numerical studies of compressible MHD turbulence (see Cho \\& Lazarian 2003; Kowal \\& Lazarian 2010 and ref. therein) and provides the mathematical formalism of the second-order Fermi acceleration of charged grains interacting with MHD turbulence. The acceleration due to the resonant interactions of MHD turbulence with charged grains decreases with decreasing grain size. Such decrease arises from the fact that the Larmor radius of charged grains becomes smaller as the grain mass decreases, and correspondingly, grains have to interact with smaller and less powerful turbulent fluctuations. In addition, compressible fluctuations (i.e., fast modes), which were identified in Yan \\& Lazarian (2003) as the most efficient source of acceleration, get suppressed at small scales due to plasma viscous damping, while the Alfv\\'{e}nic mode gets inefficient for acceleration at the small scales due to anisotropy (see Yan \\& Lazarian 2003 for more discussion). The resonant acceleration for grains with sizes $\\leq 10^{-5}$~cm becomes rather inefficient in most media discussed in Yan \\& Lazarian (2003) and Yan et al. (2004, hereafter YLD04). These conclusions were also confirmed by Hoang et al. (2011b) who considered the non-linear effects of resonant acceleration (Yan \\& Lazarian 2008) and the transit time damping acceleration (TTD) of charged grains. A new mechanism of astrophysical grain acceleration was sketched in Ivlev et al. (2010), where rough estimates of grain velocities arising from charge fluctuations in the ISM were provided. The study employed Fokker-Planck equation and assumed that (1) charge fluctuations are fast, i.e., the relaxation timescale for charge fluctuations is much shorter than the gaseous damping time and the Larmor precession period, and (2) charge fluctuations have small amplitude (i.e., $\\delta Q \\ll Q_{0}$) such that these fluctuations can be approximated as being continuous and modeled by a Gaussian distribution. These assumptions are usually not realistic for very small grains for which the charge fluctuations are substantial. Therefore, the unrealistically high velocities of very small grains were obtained. This motivates our present study which is intended to determining the actual velocities of PAHs in typical astrophysical conditions. In this paper, we aim to quantify the acceleration of very small dust grains induced by grain charge fluctuations using numerical simulations. The structure of the paper is as follows. We first discuss major processes responsible for grain charging and charge fluctuations in \\S 2. In \\S 3 we present an algorithm to simulate the charge fluctuations using Monte Carlo method. In \\S 4 we present a numerical simulation method to investigate grain acceleration, including dynamical equations and Monte-Carlo simulations. Velocities of very small grains in the ISM from numerical simulations are presented in \\S 5. Discussions and summary are presented in \\S 6 and 7, respectively. ", "conclusions": "\\subsection{Grain Acceleration from Various Mechanisms} Grain acceleration by hydrodrag (see Draine 1985) was discussed as the dominant mechanism driving grain motions in the ISM as well as in protoplanetary disks. Recently, thanks to significant progresses in understanding of MHD turbulence, Lazarian \\& Yan (2003), YLD04, and Yan (2009) identified the gyro-resonant interactions of grains with fast modes in MHD turbulence as the dominant mechanisms to accelerate large grains ($a>10^{-5}$ cm) to super-Alfv'{e}nic velocities. For very small grains (e.g., PAHs and nanoparticles), the gyro-resonance acceleration is inefficient for most phases of the ISM, because PAHs and nanoparticles with smaller gyro radii only resonantly interact with small scale and weak fluctuations. Hoang et al. (2011b) revisited the treatment of gyroresonance acceleration by MHD turbulence by accounting for the fluctuations of the grain guiding center with respect to the mean magnetic field. They improved estimates of the previous authors, and proposed a new way of acceleration through transit time damping (TTD) of fast modes, which is efficient for super-Alfv\\'{e}nic grains (see solid lines in Fig. \\ref{fig_TTD}). \\subsection{Acceleration induced by Charge Fluctuations} In the present paper, we have numerically investigated the grain acceleration mechanism induced by charge fluctuations for very small grains. The combination of MC simulations for charge fluctuations with direct simulations of electrostatic interactions between grains with fluctuating charge allows us to follow the evolution of grain velocities in time and estimate their terminal velocities. We showed that the acceleration induced by charge fluctuations is indeed an important mechanism to accelerate PAHs. Resulting grain velocities can be several times higher than their thermal velocities, which is obviously important. \\begin{figure} [h] \\includegraphics[width=0.49\\textwidth]{f9.ps} \\caption{Grain velocity as a function of grain size $a$ due to charge fluctuations (shaded area) and due to resonance acceleration by MHD turbulence (Hoang et al. 2011b) for the CNM, WNM, and WIM. MHD turbulence is the efficient mechanism for large grains, and grain charge fluctuations work efficiently at the small end of grain size. Dotted and solid lines correspond to the acceleration from gyroresonance and gyroresonance plus transit time damping.} \\label{fig_TTD} \\end{figure} We also found that grains velocities obtained from MC simulations are between two and three orders of magnitude lower than the prediction by the FP equation. Such overestimate of the FP approach arises because the charge fluctuations of very small grains in the ISM are slow, in contrast to the assumption of fast charge fluctuations assumed in the FP treatment, which makes electric fields fluctuating rather slowly. In addition, the FP treatment disregarded the screening effect of plasma. In the ISM, the mean separation between two grains is comparable to the Debye length. Thus, the Coulomb interaction is suppressed considerably, and the grain acceleration is decreased accordingly. Figure \\ref{fig_TTD} compares grain velocities induced by charge fluctuations (shaded area) with those due the resonance interactions in MHD turbulence. As shown, MHD turbulence is efficient for large grains, while charge fluctuations are important for very small grains. \\subsection{Grain coagulation and shattering} The grain coagulation is a process in which small dust particles collide with each other to form aggregates of submicron size (Spitzer 1978). The grain coagulation can occur in dense clouds (e.g., Ossenkopf 1993; Ormel et al. 2009) as well as in protoplanetary disks (see van Boekel et al. 2003). Grain coagulation and shattering result from grain-grain collisions, which depends on grain relative velocity, denoted by $v_{dd}$. The threshold velocity for the grain shattering is a function of the grain size: \\bea v_{\\rm shat}=2.7\\left(\\frac{a}{10^{-7}~\\cm}\\right)^{-5/6} {\\rm km}~\\s^{-1},\\label{v_cri} \\ena (Chokshi et al. 1993). If $v_{dd}< v_{\\rm shat}$, the grains collide and stick together. When $v_{dd} > v_{\\rm shat}$, the collisions with high velocity produce shock waves inside the grain, and shatter them in smaller fragments. For $v_{dd}\\rightarrow 20$ km$\\s^{-1}$, the evaporation of dust grains occurs and grains are destroyed. We note that due to resonant acceleration by MHD turbulence, large grains $a>10^{-5}$ cm move frequently with velocity larger than $v_{\\rm shat}$ (see Fig. \\ref{fig_TTD}). Thus, the shattering of large grains can be efficient to replenish very small grains into the ISM. The effect of grain acceleration by MHD turbulence on grain coagulation and shattering was studied in Hirashita \\& Yan (2009). In some phases, they found that grain acceleration can modify grain size distribution. At the same time, very small grains move faster than thermal speed due to the acceleration discussed in the present paper and enhance grain coagulation. Therefore, future models of grain evolution should take into account the acceleration of very small grains. \\subsection{Effects of PAH acceleration in protoplanetary disks} Grain coagulation is the first step to planetesimal formation in protoplanetary disks (PPDs). It is also believed that the coagulation from very small grains to micron size occurs rapidly (see e.g., van Boekel et al. 2003), so that PAHs and nanoparticles are rapidly depleted in PPDs. However, recent observations by Infrared Space Observatory (ISO) and {\\it Spitzer} satellite reveal PAH emission features, which indicates the existence of PAHs in young intermediate mass (Herbig Ae/Be) stars (Acke \\& van den Ancker 2004) and in the surface layers exposed to UV radiation in young low mass (T-Tauri) stars (Geers et al. 2006; Oliveira et al. 2010; Bern\\'{e} et al. 2009). PPDs have highly stratified structure, that means, there exists some gradient of grain size from the midplane to the surface. When PAH emission is seen, it is probably only from PAHs in the surface layers that \"see\" direct stellar radiation. The abundance of PAHs and other nanoparticles below the surface is rather uncertain. {\\it (a) Effects of acceleration on dust coagulation and planet formation} Current models of grain growth in PPDs disregarded grain charges and their fluctuations. Recently, Okuzumi (2009) took into account the effects of grain charge and showed that Coulomb potential barrier can suppress grain coagulation. Okuzumi (2009) appealed to turbulence as an energy source to overcome the potential barrier. Below, we show that charge fluctuations can induce very small grains to overcome the Coulomb barrier assuming that PPDs have a sufficiently high degree of ionization for which the charge fluctuations are important. Assuming that the charge fluctuations are Gaussian, then, the mean charge and charge dispersion are respectively given by \\bea \\langle Z\\rangle=z\\frac{ak_{\\B}T_{\\gas}}{e^{2}}\\simeq-0.1\\left(\\frac{a}{10^{-6}\\cm}\\right) \\left(\\frac{T_{\\gas}}{100~\\K}\\right),~~~\\label{zmean} \\ena where $z=2.5$ is adopted (see Morfill \\& Ivlev 2009). The dust mass density in PPDs is related to the gas density as \\bea \\rho_{d}=f_{dg}\\rho_{g},\\label{rhod} \\ena where $f_{dg}=0.014$ for the PPD (see Tanaka et al. 2005). The dust density of the grain size $a$ is then \\bea n_{d}=10^{1}\\left(\\frac{a}{10^{-6}~\\cm}\\right)^{-3} \\left(\\frac{n_{\\gas}}{10^{10}~\\cm^{-3}}\\right).\\label{nd_ppd} \\ena \\\\ Using above equations for Equation (\\ref{Tterm}), we obtain \\bea \\frac{T_{d}^{\\infty}}{T_{\\gas}}=9.31\\times 10^{17}\\epsilon^{2} \\left(\\frac{T_{\\gas}}{100~\\K}\\right)^{2} \\left(\\frac{a}{10^{-6}~\\cm}\\right)^{-4}\\nonumber\\\\ \\times x_{i} \\left(\\frac{0.1~\\cm^{-3}}{n_{i}}\\right)^{2}\\left(\\frac{n_{\\gas}}{10^{10}~\\cm^{-3}}\\right),~~~~~\\label{vd_ppd} \\ena where $\\epsilon<1$ is introduced to account for the overestimate of grain velocities obtained using the FP approach, $n_{i}$ is the ion density, $x_{i}=n_{i}/n_{\\gas}$ is the ionization fraction of gas, and PPDs are assumed to be isothermal plasma with $T_{\\gas}=T_{i}$. Now, let us consider whether the charge fluctuations can help grain grow against the Coulomb potential barrier. The ratio of the grain kinetic energy arising from charge fluctuations to electrostatic potential at a distance equal to $2a$ with $a$ being the grain size is \\bea \\frac{E_{\\kin}}{E_{\\rm Coul}}=\\frac{3a k_{\\B}T_{d}^{\\infty}}{\\langle Z\\rangle^{2}e^{2}}.\\label{ratio} \\ena From Equations (\\ref{zmean}), (\\ref{vd_ppd}), and (\\ref{ratio}) we obtain \\bea \\frac{E_{\\kin}}{E_{\\rm Coul}}\\simeq 2.3\\times10^{1}\\epsilon^{2}\\left(\\frac{a}{10^{-6}~\\cm}\\right)^{-5} \\left(\\frac{T_{\\gas}}{100~ \\K}\\right)\\nonumber\\\\ \\times\\left(\\frac{10^{10}~\\cm^{-3}}{n_{\\gas}}\\right)\\left(\\frac{10^{-4}}{x_{i}}\\right),~~~\\label{ekin_coul} \\ena where the ionization fraction $x_{i}\\ne 0$ was assumed to be sufficiently high so that the charge fluctuations are important. Grain growth occurs only if $E_{\\kin}/E_{\\rm Coul}\\ge1$. Thus, from Equation (\\ref{ekin_coul}) we can derive the range of grain size with the grain growth. Using parameters for PPDs conditions, $T_{\\gas}=300~\\K$, $n_{\\gas}=10^{10}~\\cm^{-3}$ and $x_{i}=10^{-4}$ and taking $\\epsilon=10^{-1}$, we found that very small grains with size $a\\leq 6\\times 10^{-7}~\\cm$. Therefore, the growth of larger grains can be induced by other processes (e.g., turbulence as suggested in Okuzumi 2009). The presence of PAHs and nanoparticles in PPDs can arise from the fragmentation and shattering of micron graphite and silicate grains (see e.g., Dullemond \\& Dominik 2008). The level of MHD turbulence in PPDs is very uncertain, but it is believed that the layers near the surface are highly turbulent. As a result, large grains can be accelerated by the resonant interactions of fast modes to super-Alfv\\'{e}nic speed (Yan et al. 2004; Hoang et al. 2011b). Thus, the grain shattering may be efficient that return very small grains to PPDs. {\\it (b) Effects of PAH acceleration on magnetorotational instability} PAHs play an important role in the ionization equilibrium in PPDs. Bai (2011) found that PAHs can favor the accretion due to magneto-rotational instability (MRI) by decreasing ambipolar diffusions in outer layers of PPDs. In fact, if charged PAHs are more abundant than electron ($n_{\\rm PAH}> n_{e}$), then the ambipolar diffusion is less important. However, since the electrical conductivity by charged PAHs is more efficient than electron, the MRI becomes more efficient. Suprathermal velocities of PAHs due to charge fluctuations in PPDs can enhance the accretion rate of electron and ion onto grains, which certainly affect the ambipolar diffusion\\footnote{The ambipolar diffusion is usually invoked to describe the removal of magnetic field during star formation. Another powerful mechanism is reconnection diffusion based on the ability of turbulent magnetic field to reconnect and redistribute the entrained matter (Lazarian 2005; see also Lazarian 2011).} and MRI. Thus, understanding grain motions in PPDs is important for understanding the dynamics of PPDs and planetesimal formation." }, "1112/1112.1898_arXiv.txt": { "abstract": "Compact binary coalescences, such as binary neutron stars or black holes, are among the most promising candidate sources for the current and future terrestrial gravitational-wave detectors. While such sources are best searched using matched template techniques and chirp template banks, integrating chirp signals from binaries over the entire Universe also leads to a gravitational-wave background (GWB). In this paper we systematically scan the parameter space for the binary coalescence GWB models, taking into account uncertainties in the star formation rate and in the delay time between the formation and coalescence of the binary, and we compare the computed GWB to the sensitivities of the second and third generation gravitational-wave detector networks. We find that second generation detectors are likely to detect the binary coalescence GWB, while the third generation detectors will probe most of the available parameter space. The binary coalescence GWB will, in fact, be a foreground for the third-generation detectors, potentially masking the GWB background due to cosmological sources. Accessing the cosmological GWB with third generation detectors will therefore require identification and subtraction of all inspiral signals from all binaries in the detectors' frequency band. ", "introduction": "The ground-based gravitational-wave detectors are rapidly increasing their sensitivities. The first generation detectors LIGO \\cite{LIGOinstr,S5detector} and Virgo \\cite{Virgo1,Virgo2} have reached their design sensitivities and collected excellent data over several years of exposure. The second generation detectors, Advanced LIGO \\cite{aLIGO, aLIGO2}, Advanced Virgo \\cite{aVirgo}, GEO-HF \\cite{GEOHF}, and LCGT \\cite{CLIO,LCGT} are currently being built and commissioned. With 10 times better strain sensitivity, these detectors are expected to yield first direct detections of gravitational-wave signals, and their first data is expected as early as 2014. Furthermore, there are already efforts under way to design the third-generation gravitational wave detectors, with another factor of 10 improvement in sensitivity. This includes the Einstein Telescope project \\cite{ET,ET2}, for which the design study was recently completed. These detectors are expected to open a new era in astronomy and astrophysics, providing new observations of various events and objects in the Universe, complementary to the standard electromagnetic observations. Among the many sources of gravitational waves, the coalescences of binary systems, such as binary neutron stars (BNS), binary black holes (BBH), or a black hole and a neutron star (BHNS) stand out as the most likely candidates for first detections. These systems generate well understood \"chirp\" gravitational-wave signals, which have been computed using post-Newtonian approximation \\cite{phenom_waveform} or numerical relativity simulations \\cite{numrel}. One can then search for the chirp signals using matched template techniques - indeed a number of such searches have been performed using LIGO and Virgo data \\cite{CBC1,CBC2,CBC3}. It has also been argued that adding the gravitational-wave signals from all binaries in the Universe will produce a gravitational-wave background (GWB) - for example, see \\cite{regman,regfrei,regchauv,regreview,ros11,mar11,BBHstoch} for the most recent studies in the context of terrestrial detectors. The LIGO and Virgo collaborations have developed techniques for searching for GWB by cross-correlating data from pairs of gravitational wave detectors \\cite{allenromano}. Such searches have also been performed using LIGO and Virgo data \\cite{S3stoch,S4stoch,S5stoch}, and have produced competitive upper limits on the energy density carried by gravitational waves. The goal of this paper is to perform a detailed study of the accessibility of the GWB produced by the binary coalescences to the second and third generation gravitational-wave detectors. Our study follows the work of Regimbau and Mandic \\cite{regman}, and includes detailed scans of the parameter space in these models, as well as possible effects due to the uncertainty in the star formation rate and in time-delays associated with the formation of the binaries. We will show that this background is likely to be observed by the network of second generation detectors, and that the third-generation detectors will likely be able to explore most of the parameter space for these models. In Section 2 we summarize the calculation of the energy density for these models. In Section 3 we present results of our systematic study, and we include concluding remarks in Section 4. ", "conclusions": "In this paper we computed the gravitational wave background due to coalescences of binary neutron stars, binary black holes, and black hole - neutron star binaries, following the approach of \\cite{regman}. While such computations have been done in the past, in this study we performed a systematic scan of the parameter space, taking into account the possible variations in the result due to the choice of the star formation rate, and due to the choice of the distribution $P(t_d)$ of the delay time between the formation and coalescence of the binary. For each point in the parameter space, we compare the model prediction to the expected sensitivities of the second and third generation gravitational-wave detector networks. We find that models corresponding to the optimistic and realistic (in the case of BNS and BHNS models) local coalescence rates will be accessible to the second generation detector network. We also find that models corresponding to the pessimistic local coalescence rates will be accessible to the third-generation detector network. The binary coalescence GWB will, in fact, be a foreground for the third-generation detectors, and it will mask the GWB background due to early-Universe sources. Accessing the cosmological GWB with third generation detectors will therefore require identification and subtraction of all inspiral signals from all binaries in the relevant frequency band. {\\it Acknowledgments:} The authors thank C. Belczinski, M. Dominik, and T. Bulik for informative discussions about the probability distribution for delay time, $P(t_d)$. CW and VM were supported in part by NSF grant PHY0758036." }, "1112/1112.4473_arXiv.txt": { "abstract": "We present the discovery of a long-period, rapidly oscillating Ap star, \\hd. Using high-resolution time-series observations obtained with UVES at the ESO VLT telescope, we found radial velocity variations with amplitudes 7--150~\\ms\\ and a period of 23.6~min, exceeding that of any previously known roAp star. The largest pulsation amplitudes are observed for Eu\\iii, Ce\\iii\\ and for the narrow core of H$\\alpha$. We derived the atmospheric parameters and chemical composition of HD\\,177765, showing this star to be similar to other long-period roAp stars. Comparison with theoretical pulsational models \\mybf{indicates} an advanced evolutionary state for HD\\,177765. \\mybf{Abundance analyses of this and other roAp stars suggest} a systematic variation with age of the rare-earth line anomalies seen in cool Ap stars. ", "introduction": "\\label{intro} The rapidly oscillating (roAp) stars are magnetic, chemically peculiar stars which pulsate in high-overtone acoustic modes with typical periods of $\\approx$\\,10~min. These stars are located close to the instability strip crossing the main sequence between the early F and late A spectral types. First roAp pulsators were discovered by \\citet{1982MNRAS.200..807K}. Currently, about 40 of such stars are known. Several excitation mechanisms were suggested in the past to drive pulsations in roAp stars \\citep{1988IAUS..123..291D,1984trss.conf..346D,1983ApJ...275L...5S,1988MNRAS.235P...7M}. Currently, it is widely accepted that the high frequency oscillations observed in these stars are excited by the opacity mechanism working on the hydrogen ionization region. However, full non-adiabatic calculations show that the excitation of high frequency acoustic oscillations by this mechanism can only be achieved in non-standard models, such as models with a modified T-tau relation \\citep{1998MNRAS.301...31G}, or models with envelope convection partially or fully suppressed \\citep{2001MNRAS.323..362B,2005MNRAS.360.1022S}.Among these, models with convection suppressed seem to reproduce better the observed instability strip, however, the predicted red edge remains significantly hotter than the observed one. The majority of roAp stars have been identified and analysed using high-speed photometric methods \\citep{2000BaltA...9..253K}. However, it has been realised that the ground-based photometry is not particularly suitable for discovering roAp stars. Time-resolved spectroscopy has a major advantage for detection of smaller-amplitude and longer-period pulsations. This has been demonstrated by, for example, the discovery of very low-amplitude pulsations in HD\\,75445 \\citep{2009A&A...493L..45K} and by the detection of long-period oscillations in HD\\,137909 \\citep[\\bCrB,][]{2004MNRAS.351..663H} and HD\\,116114 \\citep{2005MNRAS.358..665E}, which have been repeatedly classified as non-pulsating by photometric observations \\citep{1994MNRAS.271..129M,2005IBVS.5651....1L}. The two latter objects, along with a faint roAp star KIC\\,10195926 found by the Kepler satellite \\citep{2011MNRAS.414.2550K}, form an unusual sub-group with pulsation periods of 16--21~min and spectroscopic properties different from the ``classical'', shorter-period roAp stars. Here we present the spectroscopic discovery of another member of this class, HD\\,177765, whose pulsation period of 24~min is the longest known for any roAp star. ", "conclusions": "\\label{concl} We have analysed time-resolved spectra of the cool Ap star \\hd\\ obtained with the UVES instrument at VLT. The radial velocity and frequency analysis of these data reveals this object to be a roAp star with the pulsation period of 23.6~min. These oscillations, clearly present with amplitudes 40--150~\\ms\\ in the combined radial velocity curves of Ce\\ii, Eu\\ii\\ and Gd\\ii\\ lines, as well in the \\ha\\ core, occur with the longest known pulsational period for any roAp star. This discovery makes \\hd\\ a key object for testing predictions of pulsation theories because the frequency limits help in distinguishing alternative driving mechanisms and can provide useful asteroseismic constraints on the atmospheric and interior stellar structure. No trigonometric parallax measurement is availble for \\hd. Therefore, we can only approximately place it in the HR-diagram for a comparison with pulsational models. Using the effective temperature, derived from photometry and spectroscopy, the pulsation frequency $\\nu=0.7$ mHz, and the evolutionary tracks from \\citet{2002MNRAS.333...47C}, which provide the frequency of the most unstable mode for models with given effective temperature and luminosity, we obtain a mass $M$\\,$\\approx$\\,2.2$M_\\odot$ and a luminosity $\\log L/L_\\odot\\approx$\\,1.5. These parameters imply that the star is significantly evolved from the zero age main sequence, making it similar to the three other long-period, evolved roAp stars: \\bCrB, HD\\,116114, and KIC\\,10195926. We have carried out an abundance analysis of \\hd\\ using the equivalent width and spectrum synthesis methods. A slow variation of the mean field modulus suggests a long rotational period. The diversity of Zeeman splitting patterns and the dependence of the abundance on the line strength and excitation potential (most clearly seen for Fe) indicates strong horizontal and vertical inhomogeneities. We found that \\hd\\ is chemically similar to \\bCrB. Both stars are distinguished by a high Ce abundance and discordant abundances inferred from different ions of \\mybf{Eu and Ce}. But both stars lack a strong PrNd ionisation anomaly which is characteristic of higher frequency roAp stars. This anomaly is also absent in HD\\,116114. Thus, instead of being a spectroscopic signature of roAp stars as suggested by \\citet{2004A&A...423..705R}, the PrNd anomaly is probably related to the evolutionary state of roAp stars, distinguishing evolved, long-period pulsators from the shorter-period stars closer to the zero age main sequence. This finding represents one of the most compelling evidence for a systematic variation of surface chemical composition of Ap stars with age. Longer time-series observations of \\hd\\ are highly desirable to determine additional pulsation frequencies which are necessary for an astroseismic analysis. Observations during a single night might be sufficient to resolve the expected large frequency separation of $\\Delta\\nu_0\\sim$\\,50~$\\mu$Hz." }, "1112/1112.6247.txt": { "abstract": "We construct a simple model of the star-formation- (and resultant supernova-) driven mass and energy flows through the inner $\\sim$200 pc (in diameter) of the Galaxy. % Our modelling is constrained, in particular, by the non-thermal radio continuum and $\\gamma$-ray signals detected from the region. % The modelling points to a current star-formation rate of $0.04-0.12 \\msun$/year at $2\\sigma$ confidence within the region with best-fit value in the range $0.08-0.12 \\msun$/year which -- if sustained over 10 Gyr -- would fill out the $\\sim 10^9 \\msun$ stellar population of the nuclear bulge. % Mass is being accreted on to the Galactic centre (GC) region at a rate $\\dot{M}_{IN} \\sim 0.3 \\msun/$year. % The region's star-formation activity drives an outflow of plasma, cosmic rays, and entrained, cooler gas. % Neither the plasma nor the entrained gas reaches the gravitational escape speed, however, and all this material fountains back on to the inner Galaxy. % The system we model can naturally account for the recently-observed $\\gtrsim 10^6 \\msun$ `halo' of molecular gas surrounding the Central Molecular Zone out to 100-200 pc heights. % The injection of cooler, high-metallicity material into the Galactic halo above the GC may catalyse the subsequent cooling and condensation of hot plasma out of this region and explain the presence of relatively pristine, nuclear-unprocessed gas in the GC. % This process may also be an important ingredient in understanding the long-term stability of the GC star-formation rate. % The plasma outflow from the GC reaches a height of a few kpc and is compellingly related to the recently-discovered Fermi Bubbles by a number of pieces of evidence. % These include that the outflow advects precisely i) the power in cosmic rays required to sustain the Bubbles' $\\gamma$-ray luminosity in saturation; ii) the hot gas required to compensate for gas cooling and drop-out from the Bubbles; and iii) the magnetic field required to stabilise the walls of these structures. % Our modelling demonstrates that $\\sim 10^9 \\msun$ of hot gas is processed through the GC over 10 Gyr. % We speculate that the continual star-formation in the GC over the age of the Milky Way has kept the SMBH in a quiescent state thus preventing it from significantly heating the coronal gas, allowing for the continual accretion of gas on to the disk and the sustenance of star formation on much wider scales in the Galaxy. % In general, our investigations explicitly reveal the GC's important role in the Milky Way's wider stellar ecology. ", "introduction": "The inner 200 pc (in diameter) of the Milky Way features a spectacular confluence of unusual and energetic astrophysical phenomena. % Within this region of the Galaxy -- circumscribed by the Inner Lindblad Resonance associated with the non-axisymmetric gravitational potential of the Galactic bar -- the distribution of stars cusps sharply into the distinct population of the so-called nuclear bulge \\citep{Serabyn1996}. % Correspondingly, over the same region the current, inferred areal density of star formation, $\\dot{\\Sigma}_*$, sharply peaks to $\\sim 200 \\msun/$kpc$^2$/yr. % This is approximately three orders of magnitude higher than the mean value in the Galactic disk. % With such a high $\\dot{\\Sigma}_*$, observations of the nuclei of external, star-forming galaxies tell us to expect a star-formation-driven outflow; there is much empirical and theoretical evidence that such an outflow exists in the GC as we have explored in a number of recent papers \\citep{Crocker2010b,Crocker2011a,Crocker2011b}. % This paper adds significantly to that evidence. The very high star formation rate (SFR) density likewise sustains a very high energy density in all phases of the GC ISM. % Most directly, the optical and UV output of the many young, hot stars in the region is reprocessed by thick, ambient dust into a dominantly infrared photon background of $\\sim$100 eV cm$^{-3}$. % Radio continuum and $\\gamma$-ray observations allow one to place a robust lower limit of $\\sim$50 $\\mu$G on the typical magnetic field throughout the entire inner $\\sim$800 pc (in diameter) of the Galaxy \\citep{Crocker2010a}; modelling \\citep{Crocker2010b,Crocker2011b} points to a magnetic field in the inner 200 pc that is at least 100 $\\mu$G. % In association with and, in fact, as a necessary precondition to, the high SFR, observations reveal an enormous agglomeration of hot, dense, and highly-turbulent molecular gas of mass $3 \\times 10^7 \\msun$ \\citep{Dahmen1998,Molinari2011}, 5-10\\% of the Milky Way's entire $H_2$ allocation. % This gas forms an asymmetric distribution extended along the plane referred to as the Central Molecular Zone \\citep[CMZ:][]{Serabyn1996}. % Recent infrared observations of the CMZ by $Herschel$ \\citep{Molinari2011} place much of this $H_2$ on a $\\sim 100 $ pc-radius ring, apparently akin to the star-forming rings observed in the nuclei of many face-on galaxies. The CMZ $H_2$ is constantly bombarded by an extended, hard-spectrum cosmic ray ion population which results in a diffuse glow of hard-spectrum, $\\sim$TeV $\\gamma$-rays coextensive with the gas \\citep{Aharonian2006}. % GC X-ray observations \\citep{Koyama1989} {\\it apparently} reveal the existence of a very hot ($\\sim$ 7 keV), extended thermal plasma which would have an energy density similar to the lightfield, turbulent molecular gas, and magnetic field \\citep{Spergel1992}. % It must also be remarked that the GC hosts the Milky Way's resident supermassive ($\\sim4.3 \\times 10^6 \\msun$: \\citealt{Gillessen2009}) black hole (SMBH). % Though currently in a state of apparently unusual quiescence, this must certainly have been much more active at various times in the past \\citep[e.g.][]{Ponti2010}. % On the other hand, we have found from our recent work that the mechanical power delivered by supernovae -- occurring at a rate consistent with that pointed to by the region's current star-formation as traced by FIR emission -- is completely sufficient to sustain the currently-observed non-thermal emission ($\\sim$GHz radio continuum and $\\sim$TeV $\\gamma$-ray) from the $\\sim$200 pc scales of interest here. % Thus, from the point of view proffered by the non-thermal data, it is not {\\it necessary} that the SMBH have any significant role beyond the inner few pcs; our investigations below confirm this in general. Finally, one of the most interesting recent discoveries in high energy astrophysics is of the `Fermi Bubbles', so-called because these structures were revealed \\citep{Su2010} in $\\sim$ GeV $\\gamma$-ray data collected by the $Fermi$-LAT. % The Bubbles are north-south symmetric about the Galactic plane and centred on the Galactic nucleus. % Given this morphology they are compellingly associated with some sort of activity in the GC. % Given, then, the Bubbles' large angular extent (they rise to $\\pm 50^\\circ$ in $b$) they are enormous structures extending 10 kpc from the plane. % The Bubbles' $\\gamma$-ray emission might be due to inverse Compton emission from a rather young (given short energy loss time) population of cosmic ray electrons. % Alternatively, the emission might be due to hadronic collisions experienced by a hard-spectrum cosmic ray proton and heavier ion population \\citep{Crocker2011a,Crocker2011c,Zubovas2011}. % We set out the evidence connecting the Fermi Bubbles with multi-Gyr-scale GC-star-formation-driven injection of cosmic ray {\\it protons} into the Galactic halo below. \\subsection{Motivating Questions} The preceding tour of GC and inner Galaxy phenomenology helps motivate a number of questions which we seek to address in this paper: \\begin{enumerate} \\item{\\bf Gas accretion:} What is the rate at which the GC region typically accretes gas through the Galactic plane? % How do we understand the presence \\citep{Lubowich2000,Riquelme2010} of relatively pristine gas (i.e., that has undergone relatively little nuclear processing) in the GC? \\item {\\bf Star formation:} What is the efficiency with which the GC converts gas into stars? % Given the unusual conditions in the GC environment, is GC star-formation biased towards the production of massive stars? % \\item {\\bf GC ISM conditions:} Is the very hot plasma putatively revealed by X-ray observations real or not? % What contribution do the non-thermal ISM phases, in particular the cosmic rays, make to the overall energy density in the region? % What is the dynamical importance, if any, of the cosmic rays? \\item {\\bf Outflows:} There is multi-wavelength evidence (reviewed below) for outflows from the GC over size scales from pcs to 10 kpcs. % Are these outflows all different aspects of the same overarching phenomenon and how are they driven? % What is the wider importance of the GC outflow(s) to the Galactic ecology? % Is the material expelled from the nucleus lost to extra-galactic space or does it fountain back on to the Galactic disk? % How do the recently-discovered Fermi Bubbles \\citep{Su2010} relate to activity in the GC? \\end{enumerate} More broadly, we aim in this paper to produce a first draft of a coherent explanation of all the disparate phenomena listed above that is itself physically plausible and motivated. % Overall, we shall see that it is star-formation (driven by secular accretion processes over long timescales) -- rather than processes associated directly with the SMBH -- that seems to control the overall dynamics of the GC. % \\subsection{Conventions and Assumptions} We assume a distance to the Galactic centre of 8 kpc in this work. % We use {\\tiny MATHEMATICA} notation: $f[x]$ denotes $f$ in a function of parameter $x$. % Formally, the region we are investigating and attempting to model is that centred on $(l,b) = (0,0)$ and extending to $\\pm 0.8^\\circ$ in Galactic longitude and $\\pm 0.3^\\circ$ in Galactic latitude; this is the region for which the HESS telescope reported \\citep{Aharonian2006} a diffuse flux of $\\sim$TeV $\\gamma$-rays. ", "conclusions": "\\label{sctn_Conclusions} We have modelled the mass and energy flows through the Galactic centre in a one-zone model and shown that star-formation -- and resultant supernovae -- in the GC are well-sufficient to drive the gross dynamics of the system and to explain its non-thermal phenomenology. % None of this is to deny the importance of, e.g., stellar-radiation driving of gas dynamics in particular regions of the GC, but radiation pressure is not required to explain the large-scale mass movements we infer here. Current evidence does not seem strong enough for us to promote the empirically-inspired scaling of supernova mechanical energy with zero-age, main sequence progenitor mass that we have tested above; if anything, the `standard' $E_{SN} = 10^{51}$ erg assumption seems to work better in a number of cases (e.g., in suggesting a mass loading factor closer to expectation for star-burst-like environments; a SFR closer to other, independent estimates; supplying the total stellar mass of the nuclear stellar bulge for less extreme values of $\\dot{M}_{IN}$). % In any case, as emphasised above, the $E_{SN}$ scaling we adopted certainly constitutes an upper limit to the true, population-averaged evolution of $E_{SN}$ with stellar mass. % Likely, further modelling, probably involving other constraints, will be needed before a better handle on $E_{SN}[M_{ZAMS}]$ can be arrived at. % Given these considerations, we now particularise our discussion to the case of $E_{SN} = 10^{51}$ erg. A number of indicators come together to suggest that our control parameter, $\\dot{M}_{IN}$ -- the total mass being fed into the system -- has a lower limit at around 0.4 $\\msun$/year.% We find a 2$\\sigma$ upper limit on $\\dot{M}_{IN}$ (for the $E_{SN} = 10^{51}$ case) at 1.8 $\\msun$/year. % We emphasise, however, that $\\dot{M}_{IN}$ is an overestimate of the total mass flux accreting out of the plane of the Galaxy on to the GC. % Our modelling suggests that there is an outflow of plasma and cosmic rays from the system that entrains cool gas. % This entrained mass constitutes most of the mass flux but will fountain back on the GC. % This naturally accounts for the recently-observed halo of warm molecular hydrogen found to be surrounding the CMZ and may self-catalyse the accretion of relatively pristine corona gas into the system. We find that our modelling robustly predicts an almost invariant $10^{39}$ erg/s for the power going into the freshly accelerated cosmic ray proton population in the GC region. % We nowhere constrain our model to produce this result: it emerges from the numerical minimisation procedure. % As we have previously emphasised \\citep{Crocker2011a,Crocker2011b} this power is {\\it precisely} enough to sustain the $\\gamma$-ray emission from the Fermi Bubbles in a hadronic saturation scenario and allow for the inflation of the Bubbles against the pressure of the external medium in a few Gyr (probably assisted by the injected magnetic field and injected plasma). % Equally, we find that the modelling robustly predicts that the GC system injects $\\sim 10^{38}$ erg/s into hard-spectrum cosmic ray electrons; this is sufficient \\citep{Crocker2010a,Crocker2011b} to explain the non-thermal synchrotron radiation detected from the GC lobe and wider diffuse, non-thermal source region detected around the GC \\citep{LaRosa2005,Crocker2010a}. Together with the evidence that the GC SFR has been quasi-stationary for Gyr timescales and that the outflow from the GC advects most non-thermal particles out of the acceleration region before they lose much energy, we consider our finding that the GC accelerates $10^{39}$ erg/s in cosmic rays as a very strong indication that GC star-formation essentially explains the Fermi Bubbles. % This argument is strengthened by the facts that the same star-formation processes can inject the plasma mass and thermal power required to fill-out the Bubbles and sustain their X-ray emission {\\it and} to inject the magnetic fields that can stabilise the Bubble surfaces against fluid instabilities, trap their cosmic ray and plasma contents for long timescales and explain their microwave synchrotron emission. \\vspace{1 cm} The Galactic centre is not particularly distinguished in the night sky -- but this belies its true activity: the many orders of magnitude of visual extinction arising from the column of dust we view it through is the reason for this $extrinsic$ dimness. % Equally, the GC is not particularly impressive as a non-thermal radiation source; its $\\sim$ GeV $\\gamma$-ray luminosity, at few $\\times 10^{36}$ erg/s, is an order of magnitude short of the 5-10\\% of the Galaxy's output one might guess on the basis of the amount of massive star-formation happening in the system \\citep[the Galactic $\\sim$ GeV $\\gamma$-ray luminosity is about $3 \\times 10^{38}$ erg/s as inferred from fig.~1 of][]{Strong2010}. % As we have emphasised previously \\citep{Crocker2010b,Crocker2011a,Crocker2011b} {\\it this is because the cosmic rays accelerated in the region are mostly leaving before they can radiate}. % The radiation that these particles finally do emit is writ large in the Fermi Bubbles: it is from these structures that we detect the $\\sim$ 10\\% of Galactic $\\gamma$-ray luminosity (i.e., $4 \\times 10^{37}$ erg/s) that we expect on the basis of the GC's share of Galactic star-formation. The fact that we can detect the Fermi Bubbles at all is testament to the long-term stability of the GC as star-forming system. % We have seen hints above as to how this stability can be established, in particular, how GC star-formation activity can be insulated from the vicissitudes of conditions in the Galactic disk: it seems that a minimal level of accretion in the GC system is self-catalysed. % This is consistent with the presence of relatively pristine gas accreted out of the halo via a mechanism or mechanisms directly related to the star-formation-driven outflow (i.e., injection of cool, high metallicity gas and/or dust and/or shocks into the halo plasma). We find complete consistency between the long-timescale-averaged power required to drive the GC system and the power injected by the star-formation we can infer is currently taking place in the system. % We need not invoke periods of very bright AGN-type activity of the supermassive black hole to explain the dynamics of the GC. % In fact, as a final speculation: these studies hint that the importance of sustained, GC star-formation is that it effectively erects a curtain wall around the SMBH, either using-up gas directly by creating new stars or blowing it away before much of it can reach Sgr A$^*$. % This prevents mass accretion at rates that would allow the SMBH to undergo phases of activity sufficient to heat the Galaxy's coronal gas to such high temperatures that further accretion on the Galactic disk would become impossible\\footnote{We emphasise that this certainly does not imply there is {\\it no} accretion on to the SMBH or that it must be in a state of absolute quiescence.}. % This, in turn, enables the long-term sustenance of {\\it disk} star formation \\citep[cf.][]{Binney2011}. % Such a mechanism would explain the emerging finding \\citep{Erwin2011} -- to which the Milky Way adheres -- that the mass of a nuclear star cluster correlates with the {\\it total} stellar mass of its host galaxy rather than the galaxy's bulge." }, "1112/1112.5955.txt": { "abstract": "We report on a Suzaku measurement of the shock feature associated with the western radio relic in the merging cluster A~3376\\@. The temperature profile is characterized by an almost flat radial shape with $kT \\sim 4$ keV within $0.5 r_{200}$ and a rise by about 1 keV inside the radio relic. Across the relic region ($0.6-0.8 r_{200}$), the temperature shows a remarkable drop from about 4.7 keV to 1.3 keV\\@. This is a clear evidence that the radio relic really corresponds to a shock front possibly caused by a past major merger. The observed sharp changes of the temperature and electron density indicate the Mach number ${\\cal M} \\sim 3$. The radial entropy profile is flatter than the prediction ($r^{1.1}$) of numerical simulations within $0.5 r_{200}$, and becomes steeper around the relic region. These observed features and time-scale estimation consistently imply that the ICM around the radio relic has experienced a merger shock and is in the middle of the process of dynamical and thermal relaxation. ", "introduction": "\\label{sec:intro} %%\\textcolor{red}{xx\u0082\u00cd\u0081AA~3667 paper\u0082\u00c9\u0082\u00c8\u0082\u00e8\u0082\u00dc\u0082\u00b7. \u0089\u00fc\u0092\u00f9\u0081A\u0092\u00c7\u0089\u00c1\u0082\u00b5\u0082\u00bd\u0095\u00b6\u008f\u00cd\u0082\u00cd\u0081A\u0090\u00d4 %% \u008e\u009a\u0082\u00c5\u008e\u00a6\u0082\u00b5\u0082\u00c4\u0082\u00a2\u0082\u00dc\u0082\u00b7} \\textcolor{blue}{ \u008d\u00c4\u0089\u00fc\u0092\u00f9\u0082\u00b5\u0082\u00bd\u0095\u00aa\u0082\u00cd\u0081A\u0090\u00c2\u008e\u009a\u0082\u00c9\u0082\u00c8\u0082\u00e8\u0082\u00dc\u0082\u00b7\u0081B} %% \\textcolor{green}{ \u008d\u00c4\u008d\u00c4\u0089\u00fc\u0092\u00f9\u0082\u00b5\u0082\u00bd\u0095\u00aa\u0082\u00cd\u0081A\u0090\u00c2\u008e\u009a\u0082\u00c9\u0082\u00c8\u0082\u00e8\u0082\u00dc\u0082\u00b7\u0081B} Clusters of galaxies are believed to grow through gas accretion from large-scale filaments and mergers of subclusters. Cluster mergers are the most energetic events in the Universe after the Big Bang with the total kinetic energy of the colliding subclusters reaching $10^{65}$ ergs. The kinetic energy is converted to thermal energy by driving shocks and turbulence. Existence of mega-parsec scale radio emission (halos and relics) in merging clusters indicates that those shocks and turbulence operate as main mechanisms of particle acceleration (see \\cite{ferrari08} and reference therein). Radio relics are considered to be the synchrotron emission generated through the interaction between relativistic electrons, accelerated by a shock, and amplified magnetic field in the intracluster medium (ICM)\\@. Although these radio features provide direct evidence of particle acceleration, the detection of non-thermal (hard) X-ray emission caused by relativistic particles is still controversial \\citep{nevalainen04,ajello09}. Currently, hard X-ray observations set lower limits of magnetic fields in ICM~\\citep{wik09,sugawara09, clarke06}. Based on recent studies, it is believed that magnetic field of 0.1--10 $\\mu$G exists in ICM\\@. %To understand cluster formation, gas heatings and particle accelerations, %cluster mergers are very important fundamental process. Theoretical studies of cluster merger shocks provide rich information about the magnetic field, turbulence, particle acceleration and their time evolution (see e.g.\\ \\cite{brueggen11} for a review). Merger shocks are characterized by a Mach number ${\\cal M} \\lesssim 3$, causing amplification of magnetic fields to $B \\sim 10\\ \\mu$G and production of relativistic particles \\citep{takizawa00,takizawa08}. Also, turbulence is considered to give a significant pressure support in the outer regions, with around 10\\% of the thermal pressure at least for a few Gyr after a major merger \\citep{parrish11}. Recent numerical simulations include interaction of merger shocks with the filamentary cosmic web structures outside of clusters, and the results indicate that shock fronts are more enhanced in the filament direction \\citep{paul11}. Though past X-ray observations have shown images and temperature structures of cluster mergers, there are only few clusters for which clear evidence of the shocks has been obtained (1E0657-56, A520, and A2146: \\cite{markevitch05, clowe06,russell10}). The low X-ray surface brightness in the outer regions has hampered precise measurements of gas temperature and density in the pre-shock region in particular. Recently, existence of shock fronts with a Mach number of about 2 was confirmed in the radio relic region of A~3667 \\citep{finoguenov10,akamatsu11b}. Additionally, a remarkable radio relic indicated a shock front with a Mach number of 4.6 based on the measurement of radio spectral index (CIZA2242: \\cite{weeren10}). Thus, radio relics are the good probe to identify merger shocks and combination with X-ray data will yield important information about the gas dynamics associated with cluster mergers. Since the radio relics are mainly found in the outskirts of clusters where the gas density is low, sensitive X-ray observations are needed, especially in the upstream side of the shocks. The X-ray data will allow us to estimate the parameters of the gas and the magnetic field, and enable us to look into the actual process occurring during the cluster evolution. In this paper, we present a new X-ray evidence of the shock associated with the radio relic in Abell 3376, a well-known merging cluster with irregular morphology at $z=0.046$. Previous studies showed the global mean temperature to be 4.0 keV and the existence of a pair of Mpc-scale radio relics revealed by 1.4 GHz VLA NVSS observations~\\citep{bagchi06}. Another feature of A~3376 is the report of the hard X-ray signal with BeppoSAX/PDS~\\citep{nevalainen04}. %However, because of the Suzaku/HXD could't detect those signal\\citep{kawano09}, the detection of hard X-ray signal were still controversial. However, Suzaku HXD observation gave an upper limit which did not exclude the BeppoSAX flux \\citep{kawano09}. We need to await further sensitive observations in the hard X-ray band. We use $H_0 = 70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\rm M}=0.27$ and $\\Omega_\\Lambda = 0.73$, respectively, which give 54 kpc per arcminute at $z = 0.046$. The virial radius is approximated by $r_{200} = 2.77 h_{70}^{-1} (\\langle T\\rangle /10 {\\rm keV})^{1/2}/E(z)\\ {\\rm Mpc} $, where $E(z)=(\\Omega_{\\rm M}(1+z)^{3}+1-\\Omega_{\\rm M})^{1/2}$ \\citep{henry09}. For our cosmology and redshift, $r_{200}$ is 1.86 Mpc ($= \\timeform{34.6'}$) with $kT = 4.7$ keV\\@. We employ solar abundance defined by \\citet{anders89} and Galactic absorption with $N_{\\rm H}=5.8 \\times10^{20}$cm$^{-2}$~\\citep{dickey90}. Unless otherwise stated, the errors correspond to 90\\% confidence for a single parameter. \\begin{table*}[t] \\label{tab:obslog} \\begin{center} \\caption{Suzaku observations of A~3376 and the background region.} \\begin{tabular}{cccccccccc} \\hline Target (Obs.\\ ID) & Position &Start time &Exposure$^\\ast$ &Exposure$^\\dagger$ \\\\ & (R.A., Decl.) & (UT) & (ks) & (ks) \\\\ \\hline A~3376c (100034010) & $(90.56, -39.94)$ & 2005/10/06 14:46:08 & 110.4 & 78.0 \\\\ A~3376w (800011010) & $(90.05, -39.98)$ & 2005/11/07 14:15:05 & 120.0 & 99.9 \\\\ Q~0551-3637 (703036020) & $(88.19, -36.63)$ & 2008/05/14 13:25:55 & 18.8 & 15.3\\\\ \\hline $\\ast$: no selection with COR2 &\\multicolumn{2}{l}{ ${\\dagger}$: COR2 $> 8$ GV}\\\\\\ \\end{tabular} \\end{center} \\end{table*} \\begin{figure}[t] \\begin{center} \\includegraphics[scale=0.45]{./a3376_0.5_8.0.ps} \\end{center} \\caption{ X-ray image of A~3376 in the energy band 0.5-8.0 keV, after subtraction of the NXB with no vignetting correction and after smoothing by a 2-dimensional Gaussian with $\\sigma =16$ pixel =$\\timeform{17''}$. The large blue circle shows the virial radius of A~3376 with $r=34'.6$, and white dotted circles show the annular regions used for the spectral analysis. Small magenta circles show point sources which are detected by {\\it wavdetect} (see text). The VLA 1.4 GHz radio image is shown with green contours. } \\label{fig:suzaku_image} \\end{figure} ", "conclusions": "\\label{sec:discussion} Suzaku performed 2 pointing observations of Abell 3376 along its merger axis. The ICM temperature was found to be fairly flat at about 4 keV with a slow rise with radius in the region from the center to $0.6 r_{200}$~(1 Mpc). This is followed by a sharp drop in the radio relic region, which implies there is a shock front. We will attempt to evaluate the gas properties related with the shock (temperature, Mach number, magnetic field, and entropy) and discuss their implications. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %Temperature profiles %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \\subsection{Temperature Profiles}\\label{sec:kt_comp} %\\textcolor{red}{ %In most of the relaxed clusters, the ICM temperature % structures reflect the gravitational % potential and can be used to constrain the dark matter distribution. %} The ICM temperature is an important observational parameter which reflects the depth of the gravitational potential of the cluster. Their profiles also enable us to look into the history of the ICM heating and the growth of clusters. Previous Chandra, XMM-Newton results for other clusters indicated a significant temperature decline toward the outer regions. Recent studies with Suzaku showed temperature profiles of ICM to the virial radius ($r_{200}$) for several clusters \\citep{george08, reiprich09, bautz09, hoshino10, kawaharada10, simionescu11}. These profiles for relaxed clusters are approximated by a ``universal'' profile which declines by a factor of $\\sim 3$ at $r_{200}$ from the center. On the other hand, merging clusters show complex temperature distribution depending on the mass ratio and collision geometry. In the outer regions toward $r_{200}$, radial temperature profiles show some excesses and sharp drops which are not seen in relaxed systems (\\cite{akamatsu11a,akamatsu11b}). However, very little systematic studies have been made about the ICM properties in the outer regions of merging clusters. We normalized the radial temperature profile by the average temperature determined within $0.4 r_{200}$. Since A~3376 does not show a strong cool core, this should give a reasonable average temperature. %As mentioned above section, the $r_{200}$ derived from \\citet{henry09}. The resulting scaled temperature profile is shown in~Fig.\\ \\ref{fig:comp} along with the previous Suzaku results (black crosses and diamonds: taken from \\cite{akamatsu11a}). Clearly, there is a marked difference between the profiles of relaxed and merging clusters. The relaxed clusters (black crosses) generally show a smooth decline from the center to the outer regions, consistent with the result of hydrodynamic simulations~\\citep{burns10}. Compared with these profiles, the A~3376 profile shows a large excess within $0.7 r_{200}$ before the sharp drop at $0.7-0.8 r_{200}$. Numerical simulation of merging clusters by \\citet{paul11}shows a shock feature very similar to the one in A~3376. This simulation after about 3 Gyr after a major merger shows an arc-like high temperature region whose width is about 200 kpc, very similar to the observed temperature enhancement in Fig.\\ \\ref{fig:comp}. This suggests that A~3376 is still evolving after several Gyrs elapsed from the last major merger. The $\\Lambda$CDM model predicts that clusters grow through subcluster mergers and matter accretion from large-scale structures. Those dynamical events should give strong impacts on the ICM temperature structure, such as seen in A~3376. Since the ICM density is very low in the cluster outskirts ($n_e = 10^{-5}-10^{-3}\\rm\\ cm ^{-3}$), the radiative cooling and conduction times are expected to be very long with the cooling time largely exceeding the age of the Universe. We can naturally expect that the temperature structure carries important information about the history of the cluster growth. However, as mentioned above, there are no clear signs of dynamical evolution in relaxed clusters. This indicates that the observed peculiar temperature structure in the merging clusters will be settled to the universal profile in a fairly short time scale, much shorter than the Hubble time. Since cooling time is too long, we have to consider other possibilities such as (i) ICM cooling by adiabatic expansion, and (ii) ICM diffusion caused by pressure gradient. We will discuss these time scales in section~\\ref{sec:timescale}. A fraction of the merger energy would be channeled into the acceleration of ultra-relativistic particles and amplification of magnetic fields. However, it is unknown that how these processes actually take place and distribute the merger energy to other forms in an efficient way. The turbulent motion can be detected through line broadening by X-ray microcalorimeters such as with SXS instrument on ASTRO-H \\citep{mituda10} and Athena, since X-ray calorimeters provide superior energy resolution by a factor of 20--30 better than that with the CCD instruments. The wealth of information will open a new window in our understanding of heating and particle acceleration caused by cluster merger shocks. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%Mach number!! %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \\subsection{Mach Number} \\label{sec:shock} Recent studies with XMM-Newton and Suzaku showed sharp temperature drops at the region of the radio relic in A~3667~\\citep{finoguenov10, akamatsu11b}. \\citet{finoguenov10} reported a sharp edge in the surface brightness at the outer boundary of the NW radio relic in A~3667\\@. \\citet{akamatsu11b} reported a significant jump in the ICM parameters (temperature, surface brightness, electron density, and pressure) across the same relic. The derived Mach number is ${\\cal M} \\sim 2$. We estimate the Mach number based on the Suzaku data in the same way as \\citet{akamatsu11b}. The Mach number can be obtained by applying the Rankine-Hugoniot jump condition, \\begin{equation} \\frac{T_2}{T_1}=\\frac{5{\\cal M}^4+14{\\cal M}^2-3}{16{\\cal M}^2}, ~~ \\frac{1}{C}=\\frac{3}{4{\\cal M}^2}+\\frac{1}{4}, \\end{equation} where $C=\\frac{n_{e1}}{n_{e2}}$ is the shock compression and subscripts 1 and 2 denote pre-shock and post-shock values, respectively, assuming the ratio of specific heats as $\\gamma=5/3$. Here, we assume the regions \\timeform{21'}-\\timeform{24'} and \\timeform{27'}-\\timeform{31'} to be the post and pre shock regions, respectively. Because the compression factor, $C$, exceeds the value for the strong shock limit, we only derived mach number from temperature jump. Table~\\ref{tab:mach} shows the resultant Mach number, whose values are ${\\cal M}=2.94\\pm0.77$ based on the jumps of temperature. Compared with other clusters which also show shock fronts ~(A520, 1E0657-558, A2246, A~3667: ~\\cite{markevitch05,clowe06, russell10, akamatsu11b}), the Mach number in A~3376 is slightly larger than the other cases. We derive the compression factor, $C = 7.8 ~\\pm~4.7$, which almost exceeds the value for the strong shock limit, even though with a large error. The electron density estimated here is based on the assumption of spherical symmetry, which is not strictly correct when the shock propagates along the merger axis. Since the spherical assumption relates the observed flux to a shell-like volume with a larger line-of-sight depth than the case of a relic volume, the density in particular for the pre-shock region is underestimated by a factor of roughly 1.5. In addition to this uncertainty, we checked the dependence on the centroid position of the spherical shock. We set the center of the sphere at the secondary X-ray peak which is closer to the radio relic. The resultant compression parameter is $C=5.01$. This indicates that the determination of the centroid have a significant impact on the shock parameter, which is mainly resulted from the estimated line of sight depth of the shocked region. The shock front in A~3376 is located far from the cluster center~(1.2 Mpc $\\sim~0.6~r_{200}$) just as in A~3667\\@. The temperature in the upstream region of the shock is $\\sim 1\\rm ~keV$, indicating the pre-shock sound speed to be $v_{\\rm ss}\\sim 520$ km~s$^{-1}$. Combining this with the shock compression $C$, we can evaluate the shock speed by $v_{\\rm shock}=C\\cdot v_{\\rm ss}$ to be $>2080$ km s$^{-1}$ assuming $C=4.0$ from the above discussion. The estimated shock speed well agrees with those in other clusters (1E0657-558: 4500 km s$^{-1}$, A520: 2300 km s$^{-1}$), but twice higher than the A~3667 case ($1360\\pm 120$ km s$^{-1}$). The shock speed is large enough to account for the radio relic by a merger shock model \\citep{takizawa00,ricker01, mathis05}. On the other hand, this $v_{\\rm shock}$ seems too high considering the ICM temperature of 4 keV\\@. We may need to consider a possibility that the true gas temperature is much higher than the measured electron temperature. \\begin{figure*}[t] \\begin{tabular}{ccc} \\begin{minipage}{0.33333\\hsize} (a) Normalized temperature profiles \\begin{center} \\includegraphics[scale=0.32,angle=-90]{./comp-latest.ps} \\end{center} \\end{minipage} \\begin{minipage}{0.33333\\hsize} (b) Entropy profiles \\begin{center} \\includegraphics[scale=0.32,angle=-90]{./entropy.ps} \\end{center} \\end{minipage} \\begin{minipage}{0.33333\\hsize} (c) Equilibrium time scale \\begin{center} \\includegraphics[scale=0.32,angle=-90]{./eqtime.ps} \\end{center} \\end{minipage} \\end{tabular} \\caption{\\label{fig:comp} Radial profiles of normalized temperature, entropy and equilibration time scale. (a) Scaled projected temperature profiles compared with relaxed clusters~\\citep{akamatsu11a}. The profiles have been normalized to the mean (within 0.4 $r_{200}$) temperature. The $r_{200}$ value is derived from \\citet{henry09}. Dotted line shows simulation result~\\citep{burns10}. Two gray dashed lines show standard deviation. Black crosses are for the relaxed clusters. Diamonds show merger clusters (black:~A2142, blue:~A3667, red:~A3376). In entropy profile (b), black solid line shows $K \\propto r^{1.1}$ given by~\\cite{voit03}. (c) Ion-electron equilibration time scale. Horizontal scale is normalized by the radius of the radio relic. Solid curve shows the time after a shock heating assuming a constant shock speed $v=2000$ km s$^{-1}$. Red diamonds show the electron-ion equilibration time $t_{ie}$. Gray dashed diamond show the ionization time for $n_{e}t=3\\times10^{12} \\rm cm^{-3}s$. } \\end{figure*} \\begin{table}[t] \\begin{center} \\caption{Mach number estimation for A~3376 assuming $\\gamma=5/3$.} \\label{tab:mach} \\begin{tabular}{cccc}\\hline \\hline Parameter & Pre shock & Post shock & Mach number\\\\ \\hline $kT$ [keV] & $1.34\\pm 0.69$ & $4.68 \\pm 0.48$ & $2.97\\pm 0.77$ \\\\ \\hline \\end{tabular} \\end{center} \\end{table} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %Entropy profiles %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \\subsection{Entropy Profile} \\label{sec:entropy} Theoretical studies of the ICM~\\citep{tozzi01,voit03} predicts the entropy profile approximated by $r^{1.1}$, assuming that the gravitational energy of the accreting gas is efficiently converted into the thermal energy. However, Chandra observations revealed that the entropy in the central regions of relaxed clusters often showed a flat profile \\citep{cavagnolo09}. The origin of this behavior is considered to be the feedback of star formation and central AGN activities, which generate extra entropy \\citep{peterson06}. Entropy profile is thus sensitive to the non-gravitational energy injection and can be used to explore such a process even in the outer regions of clusters. We will define the ``entropy'' of the ICM by $K = kT\\cdot n_e^{-2/3}$. Fig.\\ \\ref{fig:comp} shows the resultant entropy profile of A~3376. The slope shows a large deviation from the standard value of 1.1 and close to 0.4 in the radius range \\timeform{0.5'}-\\timeform{20.0'}. The solid black line shows $K\\propto r^{0.4}$, compared with the solid gray line indicating $K\\propto r^{1.1}$. Recent XMM-Newton study reported entropy profiles of 31 clusters \\citep{pratt10}, and the distribution of the slope indicated a peak value of 0.98. Therefore, the slope of 0.4 is very unusual and can be regarded as a transient value occurring in a certain phase of cluster mergers. The slope shows a steepening and flattening across the radio relic, which itself indicates the shock heating. A similar feature was also reported in A~3667 \\citep{akamatsu11b}. The slope of A~3667 showed a marginal increase and a sudden drop across the radio relic region. In A~3376, the increase of the entropy slope is more significant, which suggests a stronger shock than in A~3667\\@. This probably corresponds to the feature that the estimated Mach number in A~3376 is factor of 2 higher that in A~3667\\@. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %Dynamical time scale %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \\subsection{Relaxation Time Scales} \\label{sec:timescale} As shown in the previous sections, A~3376 exhibits distinct ICM properties which are significantly different from those seen in the relaxed clusters. The density and temperature structures around the shock region, along with the irregular morphology of this cluster, suggest that A~3376 is a young system regarding the relaxation of the ICM\\@. The fact that most of the relaxed clusters now attain almost universal temperature profile indicates that the time needed for the ICM to reach thermal and dynamical equilibrium is much shorter than the cluster life. In this section we evaluate several relevant time scales based on the present observational quantities. We first estimate the time in which the shock front reaches the current position. Assuming that the shock has traveled the distance between the cluster center and the relic position with a constant velocity ($v = 2000$ km s$^{-1}$), the time required to propagate this length of $\\sim 1.5$ Mpc is 0.32 Gyr. Next, we estimate the time scale to attain thermal equilibrium after the shock heating. The shock heating should first act on ions rather than electrons, and then the thermal energy is transferred from ions to electrons through Coulomb collisions. The equilibration time scale between ion and electron is 3 orders of magnitude longer than the electron-electron timescale. Therefore, the cluster thermal time scale after the shock heating is limited by the electron-ion time scale. According to \\citet{takizawa98}, the electron-ion equilibration timescale is given by \\begin{equation} t_{\\rm ie}= 2\\times 10^{8}{\\rm\\ yr} \\left(\\frac{n_{e}}{10^{-3}{\\rm\\ cm^{-3}}}\\right)^{-1} \\left(\\frac{T_{e}}{10^{8}\\rm\\ K}\\right)^{3/2} \\left(\\frac{\\ln \\Lambda}{40}\\right), \\end{equation} where $\\ln \\Lambda$ denotes the Coulomb logarithm \\citep{spitzer56}. In the central region, this gives 0.03 Gyr, which is about 10 times shorter than the elapsed time after the shock passage. We calculated the elapsed time after the shock passage and the time for electron-ion equilibration $t_{\\rm ie}$ as a function of radius. The resultant time scales are shown in Fig.\\ \\ref{fig:comp} (c). Here, red diamonds show $t_{\\rm ie}$, and gray dashed diamonds are the ionization time scale $t_i$ of Fe-K$\\alpha$ assuming $n_et=3\\times10^{12}$ cm$^{-3}$~s. Solid line shows the elapsed time after the shock passage, assuming the shock propagation from inner to outer regions. We can see that the elapsed time is shorter than $t_{\\rm ie}$ outside of $0.6 r_{\\rm shock}$ corresponding to $17'$ (920 kpc). This suggests that the electron temperature is likely to be lower than the ion temperature in the outer region including the radio relic \\citep{akahori08, akahori11}. Those sign already reported another merging cluster RX J1347.5-1145~\\citep{ota08}. In the region that $t_{i}$ exceed the sound crossing time $t_{sc}$ ( $t_{i} >~t_{sc}$), there will be non-equilibrium ionization state. The future high resolution X-ray spectroscopy can reveal those dynamical ionization states. Finally, we estimate the time scale in which the sharp change of the temperature disappears. Those high temperature region will diffuse out due to the high pressure in the post-shock region and reach equilibrium in the sound crossing time ($t_{sc}=R/v_{s}$), where $R$ is the width of the high temperature region and $v_{s}$ is the sound speed, respectively. Using the observed temperature ($kT= 4.7$ keV which corresponds to $v_{s}=1100$~km s$^{-1}$) and the width of the high temperature region ($R=320$ kpc), we can calculate the sound crossing time. The resultant time scale is $t_{sc}=0.28$ Gyr. In summary of this section, the observed features of temperature change, entropy slope and the time scale consideration all point to that the outer region ($> 1$ Mpc) of A~3376 is not reaching the equilibrium. In this sense, A~3376 is still a very young system in view of the cluster evolution. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %Summary!!!! %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%" }, "1112/1112.3645_arXiv.txt": { "abstract": "LHC-7 has narrowed down the mass range of the light Higgs boson. This result is consistent with the supergravity unification framework, and the current Higgs boson mass window implies a rather significant loop correction to the tree value, pointing to a relatively heavy scalar sparticle spectrum with universal boundary conditions. It is shown that the largest value of the Higgs boson mass is obtained on the Hyperbolic Branch of radiative breaking. The implications of light Higgs boson in the broader mass range of 115 GeV to 131 GeV and a narrower range of 123 GeV to 127 GeV are explored in the context of the discovery of supersymmetry at LHC-7 and for the observation of dark matter in direct detection experiments. ", "introduction": "} In models based on supersymmetry the light Higgs boson~\\cite{HiggsBoson} has a predictive mass range, and recently LHC-7 has stringently constrained the light Higgs boson to lie in the $115\\GeV$ to $131\\GeV$ range~(ATLAS) and the $115\\GeV$ to $127\\GeV$ range~(CMS) at the $95\\%$~C.L.~\\cite{dec13} with possible hints of evidence within a few GeV of $125\\GeV$. This mass window lies in the range predicted by supergravity unification {(SUGRA)}~\\cite{can} (for reviews see~\\cite{Nath,IbanezRoss,BSM}). In this work we investigate supergravity model points that are consistent with the mass range given by the new LHC-7 data~\\cite{dec13} (for a previous work on the analysis of the Higgs boson in SUGRA and string models pointing to a heavier Higgs in the $120\\GeV$ range see~\\cite{fln-higgs}). LHC-7 has made great strides in exploring the parameter space of supersymmetric models. Indeed, early theoretical projections for the expected reach in sparticle masses and in the $m_0-m_{1/2}$ plane for LHC-7~\\cite{arXiv:0912.4217,arXiv:1002.2430,Baer:2010tk,arXiv:1008.3423} have been met and exceeded by the $1\\fb$ and $2\\fb$ LHC-7 data~\\cite{cmsREACH,AtlasSUSY,atlas0lep,atlas165pb,atlas1fb}. { The implications of the new LHC results have been analyzed} by a number of authors in the context of lower limits on supersymmetric particles and in connection with dark matter~\\cite{LHC-7,Akula:2011zq,Akula:2011dd,Akula:2011ke,Akula2,Grellscheid:2011ij,Ellwanger:2011mu}. Now the most recent results from CERN~\\cite{dec13} indicate that { the two detectors, ATLAS and CMS, have collected as much as $5\\fb$ of data.} One of the most interesting implications of the LHC-7 data concerns the constraints it imposes on the Higgs boson mass. As mentioned above we will work within the framework of a supergravity grand unification model with universal boundary conditions~\\cite{can,hlw,nac}. Here we discuss the dependence of the light Higgs boson mass on the parameter space, i.e., on $m_0, m_{1/2}, A_0, \\tan\\beta$~\\cite{an1992}, where $m_0, m_{1/2}$ and $A_0$ are the parameters at the GUT scale, where the GUT scale, { $M_\\mathrm{GUT} \\sim 2 \\times10^{16}\\GeV$} is defined as the scale at which the gauge couplings unify, and where $m_0$ is soft scalar mass, $m_{1/2}$, the gaugino mass, $A_0$, the trilinear coupling and $\\tan\\beta$, the ratio of the two Higgs VEVs in the minimal supersymmetric standard model. An important aspect of SUGRA models is that the radiative electroweak symmetry breaking, REWSB, is satisfied for $A_0/m_0$ {typically in the $-5$ to $5$ range}. The renormalization group evolution then leads to a value of the trilinear coupling, $A_t$, at the electroweak scale to also be $\\mathcal{O}({\\rm TeV})$. The relevance of this observation is that quite generically supergravity unification { leads } to a sizable $A_t$ which is needed to give a substantial leading order loop correction to the Higgs Boson mass for any fixed $\\mu, \\tan \\beta$ and $m_0$, where $\\mu$ is the Higgs mixing parameter in the superpotential. Thus a generic prediction of SUGRA models under radiative electroweak symmetry breaking for a sizable $A_0/m_0$ is that there would be a substantial loop correction to the Higgs boson mass, and it is well known that the light Higgs mass at the tree level has the value $\\mh \\leq M_Z$ and there is a significant loop correction $\\Delta \\mh$ to lift it above $M_Z$~\\cite{1loop,hep-ph/9210242,Casas:1994us,Carena:1995wu,Espinosa:2000df,Carena:1998wq,hep-ph/0503173}. The dominant one loop contribution arises from the top/stop sector and is given by \\beqn \\Delta \\mh^2\\simeq \\frac{3m_t^4}{2\\pi^2 v^2} \\ln \\frac{M_{\\rm S}^2}{m_t^2} + \\frac{3 m_t^4}{2 \\pi^2 v^2} \\left(\\frac{X_t^2}{M_{\\rm S}^2} - \\frac{X_t^4}{12 M_{\\rm S}^4}\\right)~, \\label{tloop} \\eeqn where $v=246\\GeV$, $M_{\\rm S}$ is an average stop mass, and $X_t$ is given by \\beqn X_t\\equiv A_t - \\mu \\cot\\beta~. \\label{xt} \\eeqn From Eq.~(\\ref{tloop}) one finds that the loop correction is maximized when \\be X_t \\sim \\sqrt 6 M_{\\rm S} ~. \\label{xt} \\ee We note that there can be important loop corrections also from the $b$-quark sector and a correction similar to Eq.~(\\ref{tloop}) can be written where $X_t$ is replaced by $X_b= A_b- \\mu \\tan\\beta$ along with other appropriate replacements. Thus when $\\mu \\tan\\beta$ becomes large, the $b$-quark contribution to the loop correction, which is proportional to powers of $X_b$, becomes large and is comparable to the top contribution which implies that a high Higgs mass can also result in stau-coannihilation models where typically $\\mhf$ is large and $m_0$ is relatively small. Further, we note that the approximation of Eq.~(\\ref{xt}) would not hold if the off-diagonal elements of the stop mass squared matrix are comparable to the diagonal elements which can happen for very large $A_t$. In addition, it is well known that the two loop corrections are substantial (see e.g.~\\cite{Slavich2} for a numerical analysis). While the correction at the one loop level has the symmetry $X_t \\to - X_t$, this symmetry is lost when the two loop corrections are included and then $\\sgn\\left(A_0/m_0\\right)$ plays an important role in the corrections to the Higgs boson mass. As seen later this observation is supported by the full numerical analysis which includes the two loop corrections. We note in passing that the theoretical predictions for the light Higgs boson mass depend sensitively on the input parameters which include the gauge coupling constants as well as the top mass with their experimental errors. Additionally, there are also inherent theoretical uncertainties which together with the uncertainties of the input parameters allow theoretical predictions of the light Higgs boson mass accurate to only within an error corridor of a few GeV (see e.g.~\\cite{Slavich2}). Since the loop corrections involve the sparticle spectrum, a large loop correction implies a relatively heavy sparticle spectrum and specifically heavy scalars. Such a possibility arises in REWSB which allows for scalars heavier than $10\\TeV$~\\cite{Chan:1997bi}. Specifically, with scalars approaching $10\\TeV$, the Higgs boson mass can remain heavy while the gaugino sector is free to vary. This occurs within the minimal SUGRA framework and similar situations arise in other works of radiative breaking~\\cite{Feldman:2011ud,Baer}. Indeed, quite generally in SUGRA and string models with the MSSM field content, the analysis of the Higgs mass with loop {corrections} under the constraints of REWSB { gives an upper} limit on the light Higgs boson mass of about { $135\\GeV$ for a wide range of input parameters}.\\footnote{ We note that heavier Higgs boson masses can be obtained in a variety of different models such as hierarchical breaking models~\\cite{Wells1,adgr,Kors:2004hz} (for recent work see~\\cite{Cabrera:2011bi,arXiv:1108.6077}) or by addition of vector like multiplets~\\cite{Babu:2008ge}.} {A very} interesting aspect of the recent LHC-7 data concerns the fact that a large portion of the Higgs boson mass window has been excluded and what remains is consistent with the range predicted by the SUGRA models. ", "conclusions": "Recent data from LHC-7 indicates a narrow window on the light Higgs mass. This allowed mass window is consistent with the range predicted by SUGRA models and specifically by the mSUGRA model. Here we discussed the implications of the indicated mass range for the light Higgs mass for the sparticle mass spectrum and for dark matter. Using the allowed Higgs mass range above $115\\GeV$ the corresponding ranges for the soft masses and couplings, as well as the ratio of the vacuum expectation values of the Higgs doublets and the Higgsino mass parameter were found. We then investigated the ranges for the sparticle masses correlated to the predicted value of the Higgs Boson mass, specifically for the chargino, the neutralino, the gluino, the stop, the stau, for the first and second generation squarks and sleptons {and for the heavier Higgs of the minimal supersymmetric standard model, i.e., the CP odd Higgs $A^0$, the CP even Higgs $H^0$, and the charged Higgs $H^{\\pm}$.} {Our conclusions are} that the largest Higgs masses are realized on the Focal Surface {of the Hyperbolic Branch} of radiative electroweak symmetry breaking. {We also point out that low values of $\\mu \\sim 150 \\GeV$ are consistent with heavy squarks and sleptons in the 10 TeV region or larger. We find that $\\mh~\\in (123-127)\\GeV$ does allow for light third generation stop as low as $m_{\\tilde t_1} > 230\\GeV$, though the second generation squarks are at least $m_{\\tilde q} > 1.5\\TeV$ and second generation sleptons are at least 475 GeV.} { Thus, the restriction of the light Higgs boson to the mass window $\\mh~\\in (123-127)\\GeV$ provides further constraints on the sparticle spectrum that are complimentary to the direct searches for sparticles at the LHC. } { Further, we find precise predictions for dark matter if the light Higgs boson mass lies between $123\\GeV$ and $127\\GeV$.} For these light Higgs boson masses, the corresponding range of the lightest neutralino mass would be accessible in the next generation of direct detection dark matter experiments. The light Higgs boson in the $123\\GeV$ and $127\\GeV$ range was shown to be generic for the case of heavy scalars in minimal supergravity with {$|A_0/m_0|\\sim\\mathcal{O}(1)$.}" }, "1112/1112.1689_arXiv.txt": { "abstract": "\\noindent A stochastic gravitational wave background (SGWB) would gravitationally lens the cosmic microwave background (CMB) photons. We correct the results provided in existing literature for modifications to the CMB polarization power spectra due to lensing by gravitational waves (GW). Weak lensing by gravitational waves (GW) distorts all the four CMB power spectra, however its effect is most striking in the mixing of power between the E-mode and B-mode of CMB polarization. This suggests the possibility of using measurements of the CMB angular power spectra to constrain the energy density ($\\Omega_{GW}$) of the SGWB. Using current data sets (QUAD, WMAP and ACT), we find that the most stringent constraints on the present $\\Omega_{GW}$ come from measurements of the angular power spectra of CMB temperature anisotropies. In near future more stringent bounds on $\\Omega_{GW}$ can be expected with improved upper limits on the B-modes of CMB polarization. Any detection of B-modes of CMB polarization above the expected signal from large scale structure(LSS) lensing could be a signal for a SGWB. ", "introduction": " ", "conclusions": "" }, "1112/1112.4586_arXiv.txt": { "abstract": "We study shear viscosities of different species in hot and neutrino-trapped dense matter relevant to protoneutron stars. It is found that the shear viscosities of neutrons, protons and electrons in neutrino-trapped matter are of the same orders of magnitude as the corresponding shear viscosities in neutrino-free matter. Above all, the shear viscosity due to neutrinos is higher by several orders of magnitude than that of other species in neutrino-trapped matter. Next we investigate the effect of shear viscosity in particular, neutrino shear viscosity on the thermal nucleation rate of droplets of antikaon condensed matter in protoneutron stars. The first-order phase transition from hadronic matter to antikaon condensed matter is driven by the thermal nucleation process. We compute the equation of state used for the calculation of shear viscosity and thermal nucleation time within the relativistic mean field model. Neutrino shear viscosity enhances the prefactor in the nucleation rate by several orders of magnitude compared with the $T^4$ approximation of earlier calculations. Consequently the thermal nucleation time in the $T^4$ approximation overestimates our result. Furthermore, the thermal nucleation of an antikaon droplet might be possible in neutrino-trapped matter before neutrino diffusion takes place. \\pacs{97.60.Jd, 26.60.-c,52.25.Fi,64.60.Q-} ", "introduction": "A first order phase transition from nuclear matter to some exotic form of matter might be possible in (proto)neutron stars. It could be either a nuclear to quark matter transition or a first order pion/kaon condensation. Consequently, it might have tremendous implications for compact stars \\cite{Gle} and supernova explosions \\cite{Sag}. Here the focus is the first order phase transition proceeding through the thermal nucleation of a new phase in particular, antikaon condensed phase in hot and neutrino-trapped matter. After the pioneering work by Kaplan and Nelson on antikaon ($K^-$ meson) condensation in dense baryonic matter formed in heavy ion collisions as well as in neutron stars \\cite{Kap}, several groups pursued the problem of antikaon condensation in (proto)neutron stars \\cite{Bro,Tho,Ell,Lee,Pra97,Gle99,Kno,Sch,Pal,Bani1,Bani2,Bani3,Bani4,Pons,Bani5}. In most cases, the phase transition was studied using either Maxwell construction or Gibbs' rules for phase equilibrium coupled with global baryon number and charge conservation \\cite{Gle92}. The first order phase transition driven by nucleation of antikaon condensed phase was considered in a few cases \\cite{Nor,Bani7}. In particular, the calculation of Ref.\\cite{Bani7} dealt with the role of shear viscosity on the the thermal nucleation of antikaon condensed phase in hot and neutrino-free compact stars \\cite{Bani7}. It is to be noted here that the first order phase transition through the thermal nucleation of quark matter droplets was also investigated in (proto)neutron stars \\cite{Nor,San,Sat,Hei,Bom,Min,Bom2} using the homogeneous nucleation theory of Langer \\cite{Nor,San,Lan}. The thermal nucleation is an efficient process than the quantum nucleation at high temperatures \\cite{Bom,Bom2}. We adopt the homogeneous nucleation theory of Langer \\cite{Lan,Tur} for the thermal nucleation of antikaon condensed phase. Nuclear matter would be metastable near the phase transition point due to sudden change in state variables. In this case thermal and quantum fluctuations are important. Droplets of antikaon condensed matter are formed because of thermal fluctuations in the metastable nuclear matter. Droplets of the new and stable phase which are bigger than a critical radius, will survive and grow. The transportation of latent heat from the surface of the droplet into the metastable phase favours a critical size droplet to grow further. This heat transportation could be achieved through the thermal dissipation and viscous damping \\cite{Tur,Las,Raj}. A parametrised form of the shear viscosity was used in earlier calculations of the nucleation of quark matter \\cite{Bom}. Recently, the influence of thermal conductivity and shear viscosity on the thermal nucleation time was studied in a first-order phase transition from nuclear to antikaon condensed matter in hot neutron stars \\cite{Bani7}. The shear viscosity due to neutrinos was not considered in that calculation. It would be worth studying the effect of shear viscosity on the thermal nucleation rate of droplets of antikaon condensed matter in neutrino-trapped matter relevant for protoneutron stars. Besides shear viscosities due to neutrons, protons and electrons, this involves the contribution of neutrinos to the total shear viscosity. Shear viscosities of pure neutron and neutron star matter were calculated by several groups \\cite{Flo1,Flo2,Cut,Ben,Yak,Glam}. We also performed the calculation of shear viscosity in neutron star matter using the equation of state (EoS) derived from relativistic field theoretical models \\cite{Bani6}. Transport properties of degenerate neutrinos in dense matter were estimated by Goodwin and Pethick \\cite{Good}. We organise the paper in the following way. We describe models for EoS, shear viscosities of different species including neutrinos and the calculation of thermal nucleation rate in Sec. II. Results of this calculation are discussed in Sec. III. Sec. IV gives the summary and conclusions. ", "conclusions": "We have studied shear viscosities of different particle species in neutrino-trapped $\\beta$-equilibrated and charge neutral nuclear matter. We have derived equations of state of nuclear and antikaon condensed phases in the relativistic mean field model for the calculation of shear viscosity. It is noted that neutrons, protons and electrons come into thermal equilibrium in the weak interaction time scale. The shear viscosity due to neutrinos is calculated treating other particles as background and found to dominate the total shear viscosity. Next we have investigated the first-order phase transition from neutrino-trapped nuclear matter to antikaon condensed matter through the thermal nucleation of a critical droplet of antikaon condensed matter using the same relativistic EoS as discussed above. Our emphasis in this calculation is the role of the shear viscosity due to neutrinos in the prefactor and its consequences on the thermal nucleation rate. We have observed that the thermal nucleation of a critical antikaon droplet might be possible well before the neutrino diffusion takes place. Furthermore, a comparison of our results with that of the calculation of thermal nucleation time in the $T^4$ approximation shows that the latter overestimates our results of thermal nucleation time computed with the prefactor including the neutrino shear viscosity. Though we have performed this calculation with antikaon optical potential depth of $U_{\\bar K} (n_0) = -120$ MeV, we expect qualitatively same results for other values of the antikaon optical potential depth \\cite{Bani8}." }, "1112/1112.1476_arXiv.txt": { "abstract": "Cosmic rays fill up the entire volume of galaxies, providing an important source of heating and ionisation of the interstellar medium, and may play a significant role in the regulation of star formation and galactic evolution. Diffuse emissions from radio to high-energy \\grays{} ($> 100$ MeV) arising from various interactions between cosmic rays and the interstellar medium, interstellar radiation field, and magnetic field, are currently the best way to trace the intensities and spectra of cosmic rays in the Milky Way and other galaxies. In this contribution, I describe our recent work to model the full spectral energy distribution of galaxies like the Milky Way from radio to \\gray{} energies. The application to other galaxies, in particular the Magellanic Clouds and M31 that are detected in high-energy \\gray{s} by the Fermi-LAT, is also discussed. ", "introduction": "The luminosity of a star-forming galaxy like the Milky Way (MW) is dominated by the relatively narrow frequency range of the spectral energy distribution (SED) from the ultraviolet (UV) to far infrared (FIR), which is due to stellar emission and dust reprocessing in the interstellar medium (ISM). Related to the birth and death of massive stars, cosmic rays (CRs) are pervasive throughout the ISM (see, e.g.,~\\cite{Strong2007}~for a recent review). The diffuse emissions arising from various interactions between CRs and the ISM, interstellar radiation field (ISRF -- the UV--FIR component of the galactic SED), and magnetic field span radio frequencies to high-energy \\grays{} ($>$100 MeV), but at a lower level of intensity compared to the stellar and dust component. These broadband emissions are currently the best way to trace CR intensities and spectra throughout the MW and other galaxies. Gamma rays are particularly useful in this respect because this energy range gives access to the dominant hadronic component in CRs via the observation of $\\pi^0$-decay radiation produced by CR nuclei inelastically colliding with the interstellar gas. Understanding the global energy budget of processes related to the injection and propagation of CRs, and how the energy is distributed across the electromagnetic spectrum, is essential to interpret the radio/far-infrared relation (\\cite{Helou1985,Murphy2006}), galactic calorimetry (e.g.,~\\cite{Volk1989}), and predictions of extragalactic backgrounds (e.g.,~\\cite{Thompson2007,Murphy2008}), and for many other studies. ", "conclusions": "" }, "1112/1112.1979_arXiv.txt": { "abstract": "The Large Area Telescope on board the \\textit{\\textit{Fermi}} satellite observed a gamma-ray flare in the Crab nebula lasting for approximately nine days in April of 2011. The source, which at optical wavelengths has a size of $\\approx$11 ly across, doubled its gamma-ray flux within eight hours. The peak photon flux was $(186 \\pm 6) \\times 10^{-7}$ cm$^{-2}$ s$^{-1}$ above 100 MeV, which corresponds to a 30-fold increase compared to the average value. During the flare, a new component emerged in the spectral energy distribution, which peaked at an energy of (375 $\\pm$ 26) MeV at flare maximum. The observations imply that the emission region was likely relativistically beamed toward us and that variations in its motion are responsible for the observed spectral variability. ", "introduction": "The Crab nebula is the remnant of a supernova observed in 1054 AD. The explosion left behind a rotating neutron star emitting electromagnetic radiation pulsed at the rotation period, that powers a wind of relativistic particles. These particles interact with the remnant gas and magnetic field, causing the nebula to glow brightly at all wavelengths, predominantly by synchrotron radiation. The spectral energy distribution (SED) of the nebula is, accordingly, dominated by a synchrotron component extending from radio wavelengths into the gamma-ray band \\citep{hester_crab_2008}. Above 450 MeV a second component emerges, attributed to inverse-Compton scattering by the same relativistic particles \\citep{gould_1965,dejager_1992,atoyan_fluxes_1996}. The angular size of the Crab nebula is $\\approx$0.1$^\\circ$ in the optical and smaller at higher energies. This corresponds to 3.5 pc, or 11 ly, at its estimated distance of 2 kpc \\citep{trimble_distance_1973}. Today, the pulsar and nebula (henceforth referred to together as the Crab) are considered prime examples of non-thermal sources in the Universe and serve as a laboratory for relativistic plasma physics. New puzzles for our understanding of the Crab have been posed by the detection of three bright gamma-ray flares by the AGILE satellite and the Large Area Telescope (LAT) on board the \\textit{Fermi} satellite between 2007 and 2010 \\citep{abdo_crab_2011,tavani_crab_2011}. During these flares the unpulsed component of the gamma-ray flux increased by a factor of {$\\approx$10} on time scales as short as 12 hours \\citep{balbo_crab_2011}, while the period and flux of the pulsed component remained stable. More recently, in April of 2011 the LAT detected a fourth flare, three times brighter than any of the previous ones \\citep{buehler_crab_2011,hays_crab_2011}. The flare was swiftly confirmed by the AGILE satellite \\citep{striani_crab_2011}. Observations at lower energies, in particular by the Chandra X-ray observatory, have not yet revealed any variability correlated with the gamma-ray flares \\citep{tennant_2011}. However, analysis of these observations is ongoing and will be discussed elsewhere. Here we present the LAT gamma-ray results obtained during the flare, together with a broader analysis of the first 35 months of Crab observations by \\textit{Fermi}. ", "conclusions": "A year after first being reported, the gamma-ray flares from the Crab remain enigmatic. Where within the nebula does the emission come from? What produces the flux variations? How were the emitting particles accelerated? How are the flares related to the variability observed on yearly and monthly time scales? Although several ideas have been proposed, no certain answers can be given today. The observations presented here give us the most precise look into the flare phenomenon to date, during the brightest outburst detected so far. We will proceed to discuss some of the implications and challenges posed by these observations. One striking property of the Crab nebula flares is their rapid flux variations, doubling within {t$_{d} < 8$ hours} at the rise of the 2011 April flare. Causality arguments imply that the emission region is compact, with a length {$L < c t_{d} \\approx 2.8 \\times 10^{-4}$ pc}. The emitted isotropic power at the peak of the flare of $\\approx4 \\times 10^{36}$ erg s$^{-1}$ corresponds to $\\approx$1\\% of the total spin-down power of the pulsar, the ultimate energy source of the nebula. It is difficult to explain how this energy is focused into such a small emission volume. The focusing is generally easier to explain when the emission site is closer to the pulsar. The absence of pulsation in the flare signal implies that the emission region is at least located outside the light cylinder of the pulsar. Another possibility to explain the flare brightness is that the emission is highly anisotropic, as would be expected if the emission region moves relativistically toward us. While only mildly relativistic motion with velocities of {$\\approx$0.5c} are observed inside the nebula \\citep{scargle_activity_1969,hester_hubble_2002,melatos_2005}, relativistic motion is expected in the pulsar wind and in the downstream medium behind the wind termination \\citep{camus_2009}. Relativistic bulk motion is particularly expected at the ``arch shock'' of the wind termination, which has been proposed as the main site of gamma-ray emission \\citep{komissarov_2011}. The flare emission is expected to result from synchrotron radiation by relativistic electrons and positrons (henceforth referred to together as electrons) \\citep{abdo_crab_2011}. A new spectral component emerges in the SED during the flare. The hard photon spectrum ($\\gamma \\approx 1.3$) of the flaring component implies that most of the electron energy is carried by the highest energy electrons. If the electron \\emph{particle} density $n(\\epsilon)$ per energy at an energy $\\epsilon$ is characterized by a power law $n(\\epsilon) \\sim \\epsilon^{-p}$, the spectral index is $p \\le 2 \\gamma - 1 \\approx 1.6$, in a random magnetic field in which the electrons are isotropically distributed. The energy per logarithmic energy interval, which is proportional to $\\epsilon^2 \\times n(\\epsilon)$, is therefore rising with energy. We note that such a spectrum is also inferred for the radio emitting electrons in the Crab nebula, and more generally in pulsar wind nebulas \\citep{gaensel_2006,sironi_2011}. While the radio emission is produced by a different electron population than the gamma-ray flares, efficient particle acceleration appears to be a common feature in these systems. It is an interesting question how such a hard electron spectrum is produced. Standard diffusive shock acceleration typically results in spectra with $p \\ge 2$ \\citep{gallant_1992,kirk_2000}. Even though it has been shown that harder spectra can be produced in certain field configurations with low-level turbulence \\citep{kirk_1989,summerlin_2011}, these conditions are not expected at termination shocks of pulsar winds. Additionally, shock acceleration appears to be inefficient at highly oblique shocks that are representative of the pulsar wind termination discontinuity \\citep{ellison_2004,summerlin_2011,sironi_2011b}. One alternative is that magnetic reconnection in the striped pulsar wind might accelerate particles \\citep{lyubarsky_2003,kirk_2004,yuan_2011,bednarek_2011}. However, simulations show that reconnection behind the pulsar termination probably does not provide the required electron energies to produce gamma-ray emission \\citep{sironi_2011}. Another interesting possibility is that acceleration is occurring directly in the electric field induced by the pulsar, as discussed by \\citet{abdo_crab_2011}. The observation of a peak synchrotron energy of $\\approx$380 MeV is among the highest yet seen from astrophysical source today. The observation is surprising as particle acceleration in the presence of synchrotron cooling is expected to limit synchrotron emission to photon energies below $\\approx$150 MeV \\citep{guilbert_1983,de_jager_gamma-ray_1996,komissarov_2011}. Two solutions to this problem have been proposed recently in this context: \\begin{enumerate} \\item The electric field at the acceleration site is larger in magnitude than the magnetic field. This is generally an unstable state in plasma, as charges will short out the electric field; however, temporarily such a configuration is expected, e.g. in magnetic reconnection events. The gamma-ray emission might occur afterward, when the accelerated electrons enter a region of enhanced magnetic fields \\citep{uzdensky_2011,cerutti_2011}. More detailed studies are required to assess whether such a scenario is plausible in the nebula environment and can be sustained for the duration of the flares. \\item The gamma rays are emitted in a region of bulk relativistic motion, and are therefore Doppler boosted toward the observer \\citep{komissarov_2011}. For a flow moving directly toward us a Lorentz factor $\\gtrsim 2$ is sufficient to accommodate the observed peak energy. \\end{enumerate} Variations in the Doppler boosting can naturally account for the observed flux variation \\citep{lyutikov_2011}. The observed spectral evolution is compatible with such an interpretation: the energy flux of the emission varies approximately as a power of {$\\alpha$ = 3.42 $\\pm$ 0.86} with the cutoff energy; a correlation with {$\\alpha \\approx$ 3} is indeed expected for variations produced by changes in relativistic beaming \\citep{lind_1985}. The flare brightness, the high frequency of the observed peak of the gamma-ray emission, and the spectral evolution during the flare all suggest the presence of relativistic beaming. We therefore conclude that, independent of the location of the emission region and the physical processes responsible for the flares, the emission region is moving relativistically toward us, and changes in its motion are likely the predominant mechanism responsible for the observed flux variations. Such a kinematic explanation does however not address the issue of how a moving source can be created dynamically and sustained radiatively in the face of strong losses. The Crab Nebula still has much more to teach us." }, "1112/1112.3184_arXiv.txt": { "abstract": "We estimate distances to the spherical planetary nebula Abell\\,39 and the bipolar planetary nebula NGC\\,7027 by interpolating from a wide grid of photoionization models using the 3-D code, MOCASSIN. We find preliminary distances of 1.5\\,kpc and 0.9\\,kpc respectively, with uncertainties of about 30\\%. ", "introduction": "Accurate distances to planetary nebulae (PNe) are crucial in unraveling the connection between the physical properties of the nebulae with those of their central stars (CSs). Reliable distances facilitate the accurate estimation of fundamental parameters such as the CS mass and luminosity, and the nebular mass and age. We have begun a program of using photoionization modeling to refine the distances to a sample of nearby PNe (see also \\cite[Danehkar \\etal\\ 2011]{Danehkar_etal11}). In this work, we study two very different PNe as a proof of concept: Abell\\,39, a simple spherical shell with no microstructures, and NGC\\,7027, a well-known, young, luminous bipolar PN with a massive molecular envelope. Using the 3-D photoionization code (MOCASSIN; \\cite[Ercolano et al. 2003]{Ercolano_etal03}), our ultimate aim is to constrain the distance to individual PNe, utilizing the physical PN radius we have calculated and the angular size. ", "conclusions": "" }, "1112/1112.3651_arXiv.txt": { "abstract": "We are building a new spectral library with the X-Shooter instrument on ESO's VLT: XSL, the X-Shooter Spectral Library. We present our progress in building XSL, which covers the wavelength range from the near-UV to the near-IR with a resolution of $R\\sim10000$. At the time of writing we have collected spectra for nearly 240 stars. An important feature of XSL is that we have already collected spectra of more than 100 Asymptotic Giant Branch stars in the Galaxy and the Magellanic Clouds. ", "introduction": "Stellar population models are powerful tools which are widely used to study galaxy evolution. Using these models, one can determine galaxy ages, metallicities and abundances. Spectral libraries are an integral component of stellar population models (e.g. \\cite{BC03}). A spectral library gives the behavior of individual stellar spectra as function of effective temperature ($T _{\\mathrm{eff}}$), gravity ($\\log g$) and metallicity ([Fe/H]). A stellar population model integrates these spectra together with a set of stellar isochrones and an initial mass function to produce a model spectrum of an entire population. In order to reproduce galaxy spectra as precisely as possible, one requires a comprehensive stellar spectral library that covers the entire desired parameter space of $T _{\\mathrm{eff}}$, $\\log g$, and [Fe/H]. Moreover, extended wavelength coverage is strongly desirable, because different stellar phases contribute their light in different bands. For instance, bright giants contribute more light in the near-infrared than the faint giant and subgiant stars, while in the optical, the situation is reversed \\cite{Frogel88}. These Asymptotic Giant Branch (AGB) stars dominate the light of intermediate-aged and old stellar populations in the near-infrared but are unimportant in the optical (e.g., \\cite{Maraston05,Conroy10}). Detecting their presence requires \\emph{broad wavelength coverage} in both the target and model spectra. Stellar spectral libraries can be classified into empirical and theoretical libraries, depending on how the library is obtained. Both theoretical and empirical libraries have improved in recent years. The most widely used theoretical libraries in stellar population models are those of \\cite{Kurucz93}, \\cite{Munari05}, \\cite{Gustafsson08}, \\cite{Coelho05}, \\cite{Coelho07}, and \\cite{Martins05}. Theoretical libraries have the advantage of (nearly) unlimited resolution and selectable abundance patterns -- not only scaled-solar abundances but also non-solar patterns. Unfortunately, theoretical libraries suffer from systematic uncertainties, as they rely on model atmospheres and require a reliable list of atomic and molecular line opacities \\cite{Coelho05}. Empirical stellar libraries, on the other hand, have the advantage of being drawn from real, observed stars and therefore do not suffer this limitation; however they frequently have relatively low resolution (with a few exceptions; see below) and are unable to reproduce the indices measured in giant elliptical galaxies \\cite{Reynier89, Worthey92}, because they are based on local stars with typical Milky Way disk abundance patterns. In this proceeding we present the X-Shooter Stellar Library, XSL. XSL is being obtained using the new X-Shooter three-arm spectrograph on ESO's VLT \\cite{Vernet10}. XSL has the unique advantage of simultaneously acquiring spectra covering the near-ultraviolet up to the near-infrared. Furthermore, the fact that X-Shooter is mounted on an 8.2-m telescope allows us to include faint objects in stars in the Galactic bulge and the Magellanic Clouds for the first time, along with stars in the Galactic disk and halo, at moderate spectral resolution ($R\\sim10000$). \\subsection{Previous stellar libraries} \\begin{table} \\caption{\\label{tablib}Some previous stellar libraries} \\begin{center} \\lineup \\begin{tabular}{*{5}{l}} \\br Library&Resolution&Spectral range&Number &Reference\\cr &R=$\\lambda/\\Delta\\lambda$&(nm)&of stars&\\cr \\mr STELIB & 2000 & 320-930 & 249 &\\cite{LeBorgne03}\\cr ELODIE & 10000 & 390-680 & 1300 &\\cite{Prugniel01,Prugniel04,Prugniel07}\\cr INDO-US & 5000 & 346-946 & 1237&\\cite{Valdes04}\\cr MILES & 2000 & 352-750 & 985 &\\cite{miles06}\\cr IRTF-SpeX & 2000 & 800-2500& 210 &\\cite{Rayner09}\\cr NGSL & 1000 & 167-1025& 374 &\\cite{Gregg06}\\cr UVES-POP & 80000 & 307-1030& 300 &\\cite{Bagnuolo03}\\cr LW2000 & 1100 & 500-2500& 100 &\\cite{LW2000}\\cr \\br \\end{tabular} \\end{center} \\end{table} We begin with a review of several previous empirical stellar libraries and their principal features, listed in Table~\\ref{tablib}. In the optical, we have (among others) Lick/IDS \\cite{Worthey94}, MILES \\cite{miles06}, ELODIE \\cite{Prugniel01,Prugniel04,Prugniel07}, STELIB \\cite{LeBorgne03}, NGSL \\cite{Gregg06}, and the Pickles library \\cite{Pickles85}. Libraries in the near-IR are a challenging task, but pioneering work has been done by Lan\\c{c}on and Wood \\cite{LW2000} (LW2000), Rayner et al.\\ \\cite{irtf09} (IRTF-SpeX), and M\\'{a}rmol-Queralt\\'{o} et al.\\ \\cite{Marmol-Queralto}. However, extended-wavelength-coverage spectral libraries at moderate resolution are still largely missing. STELIB \\cite{LeBorgne03} has been used to construct the widely used stellar population synthesis models of Bruzual \\& Charlot 2003 (\\cite{BC03}, here after BC03). This library has a wide wavelength range, 3200--9300 \\AA, at modest resolution ($R\\sim2000$). The spectra of local galaxies have been reproduced by the models of BC03 through this library \\cite{Gallazzi05}. However, stars with low and high metallicites are sparse in this library, leading to problems in some regimes, like modeling old, metal-poor globular cluster spectra \\cite{Koleva08}. The ELODIE library \\cite{Prugniel01,Prugniel04,Prugniel07} is a moderate-resolution spectral library with $R\\sim10000$. It has been applied in the P\\'{E}GASE-HR synthetic model \\cite{LeBorgne04}. The atmosphere parameter coverage of this library had been improved through its updated version, but the wavelength range is still limited in the optical region (4100--6800 \\AA). The INDO-US library \\cite{Valdes04} includes a large number of stars (1273), covers a fair range in atmosphere parameters with moderate-resolution ($R\\sim5000$) . Atmospheric parameters of its stars are given in Wu et al. \\cite{Wuyue11}. The data have been used for automated spectral classification of high-resolution spectra over a wide wavelength range \\cite{Pickles07}. The problem of this library is the lack of accurate spectrophotometry, making inclusion in stellar population models problematic. The MILES library \\cite{miles06} is widely used in stellar population models. This library profits from its good physical parameter coverage, careful flux calibration, and a large number of stars (985), enabling stellar population models to predict metal-poor or metal-rich systems. The MILES library's wavelength range covers 3500--7500 \\AA\\ at modest resolution ($R\\sim2000$) \\cite{Falcon11}. The Next General Spectral Library (NGSL\\footnote{See \\url{http://archive.stsci.edu/prepds/stisngsl/}}, \\cite{Gregg06}) is a low resolution ($R\\sim1000$) stellar library obtained with STIS on HST. It contains $\\sim400$ stars having a wide range in metallicity and age. Since its spectra were observed in space, this library does not suffer from telluric absorption or seeing variations, and hence has excellent absolute spectrophotometry. Other features of NGSL are its broad wavelength coverage, from the space UV to 1 $\\mu$m, and high signal-to-noise. There is also a high-resolution ($R\\sim40000$, covering 3600--11000 \\AA) extension of NGSL's southern stars taken with the UVES spectrograph of ESO's VLT (Hanuschik et al., in prep.). The UVES Paranal Observatory Project (UVES-POP, \\cite{Bagnuolo03}) is a library of spectra of $\\sim300$ nearby bright stars taken with UVES. The spectra cover the optical region at $R\\sim80000$ and have a typical S/N ratio is 300--500 in the $V$ band. The library has been flux calibrated but has not been corrected for the severe telluric absorption in the red. For stellar population models the sample is incomplete, since it only contains bright stars around solar metallicity. In the near-infrared range, the IRTF Spectral Library (IRTF-SpeX, \\cite{Rayner09}) contains 200 stars observed with the cross-dispersed infrared SpeX spectrograph on IRTF at a resolving power of $R\\sim2000$. The largest spectral library of very cool supergiants and giants over the wavelength range $\\lambda=0.5$--2.5 $\\mu$m was published by Lan\\c{c}on \\& Wood (LW2000, \\cite{LW2000}), which contains $\\sim100$ stars taken at a resolution of $R=1100$, including observations over multiple phases. Unfortunately, cool giants are variable in optical and NIR, so simultaneous observations are required. Libraries with more limited spectral coverage in the near-infrared, like the 2.3 $\\mu$m library of Marm\\'ol-Queralt\\'o et al.\\ \\cite{MQ08}, also exist, but are limited to a small number of spectral features (like the CO bandhead). ", "conclusions": "XSL, the X-Shooter Stellar Library, is intended to be the largest stellar library with complete wavelength coverage from 320--2480 nm and covering as large a range in stellar atmospheric parameters as possible. We will soon have a first version of XSL with spectra of $\\sim240$ stars. The final library will contain $\\sim600$ stars at moderate resolution ($R\\sim10000$). The stellar parameter coverage will be enforced through carefully selected samples, so that not only stars with solar metallicities and abundances are included but also metal-poor or metal-rich stars from the Bulge and Magellanic Clouds. With X-Shooter's unique capability, the variable stars will be observed consistently, which will yield more reliable stellar population models. \\subsection*" }, "1112/1112.1712_arXiv.txt": { "abstract": "Whether analytic exact vacuum(electrovacuum) solutions of the Einstein(Einstein-Maxwell) field equations can accurately describe or not the exterior spacetime of compact stars remains still an interesting open question in Relativistic Astrophysics. As an attempt to establish their level of accuracy, the radii of the Innermost Stable Circular Orbits (ISCOs) of test particles given by analytic exterior spacetime geometries have been compared with the ones given by numerical solutions for neutron stars (NSs) obeying a realistic equation of state (EoS). It has been so shown that the six-parametric solution of Pach\\'on, Rueda, and Sanabria (2006) (hereafter PRS) is more accurate to describe the NS ISCO radii than other analytic models. We propose here an additional test of accuracy for analytic exterior geometries based on the comparison of orbital frequencies of neutral test particles. We compute the Keplerian, frame-dragging, as well as the precession and oscillation frequencies of the radial and vertical motions of neutral test particles for the Kerr and PRS geometries; then we compare them with the numerical values obtained by Morsink and Stella (1999) for realistic NSs. We identify the role of high-order multipole moments such as the mass quadrupole and current octupole in the determination of the orbital frequencies especially in the rapid rotation regime. The results of this work are relevant to cast a separatrix between black hole (BH) and NS signatures as well as probe the nuclear matter EoS and NS parameters from the Quasi-Periodic Oscillations (QPOs) observed in Low Mass X-Ray Binaries. ", "introduction": "One of the greatest challenges of the general theory of re\\-la\\-tivity has been the construction of solutions to the Einstein-Maxwell field equations representing the gravitational field of compact stars such as neutron stars (NSs). Stationary axially symmetric spacetimes satisfy basic properties one expects for rotating objects, namely time symmetry and reflection symmetry with respect to the rotation axis \\citep[see e.g.][]{2006CQGra..23.3251P}. The simplest stationary axially symmetric exact exterior vacuum solution describing a rotating configuration is the well-known Kerr metric \\citep{1963PhRvL..11..237K}. The Kerr metric is fully described by two free parameters: the mass $M$ and the angular momentum $J$ of the object. However, it is known from numerical models that the quadrupole moment of rotating NSs deviates considerably from the one given by the Kerr solution $Q_{\\rm Kerr}=-J^2/(M c^2)$ \\citep[see e.g.][for details]{1999ApJ...512..282L}. In the mean time, a considerable number of analytic exterior solutions with a more complex multipolar structure than the one of the Kerr solution have been developed \\citep[see e.g.][]{1995JMP....36.3063M,2000PhRvD..62d4048M,exactsolbook}. Whether analytic exterior solutions are accurate or not to describe the gravitational field of compact stars is an interesting and very active topic of research \\citep[see e.g.][and references therein]{2002MNRAS.336..831S,2004MNRAS.350.1416B,2006PhRvD..73j4038P}. The accuracy of analytic solutions to describe the exterior geometry of a realistic rotating compact star has been tested by comparing physical properties, e.g. the radius of the Innermost Stable Circular Orbit (ISCO) on the equatorial plane and the gravitational redshift \\citep[see][for details]{1998AstL...24..774S,2004MNRAS.350.1416B,2006PhRvD..73j4038P}. In order to do such a comparison, the free parameters (i.e. the lowest multipole moments) of the analytic exterior spacetime, are fixed to the corresponding lowest multipole moments given by numerical interior solutions of the Einstein equations, for NS realistic models \\citep[see e.g.][]{2004MNRAS.350.1416B}. Following such a procedure, the solution of \\citet{2000PhRvD..62d4048M} has been compared by \\citet{2002MNRAS.336..831S} and by \\citet{2004MNRAS.350.1416B} with the numerical solutions for NSs calculated by \\citet{1994ApJ...424..823C} and with those derived by \\citet{2004MNRAS.350.1416B}, respectively. However, being a generalization of the solution of \\citet{1972PhRvL..29.1344T}, it cannot describe slowly rotating compact stars \\citep[see e.g.][]{2004MNRAS.350.1416B}, but the dynamics of astrophysical objects with anisotropic stresses \\citep[see][for details]{2007PhRvD..75b3008D}. Following a similar procedure, based on tests of the ISCOs radii on the equatorial plane of the rotating neutron stars obtained by \\cite{2004MNRAS.350.1416B}, it has been shown that the six-parametric solution of \\cite{2006PhRvD..73j4038P} (hereafter PRS solution, see Sec.~\\ref{sec:2} for details) is more accurate than the model of \\cite{2000PhRvD..62d4048M}. In addition, being a generalization of the Kerr solution, this solution can be used for arbitrary rotation rates. Besides the ISCOs radii, there are additional physical properties that can be computed with analytic and numerical models and thus useful to compare and contrast the accuracy of analytic exact models. The aim of this article is to analyze the properties of orbital frequencies of neutral test particles in the PRS and in the Kerr geometries with especial focus on the Keplerian $\\nu_{\\rm K}$, frame-dragging (Lense-Thirring) $\\nu_{\\rm LT}$, as well as the precession(oscillation) frequencies of the radial and vertical motions, $\\nu^{\\rm P}_{\\rho}$($\\nu^{\\rm OS}_{\\rho}$) and $\\nu^{\\rm P}_{z}$($\\nu^{\\rm OS}_{z}$), respectively. The relevance of these frequencies relies on the fact that they are often invoked to explain the Quasi-Periodic Oscillations (QPOs) observed in some relativistic astrophysical systems such as Low Mass X-ray Binaries (LMXBs), binary systems harboring either a NS or a black hole (BH) accreting matter from a companion star. For instance, within the Relativistic Precession Model (RPM) introduced by \\cite{1998ApJ...492L..59S,1999ApJ...513..827M,1999ApJ...524L..63S,1999PhRvL..82...17S}, the kHz QPOs are interpreted as a direct manifestation of the modes of relativistic epicyclic motion of blobs arising at various radii $r$ in the inner parts of the accretion disk around the compact object (see Sec.~\\ref{sec:RPM}, for details). In addition to the RPM, the Keplerian, precession and oscillation frequencies are used in other QPO theoretical models \\citep[see e.g.][for a recent comparison of the existing models]{2011ApJ...726...74L}. Due to the influence of general relativistic effects in the determination of such frequencies, an observational confirmation of any of the models might lead to an outstanding test of general relativity in the strong field regime. In this line, it is of interest to compare and contrast the orbital frequencies given by the Kerr solution and by the PRS solution (see Sec.~\\ref{sec:3}), which help to establish the differences between possible BH and NS signatures. We emphasize in this article the major role of the quadrupole moment as well as of the octupole moment of the object, whose possible measurement can be used as a tool to test the no-hair theorem of BHs \\citep[see e.g.][]{2011ApJ...726...11J} and to discriminate between the different theoretical models proposed to explain the physics at interior and exterior of the Neutron Stars. Additionally, in the case of NSs, the interpretation of QPOs as the manifestation of orbital motion frequencies might lead to crucial information of the NS parameters such as mass, angular momentum \\citep[see e.g.][]{1998ApJ...492L..59S,2010ApJ...714..748T}, and quadrupole moment \\citep[see e.g.][]{1999ApJ...513..827M}. These parameters reveal, at the same time, invaluable information about the EoS of nuclear matter. The article is organized as follows. In Sec.~\\ref{sec:2} we recall the properties of the PRS solution. The computation of the orbital frequencies as well as the comparison of their features in the Kerr and in the PRS spacetimes, is shown in Sec.~\\ref{sec:3}. In Sec.~\\ref{sec:4} we study the accuracy of the analytic formulas of the periastron and nodal frequencies derived by \\cite{1995PhRvD..52.5707R} for stationary axially symmetric spacetimes. \\textcolor{black}{In Sections 5 and 6 we discuss the accuracy of the PRS solution in describing the frequencies of realistic NS models and its relevance in the Relativistic Precession Model, respectively.} The conclusions of this work and a discussion on possible additional effects to be accounted for in the determination of the orbital frequencies, e.g. the effect of magnetic dipole moment, are outlined in Sec.~\\ref{sec:6}. \\section[]{The PRS analytic exact solution}\\label{sec:2} We first recall the PRS analytic model \\citep{2006PhRvD..73j4038P}, for the exterior gravitational field of a compact object\\footnote{Mathematica 8.0 scripts with the solution, some limiting cases as well as the the calculations presented in this paper are available at \\href{http://www.chem.utoronto.ca/~lpachon/scripts/nstars} {http://www.chem.utoronto.ca/$\\sim$lpachon/scripts/nstars}}. In the stationary axisymmetric case, the simplest form of the metric can be written as \\citep{1953AnP...447..309P} \\begin{equation} \\label{Papapetrou} ds^2=-f(dt-\\omega d\\phi)^2+f^{-1}\\left[e^{2\\gamma} (d\\rho^2+dz^2)+\\rho^2d\\phi^2 \\right]\\, , \\end{equation} where $f$, $\\omega$ and $\\gamma$ are functions of the quasi--cylindrical Weyl coordinates $(\\rho,z)$. Thus, the components of the metric tensor $g_{\\mu\\nu}$ are \\begin{align} g_{\\phi \\phi} &= \\frac{\\rho^2}{f(\\rho,z)} - f(\\rho,z) \\omega(\\rho,z)^2, \\\\ g_{tt} &= -f(\\rho,z), \\\\ g_{t\\phi} &= f(\\rho,z) \\omega(\\rho,z), \\\\ g_{zz} &= g_{\\rho \\rho} = \\frac{{\\rm e}^{2\\gamma(\\rho,z)}}{f(\\rho,z)} = \\frac{1}{g^{zz}} = \\frac{1}{g^{\\rho\\rho}}. \\end{align} Using the above line element, the Einstein-Maxwell equations can be reformulated, via Ernst's procedure in terms of two complex potentials ${\\cal E}(\\rho,z)$ and $\\Phi(\\rho,z)$ \\citep{1968PhRv..167.1175E,1968PhRv..172.1850E}. By means of Sibgatullin's integral method \\citep{1991owsg.book.....S,1993CQGra..10.1383M} this system of equations can be solved va \\begin{align} \\label{Ernst1} {\\cal E}(z,\\rho)&=\\int\\limits_{-1}^1\\frac{d\\sigma}{\\pi} \\frac{e(\\xi)\\mu(\\sigma)}{\\sqrt{1-\\sigma^2}}, \\\\ \\Phi(z,\\rho)&=\\int\\limits_{-1}^1 \\frac{d\\sigma}{\\pi} \\frac{f(\\xi)\\mu(\\sigma)}{\\sqrt{1-\\sigma^2}}, \\end{align} where $e(z):={\\cal E}(z,\\rho=0)$ and \\mbox{$f(z):=\\Phi(z,\\rho=0)$}. The unknown function $\\mu(\\sigma)$ must satisfy the singular integral equation \\begin{equation} \\dashint_{-1}^{1}\\frac{\\mu(\\sigma)[e(\\xi)+\\tilde e(\\eta)+2f(\\xi)\\tilde f(\\eta)]d\\sigma}{(\\sigma-\\tau)\\sqrt{1-\\sigma^2}}=0 \\end{equation} and the normalizing condition \\begin{equation} \\int_{-1}^1\\frac{\\mu(\\sigma)d\\sigma}{\\sqrt{1-\\sigma^2}}=\\pi, \\end{equation} where $\\xi=z+i\\rho\\sigma$, $\\eta=z+i\\rho\\tau$, $\\rho$ and $z$ being the Weyl-Papapetrou quasi--cylindrical coordinates, $\\sigma, \\tau\\in[-1,1]$, $\\tilde e(\\eta):=\\overline{e(\\bar\\eta)}$, $\\tilde f(\\eta):=\\overline{f(\\bar\\eta)}$ and the overbar stands for complex conjugation. In \\citep{2006PhRvD..73j4038P}, the Ernst potentials were chosen as \\begin{align} \\label{Potenciales eje} \\begin{split} e(z) = \\frac{z^3-z^2(m+\\textrm{i}a)-kz+\\textrm{i}s}{z^3+z^2(m-\\textrm{i}a)-kz+\\textrm{i}s}\\, , \\\\ f(z) = \\frac{q z^2 + \\textrm{i}\\mu z}{z^3+z^2(m-\\textrm{i}a)-kz+\\textrm{i}s}\\, . \\end{split} \\end{align} We calculate the multipole moments following the procedure of \\cite{1990CQGra...7.1819H}. We denote the mass multipoles by $M_i$ while, the current (rotation) multipoles, by $S_i$. The electric multipoles are denoted by $Q_i$ and the magnetic ones by ${\\cal B}_i$. Thus, for the PRS solution we have \\begin{align} \\begin{split} \\label{multipolosP} M_0 &= m\\, ,\\quad M_2 = m k - m a^2\\, ,\\quad \\ldots \\\\ S_1 &= m a\\, ,\\quad S_3 = - m a^3 + 2 m a k - m s\\, ,\\quad \\ldots \\\\ \\end{split} \\\\ \\begin{split} \\label{multipolosQ} Q_0 &= q\\, ,\\quad Q_2 = - a^2 q - a \\mu + k q\\, ,\\quad \\ldots \\\\ {\\cal B}_1 &= \\mu + a q\\, ,\\quad {\\cal B}_3 = - a^2 \\mu + \\mu k - a^3 q + 2 a k q - q s\\, ,\\quad \\ldots \\end{split} \\end{align} This allows us to identify $m$ as the total mass, $a$ as the total angular moment per unit mass ($a=J/m$, being $J$ the total angular moment); while $k$, $s$, $q$ and $\\mu$ are associated to the mass-quadrupole moment $M_2$, current octupole $S_3$, electric charge and magnetic dipole, respectively. The potentials (\\ref{Potenciales eje}) can be written in an alternative way, we mean \\begin{equation} e(z)=1+\\displaystyle \\sum_{i=3}^{3} \\frac{e_{i}}{z-\\beta_{i}}\\, ,\\qquad f(z)=\\displaystyle \\sum_{i=3}^{3} \\frac{f_{i}}{z-\\beta_{i}}\\, , \\end{equation} with \\begin{align} e_{j}&= (-1)^{j}\\frac{2 m \\beta^{2}_{j}}{(\\beta_{j}-\\beta_{k})(\\beta_{j}-\\beta_{i})}\\, , \\\\ \\label{equ:f_i} f_{j}&= (-1)^{j+1}\\frac{i \\mu + d\\beta_{j} }{(\\beta_{j}-\\beta_{k})(\\beta_{j}-\\beta_{i})}\\, , \\quad i,k \\neq j\\, . \\end{align} Then, using Eqs.~(\\ref{Ernst1}) and (\\ref{Potenciales eje}), we obtain the Ernst potentials \\begin{equation}\\label{potenciales_ernst} {\\cal E}=\\frac{A + B }{A - B}\\, , \\qquad \\Phi=\\frac{C}{A - B}\\, , \\end{equation} and the metric functions in the whole spacetime \\begin{align}\\label{eq:metricfuncs} f&=\\frac{A \\bar{A}-B \\bar{B} + C \\bar{C}}{( A - B)(\\bar{A}-\\bar{B})}\\, ,\\quad e^{2\\gamma}=\\frac{A \\bar{A} -B \\bar{B} + C \\bar{C}}{\\displaystyle{K \\bar{K}\\prod_{n=1}^{6}r_n}}\\, ,\\\\ \\omega &= \\frac{{\\rm Im}[(A + B)\\bar{H}-(\\bar{A} + \\bar{B})G - C \\bar{I}]}{A \\bar{A} - B \\bar{B} + C \\bar{C}}\\, , \\end{align} where the functions $A$, $B$, $C$, $H$, $G$, $K$, and $I$ can be found in the Appendix \\ref{app:metricfuncs}. The PRS electrovacuum exact solution belongs to the extended $N$-soliton solution of the Einstein-Maxwell equations derived by \\cite{1995PhRvD..51.4192R}, in the particular case $N=3$. In addition, the functional form of the metric functions resembles the one derived previously by \\cite{1999CQGra..16.3725B}. Besides the limiting cases discussed in \\cite{2006PhRvD..73j4038P} it is worth mentioning that, in the vacuum case $q=0$ and $\\mu=0$, for $s=0$ this solution reduces to the solution of \\cite{1995JMP....36.3063M} under the same physical conditions, namely $q=0$, $c=0$ and $b=0$ in \\cite{1995JMP....36.3063M}. ", "conclusions": "\\label{sec:6} We have done an extensive comparison of the orbital motion of neutral test particles in the PRS and Kerr spacetime geometries. In particular we have emphasized on the Keplerian and frame-dragging frequencies, as well as the precession and oscillation frequencies of the radial and vertical motions. \\textcolor{black}{We have evidentiated the differences in this respect between the Kerr and PRS solution, especially in the rapid $\\sim$kHz rotation regime. Such differences are the manifestation of the influence of the high order multipole moments such as the quadrupole and octupole.} \\textcolor{black}{The analysis of the deviations between the Kerr and PRS features for given mass and angular momentum of a source studied in this work are useful to distinguish the signatures between BHs and NSs, which relevant to establish a separatrix for the identification of the compact objects harbored in X-Ray Binaries. In the case of BH candidates, these results might become important for testing the no-hair theorem of BHs \\citep[see e.g.][]{2011ApJ...726...11J}. Equally important, the application of the precession and oscillation frequencies to the explanation of QPOs in LMXBs possessing a NS, can unveil information on the NS parameters, leading to a possible identification of the behavior of the nuclear matter EoS at supranuclear densities. In this line, the identification of the rotation frequency of NSs in LMXBs from the pulsating X-ray flux $\\nu_{\\mathrm{burst}}$, e.g.~the case of 4U 1728--34 \\citep{1998ApJ...506L..39F}, 4U 1916--053 \\citep{2001ApJ...549L..85G} and more recently the case of IGR J17191--2821 \\citep{2010MNRAS.409.1136A}, will certainly help to constrain QPO models as well as the NS parameters. Additional information coming from the photospheric radius expansion phenomena observed in these systems \\citep[see e.g.][for details]{2001ApJ...553L.157M} during their transient activity with Super-Eddington emission can become of paramount importance if combined with the QPO information.} The generalization of the present work to the electrovacuum case is important to establish the influence of the magnetic dipole and quadrupole moments on the orbital motion of particles around compact objects \\citep[see e.g.][]{2010CQGra..27d5001B,2010PhRvD..82l4014S,2012CQGra..29f5012B}. Interesting effects on the epicyclic frequencies due to the presence of the magnetic dipole have been already pointed out recently by \\cite{2010CQGra..27d5001B} and \\cite{2012CQGra..29f5012B}. These effects were predicted after neglecting the contribution of the electromagnetic field to the curvature, for $j=0$ see \\cite{2010CQGra..27d5001B} and for $j\\neq0$ \\cite{2012CQGra..29f5012B}. \\cite{2010CQGra..27d5001B} assumed the model of the star as a dipole magnetic field superposed on a Schwarzschild black hole. In the second work, they studied the case of a magnetized slowly rotating neutron stars; to build the model they superpose an dipolar magnetic field on the Lense-Thirring geometry. The effects of the magnetic dipole on the location of the ISCO, within the PRS solution, has been investigated by \\cite{2010PhRvD..82l4014S}. A complete analysis of the effects due to the emergence of electromagnetic structure on the orbital motion of charged particles is therefore of interest and deserve the appropriate attention. Recent observations have shown that for stars with strong magnetic fields the quadrupole and octupole magnetic terms make significant contributions to the magnetic field \\citep{2006MNRAS.370..629D}, which indicates that arbitrary higher order multipole components might be required in a realistic model. The presence of a magnetic quadrupole demands the breaking of the reflection symmetry \\citep[see] [for details]{2006CQGra..23.3251P}, by means of a slightly change to the Ernst electric potential over the symmetry axis \\begin{align} \\label{PotencialesPRSmod} \\begin{split} f(z) = \\frac{q z^2 + \\textrm{i}\\mu z + \\textrm{i} \\zeta}{z^3+z^2(m-\\textrm{i}a)-kz+\\textrm{i}s}\\, , \\end{split} \\end{align} a quadrupolar magnetic component $\\mathcal{B}_2 = \\zeta$ can be introduced to the PRS solution. Such a change generates just a redefinition of the coefficients $f_i$ in Eq.~(\\ref{equ:f_i}). In this way the PRS solution can be readily use to explore the effect of strong magnetic fields with non-dipolar structure." }, "1112/1112.2711_arXiv.txt": { "abstract": "We present a new census of the Orion Nebula Cluster (ONC) over a large field of view ($\\gtrsim30^\\prime\\times30^\\prime$), significantly increasing the known population of stellar and substellar cluster members with precisely determined properties. We develop and exploit a technique to determine stellar effective temperatures from optical colors, nearly doubling the previously available number of objects with effective temperature determinations in this benchmark cluster. Our technique utilizes colors from deep photometry in the $I$-band and in two medium-band filters at $\\lambda\\sim753$ and $770$~nm, which accurately measure the depth of a molecular feature present in the spectra of cool stars. From these colors we can derive effective temperatures with a precision corresponding to better than one-half spectral subtype, and importantly this precision is independent of the extinction to the individual stars. Also, because this technique utilizes only photometry redward of 750~nm, the results are only mildly sensitive to optical veiling produced by accretion. Completing our census with previously available data, we place some 1750 sources in the Hertzsprung-Russel diagram and assign masses and ages down to 0.02 solar masses. At faint luminosities, we detect a large population of background sources which is easily separated in our photometry from the bona fide cluster members. The resulting initial mass function of the cluster has good completeness well into the substellar mass range, and we find that it declines steeply with decreasing mass. This suggests a deficiency of newly formed brown dwarfs in the cluster compared to the Galactic disk population. ", "introduction": "\\label{section:introduction} Understanding the initial mass function (IMF) is one of the most important problems of stellar astrophysics. The IMF, together with the star formation history, dictates origin, evolution, and fate of the stellar populations, from individual clusters up to entire galaxies. The distribution of stellar masses has been studied in depth for more than half a century, starting with the pioneering work of \\citet{salpeter55}; the major open question is whether the IMF is universal or if it depends on the initial conditions of star formation \\citep[see, e.g.,][]{kroupascience,bastian2010}. Whereas it is commonly accepted that the IMF seems well reproduced by a power law for masses greater than several tenths of a solar mass, the shape and universality of the IMF in the substellar mass regime is still under investigation \\citep[e.g.,][]{Wang2011}. Very young clusters (few Myr old) in star-forming regions provide a unique tool to investigate the IMF across the entire mass spectrum, for a number of reasons. First, they are usually young enough that neither dynamical processes nor stellar evolution have altered the mass distribution; therefore the measured distribution of stellar masses coincides with the IMF. Moreover, young low-mass stars and brown dwarfs (BDs) are in their brightest evolutionary stage when they contract towards the main sequence, therefore these objects are more easily detected and characterized when they are young. Among the nearby Galactic star-forming regions, the Orion Nebula Cluster (ONC) is an ideal site for the study of star formation in particular for low-mass stars and BDs. This cluster counts a few thousand members, 1-3~Myr old \\citep{hillenbrand97,paperII}, spanning the entire mass spectrum ($M\\lesssim50$~M$_\\odot$). Several studies have been conducted in the past decade to measure the IMF in the ONC \\citep[e.g.,][]{hillenbrand97,hillenbrand-carpenter2000,muench2000,muench2002,lucasroche2000,lucas2005,slesnick04,paperII}; they generally find (similar to the field star population) a Salpeter-like slope above 1~M$_\\odot$, which flattens to a broad peak at 0.2-0.3~M$_\\odot$ (though the shape and position of the IMF peak is highly model dependent, \\citealt{paperII}), and the mass distribution likely decreases in the substellar mass range. Determining masses in a young region such as the ONC is not without difficulties: besides the strong nebular emission, another major impediment is caused by differential reddening, which has strongly limited the ability to derive the stellar parameters of individual sources based on photometry alone. To overcome this shortcoming, spectroscopic surveys have been carried out \\citep{hillenbrand97,lucas2001,lucas2006,slesnick04,riddick2007,weights2009}, but they are either limited to a fraction of the members, or to the very central part of the region, the Trapezium cluster. Using near-infrared (NIR) photometry it is possible to asses stellar masses down to the planetary masses ($M<13$~M$_J$), for example by de-reddening the measured color-magnitude diagrams (CMDs) on one isochrone or even to deriving directly the IMF from the NIR luminosity functions (LFs) \\citep{hillenbrand-carpenter2000,muench2002}. This second approach, however, is not very accurate: the ONC shows a significant luminosity spread, sometimes interpreted as a real age spread \\citep{reggiani2011}, and stars of different age follow different mass-luminosity relation. Moreover the NIR excess originating in the inner circumstellar disks \\citep{meyer97} alters the observed fluxes. Clearly, in order to improve our knowledge of the ONC IMF, a precise and systematic characterization of the stellar parameters of individual members is needed. To this purpose, the \\emph{Orion HST Treasury Program} \\citep{robberto2005treasury} has produced a high spatial resolution photometric survey of the ONC with three instruments onboard the Hubble Space Telescope, from $U-$band to the NIR, over a large field of view (FOV), with a sensitivity well into the BD mass range. To fully exploit this exceptional dataset, an accurate estimate of the effective temperature ($\\teff$) is needed to derive $A_V$ from the observed colors and therefore the stellar luminosities. In our previous works \\citep[][hereafter Paper~I and Paper~II]{paperI,paperII} we have collected spectral types from the literature and complemented them with new optical spectroscopy. Moreover, we have presented a new observational technique, based on optical medium-band photometry at 6200\\AA---a wavelength where the spectra of cool stars show a deep, $\\teff$-dependent TiO absorption feature---to derive the spectral types of M-type sources. This allowed us to obtain the stellar parameters of $\\sim1000$ members and derive the most complete HRD of the ONC, down to the hydrogen burning limit. In the present paper we extend our investigation to lower masses, well into the BD regime, and with higher completeness. To this purpose, we take advantage of the same observational strategy presented in Paper~I, using medium-band photometry to derive spectral types of cool sources, from which $\\teff$ and $A_V$ are derived, and place individual sources in the HRD. Here, instead of the $6200\\AA$ filter used in Paper~I, we select two bands at longer optical wavelengths, in order to increase our sensitivity for cooler sources such as BDs and further reduce the influence of extinction. In Section \\ref{section:the_data} we present our new observations, the data reduction and calibration. In Section \\ref{section:analysis} we define two spectro-photometric indices based on our medium-band photometry. Using available stellar parameters for a small sample of ONC very-low mass stars and BDs, we define empirical transformations to convert these indices into $\\teff$ and $A_V$. We also show that this technique is not significantly affected by optical excesses from mass accretion, typical of young stars and BDs. In Section \\ref{section:HRD} we derive the new HRD for the ONC, which now includes $\\sim 1800$ sources, reaching masses as low as $0.02~M_\\odot$. We study the completeness as a function of $\\teff$ and bolometric luminosity ($L_{\\rm bol}$) accounting for photometric detection as well as differential reddening; we detect a population of candidate background contaminants, which appears well separated from the ONC members in the HRD, at lower luminosity. After excluding the contaminants, in Section \\ref{section:IMF} we derive and discuss the IMF of the ONC. ", "conclusions": "" }, "1112/1112.0008_arXiv.txt": { "abstract": "We measure the real-space galaxy power spectrum on large scales at redshifts 0.5 to 1.2 using optical colour-selected samples from the CFHT Legacy Survey. With the redshift distributions measured with a preliminary $\\sim$14000 spectroscopic redshifts from the VIMOS Public Extragalactic Redshift Survey (VIPERS), we deproject the angular distribution and directly estimate the three-dimensional power spectrum. We use a maximum likelihood estimator that is optimal for a Gaussian random field giving well-defined window functions and error estimates. This measurement presents an initial look at the large-scale structure field probed by the VIPERS survey. We measure the galaxy bias of the VIPERS-like sample to be $b_g=1.38 \\pm 0.05$ ($\\sigma_8=0.8$) on scales $k<0.2\\hmpc$ averaged over $0.5<{\\rm z}<1.2$. We further investigate three photometric redshift slices, and marginalising over the bias factors while keeping other \\LCDM~ parameters fixed, we find the matter density $\\Omega_m=0.30\\pm0.06$. ", "introduction": "The shape of the galaxy clustering power spectrum encodes the dynamical history of the Universe under the influence of baryons, dark matter and dark energy. On large scales the assumption of Gaussianity can be made, and the statistic summarises all of the cosmological information that is available. Measurements of the power spectrum at $\\z\\sim0$ have led to fundamental tests of the \\LCDM~model \\citep{Tegmark04b,Efstathiou02}. The angular distribution of galaxies on the sky, although less sensitive than the full three-dimensional view, has played an important role as well. Indeed, strong tests of the CDM model were made with the two-dimensional correlation function from the APM galaxy survey \\citep{Maddox90}. Photometric surveys have the capability to probe significantly larger volumes at higher sampling rates than targeted spectroscopic surveys. The advantages have become clear with the advancement of photometric redshift estimation methods. The loss in three-dimensional precision can be compensated for with increased statistics leading to strong cosmological constraints that are comparable to the results from spectroscopic surveys. Additionally, the projected density field is only weakly sensitive to redshift-space distortions, thus it provides a means to infer the real-space power spectrum directly. The dependence on peculiar velocities becomes important for narrow redshift slices and can be turned into a useful measure of the growth rate \\citep{Ross11}. Measurements of the baryon acoustic feature and redshift-space distortions have now been made on photometric samples taken from the Sloan Digital Sky Survey \\citep{Padmanabhan07,Blake07,Thomas11}. In this analysis, we present a new measurement of the real-space galaxy power spectrum using a photometric catalogue of galaxies at $0.5<\\rm{z}<1.2$ from the Canada-France-Hawaii Telescope Legacy Survey (CFHTLS) Wide survey. The survey consists of four fields covering a total area of 133 \\sqrdeg. The extent of the largest field, W1, is $\\sim10\\degr$ or $200~{\\rm h}^{-1}~{\\rm Mpc}$ at z=0.7, giving a maximum scale we may probe of $k_{min}\\sim0.05\\hmpc$. The dataset has been used for previous cosmological analyses, in particular for weak lensing \\citep{Fu08,Kilbinger09,Tereno09,Shan11} and galaxy correlation function measurements \\citep{Coupon11}. A key ingredient needed to interpret the projected density field and constrain the three-dimensional power spectrum is the redshift distribution of the galaxy sample. For this we use spectroscopy from the VIMOS Public Extragalactic Survey\\footnote{VIPERS website: \\href{http://vipers.inaf.it}{vipers.inaf.it}} (VIPERS)\\citep{VIPERS}. VIPERS is an ongoing spectroscopic program to target $10^5$ galaxies in the redshift range 0.5-1.2 in a total area of 24\\sqrdeg~in the CFHTLS W1 and W4 fields. The accuracy of the spectroscopic measurements from VIPERS provide an unbiased estimate of the redshift distribution. With this knowledge, we are confident that we can deproject the angular clustering signal and constrain the three dimensional power spectrum. The primary advantage of studying the deprojected power spectrum $P_k$ is its closeness to theory. The shape of the angular power spectrum is complicated by its dependence on survey properties, its depth and geometry. Furthermore, in projection, scales are mixed. Ideally we would like to separate the power on large scales in the linear regime from power on small scales that is influenced by complex astrophysical processes. \\begin{figure} \\includegraphics[scale=1]{plots/fig1} \\caption{Galaxy count maps in the CFHTLS-Wide fields with the VIPERS-like colour selection. We use {\\sc Healpix} cells with size 7\\arcmin. Gaps in the survey coverage are left as blank pixels. The grid overlay has spacing of 1\\degr.\\label{fig:fields}} \\end{figure} How to derive the three-dimensional power spectrum from the two-dimensional density field is a problem of deconvolution. A good inversion method should be stable against noise in the data. Preliminary work done by \\citet{Baugh93,Baugh94} and \\citet{Gaztanaga98} used the Lucy deconvolution method that is known to be robust. To further derive cosmological constraints, we must be able to estimate the covariance of the deprojection. Methods of propagating the error from the angular correlation function to the three-dimensional power spectrum were developed by \\citet{Dodelson00} who perform the inversion with a prior on the smoothness of the power spectrum and compute a covariance matrix of the estimate. Further work by \\citet{Eisenstein01}, \\citet{Dodelson02} and \\citet{Maller05} made use of the singular-value decomposition technique to remove modes that destabilise the inversion. Importantly, the deprojection method should produce well-defined window functions that describe the mode mixing. The aim is to separate the small and large scales that are mixed in projection, and the residual leakage should be understood. A similar problem was solved with the maximum likelihood methods developed for the cosmic microwave background angular power spectrum \\citep{Tegmark97,Bond98} and then later applied to galaxy surveys \\citep{Huterer01,Tegmark02}. Applications to the deprojection of the power spectrum were presented by \\citet{Efstathiou01} and \\citet{Szalay03}. In this work, we adopt the maximum-likelihood technique to construct an estimator for the power spectrum. The result is optimal under the assumption that the density is represented by a Gaussian random field. This is a reasonable assumption for the galaxy distribution on large scales. Moreover, the estimator also simultaneously gives the covariance of the estimate as well as the window functions. For small surveys, where the window functions must be handled carefully, the approach is especially useful. We pay close attention to the window functions for the results presented here. Maximum likelihood estimates are computationally expensive. However, because of the relatively small field sizes we consider, we can perform all computations on a consumer level four-core desktop computer. In this article, we first introduce the CFHTLS and VIPERS datasets used. In Sections 3 and 4 we review the angular power spectrum formalism and the maximum likelihood deprojection using a quadratic estimator. We then apply the method to Gaussian simulations and investigate potential biases due to uncertainty in the redshift distribution and the fiducial cosmology. Lastly, we measure the power spectrum with CFHTLS data and constrain the linear galaxy bias and matter density. We report magnitudes using the AB magnitude convention in the CFHT $u^*g'r'i'z'$ photometric system. We assume a flat \\LCDM~cosmology with $H_0=70.4~{\\rm km~s^{-1}~Mpc^{-1}}$, $\\Omega_m=0.272$, $\\Omega_b=0.0456$, $n_s=0.963$ and $\\sigma_8=0.8$ \\citep{Larson11}. ", "conclusions": "The CFHTLS-Wide fields probe a significant cosmological volume at redshifts not reached by other galaxy surveys to date. We use the projected density field from photometric redshift samples to constrain the real-space power spectrum and derive constraints on the matter density and linear galaxy bias factors. These results are made possible by precise knowledge of the redshift distributions provided by preliminary results from the VIPERS survey. The primary advantage of computing the power spectrum directly from the angular distribution, instead of using conventional spherical harmonics $\\Cl$ is that we may construct window functions in Fourier space. By optimising this, we achieve sharper constraints on the power spectrum than when we are limited to $\\ell$ bands. This approach comes with the cost that we must adopt a fiducial power spectrum. We showed that using the wrong fiducial power spectrum, although leading to sub-optimal weights, does not significantly bias the estimate. This is true even on small scales, and we can effectively use this method to deconvolve small and large scales in Limber's equation. Residual systematic error on the derived power spectrum is at the 1\\% level, well below the sensitivity of the measurement. The deprojection does strongly depend on the assumed redshift distribution of the galaxy sample as well as the cosmology used to compute the redshift-distance relation. The cosmology dependence of the measurement makes the interpretation difficult, but to a first approximation, only the amplitude is affected; the shape of the power spectrum is recovered correctly. Thus, a converging iterative procedure can be implemented by updating the fiducial model and repeating the analysis. There is a degeneracy between a shift in the assumed redshift distribution and the cosmological model. This is unavoidable when studying a field in projection. However, the constraints on the redshift distribution can always be improved with further observation. In our analysis, from the sampling biases present in the VIPERS spectroscopy, we estimate the uncertainty in the mean redshift to be at the 1\\% level. Thus, we do not expect a strong systematic error in the derived galaxy bias parameters. We do note that the observed trend of low $\\chi^2$ values for the best-fit models in the higher redshift samples can arise if the covariance is overestimated. This could be a weak hint that the true mean redshift is lower than what we assume or that a modification is needed in the fiducial cosmology. Recently, the galaxy bias was measured from the CFHTLS-Wide fields in the context of the halo model by \\citet{Coupon11}. Our final two photometric redshift bins, S7 and S8, correspond with samples constructed by Coupon \\etal so we are able to compare the resulting bias values. Coupon \\etal constructed volume-limited samples using luminosity cuts resulting in a selection of brighter galaxies, thus we may expect their bias values to be larger. The halo model constraints of \\citet{Coupon11} give for S7, $b_g=1.44\\pm0.01$, and for S8, $b_g=1.79\\pm0.03$. These values have been scaled by 1.03 to transform from a cosmology with $\\Omega_m=0.25$ to $\\Omega_m=0.272$ which is assumed here. Our value of $b_g$ for the S7 sample is higher, while for the S8 sample it is lower, although both are in agreement with Coupon \\etal within the 2$\\sigma$ confidence limit. The measurements are based on different physical scales (Coupon \\etal restrict the correlation function to angular scales $<1.5\\degr$) and different model assumptions have been used. Thus it is reasonable to consider the measurements as independent estimates. Our results provide a preliminary look at the large-scale structure field probed by the VIPERS colour selection and demonstrate the strengths of the VIPERS sample for clustering studies at $\\z>0.5$. We anticipate promising results with the full VIPERS spectroscopic sample." }, "1112/1112.2657_arXiv.txt": { "abstract": "We perform a pixel-by-pixel analysis of 467 galaxies in the GOODS-VIMOS survey to study systematic effects in extracting substructure and properties of stellar populations (age, dust, metallicity and star formation history) from the pixel colors using the pixel-$z$ method. Several systematics are examined in this paper, including the effect of the input stellar population synthesis models whose SEDs are fitted to each pixel's colors, the effect of passband limitations and differences between the individual SED fits to pixels and global SED fitting to a galaxy's colors. We find that with optical-only colors ($bviz$), the systematic uncertainties due to differences among stellar population synthesis codes are well constrained. The largest impact on the stellar population age and SFR e-folding time estimates in the pixels arises from differences between the Maraston~(2005) models on one hand and the Bruzual \\& Charlot~(2003) and Charlot \\&Bruzual~(2007) models on the other, when optical-only ($bviz$) colors are used. This results in systematic differences larger than the $2\\sigma$ uncertainties in over 10 percent of all pixels in the galaxy sample. The effect of varying the number and choice of available passbands is more severe. In 26 percent of the pixels in the full sample, these limitations result in systematic biases in the age determination which are larger than the $2\\sigma$ uncertainties in the measurements. Robust results can, however, still be obtained with a minimum of 3 optical filters provided they span the 4000\\,\\AA \\ break. Near-IR data is also added to a subsample of 46 galaxies from the GOODS-NICMOS survey and systematics arising from model differences are again investigated. Differences among the models in their predicted rest-frame red/NIR colors manifest themselves as follows. For $z > 1$ galaxies the observed optical/NIR colors span the rest frame UV-optical SED, and the use of different population synthesis models does not significantly bias the estimates of the stellar population parameters compared to using optical-only colors. However, for $z < 1$, where the rest-frame NIR is still probed, there is a larger discrepancy between models when using optical-only colors compared to optical/NIR colors. This affects in particular the age determination in the pixels. With this characterization of the systematic errors, we illustrate how pixel-$z$ can be applied robustly to make detailed studies of substructure in high redshift objects such as (a) radial gradients of properties such as age, SFR and dust and (b) the distribution of these properties within subcomponents such as spiral arms and clumps. Finally, we show preliminary results from applying pixel-$z$ to galaxies in the CANDELS survey, illustrating how the new HST/WFC3 data can be exploited to probe substructure and stellar populations in $z\\sim1-3$ galaxies. ", "introduction": "} \\subsection{Clumpy Disks: Resolved Stellar Populations in $z\\sim 1-3$ Objects as a Probe of High-z Disk Galaxies} Many galaxies at $z\\sim1-3$ have morphologies that are very different from the galaxy population at $z\\sim0$ (Lotz~et~al.~2004; Lotz~et~al.~2006; Law~et~al.~2007; Scarlata~et~al.~2007; Elmegreen~et~al.~2007; Pannella~et~al.~2009). However, the morphologies of these objects only tell one part of the story. At $z>1$, rest-frame UV wavelengths, which are most impacted by dust obscuration and star formation, can be traced by high resolution optical imaging (Toft~et~al.~2007; Overzier~et~al.~2010; Cameron~et~al.~2010). The location and distribution of stellar populations within these galaxies, particularly sites of active star formation, is likely to be affected by various physical and dynamical mechanisms which are present in earlier epochs in the galaxy's evolution. The pixel-$z$ technique (Conti~et~al.~2003; Welikala~et~al.~2008; Welikala~et~al.~2009) can measure spatially resolved stellar populations for large samples of $z\\sim1-3$ galaxies using their multi-band HST images. With this approach, one can study asymmetries that exist in the star formation, age and dust distribution across these galaxies. In parallel with the detailed kinematical studies of these objects from spatially resolved spectroscopy, this will enable insights into the formation mechanisms of these galaxies. One such mechanism that could be responsible for the formation of these clumps on kiloparsec scales is disk fragmentation in Toomre-unstable gas-rich disks (Elmegreen~et~al.~2007, Elmegreen~et~al.~2009). There is now direct kinematic evidence that high-z disks are much more turbulent than their counterparts at $z \\sim 0$ (Bournaud~et~al.~2007; Bournaud~et~al.~2008; Genzel~et~al.~2008; Genzel~et~al.~2011). For example, high-redshift disks have been found to have local intrinsic gas velocity dispersions of $20-90 \\,$km\\,s$^{-1}$ as well as high gas-to-total baryonic mass fractions (Daddi~et~al.~2010; Tacconi~et~al.~2010). Numerical simulations of gas-rich turbulent disks indicate that massive kpc-sized clumps can form in-situ through gravitational instabilities, a phenomenon known as a `clump-cluster' phase. (Noguchi~et~al.~1999; Immeli~et~al.~2004a,b; Bournaud et al.~2007; Elmegreen et al.~2008; Dekel, Sari, \\& Ceverino~2009; Agertz~et~al.~2009; Ceverino~et~al.~2010; Aumer et~al.~2010; Genzel~et~al.~2010). According to these simulations, the clumps can migrate towards the gravitational center as a result of both their mutual interactions and dynamical friction against the host disk, so that they can eventually coalesce into a young bulge on the order of a few dynamical times. Alternatively, these high-redshift clumpy disk and irregular galaxies may form via mergers. In particular, numerical simulations of gas-rich mergers result in remnant disks that show a low velocity field asymmetry that also satisfies the criteria necessary to be classified as a high-redshift disk galaxy observationally. For example, Robertson \\& Bullock (2008) compared one of the merger remnants to the bulk properties of the well-studied $z=2.38$ galaxy BzK-15504 and showed that it has a star formation rate (SFR), gas surface density, and a circular velocity-to-velocity dispersion ratio that is in excellent agreement with BzK 15504. Such numerical simulations suggest that gas-rich mergers can play a prominent role in the formation of disk galaxies at high redshift. Recent and upcoming larger IFU programs such as the SINS survey (F\\\"{o}rster-Schreiber et al.~2006, Genzel et al.~2006), the SAURON project (Bacon et al.~2001) and the MASSIV survey (Contini et al.~2011; Epinat et al.~2011; Queyrel et al.~2011) are expected to shed further light on these formation mechanisms by performing detailed studies of the gas kinematics of $z \\sim 2$ galaxies. Other large IFU programs include the CALIFA project (Sanchez et al.~2011), which will target approximately 600 objects with PMAS IFUs on the Calar Alto Telescope, and the SAMI program (Croom et al.~2011) which is a planned low redshift survey of serveral thousand galaxies. In addition to kinematic studies, a complementary study of the spatial distribution of the stellar populations within these galaxies can provide additional insights into whether these objects are formed through mergers or are, in fact, disks formed in-situ. F\\\"{o}rster-Schreiber et al.~(2011a, b) used a combination of near-infrared integral field spectroscopy from the SINFONI instrument on the VLT, combined with deep high resolution HST NIC2/F160W imaging of six $z \\sim 2$ star-forming galaxies to characterize the properties of kpc-scale clumps and their contribution to the rest-frame optical emission. In our next papers, we aim to extend these previous studies and make detailed spatially-resolved measurements of the properties of large samples of disk galaxies. These studies will also make use of new fully reduced NIR data from the HST Wide Field Camera 3 (WFC3) which will become available for the GOODS-South field in early 2012. \\subsection{Spatially Resolved Star Formation and The Role of Environment in Galaxy Evolution} The relation between galaxy environment and galaxy properties have been extensively studied in the local Universe (Gomez et al.~2003, Blanton et al.~2005, Hogg et al.~2004). A number of studies have observed that the high-$z$ ($z\\gtrsim 1$) SFR-density relation is either reversed or weaker at $z \\sim 1$ than that seen locally (Cucciati et al.~2006; Cooper et al.~2007; Elbaz et al.~2007; Ilbert et al.~2006; Poggianti et al.~2006; Ideue et al.~2009; Salimbeni et al.~2009; Scodeggio et al.~2009; Tran et al.~2010; Gr\\\"{u}tzbauch et al.~2011). These examples are consistent with the scenario that galaxies in dense environments form stars rapidly at early times, quickly building up mass and becoming quiescent, while galaxies in less dense environments form stars at a more sedate pace but over longer timescales, a phenomenon we refer to as {\\em in-situ\\/} evolution (Wijeshinge et al.~2011). This is distinct from the \u2018\u2018infall and quench\u2019\u2019 scenario where the properties of galaxies are impacted directly by their local environment through physical mechanisms such as ram-pressure stripping and galaxy harassment. In addition, Peng at al.~(2010) makes a distinction between this type of quenching which is directly related to the environment of galaxies and {\\em mass quenching} which dominates at high masses and early cosmic times. A few recent studies have, however, continued to observe the same SFR-density relation upto $z\\sim1$ as seen locally (Patel et al.~2011). A few studies have investigated the relation beyond $z\\sim1$. Quadri et al.~(2011) showed that galaxies with quenched SF tend to reside in dense environments out to at least $z \\sim 1.8$ and that the the SFR-density relation holds even at fixed stellar mass. Their density estimates, however, have relatively high uncertainties as they are derived from photometric redshifts. While the SFR-density relation at high redshifts is a matter of some debate, it is clear that an {\\em in-situ\\/} evolution makes a distinct prediction compared to the \u2018\u2018infall and quench\u2019\u2019 models. The latter suggests that galaxies in dense environments should show a SFR distribution that is progressively suppressed from the outside in, as the outer regions are those which will be affected first by their rapidly changing environment. The former, on the contrary, suggests that the suppression should either happen uniformly as a galaxy ages, or that the inner regions should be suppressed first. By studying the spatial distribution of SFR in star-forming galaxies as a function of environment, we should be able to distinguish clearly between these two scenarios. Studies by Welikala et al.~(2008; 2009) at $z \\sim 0.1$ confirm the global SFR-density relation observed locally but also show that the SFR at $z \\sim$ 0.1 is centrally concentrated relative to the outskirts, and in more dense environments it is this central SFR that is supressed, favouring the {\\em in-situ\\/} evolution scenario. Park et al.~(2007) also studied the color gradients of galaxies as a function of the local galaxy density and galaxy morphology in the SDSS. They found no environmental dependence of the color gradient for early-type galaxies at a given luminosity, and only a weak dependence on environment for faint late-type galaxies which are observed to become bluer in their outskirts (relative to the galaxy center) in low density environments. It thus becomes extremely interesting to determine whether the radial variation of star formation with environment seen locally is also in place at high redshifts. The development of MOS IFU instruments and surveys with large photometric and spectroscopic samples, such as the Galaxy And Mass Assembly (GAMA) survey (Driver et al.~2011), will allow such a statistical analysis for a large sample of galaxies. There are two complementary approaches to this. While spatially resolved spectroscopy provides detailed measures of the current spatial distribution of physical properties within galaxies, pixel-$z$ adds the ability for the star formation history to be probed. While pixel-$z$ also provides the current SFR and other properties of the stellar populations in subcomponents of galaxies, these measurements will have higher uncertainties than spectroscopic measures, and cannot provide detailed kinematic information that spectroscopic observations can. However, until the advent of highly-multiplexed MOS IFUs, even IFU surveys will be limited to several thousands of targets, while pixel-$z$ can be applied to tens of thousands and up to millions of objects in order to determine the radial variation of stellar population parameters within them and the evolution of this radial variation with redshift. We therefore aim to extend studies performed by Welikala et al.~ (2008; 2009) which applied the pixel-$z$ method to $\\sim300,000$ galaxies at $z\\sim0.1$ in the SDSS and determined the radial variation of star formation for this sample, the scatter in this relation, and its dependence on galaxy environment and galaxy morphology. There are, however, inherent systematic effects that may limit the utlity of pixel-$z$ in probing spatially resolved stellar populations in galaxies. The current analysis explores these in order to quantify the impact of such systematics on the measurements made by pixel-$z$. \\subsection{Stellar Populations Extracted from Pixel Colors: Assumptions and Limitations} A number of studies have sought to extract stellar populations for spatially resolved colors. These compare the colors in individual pixels to a library of SEDs generated by stellar population synthesis models. The majority of these utilised small samples of galaxies at $z \\sim 0$ and $z \\sim 1$ (Abraham et al.~1999; Johnston et al.~2005; Kassin et al.~2003; Lanyon-Foster et al.~2007; Zibetti et al.~2011). Given a multi-band image of a galaxy, the pixel-$z$ method (Conti et al.~2003; Welikala et al.~2008, 2009) computes the stellar population properties such as stellar population age, SFR e-folding timescale $\\tau$, dust obscuration through $E(B-V)$ and metallicity $Z$ in each pixel in the galaxy, as well as the associated uncertainties in these parameters. It does this by taking a library of SEDs generated from stellar population synthesis (SPS) codes, redshifting each of these SEDs and convolving them with the available passbands. The simulated fluxes are then compared to the observed fluxes in each pixel and a best-fitting SED is computed for that pixel. The statistical errors in each parameter are calculated by marginalizing the likelihood function from the fit over the remaining parameters, and are conservative. The pixel-$z$ approach makes two fundamental assumptions: \\begin{itemize} \\item{\\textbf{Independently evolving pixels: } The pixels in the galaxy image are assumed to be evolving independently of each other. This ignores any mixing of stellar populations between neighboring pixels which is expected to occur in real galaxies. Neighbouring pixels will also not truly be independent, due to a finite Point Spread Function (PSF), and this effect is also neglected.} \\item{\\textbf{Single stellar populations with exponentially declinining SFH: } The stellar population underlying each pixel is assumed to be one whose spectrum is derived from integrating over single stellar populations (SSPs) of a given age, weighted by a SFR which is characterized by an initial burst followed by an exponential decrease in SFR. The decrease is parameterized by a range of allowed values of e-folding times $\\tau$. In reality, galaxies are expected to have more complicated SFHs, including multiple bursts.} \\end{itemize} In addition to these two fundamental assumptions, there is a limitation in the form of degeneracies among the SED models which could impact the accuracy with which pixel-$z$ determines stellar population parameters in the pixels. SPS models show that age, metallicity and dust all tend to affect spectra in similar ways (Bruzual \\& Charlot~2003). Multiple SEDs each generated for different populations (different ages, metallicites and SFHs) can, as a result, still predict very similar colors of a galaxy or its subcomponents. Some of these degeneracies arise from well known physical correlations, such as those between age and metallicity of the stellar population. Worthey~(1994) constructed detailed models of old and intermediate stellar populations and their absorption indices to show that if $\\delta_{age}$/$\\delta{Z} \\approx 3/2$ for two populations, they would appear identicial in most absorption indices. The age-metallicity degeneracy can also affect the photometric evolution of the SSPs since optical and near-infrared (NIR) colors reflect the relative contribution of hot and cool stars to the integrated light. Bruzual \\& Charlot~(2003) followed the evolution of optical and NIR colors and the stellar mass-to-light ratio $M/L$ in their model, and found that at fixed age, increasing metallicity tends to redden the colors and increase the $M/L$ ratio. This is explained by the fact that increasing the metallicity (at fixed stellar mass) causes (a) stars to evolve at lower effective temperatures and lower luminosities (Girardi et al.~2000) and (b) changes the relative number of red and blue supergiants, which strongly impacts the color evolution of the SSP. They also found that increasing metallicity at fixed age had a similar effect to increasing age at fixed metallicity. The age-metallicity degeneracy is therefore inherent in the SSP model used. There could also be other degeneracies, including those between dust obscuration and metallicity as well as between age and dust obscuration. The redshift-evolution of the obscuration-metallicity degeneracy in the pixels was characterized by Conti et al.~2003. This work does not test the two fundamental assumptions above. As for degeneracies, exploring the different correlations between the parameters on a pixel-to-pixel basis and propagating the resulting errors are beyond the scope of this paper. Welikala et al.~2008 characterized some of these correlations using the likelihood function from the fit to the pixels in SDSS galaxies. Since the severity of the degeneracies depend on the number and type of available broadband colors, the SPS model and the chosen grid of stellar population parameters, we focus on these systematics in this paper. The validity of both the two assumptions and the impact of degeneracies will be explored in a future paper that compares the spatial distribution of stellar populations inferred from spatially resolved spectroscopy obtained from large IFU programs (Croom et al.~2011) with those inferred from applying the pixel-$z$ technique to the same galaxies. Such a detailed comparison with IFU programs will enable a robust calibration and testing of pixel-$z$, leading eventually to a possible refinement of these assumptions. In this paper, we focus on the following systematics that could potentially bias our results: \\begin{itemize} \\item{\\textbf{The SPS model: } An inevitable bias is the choice of the stellar population synthesis model. Different stellar population synthesis models predict different colors. In particular, there are significant differences in the prescriptions for Thermally Pulsating Asymptotic Giant (TP-AGB) stars among the following models: Maraston 2005 (M05), Bruzual \\& Charlot 2003 (BC03) and Charlot \\& Bruzual 2007 (CB07). In this paper, we compare the outcomes of using these different models on a pixel-by-pixel basis, for all galaxies in the sample. Similar comparisons of population synthesis models have been performed previously on the integrated colors of galaxies but these have focused primarily on the effect on their stellar mass (Maraston et al.~2006). In this work, we focus on the effect of these differences on the age, SFR e-folding time and obscuration estimated in the pixels. } \\item{\\textbf{The available passbands: }The accuracy with which the pixel-$z$ parameters are determined in the pixels depends on the the available passbands. This paper explores the systematic effects arising from passband limitations, using both optical and near-IR colors. } \\item{\\textbf{The parameter space searched by pixel-$z$ and the number of SEDs per model: } Pixel-$z$ searches a grid of ages, e-folding times, dust attenuation values and metallicities. In general, a large as possible range of values is chosen while physically unmeaningful values will be associated with high statistical uncertainties. Despite this, a large parameter space can result in degeneracies among the various SEDs searched by pixel-z which in turn may bias the pixel-$z$ estimates. Some priors are already built in to stop pixel-$z$ searching unphysical parameters e.g., the age of a stellar population in a pixel cannot exceed the age of the Universe at a given redshift. Here, we explore the systematic error arising from increasing the range of parameter values allowed on the pixel-$z$ grid. In particular, we focus on the impact of increasing the metallicity range. This has two predicted effects: (a) it increases the severity of the age-metallicity degeneracy and would bias the pixel-$z$ estimates in quantities such as age, $\\tau$ and dust and (b) it introduces an uncertainty associated with the models: while spectral synthesis models have been well established for solar metallicity stellar populations in optical photometry, the models are less well established for very sub-solar or super-solar populations. Increasing the metallicity range will test the impact of both these effects on the derived pixel-$z$ parameters. } \\item{\\textbf{Global SED-fitting versus pixel-$z$: } We examine any biases resulting from measuring stellar population quantities, particularly the SFR, in individual pixels compared to the same quantities derived from aperture photometry and SED-fitting. These biases arise because individual pixels can have different colors to the integrated color of the galaxy. This bias could give rise to a difference between (1) the mean properties of the galaxy obtained by integrating the fluxes in an aperture and SED-fitting to the total fluxes and (2) the mean properties derived from fitting to the individual pixels within the same aperture and summing the resulting fits.} \\end{itemize} ", "conclusions": "\\label{sec:results} \\subsection{Effect of SSP Model Differences using Optical-only Bands} \\label{subsec:model_tests} Here, we illustrate the effect of using different stellar population synthesis models on the values of the parameters that pixel-$z$ determines using the \\textit{bviz} passbands. This is first shown qualititatively by displaying the resulting maps, and second, quantitatively, by measuring the systematic error in each parameter and following its distribution. In the top panel of Figure~\\ref{fig:models_optical_age}, we illustrate the effect of the chosen stellar populations model on the age maps. There is a clear correlation between the pixel-$z$ maps of this galaxy and the $i$ band image. The qualitative differences from pixel-to-pixel between the various models(M05,BC03,CB07) are small but there is a difference along the right spiral arm between BC03 and CB07 on the one hand and M05 on the other. BC03 and CB07 both predict that the majority of pixels along the arm have ages $\\sim0.1$Gyr, while M05 shows a larger proportion of pixels showing younger ages ($\\sim0.01$Gyr). Otherwise, both age values and their associated uncertainties in the pixels are qualitatively similar for all 3 models. The red pixels in the inter-arm regions indicate much older populations but since these are also typically lower signal-to-noise pixels, the rms error is also significantly worse for these pixels. Pixels belonging to the sky which are artificially fit are also assigned large uncertainties. The dust obscuration, shown in Figure~\\ref{fig:models_optical_ebv} is also correlated with the spatial distribution of the stellar age. Unlike the disagreement seen in the age values of the spiral arm between the M05 models on one hand and the BC03 and CB07 models on the other hand, the obscuration maps demonstrate a high level of agreement among the various models. In all the models, the arm regions, which were predicted to have the youngest stellar populations, also have a moderately high level of dust obscuration, with $E(B-V)\\sim 0.5$. The uncertainties in the dust obscuration in the pixels are similar among the various models as well. These uncertainties correlate strongly with morphological features in the image as in the case of the stellar population age. The knots on the spiral arms, corresponding to star-forming HII regions, are clearly displayed in the obscuration uncertainty map with the lowest errors of the entire galaxy. These statistical uncertainties are similar among all the models considered. The inner arms and the outermost edges of the galaxy, which are the regions with the lowest signal-to-noise in the entire galaxy, have, predictably, the highest rms uncertainties in their pixel-$z$ quantities. In order to quantify these effects, we use pixels from all 467 galaxies in our sample. The systematic error due to different population synthesis models is measured according to equation 1, with the resulting distribution shown in Figure~\\ref{fig:models_test_optical}. The distributions are generally non-Gaussian, peaking around zero systematic error, and showing a relatively small fraction of outliers. The shape of the distribution is a result of the quantized nature of the problem, since the pixel-$z$ parameters lie on a discrete grid of values. The level of agreement and the fraction of outliers in this distribution for every pixel-$z$ parameter is summarized in Table 1 for each of the pairs of models being compared. The two models that have the highest level of agreement in all the parameters concerned (age, $\\tau$ and E(B-V)) are BC03 and CB07. In the $\\tau$ parameter, 94 percent of pixels within the galaxies are within $1\\sigma$ and the fraction of outliers ($>2\\sigma$) is 0.04. The high level of agreement between BC03 and CB07 is due to the fact that the continuum in the rest-frame UV is very similar between these two models. There is a larger systematic difference in the pixel-$z$ parameters between BC03 and M05, and between CB07 and M05. In particular, the $\\tau$ parameter is most sensitive to differences between these models. The fraction of pixels with a systematic difference in $\\tau$ less than $1\\sigma$, is 0.71 and 0.80 for the [BC03,M05] and [CB07,M05] model pairs respectively. The fraction of outliers in $\\tau$ ($>2\\sigma$) is 0.15 and 0.17 respectively for the same model pairs. The impact of model differences between BC03/CB07 and M05 is much smaller for the age and obscuration. Dust obscuration is least sensitive to differences between any of the 3 models, with only $1-3$ percent of pixels having systematic differences in $E(B-V)$ which are larger than $2\\sigma$. For the inferred stellar age, the fraction of pixels having systematic differences in their stellar population age of $<1\\sigma$ is 0.86 and 0.87 for [BC03,M05] and [CB07,M05] respectively, while the fraction of outlier pixels in the age distribution for the same pairs of models is 0.11. For [BC03,CB07], 0.94 of pixels agree in their inferred stellar population age to within $1\\sigma$, while the fraction of outlier pixels for the same models is 0.04. The comparison of the stellar population synthesis models for the metallicity are not shown in Table 1 because the small number of values of metallicity being sampled means that all fits are within $2\\sigma$. Out of the three pixel-$z$ parameters tested, it is thus in the SFR e-folding time where the largest discrepancy lies between M05 on one hand and BC03/CB07 on the other. This discrepancy is unlikely to be due only to the differing treatment of TP-AGBs stars in these models, since if that were the case, we would expect better agreement between CB07 and M05 as contrasted against BC03. The discrepancy observed in $\\tau$ is likely related to fundamental differences between M05 on one hand and BC03/CB07 on the other when the SED fitting is performed using optical colors only. In section 3.5.2, we use some example galaxies to investigate where the discrepancy in the predicted values of $\\tau$ (between M05 and BC03/CB07) occurs spatially in the galaxies. Finally, we test the effect of redshift on these results by repeating the above analysis for subsamples of galaxies in redshift intervals of 0.5 from $z=0.5-3$ and find that our results for the optical-only passbands are not significantly impacted by the redshift of the galaxy. \\subsection{Effect of Parameter Space and SED Degeneracies} \\label{subsec:paramspace} In this test, we examine the effect of changing the parameter space on the pixel-$z$ estimates using the BC03 models. The number of metallicity values used in the pixel-$z$ grid of stellar population parameters is increased from 3 to 6, to include sub-solar metallicities, as detailed in Section~\\ref{sec:sspsfromcolors}. The increase in the metallicity range means we are comparing the original 2178 SEDs with an expanded set of 4356 SEDs fitted to each pixel. The distribution of the systematic error (as a function of the statistical error) introduced by extending the allowed metallicity range into the sub-solar regime is shown in Figure~\\ref{fig:metallicity_test} for all pixels in all the galaxies in the sample. The fraction of pixels with small systematic errors ($<1\\sigma$) and the fraction of outliers in each parameter is summarized in Table 2. The largest effect of a broadening the metallicity range is, predictably, on the stellar population age but the effect is minimal: the fraction of outliers in the age systematic is only 0.06. Around 6 percent of pixels have a bias in their $\\tau$ that is larger than the $2\\sigma$ uncertainties. The obscuration is the parameter which is least affected ($< 0.01$ of pixels) by the introduction of sub-solar metallicities. This is in agreement with the findings of Conti et al.~(2003) who examined the evolution of the obscuration-metallicity degeneracy in pixels in galaxies in the HDF-North, and found that there is a strong degeneracy in galaxies at low redshifts but that the relation flattens out at $z>1$. The conclusion of this test is that while the stellar population age and $\\tau$ are marginally affected by the increased metallicity range, as expected from the age-metallicity degeneracy, the resulting systematic differences in all the pixel-$z$ quantities are still well within the uncertainties between different stellar population synthesis models. \\subsection{Effect of Passband Limitations (Optical-only)} Here we test the effect of the passband limitations on the estimates of the pixel-$z$ parameters. The stellar population synthesis model is fixed to BC03 in this test and passbands are added and removed and the pixel-$z$ output is compared to that produced using the full complement of 4 optical bands. This test is initially performed for selected individual galaxies to identify the likely trends, before a full investigation on all pixels in the full galaxy sample is performed. Figure~\\ref{fig:passbands_optical_age} illustrates the test for one example galaxy, which is the same as shown in Figures~\\ref{fig:models_optical_age} and~\\ref{fig:models_optical_ebv}. It shows the impact of passband variations on the age values in the pixels and their associated uncertainties. Starting with only 2 optical bands ($bv$), the effect of different permutations of bands are tested. From the stellar population age maps, it is clear that 2 passbands are insufficient to resolve many of the features seen in the light distribution of the galaxy. For the case of 3 bands, the use of the redder $viz$ combination of filters comes closer to the result obtained with the full $bviz$ filters than does the $bvi$ set. The $viz$ combination produces an age map that comes closer to mirroring the galaxy morphology than does the $bvi$ set, separating the arm and inter-arm regions more clearly. For this example galaxy at $z\\sim1$, most of the relevant color information for the pixel-$z$ therefore comes from the optically redder passbands. A similar conclusion can be drawn from the map of rms values in the age parameter. With only 2 optical bands, high uncertainties throughout the disk of the galaxy reflect the inability of pixel-$z$ to find the correct SED with only 1 color, thus resulting in higher statistical errors in the ages inferred for the pixels. With 3 passbands, again it is the $viz$ combination that seems to be closest to the $bviz$ values. However, the maps of dust obscuration in Figure~\\ref{fig:passbands_optical_ebv} show that the $bvi$ combination coming closest to the result predicted from the full complement of bands. Nevertheless, the $bvi$, $viz$ and $bviz$ combinations all agree that the arms have intermediate levels of obscuration ($E(B-V)\\sim0.4$) and low associated RMS errors. Moving to the full sample of all pixels in all galaxies, we use equation 2 to compute for each parameter and for each pixel, the systematic error (again in terms of the statisitical error in the parameter) due to the difference between a given filter permutation and the full optical filter set ($bviz$). The stellar population model is fixed to BC03 as before. Figure~\\ref{fig:passbands_optical_test} shows the distribution of this systematic error for all the pixels in the full galaxy sample. It is evident that the systematic error distribution has larger tails and a lower peak ($<1\\sigma$) than the systematic errors resulting from stellar population model differences. As for the example galaxy in Figures~\\ref{fig:passbands_optical_age}, it is clear also that 2 filters ($bv$) result in a large fraction of outliers ($>3\\sigma$). For the 3-filter permutations ($biz$, $bvi$ and $viz$), across the entire sample of galaxies, we find that there is only a small difference in the distribution of the systematic error among the filter permutations. We show the fraction of outliers in each case in Table 3. It is clear that the SFR e-folding time ($\\tau$) and the stellar population age are the most impacted of the 3 parameters by passband limitations. In $\\tau$ and in the stellar population age, over 20 percent of the pixels become outliers ($>2\\sigma$) when one band of the full filter set $bviz$ is dropped. There is not a very significant difference between the 3 sets of filters, but omitting the $z$ band results in an outlier fraction for the $\\tau$ parameter of 0.26 compared to 0.20 for the other bands. For the stellar population age, the outlier fraction varies from $0.22-0.28$ relative to the full filter set ($bviz$). The dust obscuration is least impacted by passband changes, as long as a minimum of 3 are used. For each of the 3 filter sets, about 80 percent of pixels in the full galaxy sample have an $E(B-V)$ which agrees with the $bviz$ result to within $1\\sigma$. Finally, as in Section~\\ref{subsec:model_tests}, we test the effect of redshift on these results by repeating the above analysis for subsamples of galaxies in different redshift intervals from $z=0.5-3$ and find that our results for the optical-only passbands are not significantly impacted by the redshift of the galaxy. \\subsection{Effect of Adding the Near-IR Data} In this section, we explore the effect of adding NIR passbands from NIC3 to the existing optical data for a subsample of 46 galaxies. Adding the NIR data allows us to explore systematic differences between various population synthesis models more fully, since they can predict different NIR colors. In particular, it is known that BC03 and M05 differ significantly in their treatment of TP-AGB stars, consequently resulting in different predicted NIR colors. This is illustrated in Figure~\\ref{fig:spectra} which shows an example SED from all three models considered in this work, for a galaxy at $z=1.0$ containing a stellar population that is 3 Gyr old, with $\\tau=10$ Gyr, $E(B-V)=0.9$ and $Z=0.008$. There is very little difference between the shape of the continuum between the the models in the optical. However, the same galaxy in the NIR shows significant differences in the shape of the continuum among the three models. These will give rise to different predicted NIR colors in the pixels. Figures~\\ref{fig:age_nir_example} and ~\\ref{fig:ebv_nir_example} show examples of applying pixel-$z$ to a galaxy at $z=2.5$ with $bviz$ and $J$ and $H$ band imaging from NIC3. There is a good agreement among the three models compared here in both their stellar population age maps and in the age rms map. All three models predict a very young population of stars throughout ($t<0.2$ Gyr). All models considered predict the core of the galaxy to have an older population of stars ($\\sim0.1$ Gyr) than the outskirts ($t\\sim0.01$Gyr). The one notable difference is that the M05 models predict a slightly larger proportion of very young stellar populations ($t\\sim0.001$Gyr) surrounding the nucleus. The models also agree very well on the age distribution within the companions of this galaxy (to the right). The models also agree well in the uncertainty maps in the stellar population age. Figure~\\ref{fig:ebv_nir_example} shows the obscuration maps of the same galaxy as predicted by the three models. Again, there is excellent agreement among the three models in the distribution of dust and its associated uncertainty. In general, all three models predict that the galaxy has moderate levels (E(B-V)$\\sim0.3$ magnitudes) of dust, but they also show pockets of high obscuration (E(B-V)$\\sim0.5$ magnitudes) within the main galaxy and also in its two companions. The location of these pockets are the same in three models. This example therefore implies that model differences do not introduce large systematic biases in the pixel-$z$ parameters for $z\\sim2.5$ galaxies with such irregular morphologies. We investigate this hypothesis fully when we measure systematic differences across all pixels in the combined optical-NIR galaxy sample for all 3 models considered in this paper. For each parameter, we compare this distribution of the systematic differences with that obtained with using only the optical data. We account for the effect of redshift, particularly because for $z>1$, the 4000\\,\\AA \\ break shifts into the $J$ and $H$ passbands. We thus split the NIR sample into two: $z<1$ and $z>1$. In Figure~\\ref{fig:nir_test_zle1}, we show the distribution of the systematic difference between models for the low redshift sample. The fraction of pixels showing $<1\\sigma$ uncertainties in their pixel-$z$ parameters for each pair of models is given in Table 4 for both the optical-only colors and the optical and NIR colors. For the low redshift sample, there is a very high level of agreement between the BC03 and CB07 models in all the pixel-$z$ parameters, with over $97$ percent of pixels showing less than $1\\sigma$ deviation between the two models. With the addition of the NIR bands however, the systematic differences between BC03 and CB07 models are exposed. This is a result of relatively small differences among the models in their predicted rest-frame UV colors but much larger differences among them in their predicted rest-frame optical (red) and rest-frame NIR colors. This affects in particular the stellar population age (with only 78 percent of pixels showing less than $1\\sigma$ difference between the two models) and $\\tau$ (88 percent of pixels showing agreement between the two models). The obscuration is less impacted by the model differences. The addition of the NIR colors also exposes differences between BC03 and M05, and this impacts the stellar population age most strongly, with the fraction of pixels showing $<1\\sigma$ difference in the age decreasing from 0.89 with optical-only colors to 0.80 with NIR colors. The differences between BC03 and M05 have less impact on $\\tau$ and the obscuration when the NIR colors are added. When comparing CB07 with M05, it is again the stellar population age in the pixels that is most affected by the addition of the NIR passbands, with only 80 percent of pixels showing an agreement in the age to within $1\\sigma$ between the two models, compared to 89 percent for the optical-only case. For the high-redshift ($z>1$) sample, the story is somewhat different, as illustrated in Figure~\\ref{fig:nir_test_zgt1} and Table 5. There is now a much higher level of agreement between the results for the optical-only colors and the optical and NIR colors combined, for all the parameters concerned. This is due to the fact that the rest-frame UV and blue wavelength optical spectra (which are sampled by the optical and the NIR filters for $z>1$) are similar among the models. This similarity in the predicted rest-frame UV/optical (blue) colors among the models thus manifests itself as follows. In the $z<1$ sample discussed above, the stellar population age showed a large discrepancy between the $bviz$ and $bvizJH$ samples in terms of the fraction of pixels that agree on their age for the different models being compared. For the high redshift sample, the pixels in both the $bviz$ and $bvizJH$ samples show similar levels of agreement in the age parameter between the different models compared. For example, for the the BC03 and M05 models, the fraction of pixels with systematic differences in their age that are smaller than $1\\sigma$ is 0.95 in the optical-only case and 0.93 in the NIR case. The conclusion of these tests is that, for $z>1$ galaxies, differences among the stellar population synthesis models in their predicted NIR colors do not significantly bias the pixel-$z$ estimates of the stellar population parameters in the pixels, compared to using optical colors only. \\subsection{Global Galaxy Properties and the Effect of Systematics } Pixel-$z$ can measure global properties of structures composed of many pixels within the same galaxy. These will be used to determine the radial variation of stellar population parameters in galaxies as a function of the galaxy environment, and also to allow a detailed study of high redshift clumpy, disk galaxies, as outlined in Section 1.1 and 1.2. Here, we investigate the effect of systematic biases that can impact these global measurements, in particular of the SFR since this will be a direct observable in spectroscopic measurements. \\subsubsection{Biases from using pixel-$z$ versus global SED-fitting} In Figure~\\ref{fig:SFR_pixelz_global}, we show the level of consistency between the sum of SED fits to individual pixels in a galaxy and the global fit to the total flux in those same pixels, as described in Section 2.3.2. We see a high level of consistency between the two approaches in their estimates of the integrated SFR. We also quantify the magnitude of any deviation between the two approaches in terms of the statistical uncertainty in the SFR estimate for each galaxy from SED-fitting. We see that for the majority of galaxies in the sample, the difference between the two SFR estimates for the galaxies are within the statistical uncertainties of the SFR measurement. However, there is a small population of outliers ($>2\\sigma$) which are galaxies with low SFR ($<1 M_{\\odot}\\,yr^{-1}$) and which show a negative residual i.e., the SFR estimates from pixel-$z$, obtained by summing the results of the individual fits in each pixel, are higher than the SFR estimates from global fits to the total flux in the pixels. In addition, there is a second small population of outliers which are galaxies with high SFRs ($\\sim10 M_{\\odot}\\,yr^{-1}$). In these, the SFR estimate from the global fit is significantly higher than that estimated from the sum of the individual SED fits. \\subsubsection{Radial gradients of stellar populations: age, dust and star formation} In Figure~\\ref{fig:radialgradient_models}, we present the radial variation of the stellar population age, SFR e-folding time, obscuration, SFR and sSFR using the $bviz$ passbands for the same galaxy at $z\\sim1$ in Figures~\\ref{fig:models_optical_age} and~\\ref{fig:models_optical_ebv}. The radial trends in each quantity are measured according to the method described in section 2.3.3. All stellar population synthesis models agree on the qualitative trends of all the parameters with radius but there are some quantitative differences that arise, particularly between the different population synthesis models All three models agree that the stellar populations in the galaxy at all radii are young ($<0.3$Gyr) and that the stellar population age decreases rapidly from the center to the outskirts. M05 predicts a mean stellar population age that is somewhat lower in the innermost part of the galaxy compared to BC03 and CB07, and this systematic difference is of the order of the statistical uncertainty in the age measurement. BC03 and CB07 also predict a somewhat sharper drop in the stellar population age between 2 and 3 kpc, but again, the difference is well within the statistical uncertainties. The models agree on a range of $\\tau\\sim0.5-1$ Gyr across the galaxy. Between 3 and 5 kpc, however, the M05 model prediction diverges by $1\\sigma$ from both BC03 and CB07 models. This region corresponds to the spiral arm and the clumps of star-forming HII regions within it. This implies that the different treatment of TP-AGB stars in M05 compared to the other models has its biggest effect in the spiral arms of the galaxy. The dust obscuration trends are quite similar among the models which show that the center of the galaxy is moderately obscured ($E(B-V)\\sim0.45$). The obscuration then decreases up to 2 kpc and then increases to a maximum ($E(B-V)\\sim0.6$) in the spiral arms. The different models all find that the mean SFR in each annulus peaks in the center of the galaxy. To explain this, we note that the current SFR is determined not only by the mean age and $\\tau$ of the stellar populations, but also by the total stellar mass formed to date, which normalizes the SFR. M05 predicts a lower stellar mass throughout the galaxy than does BC03 or CB07. When the stellar mass is normalized out to give the specific SFR (sSFR), we see that the discrepancy between M05 and BC03/CB07 narrows considerably: $1\\sigma$ at most in all radial bins. Comparing the trend in sSFR with the morphology of the galaxy, we conclude that the bar structure in the center of the galaxy is actively star-forming. The models disagree, however, on how much more star-forming the bar is relative to the spiral arms at a radius of $\\sim4$kpc. BC03 and CB07 predict a slightly higher sSFR (by around $2\\times10^{-11}\\,yr^{-1}$) in the center of the galaxy compared to the spiral arms, whereas M05 predicts a somewhat lower sSFR in the center relative to the spiral arms. Nonetheless, a general picture emerges for this galaxy, which is that star formation is not simply confined to the center but takes place throughout the galactic disk. The results, albeit for a single galaxy, imply that pixel-$z$ is robust to differences between various stellar population synthesis codes when it is used to measure radial variations across galaxies. Figure~\\ref{fig:radialgradient_passbands} shows the effect of passband limitations on measuring radial variations of stellar population parameters. The population synthesis model was fixed to BC03. Using only the $b$ and $v$ bands results in large systematic uncertainties which affects all the radial quantities but particularly the mean age and the obscuration determinations. The result of using only the $b$ and $v$ bands is that the mean age is overestimated throughout the galaxy relative to the $bviz$ filter set, by as much as $4\\sigma$. This can be explained by the fact at $z\\sim1$, the 4000\\,\\AA \\ break passes out of the bluest optical bands and, as a result, the colors predicted by the model for the pixels are redder, leading to an older stellar population for the pixels. The obscuration is underestimated throughout, and at a radius of 4 kpc the difference with respect to the $bviz$ value is as much as $4\\sigma$. The $\\tau$ and the sSFR estimates are less impacted by the loss of 2 passbands. The 3-passband combinations $bvi$ and $viz$ both come closest to reproducing the $bviz$ radial trends in all the parameters. The systematic difference between the $bvi$ and $viz$ combinations is relatively small, but some differences do emerge nevertheless. In the center of the galaxy, the $viz$ filter combination predicts a larger (by $1\\sigma$) mean stellar population age than the bluer $bvi$ bands. Elsewhere in the galaxy, the three-filter combinations give results which are consistent (within their statistical uncertainties) with each other and also with the full $bviz$ filter set. The radial variation of the dust obscuration is consistent among the $bvi$, $viz$ and $bviz$ filter sets except in the innermost annulus where the $viz$ filters underpredict the level of obscuration relative to the other two combinations. The differences in $\\tau$ are small, although there is one important difference at $r=3$kpc where the $bvi$ set predicts a higher $\\tau$ compared to the $viz$ and $bviz$ sets. The effect of filter cominbations on the sSFR is also relatively small throughout the galaxy except in its outermost part ($r\\sim5$kpc) where the three-filter combinations overestimate the sSFR relative to the full optical set. If we were to generalize this test, a minimum of 3 optical passbands are therefore needed for robustly computing radial trends of stellar populations in these high redshift galaxies. \\subsubsection{Disk galaxies at $z\\gtrsim1$: spiral arms and clumps} In Figures~\\ref{fig:spiralarms}, \\ref{fig:clumpA} and \\ref{fig:clumpB}, we illustrate the way pixel-$z$ can be used to study the stellar populations of subcomponents of galaxies such as spiral arms and clumps, and we explore how this can be biased by the choice of the SPS model. The optical filter set $bviz$ is used throughout this analysis. In the spiral arm of the disk galaxy which has been isolated, the BC03 and CB07 models predict almost identical distributions for the parameters among the pixels that make up the spiral arm. In contrast, some differences start to emerge between M05 and the other two models. The age distribution of the pixels peaks at 0.1 Gyr for the BC03 and CB07 models, and at a lower age (0.01 Gyr) for the M05 model. All three models show good agreement for the dust distribution among the pixels in the arm, which peaks at $0.5$ Gyr. There is however a systematic offset in the SFR distributions which peak at 0.003 $M_{\\odot}\\,yr^{-1}$ with M05, whereas it peaks at 0.03 $M_{\\odot}\\,yr^{-1}$ with BC03/CB07. This difference is largely due to the lower stellar mass predicted for the pixels in the arm by M05 compared to the other models. When stellar mass is normalized, we see a much closer agreement between M05 and BC03/CB07. All the models predict a sSFR distribution that is bimodal in the arm, one peaking at $10^{-9.8}yr^{-1}$ and the other at $10^{-8.3}yr^{-1}$. We also investigate clumps in the disk and the impact of model differences on the inferred properties of clumps in the disk of the galaxy. Clump A in Figure~\\ref{fig:clumpA} is part of the outer arm of the galaxy. As in the case of the spiral arm, the distribution of the stellar population age and dust obscuration among the pixels in clump A, is quite similar among the different population synthesis models. As with the full spiral arm, a substantial difference exists in the SFR distributions. The M05 models predict the clump to have a lower SFR than BC03/CB07. The distribution peaks at $\\sim0.01 M_{\\odot}\\,yr^{-1}$ with M05 and $0.3 M_{\\odot}\\,yr^{-1}$ with BC03/CB07, due to a lower stellar mass predicted for the clump by the M05 model. It is also worth nothing that the peak of the SFR distribution within the clump occurs at a higher SFR than in the spiral arm. The sSFR distributions in the clump are far more consistent among all the models, again showing a bimodal star formation in the clump, with one population peaking at $10^{-9.8}yr^{-1}$ and the other at $10^{-8.3}yr^{-1}$ as in the case of the spiral arm. The SFR and sSFR distributions imply that the star formation process in the spiral arm might be driven by the star formation in the clump itself. Finally, in Figure~\\ref{fig:clumpB}, we examine clump B on the left side of the galaxy. Again, model differences affect primarily the SFR, with M05 predicting the SFR distribution to peak at $\\sim0.001 M_{\\odot}\\,yr^{-1}$ while BC03/CB07 predict that it peaks around 0.03 $M_{\\odot}\\,yr^{-1}$. Irrespective of the model, however, it is evident that the mean SFR of clump B is significantly lower than clump A, indicating that clump A is the primary driver of star formation in the outskirts of the galaxy. Despite some differences between M05 and the other models, therefore, pixel-$z$ can be used to effectively and accurately probe mean properties of stellar populations within components of galaxies such as spiral arms and clumps. As discussed in section 1.1, this can be used to probe substructure in $z\\sim2$ galaxies in order to determine if these objects are largely driven by mergers or are disks formed in-situ. \\subsection{Incorporating WFC3 to Probe Substructure: Expectations from the CANDELS Survey} \\label{subsec:candels} Figure~\\ref{fig:CANDELS_maps} shows the results of adding the NIR WFC3 imaging in the $F125W$ and $F160W$ passbands from the CANDELS survey to the existing $bviz$ bands for an example disk galaxy at $z=0.6$, with the M05 models. The most notable difference between using the optical-only passbands and the combined optical and WFC3/NIR passbands is in the spatial distribution of the oldest ($t>4$ Gyr) stellar populations in the galaxy. In the ACS-only case, the oldest stellar populations in the galaxy are confined in highly localized regions in the galaxy, with a concentration in the center and and in pockets in the south-west and sout-east of the image. The addition of the WFC3/NIR data, however, suggests that the oldest (and thus redder) populations are not simply confined in these pockets but are present more extensively in the disk itself. In the map of the specific star formation rate in the galaxy, both the optical-only and the combined optical and NIR cases suggest a qualitatively similar trend in the spatial distribution of the sSFR i.e., the sSFR is lowest in the center of the galaxy ($\\sim10^{-11}\\,yr^{-1}$) and increases towards the outskirts. Away from the center, however, the addition of the WFC3/NIR passbands, suggests that the disk contains a higher proportion of stellar populations with a sSFR ($\\sim10^{-10}\\,yr^{-1}$) that is lower than what is predicted by the optical-only colors. The differences observed in the maps of the stellar population age and sSFR are confirmed by the radial variation of these quantities, shown in Figure~\\ref{fig:CANDELS_radialplots}. The WFC3/NIR colors reveal an older population throughout the disk of the galaxy relative to the optical-only colors, with more than a $5\\sigma$ difference in the luminosity-weighted mean age between the $bviz$ and $bvizJH$ colors and up to a radius of 6 kpc from the galaxy center, . The radial variation of the sSFR also reflects the fact that the WFC3/NIR colors are more sensitive to the older stellar population in the disk, which were otherwise not revealed by the ACS-only passbands. The WFC3/NIR colors reveal a lower sSFR at all radii compared to the ACS-only colors, although the difference is most significant between 3 and 5 kpc from the galaxy center." }, "1112/1112.0187_arXiv.txt": { "abstract": "Gaia is a European Space Agency (ESA) astrometry space mission, and a successor to the ESA Hipparcos mission. Gaia's main goal is to collect high-precision astrometric data (i.e. positions, parallaxes, and proper motions) for the brightest 1 billion objects in the sky. These data, complemented with multi-band, multi-epoch photometric and spectroscopic data collected from the same observing platform, will allow astronomers to reconstruct the formation history, structure, and evolution of the Galaxy. Gaia will observe the whole sky for 5 years, providing a unique opportunity for the discovery of large numbers of transient and anomalous events, e.g. supernovae, novae and microlensing events, GRB afterglows, fallback supernovae, and other theoretical or unexpected phenomena. The Photometric Science Alerts team has been tasked with the early detection, classification and prompt release of anomalous sources in the Gaia data stream. In this paper, we discuss the challenges we face in preparing to use Gaia to search for transient phenomena at optical wavelengths. ", "introduction": " ", "conclusions": "" }, "1112/1112.2182_arXiv.txt": { "abstract": "We present a model to self-consistently describe the joint evolution of starburst galaxies and the galactic wind resulting from this evolution. This model will eventually be used to provide a subgrid treatment of galactic outflows in cosmological simulations of galaxy formation and the evolution of the intergalactic medium (IGM). We combine the population synthesis code Starburst99 with a semi-analytical model of galactic outflows and a model for the distribution and abundances of chemical elements inside the outflows. Starting with a galaxy mass, formation redshift, and adopting a particular form for the star formation rate, we describe the evolution of the stellar populations in the galaxy, the evolution of the metallicity and chemical composition of the interstellar medium (ISM), the propagation of the galactic wind, and the metal-enrichment of the intergalactic medium. The model takes into account the full energetics of the supernovae and stellar winds and their impact on the propagation of the galactic wind, the depletion of the ISM by the galactic wind and its impact on the subsequent evolution of the galaxy, as well as the evolving distributions and abundances of metals in the galactic wind. In this paper, we study the properties of the model, by varying the mass of the galaxy, the star formation rate, and the efficiency of star formation. Our main results are the following: (1) For a given star formation efficiency $f_*$, a more extended period of active star formation tends to produce a galactic wind that reaches a larger extent. If $f_*$ is sufficiently large, the energy deposited by the stars completely expels the ISM. Eventually, the ISM is being replenished by mass loss from supernovae and stellar winds. (2) For galaxies with masses above $10^{11}\\msun$, the material ejected in the IGM always falls back onto the galaxy. Hence lower-mass galaxies are the ones responsible for enriching the IGM. (3) Stellar winds play a minor role in the dynamical evolution of the galactic wind, because their energy input is small compared to supernovae. However, they contribute significantly to the chemical composition of the galactic wind. We conclude that the history of the ISM enrichment plays a determinant role in the chemical composition and extent of the galactic wind, and therefore its ability to enrich the IGM. ", "introduction": "Galactic winds and outflows are the primary mechanism by which galaxies deposit energy and metal-enriched gas into the intergalactic medium (IGM).\\footnote{Some authors make a distinction between {\\it galactic winds}, which are generated over most of the lifetime of the galaxy and inject energy and metals at a steady rate, and {\\it galactic outflows}, which result from violent processes like starbursts, are short-lived, and eject material at large enough distances into the IGM to eventually reach other galaxies. In this paper, we use one or the other to designate any material that is ejected from the galaxy and deposited into the IGM.} This can greatly affect the evolution of the IGM, and the subsequent formation of other generations of galaxies. Feedback by galactic outflows can provide an explanation for the observed high mass-to-light ratio of dwarf galaxies and the abundance of dwarf galaxies in the Local Group, and can solve various problems with galaxy formation models, such as the overcooling and angular momentum problems (see \\citealt{benson10} and references therein). Galactic outflows can explain the metals observed in the IGM via the Lyman-$\\alpha$ forest (e.g. \\citealt{my87,schayeetal03,ph04,aguirreetal08,pierietal10a, pierietal10b}), the entropy content and scaling relations in X-ray clusters \\citep{kaiser91,eh91,cmt97,tn01,babuletal02,voitetal02}, and provide observational tests that can constrain theoretical models of galaxy evolution. Local examples of spectacular outflows in dwarf starburst galaxies include those of the extremely metal-poor I Zw 18 (\\citealt{pequi08,jamet10}) and NGC1569 \\citep{west09}. More massive spirals, such as NGC7213 \\citep{hameed01}, also show evidence of global outflows. For a review of the subject, see \\citet{vcbh05}. \\subsection{Galactic Outflow Models} Large-scale cosmological simulations have become a major tool in the study of galaxy formation and the evolution of the IGM at cosmological scales. These simulations start at high redshift with a primordial mixture of dark and baryonic matter, and a spectrum of primordial density perturbations. The algorithm simulates the evolution of the system by solving the equations of gravity, hydrodynamics, and (sometimes) radiative transfer. Adding the effect of galactic outflows in these simulations poses a major practical problem. In one hand, the computational volume must be sufficiently large to contain a ``fair'' sample of the universe, typically several tens of Megaparsecs. On the other hand, the physical processes responsible for generating the outflows take place inside galaxies, at scales of kiloparsecs or less. This represents at the very minimum 4 orders of magnitude in length and 12 orders of magnitude in mass, which is beyond the capability of current computers. Since we cannot simulate both large and small scales simultaneously, the usual solution consists of simulating the larger scales and using a {\\it subgrid physics} treatment for the smaller scale. Cosmological simulations can predict the location of the galaxies that will produce the outflow, but cannot resolve the inner structure of these galaxies with sufficient resolution to simulate the actual generation of the outflow. Instead, the algorithm will use a prescription to describe the propagation of the outflow and its effect on the surrounding material. One possible approach consists of depositing momentum or thermal energy ``by hand'' into the system, to simulate the effect of galactic outflows on the surrounding material \\citep{std01,theunsetal02,sh03,cno05,od06, kollmeieretal06}. The algorithm determines the location of the galaxies producing the outflows and calculates the amount of momentum or thermal energy deposited into the IGM based on the galaxy properties (mass, formation redshift $\\ldots$). Then, in particle-based algorithms like smoothed particle hydrodynamics (SPH), this momentum or energy is deposited on the nearby particles, while in grid-based algorithms it is deposited on the neighboring grid points. This will result in the formation and expansion of a cavity around each galaxy, which is properly simulated by the algorithm. A second approach consists of combining the numerical simulation with an analytical model for the outflows. Tegmark et al. (1993) have developed an analytical model to describe the propagation of galactic outflows in an expanding universe. In this model, a certain amount of energy is released into the interstellar medium (ISM) by supernovae (SNe) during an initial starburst. This energy drives the expansion of a spherical shell that propagates into the surrounding IGM, until it reaches pressure equilibrium. This model, or variations of it, has been used extensively to study the effect of galactic outflows on the IGM (\\citealt{fl01,mfr01,sb01,sfm02,sh04,lg05,pmg07}, hereafter PMG07; \\citealt{sss08,gbm09,pmp10}). In this approach, the evolution of the IGM and the propagation of the outflow are calculated separately, but not independently as they can influence one another. The presence of density inhomogeneities in the IGM can affect the propagation of the outflow, while energy and metals carried by the outflow can modify the evolution of the IGM. There are several limitations with this second approach. In particular, it assumes that the initial starburst, which occurred during the formation of the galaxy, is the only source of energy driving the expansion of the outflow. First, the starburst lasts for a short period of time, typically 50 million years. Hence, we would expect to observe very few galaxies having an outflow. Furthermore these galaxies would be just forming and therefore would have complex and chaotic structures. Observations show instead that outflows are ubiquitous and often originate from well-relaxed galaxies (see \\citealt{vcbh05} and references therein). Second, even though the injection rate of energy is maximum during the initial starburst, the total amount of energy which is injected afterward by all generations of SNe and stellar winds could be comparable or even more important. Even if this energy is injected slowly over a large period of time, the cumulative effect could be significant. Indeed, an initial outflow caused by a starburst could be followed by a steady galactic wind that would last up to the present. The role of stellar winds has been mostly ignored in analytical models and numerical simulations of galactic outflows. Third, there is the possibility that accretion of intergalactic gas onto the galaxy might trigger a second starburst. Finally, a recent study \\citep{sbh10} suggests that there is a significant time delay between the initial starburst and the onset of the outflow, something not considered by current models. Another important issue is the amount of metals contained in the outflow, the spatial distribution of metals in the outflow, and the relative abundances of the various elements. The metallicity of the outflow depends on the metallicity of the ISM at the time of the starburst. The metals contained in the ISM at that time can have several origins: (1) metals already present in the gas when the galaxy formed, (2) the SNe produced during the starburst, and (3) the stellar winds generated by massive stars and AGB objects. Hence, the composition of the outflows will depend on the epoch of formation of the galaxy (which determines the initial metal abundances), as well as the relative amount of metals injected into the ISM by Type~Ia SNe, core-collapse SNe (Types~Ib, Ic and II), and winds, and the timing of these various processes. As for the amount of metals injected in the IGM, models often assume that it is proportional to the mass of the galaxy, and do not provide a description of the distribution of metals in the outflow and the relative abundances of the elements (\\citealt{sb01}; PMG07). \\subsection{Objectives} {\\it Our goal is to develop a new galaxy evolution model to improve the treatment of galactic winds in cosmological simulations.} This model will describe not only an initial starburst and its resulting outflow, but the entire subsequent evolution of a galaxy up to the present. It will take into account the progressive injection of energy by SNe and stellar winds (which could cause a steady galactic wind that would follow the outflow and last up to the present), and the time-evolution of the metallicity and composition of the ISM that would directly affect the composition of the galactic wind. It will also provide a description of the metal content, metal distribution, and chemical composition of the galactic wind. This emphasis on the structure and composition of the galactic wind is what distinguish our model from recent semi-analytical models of galaxy formation, which tend to focus on reproducing the properties of the galaxies themselves (luminosity and mass functions, halo properties, disk sizes, $\\ldots$). For a review of the various semi-analytical models, see \\citet{baugh06}. Our approach combines a population synthesis algorithm to describe the stellar content of the galaxy, an analytical model for the expansion of the galactic wind, and a new model for the distribution of elements inside the galactic wind. The paper is organized as follows: in \\S2, we describe the method we use to calculate the mass and chemical composition of stellar winds and SN ejecta, a key ingredient of our algorithm. In \\S3, we describe the basic equations for the evolution of the ISM. In \\S4, we describe our galaxy wind model. Results are presented in \\S5, and conclusions in \\S6. ", "conclusions": "We have combined a population synthesis code, interpolation tables for the mass and composition of SN ejecta, and an analytical model for galactic winds into a single algorithm that self-consistently describes the evolution of starburst galaxies. This model describes the evolution of the stellar populations in the galaxy, the evolution of the mass and chemical composition of the ISM, the propagation of the galactic wind, and the distribution and abundances of metals inside the galactic wind. In particular, the algorithm (1) provides a detailed calculation of the energy deposited into the ISM by SNe and stellar winds, which is responsible for driving the galactic wind, (2) takes into account the time-evolution of the chemical composition of the ISM, which directly affect the composition of the galactic wind, and (3) takes into account the removal of the ISM by galactic winds, which affects the metallicity of the ISM, and the metallicity of the stellar populations to follow. \\begin{figure} \\begin{center} \\includegraphics[width=3.3in]{fig17.eps} \\caption{Metallicity profile of material ejected in the IGM at $z=0$, for a $10^9\\msun$ galaxy with an exponential SFR. The various colors represent different star formation efficiencies $f_*$. Colors and linetypes have the same meaning as in Figures~\\ref{mass2star} and~\\ref{ZISM_z}. } \\label{Z_IGM} \\end{center} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[width=3.3in]{fig18.eps} \\caption{Comoving radius of galactic wind versus redshift, for galaxies with an exponential SFR and a star formation efficiency $f_*=0.1$, for various galaxy masses $M_{\\rm gal}$ in solar masses. Solid lines: simulations with SNe only; dotted lines: simulations with SNe and stellar winds.} \\label{R_mgal} \\end{center} \\end{figure} Our first results concern the SFR for the galaxy. For a given star formation efficiency $f_*$, a longer SFR tends to produce a galactic wind that reaches a larger extent, but this wind will be less dense. By increasing the star formation efficiency, we can produce a wind that reaches a larger extent and has a higher metallicity near its front. In some cases, the energy deposited by the stars is sufficient to completely expel the ISM. When it happens, star formation is shut down, and the galactic wind enters the post-SN phase prematurely. Hence, paradoxically, an increase in the star formation rate can sometimes result in a galactic wind that reaches a smaller extend. This happens with galaxies of masses $M_{\\rm gal}=10^8\\msun$ or less, because their shallow potential well enables the complete removal of the ISM by the galactic wind. For galaxies with mass above $10^{11}\\msun$, the material ejected in the IGM always falls back onto the galaxy, no matter the value of $f_*$. Therefore, in the case of energy-driven galactic winds, lower-mass galaxies are more likely to be the ones responsible for enriching the IGM and potentially perturbing the formation of nearby galaxies. Below $10^{11}\\msun$, the extent of the galactic wind and its mass and metal content both increase with the mass of the galaxy at constant $f_*$. With different values of $f_*$, a less massive galaxy can sometimes produce a larger wind. Our current model does not take into account the effect of Type Ia SNe. These are difficult to include, because of the uncertainties on the lifetime of the progenitors. The simulations presented in this paper start at redshift $z=15$, and end between redshifts $z=9$ and 6. The corresponding time periods are shorter than $1\\,{\\rm Gyr}$, which is shorter than the lifetime of several Type Ia progenitors. The energy produced by Type Ia SNe is about 20\\% of the energy produced by Type II SNe (see Fig.~10 of \\citealt{benson10}). Hence, including the Type Ia SNe would result in a slightly larger final radius for the outflow. A Type Ia SNe can produce up to 7 times more iron than a Type II SNe (see model W7 in \\citealt{nomotoetal97}), and their contribution to the iron enrichment of the ISM become important after $1\\,{\\rm Gyr}$ \\citep{wiersma10}. Hence, the abundances of iron we present in this paper are underestimated. But because of the delay, the additional iron produced would remain in the inner parts of the galactic wind. We have assumed a minimum value of $M_i=8\\msun$ for the minimum mass of SNe progenitors. However, the correct value is actually quite uncertain. We did a few simulations with minimum masses of $6\\msun$ and $10\\msun$. Our preliminary results show differences of order 10\\% in the final radius of the outflow, and of order 20-60\\% in the total mass ejected, with the largest effect occurring when $M_i$ is reduced. We intend to study this in more detail in the future. To conclude, properties of galactic winds depend on the host galaxy properties, such as the mass or star formation efficiency. The history of the ISM enrichment plays a determinant role in the chemical composition and extent of the galactic wind, and therefore its ability to enrich the IGM. The next step will consist of implementing this galactic outflow model into large-scale cosmological simulations of galaxy formation and the evolution of the IGM. These will be the first simulation of this kind to include a detailed treatment of the stellar winds and their impact on the chemical enrichment of the IGM" }, "1112/1112.3071_arXiv.txt": { "abstract": "{Most galaxy evolutionary models require quasar feedback to regulate star formation in their host galaxies. In particular, at high redshift, models expect that feedback associated with quasar-driven outflows is so efficient that the gas in the host galaxy is largely swept away or heated up, hence suppressing star formation in massive galaxies. We observationally investigate this phenomenon by using VLT-SINFONI integral field spectroscopy of the luminous quasar 2QZJ002830.4-281706 at z=2.4. The spectra sample the optical emission lines redshifted into the near-IR. The [OIII]$\\lambda$5007 emission-line kinematics map reveals a massive outflow on scales of several kpc. The detection of narrow H$\\alpha$ emission reveals star formation in the quasar host galaxy, with $\\rm SFR\\sim 100~M_{\\odot}~yr^{-1}$. However, the star formation is not distributed uniformly, but is strongly suppressed in the region with the highest outflow velocity and highest velocity dispersion. This result indicates that star formation in this region is strongly quenched by the quasar outflow, which is cleaning the galaxy disk of its molecular gas. This is one of the first direct observational proofs of quasar feedback quenching the star formation at high redshift. } ", "introduction": "Most of the recent galaxy formation models invoke energetic outflows as a way to regulate the evolution of galaxies throughout the cosmic epochs \\citep{Silk,Bower,Springel}. In particular, quasars are expected to drive powerful outflows that eventually expel most of the gas in their host galaxies, thereby quenching both star formation and further black hole accretion \\citep[e.g. ]{Granato,DiMatteo,Menci,Bower,Hopkins,King}. According to those models, these quasar driven outflows are required to prevent massive galaxies from overgrowing, hence explaining the shortage of very massive galaxies in the local universe, and are responsible for the red color and gas poor properties of local elliptical galaxies. Massive, large-scale outflows have been detected in the hosts of local quasars \\citep[e.g. ][]{Feruglio,Fischer,Sturm,Rupke}. However, models expect that most of the quasar feedback action occurs at high redshift, when quasars reach their peak activity (z$\\sim$2) and when star formation in the most massive galaxies is observed to decline. Evidence of outflows in luminous quasars has been detected up to very high redshift \\citep[e.g.][]{Allen,Maiolino,Alexander}. Indications that the strength of these AGN driven outflows anticorrelates with the starburst contribution to the infrared luminosity has been obtained in reddened quasars by \\cite{farrah11}. However, direct observational evidence that high-z quasar driven outflows quench star formation in their host galaxies is still missing. Here we present VLT-SINFONI near-IR integral field spectra of the quasar 2QZJ002830.4-281706 (hereafter 2QZ0028-28). This object, at $z = 2.401$, was taken from the sample of strong [OIII]$\\lambda$5007 emitters discovered by \\cite{Shemmer}, and it is one of the most luminous quasars known. In this letter we show that the spatially resolved kinematics of the [OIII]$\\lambda$5007 line clearly reveals a prominent outflow on scales of several kpc. Even more interesting, the star formation traced by narrow H$\\alpha$ emission is suppressed in the region characterized by the strongest outflow. We suggest that this is the first observational evidence of feedback associated with a quasar-driven outflow that quenches star formation at high redshift, or one of the first. \\begin{figure}[!] \\centering \\includegraphics[height=0.95\\linewidth,angle=90]{fig1.ps} \\vspace{0.05cm} \\caption{Upper panel: 2QZ0028-28 H band spectrum extracted from the central 0.5 arcsec, along with the various components used for the fit (see Appendix~\\ref{app1} for details). Vertical dashed lines indicate the rest frame wavelength of each line, by taking the [OIII]$\\lambda$5007 line peak for reference. Lower panel: Residuals of the fit. The green vertical lines enclose a the spectral zone affected by strong sky line residuals.} \\label{hband} \\end{figure} \\begin{figure*}[!] \\includegraphics[height=0.4\\textwidth,angle=90]{fig2a.ps} \\hspace{1.7cm} \\includegraphics[height=0.4\\textwidth,angle=90]{fig2b.ps} \\caption{Left panel: Velocity field (first moment map) of the [OIII]$\\lambda$5007 line, showing the prominent excess of blueshifted gas with a bow-like morphology SE of the nucleus. White contours are at 330, 360, and 390 km/s. Right panel: Velocity dispersion (second moment map) of the [OIII]$\\lambda$5007 line, showing the excess of dispersion in the SE region. White contours are at 680 and 700 km/s. In both maps the black contours trace the continuum.} \\label{moments} \\end{figure*} ", "conclusions": "By using near-IR integral field spectroscopic observations we have revealed a powerful outflow in the host galaxy of the quasar 2QZ0028-28 at z=2.4. The outflow was revealed by the velocity field traced by the [OIII]$\\lambda$5007 line, redshifted into the H-band. We estimated that the outflow rate of ionized gas is about $\\rm 200~M_{\\odot}~yr^{-1}$, which is, however, a lower limit of the total gas outflow rate. Both the high outflow velocity ($\\rm > 1000~km/s$) and the fact that the wind is mostly traced by the [OIII] line (produced primarily in the NLR) strongly suggest that the outflow is mostly driven by the quasar. The outflow is not symmetric, the highest velocities and highest velocity dispersion are found in the region SE of the nucleus. In the K-band, our data clearly reveal the presence of narrow H$\\alpha$ emission tracing star formation in the host galaxy, on scales of several kpc and with a rate of about $\\rm 100~M_{\\odot}~yr^{-1}$. However, star formation is not distributed uniformly in the host galaxy, but is mostly found in the regions not directly invested by the strong outflow. Instead, star formation is heavily suppressed in the SE region where the strongest outflow is detected. This observational result supports models invoking quasar feedback to quench star formation in massive galaxies at high redshift." }, "1112/1112.3592_arXiv.txt": { "abstract": "{In this paper a new supernova catalogue containing data for 5526 extragalactic supernovae that were discovered up to 2010 December 31 is presented. It combines several catalogues that are currently available online in a consistent and traceable way. During the comparison of the catalogues inconsistent entries were identified and resolved where possible. Remaining inconsistencies are marked transparently and can be easily identified. Thus it is possible to select a high-quality sample in a most simple way. Where available, redshift-based distance estimates to the supernovae were replaced by journal-refereed distances. Examples of statistical studies that are now possible with this new catalogue are presented in this paper.} ", "introduction": "The observation and study of supernovae (SNe) have had significant scientific impact in the past 30 years. Most prominently, \\object{SN1987A} enabled probing neutrino properties such as mass \\citep{SNIa_neutrino_mass}, electric charge \\citep{SNIa_neutrino_charge} and magnetic moment \\citep{SNIa_neutrino_magnetic_moment}. Recently, SNe of type Ia have been used as ``standard candles'' to measure distances up to cosmological scales \\citep{SNIa_distances_calibration}. In contemporary scientific research SNe are still important. Young SNe for instance may be identified as cosmic-ray accelerators by measuring high-energetic gamma or neutrino radiation shortly after the SN explosion \\citep{pulsarmodel,andobeacom,rmwaxman}. Zwicky started the first systematic SN search in 1932 at Caltech, but he and his collaborators detected the first SNe only after the newly built 18\\arcsec Schmidt telescope at Palomar Observatory was put into operation \\citep{zwicky_first_systematic_search}. A first overall SN list was published by \\citet{zwicky_first_catalogue} and contained 54 SNe discovered between 1885 and 1956. This Palomar Supernova Master List has been updated and occasionally republished \\citep{zwicky1965,kowal1971,sargent1974}. Two other SN lists were published contemporaneously \\citep{karpwicz1968,flin1979}. A new and revised supernova catalogue compiled from the Palomar Supernova Master List has been published by \\citet{barbon1984}. This catalogue then became the Asiago SN catalogue \\citep[ASC;][]{barbon1989,barbon1999} and has received great recognition so far. In 1993 another catalogue, the Sternberg Astronomical Institute (SAI) SN catalogue (SSC) was first published \\citep{SSC1993,SSC2004}. Today, running SN catalogues are easily accessible over the internet. The most important ones are the list of SNe maintained by the Central Bureau for Astronomical Telegrams (CBAT)\\footnote{http://www.cfa.harvard.edu/iau/lists/Supernovae.html}, the electronic version of the ASC\\footnote{http://web.oapd.inaf.it/supern/cat/} and the electronic version of the SSC\\footnote{http://www.sai.msu.su/sn/sncat/}. The latter two can also be accessed via VizieR\\footnote{http://vizier.u-strasbg.fr/cgi-bin/VizieR}. The recently growing interest in SNe is demonstrated in Fig.~\\ref{fig:snperyear}. The number of detected SNe per year has grown almost exponentially after the discovery of \\object{SN1987A}, reaching a current discovery rate of a few hundred SNe per year. This gain can be mostly attributed to distant SNe, because the discovery rate of bright SNe has not significantly increased. The need for a new supernova catalogue has arisen during a search for high-energetic neutrinos from young SNe with the AMANDA neutrino telescope \\citep{arxiv,Diplomarbeit}. In order to enhance the sensitivity of the analysis, the SNe were stacked (i.e. adding up small signals that may not be significant individually). Nearby SNe contribute most to the expected signal, but the distance derived from the redshift of the SN host galaxy gives only a very rough estimate. Hence, a new catalogue was compiled that includes journal-refereed distances of the host galaxies and therefore allows a more realistic signal estimate. \\begin{figure}[t] \\includegraphics*[width=\\columnwidth]{SN_discoveries_per_year.pdf} \\caption{Number of detected SNe per year. The discovery year of \\object{SN1987A} is marked with a dashed line. The fraction of bright SNe, which have a magnitude at maximum $ < 15^{\\rm{m}}$, is indicated by the hatched area. Selection criteria for the data are described in Sect. \\ref{sec:studies}.} \\label{fig:snperyear} \\end{figure} A second motivation for a new catalogue was that a comparison between the three SN catalogues (CBAT, ASC, SSC) available online revealed significant inconsistencies in the listed information. In some cases these can be attributed to simple typographical errors, while others are qualitative differences in the given information. The concept of the new catalogue is to list the undisputed information and flag differences. Therefore, the new catalogue also serves as a meta-catalogue of the current online SN catalogues. A subset of high-quality SNe with more reliable information can be selected with the meta-data. ", "conclusions": "In this paper a new unified catalogue of three existing supernova catalogues is presented. During the unification procedure several inconsistencies between the catalogues were identified and errors corrected. Remaining inconsistencies are transparently marked and enable the user to select high-quality subsamples. Wherever possible, redshift-independent distance estimates were added to provide a more realistic distance than the redshift only. Several examples for statistical studies that make use of the USC capabilities are shown. The corrected and extended information contained in this unified catalogue is intended to improve the use of SN-related information, e.g. in SN-related analyses in astro-particle physics. An example is \\citet{arxiv} which, initiated this work." }, "1112/1112.1845_arXiv.txt": { "abstract": "{As of today, over 50 planetary systems have been discovered in binary star systems, some of which have binary separations that are smaller than $20$\\,AU. In these systems the gravitational forces from the binary have a strong influence on the evolution of the protoplanetary disk and hence the planet formation process.} {We study the evolution of viscous and radiative circumstellar disks under the influence of a companion star. We focus on the eccentric $\\gamma$~Cephei and $\\alpha$~Centauri system as examples and compare disk quantities such as disk eccentricity and precession rate to previous isothermal simulations.} {We performed two-dimensional hydrodynamical simulations of the binary star systems under the assumption of coplanarity of the disk, host star and binary companion. We used the grid-based, staggered mesh code \\texttt{FARGO} with an additional energy equation to which we added radiative cooling based on opacity tables.} {The eccentric binary companion perturbs the disk around the primary star periodically. Upon passing periastron, spirals arms are induced that wind from the outer disk towards the star. In isothermal simulations this results in disk eccentricities up to $ e_\\mathrm{disk} \\approx 0.2 $, but in more realistic radiative models we obtain much smaller eccentricities of about $ e_\\mathrm{disk} \\approx 0.04 - 0.06 $ with no real precession. Models with varying viscosity and disk mass indicate that disks with less mass have lower temperatures and higher disk eccentricity.} {The fairly high disk eccentricities, as indicated in previous isothermal disk simulations, implied a more difficult planet formation in the $\\gamma$~Cephei system caused by the enhanced collision velocities of planetesimals. We have shown that under more realistic conditions with radiative cooling the disk becomes less eccentric and thus planet formation may be made easier. However, we estimate that the viscosity in the disk has to very low, with $\\alpha \\lesssim 0.001$, because otherwise the disk's lifetime will be too short to allow planet formation to occur along the core instability scenario. We estimate that the periodic heating of the disk in eccentric binaries will be observable in the mid-IR regime.} ", "introduction": "Presently, about 50 planets are known to reside in binary stars systems. In all systems with solar-type stars the planet orbits one of the stars while the secondary star acts as a perturber, i.e. these are in a so-called S-type configuration \\citep{1986A&A...167..379D}. Recently, planets seem to have been detected in evolved binary star systems that are in a circumbinary (P-type) configuration \\citep[e.g.][]{2011A&A...526A..53B}. The main observational characteristic of the known planets in binary stars system have been summarized by \\citet{2004A&A...417..353E,2006ApJ...646..523R,2007A&A...462..345D}. Most planets are in binaries with very large separations with semi-major axis beyond 100\\,AU, in particular when detected by direct imaging. Observationally, there is evidence for fewer planets in binaries with a separation of less than about 100\\,AU \\citep{2007A&A...474..273E}, in accordance with the expectation that binarity constitutes a challenge to the planet formation process. Despite this, there are several systems with a quite close binary separation such as $\\gamma$~Cephei \\citep{1988ApJ...331..902C}, Gliese\\,86 \\citep{2000A&A...354...99Q,2001A&A...370L...1E}, HD\\,41004 \\citep{2004A&A...426..695Z}, and HD\\,196885 \\citep{2008A&A...479..271C}. As known from several theoretical studies, the presence of the secondary renders the planet formation process more difficult than around single stars. Owing to the gravitational disturbance, the protoplanetary disk is hotter and dynamically more excited such that the coagulation and growth process of planetesimals as well as the gravitational instability process in the early phase of planet formation are hindered \\citep{2000ApJ...537L..65N,2004A&A...427.1097T,2006Icar..183..193T}. This is particularly true for the mentioned closer binaries with an orbital separation of about $20$\\,AU. Hence this theoretical challenge has put these tighter binaries into the focus of studies on planet formation in binary stars \\citep{2007ApJ...660..807Q,2008MNRAS.386..973P,2008A&A...486..617K} and on their dynamical stability \\citep{2004RMxAC..21..222D,2006ApJ...644..543H}. The most famous example in this category is the $\\gamma$~Cephei system, which has been known for over 20 years to contain a protoplanet \\citep{1988ApJ...331..902C,2003ApJ...599.1383H}. In recent years the orbital parameter of this system have been updated \\citep{2007A&A...462..777N,2007ApJ...654.1095T} and the basic binary parameters are quoted in Table~\\ref{tab:standardmodel} (below) because $\\gamma$~Cephei is our standard model. At the end of the paper we will briefly discuss the $\\alpha$~Centauri system as well. Because planet formation occurs in disks, their dynamical structure is of crucial importance in estimating the efficiency of planetary growth processes. The prime effect of the secondary star is the truncation of the disk owing to tidal torques \\citep{1994ApJ...421..651A}. The truncation radius depends on the mass ratio of the binary, its eccentricity and the viscosity in the disk. An additional effect, namely the excitation of eccentric modes in the disk, which has been noticed previously in close binary stars \\citep{1988MNRAS.232...35W,1991ApJ...381..259L,2008A&A...487..671K}, has recently drawn attention in studies of planet harboring binary stars as well \\citep{2008MNRAS.386..973P,2008A&A...486..617K}. In these studies it was shown that despite the high eccentricity of a $\\gamma$~Cephei type binary the disk became eccentric with an average eccentricity of about $e_{\\rm disk} \\approx 0.12$ and a coherent disk precession \\citep{2008A&A...486..617K}. However, this leaves the question of numerical effects \\citep{2008MNRAS.386..973P}. Subsequently the influence of self-gravity of the disk has been analyzed by \\citet{2009A&A...508.1493M}, who concluded that the inclusion of self-gravity leads to disks with lower eccentricity on average. Interestingly, there may be observational evidence for tidal interactions between a companion on an eccentric orbit and a circumstellar disk around the primary star \\citep{2010A&A...517A..16V}, as inferred from variable brightness on longer timescales. The possibility to detect an eccentric disk through spectral line observations has been analyzed recently by \\citet{2011A&A...528A..93R}. In all simulations mentioned above, the disk has typically been modeled using a fixed radial temperature distribution. This simplifies the numerics such that no energy equation has to be solved but of course it lacks physical reality. In this work we will extend previous studies and consider a much improved treatment of the energy equation. We will do so within the two-dimensional, flat-disk approximation following \\citet{2003ApJ...599..548D} and \\citet{2008A&A...487L...9K}. Here, the viscous heating of the disk and the radiative losses are taken into account. We will study the structure and dynamical evolution of disks with different masses and binary parameter, and will analyze the influence of numerical aspects, in particular boundary conditions. Our first target of interest will be the well-studied system $\\gamma$~Cephei and we will present results on the $\\alpha$~Centauri system subsequently. ", "conclusions": "We have investigated the dynamics of a protostellar disk in binary star systems using specifically the orbital parameter of $\\gamma$~Cephei and $\\alpha$~Centauri. We assumed a coplanar system and used a two-dimensional hydrodynamical code to evolve the non-self-gravitating disk. Extending previous simulations, we included internal viscous heating given by an $\\alpha$-type viscosity prescription, and radiative cooling from the disk surface. In a first set of simulations we investigated locally \\textit{isothermal} disks for different disk temperatures. We showed that disks in binaries of the $\\gamma$~Cephei type with a standard thickness of $H/r = 0.05$ become eccentric ($e_{\\rm disk} \\approx 0.2$) showing a coherent disk precession. This agrees well with previous simulations by \\citet{2008A&A...486..617K} and \\citet{2008MNRAS.386..973P}. Varying the temperature in the disk, we showed that the magnitude of the disk's eccentricity becomes lower when the disk thickness increases. For disks with $H/r \\gtrsim 0.065$ the mean average eccentricity has dropped below 0.08 and the disks do not show a precession anymore. Then we studied more realistic disks with internal heating and radiative cooling, varying the disk's mass. In all cases we found relatively low eccentricities and no precession. We attribute the lack of eccentricities firstly to the increased disk height, which is, in particular for the more massive disks, higher than the standard value (see Fig.~\\ref{fig:gc-hr}). Secondly, in the full radiative models the disk's dynamical behavior is more adiabatic compared to the locally isothermal case. Then, through compressional heating ($p dV$-work), kinetic energy is transferred to internal energy, which leads to a reduced growth of disk eccentricity. We have checked that purely adiabatic models show an even lower disk eccentricity than the radiative models. Hence, the radiative case lies between the adiabatic and isothermal, as expected. Because the disk's energy balance is determined via the viscosity, we changed the value of the parameter $\\alpha$ ranging from $0.005$ to $0.04$, all values that are consistent with the results of MHD-turbulent accretion disks. Here, we found that only the disk with the highest $\\alpha$ becomes eccentric. The reason for this rise is the larger disk radius, which leads to an enhancement of the tidal torques from the secondary. We note that the disk's outer radius in our models still lies well inside the 3:1 resonance with the binary. According to the linear instability model by \\citet{1991ApJ...381..259L}, the disk eccentricity is excited through the 3:1 resonance and hence, the disk should be sufficiently large, a condition which is fulfilled only for small mass ratios, $q=M_\\mathrm{secondary}/M_\\mathrm{primary}$. However, as shown by \\citet{2008A&A...486..617K}, disks in binary star systems with large mass ratios can turn eccentric as well, even though the disks are small, a feature confirmed in our simulations. The inferred short lifetime of disks with standard viscosities is slightly alarming with respect to planet formation in these systems. In the core-accretion scenario planet formation proceeds along a sequence of many steps that take a few Myr. For disks to persist this long in $\\gamma$~Cephei-type binaries a very low viscosity of $\\alpha \\lesssim 10^{-4}$ seems to be required. In the gravitational instability scenario the timescale for planet formation is much shorter and hence, this scenario may be favored by our findings. Observationally, several recent studies indeed suggest that the lifetime of disks in young binary stars is significantly reduced compared to disks around single stars \\citep{2009ApJ...696L..84C,2010ApJ...709L.114D,2011arXiv1109.4141K}. Dynamically, this behavior is expected, because the perturbation of the companion star leads quite naturally to an increased mass loss of the disk, details depending on the binary separation and eccentricity. A very critical phase in the core accretion scenario, in particular in binary stars, is the initial growth of meter to km-sized planetesimals. Here, the growth depends on the successful sticking of the two collision partners. Since the relative velocity of the bodies is increased in binary stars, planetary growth will be significantly hindered by the presence of a companion, see e.g. \\citet{2006Icar..183..193T,2011CeMDA.tmp...40T} and references therein. Here our results indicate that planetesimal growth is less negatively influenced because the disk eccentricity is reduced for more realistic radiative disks. As has been shown, eccentric disks tend to increase the mutual relative velocities of embedded objects, in particular of different sizes, because of the misaligned periastrons of the particles \\citep{2007arXiv0705.3421K,2008MNRAS.386..973P}. Hence, a radiative disk with a low viscosity could help to promote planetesimal growth. However, it remains to be seen how the inclusion of stellar irradiation (from both stars) influences the dynamics. Owing to the additional heating of the disk, we expect even more mass loss from the system and possibly higher disk eccentricities because the disks are more isothermal and will have a larger radius. Previous studies have indicated that a in mutually inclined system planetesimal growth may be enhanced because planetesimals can be size-sorted in differently inclined planes \\citep{2009ApJ...698.2066X,2009A&A...507..505M}. However, recently \\citet{2011A&A...528A..40F} showed through full 3D hydrodynamical studies that for inclined binaries the relative velocities of different sized planetesimals increases through inclinations effects. They conclude that for inclined systems planetesimal formations can take place only for very distant binary stars with $a_{\\rm bin} \\gtrsim 60$~AU. However, the simulations considered only isothermal disks, and it remains to be seen how radiative effects influence the disk. But full 3D radiative simulations are still beyond the present computational possibilities, because thousands of orbital timescales of the disk will have to be calculated." }, "1112/1112.1074_arXiv.txt": { "abstract": "{ Every massive globular cluster (GC) is expected to harbour a significant population of cataclysmic variables (CVs). In this review, I first explain why GC CVs matter astrophysically, how many and what types are theoretically predicted to exist and what observational tools we can use to discover, confirm and study them. I then take a look at how theoretical predictions and observed samples actually stack up to date. In the process, I also reconsider the evidence for two widely held ideas about CVs in GCs: (i) that there must be many fewer {\\em dwarf novae} than expected; (ii) that the incidence of magnetic CVs is much higher in GCs than in the Galactic field.} ", "introduction": "Globular clusters (GCs) are old, gravitationally bound stellar systems that typically contain $\\sim 10^6$ stars. Some of these stars are binaries, and some of these binaries are cataclysmic variables (CVs), i.e. systems in which a white dwarf (WD) accretes material from a roughly main sequence (MS) companion. These CV populations in GCs deserve special attention for at least three reasons: {\\em 1 The Globular Cluster Perspective:} the late dynamical evolution of a GC is thought to be driven largely by its close binary (CB) population (e.g. Hut et al. 1992). However, the dominant non-interacting MS-MS binaries are difficult to detect and study in GCs. CVs can therefore be used as convenient tracers of the underlying CB populations for studies of GC dynamics and evolution. {\\em 2 The Cataclysmic Variable Perspective:} in principle, GCs can provide us with sizeable samples of CVs at known distance and (to some extent) age. Such samples might allow critical tests of theoretical CV/binary evolution scenarios. An interesting complication is that not all CVs in GCs are likely to have formed and evolved in isolation: many are likely to have been produced, or at least affected, by dynamical interactions (see later). {\\em 3 The Supernova Perspective:} it has been suggested that GCs might be significant Type Ia Supernova factories (Shara \\& Hurley 2002), at least in elliptical galaxies (Ivanova et al. 2006). All SN Ia progenitor populations are thought to be close relatives of CVs (e.g. double WD systems, supersoft sources, WD + red giant binaries), so CVs can again serve as a useful tracer population for these progenitors. In fact, it is even still possible that some CVs might be SN Ia progenitors themselves (Thoroughgood et al. 2001; Zorotovic, Schreiber \\& G\\\"{a}nsicke 2011). ", "conclusions": "I hope to have shown that GC CVs should be of great interest to anybody studying the dynamical evolution of clusters, the binary evolution of CVs or the SN Ia progenitor problem. I also hope I have made it clear that great strides have been taken in recent years in finding and understanding GC CV populations, thanks mainly to the availability of Chandra and HST. Indeed, we are finally discovering significant numbers of CVs in any given massive GC, though still not the hundreds that are theoretically predicted to lurk there. I have also discussed the theoretical and observational hints that field and GC CV populations may differ systematically, although I feel that the jury is still out regarding most of these putative differences. The next big advances in the field are likely to come from one or both of the following directions. First, if we really wish to test for the presence of the predicted hundreds of CVs per cluster, we have to be sensitive to {\\em faint} CVs. How deep? Ideally, deep enough to discover systems like WZ Sge, at $M_{V} \\sim 13$ in quiescence. This is hard -- it requires reaching $m_v \\sim 27$ or so. But it is not impossible: Cohn et al. (2010) have essentially done this in NGC~6397 and uncovered two likely WZ-Sge-like CVs, the first such systems in any GC. An interesting short-cut might be to place limits on the number of faint CVs in GCs by considering the {\\em integrated} X-ray luminosity of individually undetected systems (e.g. Haggard, Cool \\& Davies 2009). Second, we need to obtain orbital periods for a significant sample of GC CVs, so that we can start to make meaningful comparisons to theoretically predicted period distributions." }, "1112/1112.4651_arXiv.txt": { "abstract": "We study the impact of the late time dynamical evolution of ejecta from core-collapse supernovae on $\\nu p$-process nucleosynthesis. Our results are based on hydrodynamical simulations of neutrino wind ejecta. Motivated by recent two-dimensional wind simulations, we vary the dynamical evolution during the $\\nu p$-process and show that final abundances strongly depend on the temperature evolution. When the expansion is very fast, there is not enough time for antineutrino absorption on protons to produce enough neutrons to overcome the $\\beta^+$-decay waiting points and no heavy elements beyond $A=64$ are produced. The wind termination shock or reverse shock dramatically reduces the expansion speed of the ejecta. This extends the period during which matter remains at relatively high temperatures and is exposed to high neutrino fluxes, thus allowing for further $(p,\\gamma)$ and $(n,p)$ reactions to occur and to synthesize elements beyond iron. We find that the $\\nu p$-process starts to efficiently produce heavy elements only when the temperature drops below $\\sim 3$~GK. At higher temperatures, due to the low alpha separation energy of $^{60}$Zn ($S_{\\alpha} = 2.7$~MeV) the reaction $^{59}$Cu$(p,\\alpha)$$^{56}$Ni is faster than the reaction $^{59}$Cu$(p,\\gamma)$$^{60}$Zn. This results in the closed NiCu cycle that we identify and discuss here for the first time. We also investigate the late phase of the $\\nu p$-process when the temperatures become too low to maintain proton captures. Depending on the late neutron density, the evolution to stability is dominated by $\\beta^+$ decays or by $(n,\\gamma)$ reactions. In the latter case, the matter flow can even reach the neutron-rich side of stability and the isotopic composition of a given element is then dominated by neutron-rich isotopes. ", "introduction": "\\label{sec:intro} Neutrino-driven winds from core-collapse supernova explosions contribute to the synthesis of elements beyond iron. After the explosion, the hot proto-neutron star cools emitting neutrinos. These neutrinos interact with the stellar matter and deposit energy in the outer layers of the proto-neutron star leading to a supersonic outflow known as neutrino-driven wind \\cite[]{duncan.shapiro.wasserman:1986}. Although neutrino-driven winds were considered the site where heavy elements are produced by the r-process \\cite[]{Woosley94}, recent simulations \\cite[]{arcones.janka.scheck:2007, Huedepohl.etal:2010, Fischer.etal:2010, Roberts.etal:2010} cannot reproduce the extreme conditions required for producing heavy r-process elements \\cite[see e.g.,][]{hoffman.woosley.qian:1997, Otsuki.Tagoshi.ea:2000, Thompson.Burrows.Meyer:2001}. The wind entropy is too low (less than $100\\,k_{\\mathrm{B}}/\\mathrm{nuc}$) and, even more significant, the ejecta is proton rich (the electron fraction $Y_e$ remains above 0.5 during seconds, see \\cite{Huedepohl.etal:2010, Fischer.etal:2010}). Even if the r-process does not take place in every neutrino-driven wind, lighter heavy elements (e.g., Sr, Y, Zr) can be synthesized in this environment as suggested by \\cite{Qian.Wasserburg:2001}. In proton-rich conditions \\cite{Froehlich06} showed that elements beyond $^{64}$Ge can be synthesized. \\cite{Wanajo.Janka.Mueller:2011} found Sr, Y, Zr in small pockets of neutron rich material ejected after the explosion of low mass progenitors. Recently, \\cite{Arcones.Montes:2011} performed a systematic nucleosynthesis study that strongly supports the production of lighter heavy elements in proton- and neutron-rich neutrino-driven winds. In proton-rich winds, charged particle reactions (alpha and proton captures) build nuclei up to $^{56}$Ni and even up to $^{64}$Ge once the temperature drops below 3~GK. Due to their long beta-decay lifetimes and low thresholds for proton capture, the nuclei $^{56}$Ni and $^{64}$Ge act as bottlenecks that inhibit the production of heavier elements. In the $\\nu p$-process, their decay is sped up by $(n,p)$ reactions, with the neutrons produced by antineutrino absorption on the abundant free protons. This allows for the production of elements beyond iron and may explain the origin of light p-nuclei \\cite[]{Froehlich06, Pruet.Hoffman.ea:2006, Wanajo:2006, Wanajo.etal:2011}. The synthesis of elements by the $\\nu p$-process depends thus on neutrino spectra and luminosities but also on the dynamical evolution as matter expands through the slow, early supernova ejecta. This produces a wind termination shock or reverse shock where kinetic energy is transformed into internal energy \\cite[]{arcones.janka.scheck:2007}. The reverse shock has a big impact on the nucleosynthesis because temperature and density increase and the expansion is strongly decelerated. This hydrodynamical feature has been extensively studied for r-process nucleosynthesis \\cite[]{Qian.Woosley:1996, Sumiyoshi00, Wanajo:2007, Panov.Janka:2009, Arcones.MartinezPinedo:2011}. Recently, \\cite{Wanajo.etal:2011} have also explored the relevance of the reverse shock on the $\\nu p$-process. Motivated by their work and by new 2D hydrodynamical simulations of the neutrino-driven wind \\cite[]{Arcones.Janka:2011}, we investigate here the wind termination shock to gain further insights on the dynamical evolution relevant for the $\\nu p$-process. Here we use a trajectory from hydrodynamical simulations (Sect.~\\ref{sec:dynamics}) combined with a complete nucleosynthesis network (Sect.~\\ref{sec:netw}). In Sect.~\\ref{sec:results} we present our results where we analyze the impact of the wind termination temperature (Sect.~\\ref{sec:constT}), the temperature jump at the reverse shock (Sect. \\ref{sec:rs}), and the influence of late temperature evolution and the decay to stability (Sect~\\ref{sec:long}). Our conclusions are summarized in Sect.~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} We have studied the impact of the supernova dynamical evolution on the $\\nu p$-process and how the wind termination affects the synthesis of elements beyond iron. For a robust and strong production of nuclei with $64 \\lesssim A \\lesssim 110$ there is an optimal wind termination temperature at $\\approx$2~GK as it was found by \\cite{Wanajo.etal:2011}. Also in agreement with their work, we have shown that if the wind termination occurs at low temperatures ($T_{\\mathrm{wt}}\\approx 1$~GK), matter stays too short time in the optimal temperature range for $\\nu p$-process nucleosynthesis, $\\sim 1 -2 $~GK. Moreover, the electron antineutrino flux quickly becomes too small to produce the necessary neutrons to overcome the $\\beta^{+}$-decay waiting points. This hinders the $\\nu p$-process and consequently the efficient synthesis of heavy nuclei. Therefore, the wind termination temperature determines the heaviest elements produced. We have identified an end point cycle that is key at high temperatures (around 3~GK). The close NiCu cycle inhibits the production of elements heavier than 56Ni. The reactions in this closed NiCu cycle are the following: \\begin{eqnarray} && ^{56}Ni (n,p) ^{56}Co (p, \\gamma ) ^{57}Ni (n,p) ^{57}Co (p, \\gamma ) ^{58}Ni \\nonumber \\\\ & &\\longrightarrow \\, ^{58}Ni (p, \\gamma ) ^{59} Cu(p, \\alpha ) ^{56} Ni \\, . \\nonumber \\end{eqnarray} At high temperatures the reaction $^{59}$Cu$(p,\\alpha)$ $^{56}$Ni prevents the synthesis of elements beyond the iron group. When the temperature drops below $\\approx 3$~GK, $^{59}$Cu$(p,\\gamma)$ $^{60}$Zn becomes more effective than $^{59}$Cu$(p,\\alpha)$ $^{56}$Ni. This leads to breakout from the NiCu cycle and the flow of matter can continue towards heavier nuclei. The temperature dependence of these two reactions is critical because it sets the temperature at which the $\\nu p$-process can start to synthesize elements beyond Nickel. If this cross-over temperature were slightly higher, matter would be closer to the proto-neutron star and thus under higher neutrino flux when the path starts to move towards heavier nuclei. This will significantly increase the efficiency of the $\\nu p$-process. Therefore, our results clearly motivate further investigation of these two key reactions: $^{59}$Cu$(p,\\alpha)$ $^{56}$Ni and $^{59}$Cu$(p,\\gamma)$$^{60}$Zn. Particularly relevant are the branching ratios for the decay by alpha and gamma emission of compound states in $^{60}$Zn above the proton separation energy. We have also explored for the first time the impact of the dynamical evolution after the wind termination on the synthesis of heavy elements by the $\\nu p$-process. Hydrodynamical simulations show that after the wind termination there is a transition from an expansion with almost constant temperature and density to a phase with almost constant velocity. The duration of the constant temperature phase, $\\Delta t$, strongly affects the final abundances of heavy nuclei. We have found that for $\\Delta t=0.0$~s the final flow of matter to stability occurs by neutron capture reactions, i.e\\ $(n,\\gamma)$ and $(n,p)$, and not by $\\beta^+$-decays. When the phase of constant temperature is very short ($\\Delta t < 0.5$~s), the abundances of nuclei with $640$ are of low statistical significance ($\\lsim3\\sigma$) and/or controversial. In this work, we aim to establish the detection limits of current X-ray observatories and explore requirements for next-generation X-ray telescopes for studying the WHIM through X-ray absorption lines. We analyze all available grating observations of Mrk~421 and obtain spectra with signal-to-noise ratio (S/N) of $\\sim90$ and 190 per 50 m\\AA\\ spectral bin from \\chandra\\ and \\xmm\\ observations, respectively. Although these spectra are two of the best ever collected with \\chandra\\ and \\xmm, we cannot confirm the two WHIM systems reported by Nicastro et al. in 2005. Our bootstrap simulations indicate that spectra with such high S/N can\\emph{not} constrain the WHIM with \\ovii\\ column densities $N_{\\rm OVII}\\approx10^{15}~{\\rm cm^{-2}}$ (corresponding to an equivalent widths of 2.5 m\\AA\\ for a Doppler velocity of $50~{\\rm km~s^{-1}}$) at $\\gsim3\\sigma$ significance level. The simulation results also suggest that it would take $>60$ Ms for \\chandra\\ and 140 Ms for \\xmm\\ to measure the $N_{\\rm OVII}$ at $\\ge4\\sigma$ from a spectrum of a background QSO with flux of $\\sim0.2$ mCrab (1 Crab = $2\\times10^{-8}~{\\rm erg~s^{-1}~cm^{-2}}$ at 0.5-2 keV). Future X-ray spectrographs need to be equipped with spectral resolution $R\\sim4000$ and effective area $A\\ge100~{\\rm cm^2}$ to accomplish the similar constraints with an exposure time of $\\sim2$ Ms and would require $\\sim11$ Ms to survey the 15 QSOs with flux $\\gsim0.2$ mCrab along which clear intergalactic \\ovi\\ absorbers have been detected. ", "introduction": "\\label{sec:intro} Identifying the ``missing baryons'' is one of the major tasks of modern cosmology. Observations of the microwave background (e.g., \\citealt{kom11}) and the big bang nucleosythesis model combined with the measurement of deuterium abundance (e.g., \\citealt{bur01, ome06, pet08}) are converging on the cosmological baryon density of $\\Omega_b=0.0455\\pm0.0028$. At high redshift ($z>3$) universe, baryons exist primarily in the form of cool, photoionized intergalactic medium (IGM) that is traced by Ly$\\alpha$ absorption forest lines \\citep{rau97}. In the present-day universe, matter detected in forms of photoionized IGM, stars, galaxies, intracluster medium, etc., adds up to only $\\sim50\\%$ of the baryons \\citep{shu11}, % leaving much of the inventory yet to be found \\citep{fuk98, fuk04}. Cosmological numerical simulations for large-structure formation indicate that the missing baryons are still in the IGM, but at current epoch they have been heated by gravitational shocks and galactic feedback to temperatures of $10^5-10^7$ K when they fall into the gravitational potential wells of the dark matter cosmic web filaments (e.g., \\citealt{cen99, cen06, dav01}). The IGM at these temperatures, so-called warm-hot intergalactic medium (WHIM), mainly absorbs and emits photons in the ultraviolet (UV) and X-ray wavelength bands. Searching for the missing baryons has been conducted extensively in the UV band. Indeed, besides the cool phase Ly$\\alpha$ absorbers (e.g., \\citealt{wey98, pen00, pen04}), the absorption lines of \\ciii, \\civ, \\nv, \\ovi, \\siiii, and \\siiv\\ have been routinely observed in spectra of background quasi-stellar objects (QSOs; e.g., \\citealt{tri00, tri08, pro04, leh06, dan05, dan08, dan06, thom08, yao11}), suggesting that a significant amount of baryons exist in the highly ionized absorbers. These high ions, \\ovi\\ in particular, are believed to trace the WHIM at the low-temperature end ($\\sim10^5-10^6$ K). However, all these ions can also be produced through photoionization in the intergalactic environment (\\citealt{opp09}; but see also \\citealt{tep11} and \\citealt{smi12} for a contrary view) and thus some of them may not contribute to the canonical WHIM (i.e., the shock-heated IGM expected from simulations). How many of these absorbers originate in the real WHIM is still under debate \\citep{dan08, tri08}. \\neviii\\ and broad \\hi\\ Ly$\\alpha$ absorbers (BLAs) are complementary and valuable tracers of hot absorbing WHIM in UV bands. However, there are only a few detections with marginal significance of \\neviii\\ \\citep{sav05, nar09, nar11} and detections of BLAs require very high signal-to-noise-ratio (S/N) of spectra to resolve broad and weak signals from the continua \\citep{ric06, leh07, sav11, dan10, dan11}. Converting measurement of these UV absorbers to census of baryonic matter also depends on ionization fractions and metallicity of the IGM, which in many absorbers are still poorly known. Nevertheless, BLAs and \\ovi-absorbers, although partially overlapped with photoionized Ly$\\alpha$ absorbers, are estimated to contribute an additional $\\sim25\\%$ to the baryon inventory at the current epoch (e.g., \\citealt{dan08, dan11, shu11}). X-ray observations of the intergalactic \\ovii\\ and \\oviii\\ emission/absorption features could provide essential information for establishing the existence of the WHIM and completing the baryon inventory in the local universe. At temperatures $T\\gsim10^6$ K, hydrogen and helium are nearly completely ionized \\citep{gnat07}. Without metals, the thermal gas emits/absorbs photons through bremsstrahlung, which unfortunately is optically thin even at intergalactic scales \\citep{bre07}. Because the K-shell transitions of all elements heavier than lithium and the L-shell transitions of elements heavier than iron lie in X-ray bands \\citep{pae03}, metals can greatly increase the emissivity of the gas. However, because of the density-squared dependence, the WHIM emission is expected to be weak and difficult to detect. In fact, the IGM emission has been directly observed only from the dense regions like intracluster and intragroup media (e.g., \\citealt{wang97, mul00, all02, sun09}). The WHIM emission signals were also claimed in the angular correlation of the diffuse X-ray background \\citep{sol99, urs06}, but disentangling the real WHIM signal from the unresolved point sources and local Galactic hot diffuse gas has been a major uncertainty (e.g., \\citealt{gal07}). Unlike emission lines, absorption lines measure the column densities and thus directly sample the total mass of the intervening gas along a line of sight. Oxygen is the most abundant metal element. In contrast of \\ovi\\ as a minority ionization state, \\ovii\\ is the most abundant ion for a gas at temperatures of $T\\sim10^{5.5}-10^{6.3}$ K and \\oviii\\ takes over when $T\\gsim10^{6.3}$ K \\citep{gnat07}. Thus they have advantages over \\ovi\\ in estimating the total baryon contained in highly ionized absorbers. \\ovii\\ and \\oviii, whose production ionization potentials are 138.1 eV and 739.3 eV compared to 113.9 eV of \\ovi, are also hard to produce through photoionization in the IGM environment (e.g., \\citealt{cen06a, chen03}). Theoretical calculations indicate that the \\ovii\\ column density is about 10 times higher than \\ovi\\ in a shock-heated gas (i.e., $N_{\\rm OVII}\\gsim10N_{\\rm OVI}$) whereas $N_{\\rm OVII}\\lsim3N_{\\rm OVI}$ in a photoionized gas \\citep{fur05}. Therefore observations of the IGM \\ovii\\ and \\oviii\\ absorption lines are crucial, not only for constraining the properties of the WHIM at high temperature end but also for probing the nature of the commonly observed \\ovi-absorbers. Unfortunately, most attempts at searching the X-ray WHIM absorption lines have been frustrated. \\ovii\\ and \\oviii\\ absorption lines at $z=0$ have been unambiguously detected in many background QSOs. However, all these X-ray absorptions are \\emph{inconsistent} with an intergalactic origin \\citep{fang06, bre07b, yao10}, but rather can be attributed to the Galactic diffuse hot gas (e.g., \\citealt{yao05, yao07, yao08, yao09, wang05}). Perhaps the most compelling intergalactic result is the detection of \\ovii\\ absorption at the redshift of the large galaxy structure Sculptor Wall ($z_{\\rm abs}=0.03$) along the QSO sight line H~2356-309 ($z_{\\rm QSO}=0.165$), albeit of low significance ($\\sim3\\sigma$; \\citealt{buo09, fang10}). However, because of the high number density of galaxies along the sight line, the absorption may sample the halo gas of one or more intervening galaxies with small impact distances ($\\lsim50$ kpc; \\citealt{wil10}), i.e., mimicking the \\ovii\\ and \\oviii\\ absorption lines at $z=0$. Therefore, it may not be representative of the typical WHIM. Furthermore, attempts at searching for the similar absorption features at the redshift of another large structure, Pisces-Cetus along the same sight line, failed \\citep{zap10}. All other claimed WHIM \\ovii\\ and \\oviii\\ absorptions have been highly debated. For instance, the detected \\oviii\\ absorption line at $z=0.0554$ in \\chandra\\ spectrum of PKS~2155-304 \\citep{fang02a, fang07} cannot be confirmed by the \\xmm\\ observations \\citep{cag04}, although at nearby redshift a small galaxy group and the corresponding UV \\ovi\\ and Ly$\\alpha$ absorption lines have been identified \\citep{shu98, shu03}. \\citet{nic05} obtained a spectrum with unprecedented S/N during the burst states of Mrk~421 and reported two WHIM detections at $z=0.011$ and $z=0.027$. Again, observations with \\xmm\\ cannot confirm the detections and reported significances have also been questioned \\citep{kaa06, ras07}. \\citet{wil10} recently found a galaxy filament at $z=0.027$, which makes the reported \\ovii\\ WHIM detections along the Mrk~421 sight line back to be a subject of discussions. In this work, we aim to provide detection limits of current X-ray observatories, \\chandra\\ and \\xmm, in measuring the WHIM through X-ray absorption lines, and establish requirements (e.g., effective area and spectral resolution) for next-generation X-ray telescopes. We begin our investigation by scrutinizing the controversial WHIM detections along the Mrk~421 sight line. We extensively explore all the available \\chandra\\ and \\xmm\\ observations, and then run bootstrap simulations to test the reliability of any absorption feature. The paper is organized as follows. In Section~\\ref{sec:obs} we describes the observations and our data reduction processes, and in Section~\\ref{sec:res} we report the data analysis results and compare them with previous work. In Section~\\ref{sec:sim}, we run bootstrap simulations to explore the detection limits of \\chandra\\ and \\xmm\\ and to establish requirements of next-generation X-ray telescopes. We summarize our results in Section~\\ref{sec:dis}. Throughout the paper, we adopt the atomic data from \\citet{ver96} and solar abundances from \\citet{asp09}. We conduct our data analysis within {\\sl XSPEC} (version 12.6; \\citealt{arn96}), and report the 1$\\sigma$ confidence range or 3$\\sigma$ limit for a fitting parameter. % We take the flux normalization of $1~{\\rm Crab}=2\\times10^{-8}~{\\rm erg~ s^{-1}~cm^{-2}}$ at 0.5-2.0 keV. Thus 0.2 mCrab is $2.15\\times10^{-13}~{\\rm erg~s^{-1}~cm^{-2}}$~\\AA$^{-1}$ for a flat spectrum in wavelength space, which corresponds to $2.34\\times10^{-4}~{\\rm photons~s^{-1}~cm^{-2}}$~\\AA$^{-1}$ around the \\ovii\\ K$\\alpha$ (21.602 \\AA) at the rest-frame. We define measurement significance level (SL) of an absorption line as \\begin{equation} \\label{equ:sig} SL=\\frac{EW}{\\Delta EW}, \\end{equation} where $EW$ and $\\Delta EW$ are the equivalent width and its $1\\sigma$ uncertainty for an absorption line. ", "conclusions": "\\label{sec:res} Our goal in this section is to examine the WHIM detections at redshifts $z=0.011$ and $z=0.027$. \\citet{nic05} reported the WHIM absorption from transitions of \\neix\\ K$\\alpha$, \\ovii\\ K$\\beta$, \\oviii\\ K$\\alpha$, \\ovii\\ K$\\beta$, \\nvii\\ K$\\alpha$, and \\civ\\ K$\\alpha$ (Table~2 and Figure~8 in \\citealt{nic05}), whose rest-frame wavelengths are 13.447, 18.629, 18.969, 21.602, 28.787, and 33.736 \\AA, respectively \\citep{ver96}. Therefore, we focus our attention on spectral ranges covering these transitions. Obtaining a good continuum model is crucial to absorption-line measurement. Since all coadded \\chandra\\ spectra contain the contribution from HRC observations, to account for grating order-overlapping issues (Section~\\ref{sec:obs}), we decide to fit a broad range of the spectra from 7 to 40 \\AA, which covers the transitions to be examined. For a ease of comparison of three \\chandra\\ spectra, we jointly fit the Spectra I, II, and III. We first use a Galactic-absorption modified power-law ({\\sl wabs*powerlaw} in xspec) to fit the continuum emission, with neutral hydrogen column density $N_{\\rm H}$ linked together while the power-law index $\\Gamma$ and normalization are allowed to vary. We obtain an unacceptable fit with $\\chi^2=11646$ over 3590 degrees of freedom (DOF), and we find that some ``broad'' features cannot be accounted for by this simple model. These features are mainly due to imperfect instrumental calibration around node boundaries, CCD chip gaps plus dithering effects, the oxygen absorption edge, and imperfect cross calibration between ACIS and HRC observations \\citep{mar04, nic05}. We then use Gaussian profiles to compensate the uncounted calibration residuals. To minimize the effect of the known strong ISM absorption lines on continuum modeling, we include narrow absorption lines (with widths $\\sigma\\lsim50~{\\rm km~s^{-1}}$) of \\oi, \\oii, \\cvi, \\ovi, \\ovii, and \\neix\\ K$\\alpha$ and \\ovii\\ K$\\beta$ transitions at their rest-frame wavelengths \\citep{ver96, jue06, yao09} in our spectral fitting. The centroid energies and widths of these Gaussians are linked together while the normalizations are allowed to vary in the jointly spectral fitting. We find that, besides these narrow absorption lines produced in the ISM, we need 10 broad and one narrow Gaussian profiles to obtain an acceptable fit, and locations of these Gaussian are 10.53, 13.80, 14.60, 17.90, 18.19, 19.02, 19.10, 23.24, 23.31, 23.66, and 29.89 \\AA\\ with widths ($\\sigma$) of 0.61, 2.56, 0.04, 0.00054, 0.17, 1.36, 0.10, 1.45, 1.69, 0.41, and 0.39 \\AA. The final $\\chi^2=4448.2$ with 3512 DOF and the constrained $N_{\\rm H}=1.092\\pm0.06\\times10^{20}~{\\rm cm^{-2}}$, and $\\Gamma=2.131\\pm0.001$, $2.011\\pm0.001$, $2.067\\pm0.001$ for Spectra I, II, and III, respectively. \\emph{Because the absorption line measurement depends only on the \\emph{local} continuum, and because the centroids of these Gaussian profiles do not directly superpose on the wavelengths of the putative WHIM absorption lines, these Gaussian profiles will not have any effect on the line equivalent width measurement conducted below.} \\xmm\\ spectrum (Spectrum IV) does not have the grating-order-overlapping issue, so we use the {\\sl wabs*powerlaw} model to fit the six narrow spectral ranges as plotted in Figure~\\ref{fig:rgs}. We exclude the bad pixels in our spectral analysis, and obtain $\\chi^2/DOF=1.36, 1.49, 1.49, 1.28, 1.75, 1.56$ with DOF of 137, 75, 103, 105, 116 and 104, respectively. Figures~\\ref{fig:chan-Ni}-\\ref{fig:rgs} show these best fit continua with the ISM absorption lines removed from spectral models. Now, let us examine the existence of the reported WHIM absorption lines. Figures~\\ref{fig:chan-Ni}-\\ref{fig:rgs} show the same portions of \\chandra\\ and \\xmm\\ spectra as shown in Figure~8 of \\citet{nic05}, in which they demonstrated their detections of the WHIM absorption lines at $z=0.011$ and $z=0.027$ from transitions of \\neix\\ K$\\alpha$, \\ovii\\ K$\\beta$, \\oviii\\ K$\\alpha$, \\ovii\\ K$\\beta$, \\nvii\\ K$\\alpha$, and \\civ\\ K$\\alpha$. In particular, Figure~\\ref{fig:chan-Ni} reproduces their Figure~8, as they were extracted from the same \\chandra\\ observations. Several bad pixels of the RGS accidentally fall around the 21.8 \\AA\\ region and cause an instrument feature (Figure~\\ref{fig:rgs}). Although a careful calibration could yield a correct instrumental response of the RGS, in this work we do not use the \\xmm\\ spectrum to assess the WHIM detection at $z=0.011$. Visual inspection reveals that none of the above WHIM lines reported by \\citet{nic05} is detected in the previously \\chandra\\ observations they analyzed (Figure~\\ref{fig:chan-Ni}), and in the newly available \\chandra\\ observations (Figure~\\ref{fig:chan-new}), and in the \\xmm\\ observations (Figure~\\ref{fig:rgs}). The reported WHIM detections could be due to the improper continuum placement on the spectrum. Visual inspection also reveals that there might be an absorption feature at $\\sim21.8$ \\AA\\ in Spectrum I, which corresponds to the reported \\ovii\\ K$\\alpha$ WHIM absorption line at $z=0.011$ (panel (c) in Figure~\\ref{fig:chan-Ni}). However, such an absorption feature is not visible in Spectra II and III (panel (c) in Figures~\\ref{fig:chan-new} and \\ref{fig:chan-all}), and its significance in Spectrum I depends largely on how the local continuum is placed. Please note that our spectral profiles of the local continua were obtained by jointly fitting the Spectra I, II, and III (see above), and the best-fit model seems to account for the general spectral variation reasonably well. To quantitatively assess the existence of the \\ovii\\ K$\\alpha$ WHIM line at $z=0.011$, in the joint analysis of the Spectra I, II, and III, in addition to the 10 Gaussian profiles, we add the 11th ``broad'' Gaussian to the local continuum and another narrower Gaussian with width \\footnote{Different widths yield nearly identical results as long as the line is still unresolved by the instrument. Please see Section~\\ref{sec:exp} for justifications for the line width.} $\\sigma=50~{\\rm km~s^{-1}}$ to represent the WHIM absorption line. Again, we link the line centroids and widths together in our joint fitting but allow the normalizations to vary among the Spectra I, II, and III. The best fit yields a line centroid of 21.90 \\AA\\ and $\\sigma=0.050$ \\AA\\ for the ``broad'' component and a centroid of 21.87 \\AA\\ for the narrow component. The red curves in panel (c) of Figures~\\ref{fig:chan-Ni}-\\ref{fig:chan-all} show the modified local continua. These additional components reduce the $\\chi^2$ of 26.4 by adding nine DOF in total. However, applying the best-fit models to individual spectra yields a $\\chi^2$ change of 13.7, 4.3, and 10.4 by introducing five additional DOF to spectral fitting of Spectrum I, II, and III, respectively. All these \\chandra\\ spectra can be well described by the same continuum profile with various normalizations, which has been proved to be reasonable to other wavelength ranges (panels (a)-(b) and (d)-(f) in Figures~\\ref{fig:chan-Ni}-\\ref{fig:chan-all}). Because we find essentially no improvement in fitting the Spectrum II, adding the ``broad'' Gaussian profile to the local continuum around 21.9 \\AA\\ is statistically unnessary. We next attempt to measure the equivalent widths (EW) of the putative WHIM absorption lines. Since we have not consistently detected any WHIM absorption of any transitions reported by \\citet{nic05}, we focus here on measuring the upper limit to the equivalent width (EW) of the \\ovii\\ K$\\alpha$ line, which is expected to be the most abundant transition in a broad temperature range of the WHIM (Section~\\ref{sec:intro}). At redshifts $z=0.011$ and $z=0.027$, the rest-frame \\ovii\\ K$\\alpha$ at 21.60 \\AA\\ is shifted to 21.84 and 22.19 \\AA, respectively, and therefore our measurements are obtained from the spectral range described by the continuum colored as blue in panel (c) of Figures~\\ref{fig:chan-Ni}-\\ref{fig:rgs}. To make a fair comparison with the results obtained by \\citet{nic05}, we also make similar measurements by using the continuum colored as red. We should emphasize that the latter measurements are only for comparison purposes, since the additional component to the local continuum that ``amplifies'' the significance of the absorption line at 21.8 \\AA\\ is statistically unnecessary. Our results are reported in Table~\\ref{tab:results}. In summary, our analysis and results do not support the existence of the two WHIM systems at $z=0.011$ and $z=0.027$. We analyze all the available \\chandra\\ and \\xmm\\ observations of Mrk~421, and have not seen any consistent WHIM absorption of any transition at either $z=0.011$ or $z=0.027$ reported by \\citet{nic05}. We measure the EW of the \\ovii\\ K$\\alpha$ absorption line. For the system at $z=0.027$, we obtain a firm 3$\\sigma$ upper limit of $EW<1.5$ m\\AA\\ from the spectrum (Spectrum II) extracted from the newly available \\chandra\\ observations, in contrast to $EW=2.2\\pm0.8$ m\\AA\\ obtained by Nicastro et al. For the system at $z=0.011$, we confirm the existence of a small dip in the spectrum (Spectrum I) identical to that used by \\citet{nic05}, but the similar spectral feature is absent in Spectrum II. The measurement of the line heavily depends on how the local continuum is placed, and the additional spectral component that amplifies its significance is statistically unnecessary. We obtain an upper limit of $EW<1.5$ m\\AA\\ from the new spectrum, in contrast to $EW = 3.0^{+0.9}_{-0.8}$ m\\AA\\ reported by \\citet{nic05}. The findings in this work are consistent with the previous investigation made by \\citet{kaa06}, in which the authors scrutinized the same data set used by Nicastro \\etal\\ and a subset of calibration observations of \\chandra\\ and \\xmm. Therefore, we conclude that there is no WHIM line detected at either $z=0.011$ or $z=0.027$ along the Mrk~421 sight line. There are also other important absorption lines in the spectral range of interest (Figures~\\ref{fig:chan-Ni}-\\ref{fig:rgs}). The prominent lines are \\ovii, \\neix, and \\civ\\ K$\\alpha$ and \\ovii\\ K$\\beta$ at $z\\approx0$, and the \\ovi\\ K$\\alpha$ with rest-frame wavelength 22.0 \\AA\\ at $z\\approx0$ is also visible in all four spectra albeit less significant. Again, we only measure the EWs of \\ovii\\ K$\\alpha$ and report them in Table~\\ref{tab:results}. The \\ovii\\ line at $z\\approx0$ can be well explained as absorption from the Galactic diffuse interstellar medium (ISM) \\citep{yao07}, and the \\ovi\\ line is also believed to originate from the ISM \\citep{sav05}. % % It is worth noting that the EW of \\ovii\\ K$\\alpha$ at $z\\approx0$ measured from \\xmm\\ observations is $\\gsim3\\sigma$ larger \\footnote{The differential significance is calculated as \\[|EW_1 - EW_2|/\\sqrt{(\\Delta EW_1)^2 + (\\Delta EW_2)^2}, \\] where $\\Delta EW$ is the $1\\sigma$ error of $EW$ measurement.} than that obtained from \\chandra\\ observations (Table~\\ref{tab:results}). Such a discrepancy in \\ovii\\ has also been measured by \\citet{kaa06}. The causes to this discrepancy are still under investigation, and the results will be published elsewhere. In this paper, we mainly focus on X-ray absorption measurements of the WHIM and will not discuss these non-WHIM measurements any further. \\label{sec:dis} 1. We analyzed all available (as of 2011 April) grating data and obtained two coadded spectra with S/N$\\sim$90 per 50-m\\AA\\ resolution element from \\chandra\\ observations, and S/N$\\sim$190 from \\xmm\\ observations. Neither \\chandra\\ nor \\xmm\\ observations support the existence of the two WHIM systems previously reported by \\citet{nic05}. 2. We ran bootstrap simulations for the detecting limits of the current X-ray telescopes. We find that the reported EWs of the \\ovii\\ K$\\alpha$ at $z=0.011$ and $z=0.027$ should have been measured at $\\ge3.7\\sigma$ and $\\ge2.3\\sigma$, in contrast to the fact that we only obtained upper limits to the EWs. 3. According to the numerical simulations, the WHIM absorbers with $N_{\\rm OVII} \\gsim10^{15}~{\\rm cm^{-2}}$ could sample $>30-50\\%$ of the \\ovii-bearing baryons. To systematically survey the 15 QSO sight lines along which the IGM \\ovi\\ absorbers have been detected, future X-ray telescopes should be able to facilitate the WHIM study via the X-ray absorption line spectroscopy from background QSOs with fluxes of $\\sim0.2$ mCrab to find $\\ge10$ WHIM systems. It takes impractical long (341 Ms for \\chandra\\ and 804 Ms for \\xmm) exposure time for current X-ray telescopes to complete the survey. 4. It would require $\\sim11$ Ms on-target exposures for future X-ray spectrographs equipped with spectral resolution $R\\gsim4000$ and effective area $A\\gsim100~{\\rm cm^2}$ to finish the survey." }, "1112/1112.0244_arXiv.txt": { "abstract": "Ultra High Energy Cosmic Rays (UHECR) at GZK cut off energy ($E\\geq 5.5 \\cdot 10^{19}$ eV) may keep sharp or diffused directionality wherever their composition is made by nucleon or light nuclei. Auger experiment UHECR (2007-2010) did show a mild clustering around Cen A. Two over three of the recent discovered AUGER multiplet (a dozen of events each) tail clustering at twenty EeV are pointing to primary sources very near the same UHECR crowded Cen A region. These tens EeV tail is aligned with the same UHECR events. We foresaw such possibility as fragment tails of lightest UHECR nuclei. We discuss the relevance of this correlation within a model where UHECR are mostly lightest He like nuclei. UHECR fragment multiplet clustering aligned along higher Cen A events (above $5.5 \\cdot 10^{19} $ eV energy) probe and reinforce our interpretation with an a priori probability to occur below $3 \\cdot 10^{-5}$. ", "introduction": "UHECR Astronomy and nuclear composition are more and more in severe conflict. This contradiction were already inherited in early, apparent, discover of a Super Galactic UHECR correlation \\cite{Auger-Nov07}. In that key article the needed UHECR directionality (assumed nucleon) have been , at the same time, in disagreement with the first observed nuclei composition. Indeed UHECR particle astronomy may rise, suffering however of some directionality spread by the smearing of magnetic field bending along the UHECR flight. This bending is negligible for proton and He, but sever for iron. The bending maybe coherent (just one direction) if the magnetic field is constant or it is random , if the field directions are changing , as it happens inside the galactic plane along and across the galactic arms while pointing to Cen A. The nucleon, the light nuclei may keep more or less precise directionality, a smearing astronomy, while heaviest ones may exhibit only tiny anisotropy if sources are near. In addition to these signals, UHECR in flight are making fragment secondaries nuclei as well as, because photo-dissociation, parasite gamma and neutrino tails somehow correlated. These UHECR compositions are leading to different secondary gamma and neutrino spectra; these different nature are making UHECR nucleon origination local and well directed (GZK cut off, tens Mpc distances) or even much local and smeared (a few Mpc) for our lightest UHECR nuclei considered in recent articles \\cite{Fargion2008}. Heavy nuclei may also have, by photon-disruption a gamma and neutrino secondary tail, probably so smeared to be mostly sink into background noise: if UHECR are only iron, as some authors still believe, than UHECR astronomy, in particular the extragalactic one, will be so much bent, polluted and smeared to be hopeless: only large scale iron anisotropy might occur by nearest galactic sources. On the contrary extragalactic lightest nuclei UHECR astronomy may be surrounded by a parasite (a little smeared) trace made by point source gamma, neutrinos (TeVs-PeVs) and also tens EeV energy UHECR fragments. We did suggested this possibility since few years \\cite{Fargion2011}. The lightest UHECR nuclei model \\cite{Fargion2008} is able to explain the surprising absence or paucity of events toward Virgo (the nearest extragalactic cluster of galaxy) and the angle range $10^{o}-15^{o}$, the spread directionality (orthogonal to galactic arms) of events around and along Cen A. A large number of authors discussed mainly the spectra of UHECR cut off as a key discriminator for UHECR modeling. However the mass composition role may confuse the real shape of any (apparent) GZK cut off. We address more on the UHECR anisotropy nature able we hope to correlate UHECR maps and composition, using all radio, IR, gamma MeV-GeV-TeV, maps available. \\begin{figure}[htb] \\begin{center} \\epsfig{file=Fargion_ERICE-11_fig01.eps,scale=0.16} \\epsfig{file=Fargion_ERICE-11_fig02.eps,scale=0.15} \\caption{Left: The last 2010 UHECR event map event map overlap with nearby infrared sky map; the clustering toward Cen A is the main UHECR signal. Around this source the presence of a twin collinear multiplet clustering at twenty EeV shown in next figures. Note the relevance of Infrared Virgo cluster and its absence in AUGER map. Note also a tiny correlation of Vela with an unique galactic triplet event. Right: The last Planck infrared map and labels , whose spread white noise is due to galactic dust, with UHECR events. Note the partial suppression of UHECR events in those regions where white dust is missing: this absence may hint for a galactic component of UHECR.}\\label{1r} \\end{center} \\end{figure} \\subsection{Deuterium, proton, gamma and neutrino tails } \\begin{figure}[!t] \\vspace{5mm} \\centering \\includegraphics[width=3.1 in]{Fargion_ERICE-11_fig03.eps} \\includegraphics[width=2.8 in]{Fargion_ERICE-11_fig04.eps} \\caption{Left: the AUGER 2010 UHECR event map and two of the three multiplet clustering toward Cen A; their sources as shown by dotted curve are within a tiny disk area (at radius of $10^{o}$); the dotted ellipsoid area of the UHECR and multiplet clustering is also extremely correlated, aligned and small. Right: multiplet clustering toward Cen A by random bending of the galactic spiral magnetic fields.} \\label{fig2} \\end{figure} \\begin{figure}[!t] \\vspace{5mm} \\centering \\includegraphics[width=2.8 in]{Fargion_ERICE-11_fig05.eps} \\includegraphics[width=3.2 in]{Fargion_ERICE-11_fig06.eps} \\caption{Left: The AUGER 2010 UHECR event map on radio 408 Mhz map and two of the three multiplet clustering toward Cen A. Right: the third Multiplet clustering points somehow toward Large and Small Magellanic clouds as well it overlaps with the clustering along the Magellanic stream. } \\label{fig2-3} \\end{figure} UHECR formed (mostly) by lightest nuclei may explain a partial clustering of events, as the one around CenA as well as a puzzling UHECR absence around Virgo. Light nuclei are fragile and fly few Mpc before being halted by photo-disruption \\cite{Fargion2008},\\cite{Fargion2009},\\cite{Fargion2010}. Their fragments $He + \\gamma \\rightarrow D+D,D+\\gamma \\rightarrow p+n+ \\gamma , He + \\gamma \\rightarrow He^{3}+n, He + \\gamma \\rightarrow T +p $ may nevertheless trace the same UHECR maps by a secondary clustering at half or even fourth of the UHECR primary energy \\cite{Fargion2011}. Neutrinos and gamma are tracing (both for nucleon or light nuclei) their UHECR trajectory, respectively growing at EeVs (if nucleon) or PeV-TeV (if light nuclei) energy. Gamma secondaries rays from cosmic sources are partially absorbed by microwave and infrared background making once again a very local limited UHECR-gamma astronomy. Among neutrinos $\\nu$, muons ones $\\nu_{\\mu}$, the most penetrating and easy to detect on Earth, are unfortunately deeply polluted by a rich atmospheric component (as smeared and as the isotropic parent CR nucleons and nuclei ). This atmospheric isotropy and homogeneity is probed by last TeV muon neutrino maps in a very smooth ICECUBE neutrino map. Tau neutrinos on the contrary, the last neutral lepton discovered, are almost absent in atmospheric secondaries (about five order of magnitude suppressed at TeVs). Rare $\\nu_{\\mu}\\rightarrow \\nu_{\\tau} $ neutrino oscillation at ten GeV atmospheric windows, may nevertheless arise; at tens TeV-PeV up to EeV $\\nu_{\\tau} $ neutrino might be a clean signal of UHECR-neutrino associated astronomy\\cite{FarTau}. Their tau birth in ice may shine as a double bangs (disentangled above PeV) \\cite{Learned}. In addition UHE tau, born tangent to the Earth or mountain, while escaping in air may lead, by decay in flight, to loud, amplified and well detectable tau-airshower at horizons \\cite{Fargion1999},\\cite{FarTau}. Both in atmosphere fluorescence tracks or by Cherenkov blazing, or by partial skimming ground detectors. Tau astronomy versus UHECR are going to reveal most violent sky as the most deepest probe. This tau airshowers or Earth skimming neutrino \\cite{Feng02} were considered since more than a decade and are going to be observed in AUGER or TA in a few years\\cite{FarTau},\\cite{Auger-01},\\cite{Feng02},\\cite{Auger07}. Regarding the puzzle of UHECR let us also remind that more than a decade ago we were facing and solving a how obsolete problem due to AGASA and Fly Eyes events; such events were calling for sources at distances above GZK cut off. The solution was based on the sources ejecting primary UHE neutrinos at ZeV energy scattering on relic (cosmic) ones making Z bosons in flight and, after decay, UHECR nucleons on Earth \\cite{Fargion1997}. ", "conclusions": "The history of Cosmic Rays and last UHECR discoveries (and disclaims) are exciting and surprising. The very unique correlation of UHECR with Cen A, the absence of Virgo, the hint of correlation with Vela and a mild connection with galactic plane or even Magellanic stream (see Fig.\\ref{1r},\\ref{fig2},\\ref{fig2-3}) might be solved by a lightest nuclei, mainly He, as a courier, leading to a very narrow (few Mpc) sky for UHECR. A soon answer maybe already written into predicted \\cite{Fargion2011} and now observed \\cite{Auger11} multiplet clustering (as Deuterium or proton fragments) at half UHECR edge energy aligned around or along main UHECR group seed discussed and shown above (see Fig.\\ref{fig2}). Indeed He like UHECR maybe bent by a characteristic as large as $\\delta_{rm-He} \\simeq 16^\\circ$,(while expected proton or D fragments at half fourth these energies, along tails spread at $\\delta_{rm-p} \\simeq 32^\\circ$). Future AUGER map possibly about 2-3-4 tens EeV, UHECR fragment clustering maps along higher energy events (5-6 $10^{19}$ eV) may probe and reinforce our interpretation (already tested with present multiplet clustering at $ P (3,2) \\simeq 3 \\cdot 10^{-5}$ level). Additional clustering may occur along few galactic sources as Vela. More along galactic plane, as (inspired by Comptel gamma map correlation) with some clustering along Cassiopeia A, in future UHECR Telescope Array events (see Fig.\\ref{fig3-4}). The discover of expected Neutrino astronomy at Icecube or by horizontal tau neutrinos airshower at ARGO or Auger,TA telescopes \\cite{FarTau},\\cite{Fargion1999}\\cite{Fargion09b},\\cite{Feng02},\\cite{Auger08}, may also shed light on the UHECR nature, their origination and mass composition, opening our eyes to secret UHECR sources." }, "1112/1112.6055_arXiv.txt": { "abstract": "The inner regions of barred galaxies contain substructures such as off-axis shocks, nuclear rings, and nuclear spirals. These substructure may affect star formation, and control the activity of a central black hole (BH) by determining the mass inflow rate. We investigate the formation and properties of such substructures using high-resolution, grid-based hydrodynamic simulations. The gaseous medium is assumed to be infinitesimally-thin, isothermal, and non-self-gravitating. The stars and dark matter are represented by a static gravitational potential with four components: a stellar disk, the bulge, a central BH, and the bar. To investigate various galactic environments, we vary the gas sound speed $\\cs$ as well as the mass of the central BH $\\MBH$. Once the flow has reached a quasi-steady state, off-axis shocks tend to move closer to the bar major axis as $\\cs$ increases. Nuclear rings shrink in size with increasing $\\cs$, but are independent of $\\MBH$, suggesting that ring position is not determined by the Lindblad resonances. Rings in low-$\\cs$ models are narrow since they are occupied largely by gas on $\\xtwo$-orbits and well decoupled from nuclear spirals, while they become broad because of large thermal perturbations in high-$\\cs$ models. Nuclear spirals persist only when either $\\cs$ is small or $\\MBH$ is large; they would otherwise be destroyed completely by the ring material on eccentric orbits. The shape and strength of nuclear spirals depend sensitively on $\\cs$ and $\\MBH$ such that they are leading if both $\\cs$ and $\\MBH$ are small, weak trailing if $\\cs$ is small and $\\MBH$ is large, and strong trailing if both $\\cs$ and $\\MBH$ are large. While the mass inflow rate toward the nucleus is quite small in low-$\\cs$ models because of the presence of a narrow nuclear ring, it becomes larger than $0.01\\Aunit$ when $\\cs$ is large, providing a potential explanation of nuclear activity in Seyfert galaxies. ", "introduction": "\\label{sec:intro} Stellar bars play an important role in the dynamical evolution of gas in galaxies. By introducing a non-axisymmetric torque, they produce interesting morphological substructures in the gaseous medium, including a pair of dust lanes at the leading side of the bar, a nuclear ring near the center, and nuclear spirals inside the ring (e.g., \\citealt{san76,rob79,sch81,van81,ath92b,pin95,but96,mar03a,mar03b,mar06}). They also transport gas inward which can trigger starbursts in the rings (e.g., \\citealt{but86,gar91,hel94,bar95,mao01,maz08}) and if the mass inflow extends all the way to the center, they may help power active galactic nuclei (AGN) (e.g., \\citealt{shl90,reg99,kna00,lau04,van10}). Since bar substructures represent a nonlinear response of the gas to a non-axisymmetric gravitational potential, their formation and evolution is best studied using direct numerical simulations.\\footnote{\\citet{eng00} argued that physical properties of nuclear spirals can be explained by the linear density-wave theories (see also \\citealt{mac04a}).} There have been a number of numerical studies on the gas dynamics in barred galaxies. Based on the numerical scheme employed, they can be categorized largely into two groups: (1) those using a smoothed particle hydrodynamics (SPH) technique (e.g., \\citealt{eng97,pat00,ann05,tha09}) and (2) those using a grid-based algorithm (e.g., \\citealt{ath92b,pin95,mac02,mac04b,reg03,reg04}). The numerical results from these two approaches do not always agree with each other, at least quantitatively, even if the model parameters are almost identical. For instance, \\citet{pin95} using the CMHOG code on a cylindrical grid reported that the gas near the corotation regions exhibits complex density features resulting from Rayleigh-Taylor and/or Kelvin-Helmholtz instabilities, while these structures are absent in the SPH simulations. In addition, overall shapes and structures of dust lanes and nuclear rings from CMHOG simulations are different from SPH results. Some differences in the numerical results may be attributable to relatively large numerical diffusion of a standard SPH method and its inability to handle sharp discontinuities accurately (e.g., \\citealt{age07,pri08,rea10}). However, after adopting and thoroughly testing the CMHOG code as part of this work, we have found it contained a serious bug in the way the gravitational forces due to the bar are added to the hydrodynamical equations. Thus, some of the discrepancies in the flows computed by CMHOG and other codes are likely due to this bug. We discuss this bug and its affect on the results reported in \\citet{pin95} in Section 2.2. In this paper, we revisit the gas dynamics in barred galaxies using a corrected version of the CMHOG code. Our objectives are three-fold. First, we wish to remedy the errors in \\citet{pin95}, and to compute the formation of bar substructures with an accurate shock-capturing grid code with the correct bar potential. Second, the morphology, shape, and strength of the bar substructures are likely to depend on the gas sound speed and the shape of the underlying gravitational potential (e.g., \\citealt{eng00}). Thus, we report new models in which we include a central black hole (BH) that greatly affects the gravitational potential in the central regions, and we vary both the BH mass as well as the sound speed to explore the dynamics in various galactic conditions. Third, we exploit advances in computational resources to compute models that have more than an order of magnitude higher resolution than the models in \\citet{pin95}, with a grid resolution of $0.13$ pc in the central regions. This allows us to resolve details in the flow in the nuclear regions, in particular the formation of nuclear rings and nuclear spirals. According to the most widely accepted theory, a nuclear ring forms near the inner Lindblad resonance (ILR) when there is only one ILR, as the gas outside (inside) ILR loses (gains) angular momentum and accumulates there, while it forms in between the inner ILR and outer ILR when there are two ILRs (e.g., \\citealt{shl90,com96,but96}). On the other hand, \\citet{reg03} argued that the ring formation is more deeply related to the existence of $\\xtwo$-orbits rather than the ILRs. But, the arguments relying on either ILR or $\\xtwo$-orbits do not take into account the effect of thermal pressure. Therefore, it is important to explore to what extent the concepts of ILR or $\\xtwo$-orbits are valid in describing nuclear rings, especially when the sound speed is large. The formation, shape, and nature of nuclear spirals that may channel the gas to the galaxy centers are also not well understood. Observations using the \\emph{Hubble Space Telescope} indicate that galaxies having nuclear dust spirals are quite common (e.g., \\citealt{mar03a,mar03b}). While most of such spirals are trailing, a few galaxies including NGC 1241 and NGC 6902 reportedly possess leading nuclear spirals \\citep{dia03,gro03}. Although the linear theory suggests that leading spirals are expected when there are two ILRs (e.g., \\citealt{mac04a}), they are absent in the numerical models of \\citet{pin95} computed with the CMHOG code, while the SPH models of \\citet{ann05} with self-gravity do form leading spirals. The SPH models suffer from poor spatial resolution at the nuclear regions as most particles gather around the rings. By running high-resolution simulations with a corrected version of CMHOG, we can clarify the issues of the nuclear spiral formation and related mass inflow rates to the galaxy center. We in this work treat gaseous disks as being two-dimensional, isothermal, non-self-gravitating, and unmagnetized, which introduces a few caveats that need be noted from the outset. By considering an infinitesimally-thin disk, we ignore gas motions and associated dynamics along the direction perpendicular to the disk. By imposing a point symmetry relative to the galaxy center, our models do not allow for the existence of odd-$m$ modes, although this appears reasonable since $m=2$ modes dominate in the problems involving a galactic bar. In addition, we are unable to capture the potential consequences of gaseous self-gravity and magnetic stress that may not only cause fragmentation of high-density nuclear rings but also affect mass inflow rates to the galaxy center. Nevertheless, these idealized models are useful to isolate the effects of the gas sound speed and the mass of a central BH on the formation of bar substructures and mass inflows. Also, these models allow us to correct the results of previous CMHOG calculations with incorrect bar forces. This paper is organized as follows. In Section 2, we describe the galaxy model, model parameters, and our numerical methods. In Section 3, we present the results of simulations for off-axis shocks and nuclear rings. The detailed properties of nuclear spirals are presented in Section 4. In Section 5, we study the mass inflow rates through the inner boundary obtained from our simulations. In Section 6, we conclude with a summary and discussion of our results and their astronomical implications. \\begin{deluxetable}{lcc} \\tabletypesize{\\footnotesize} \\tablewidth{0pt} \\tablecaption{Model Parameters\\label{tbl:model}} \\tablehead{ \\colhead{$\\;\\;\\;\\;\\;$Model$\\;\\;\\;\\;\\;$} & \\colhead{$\\;\\;\\;\\;\\;$$\\cs ~ ({\\rm km\\; s^{-1}})$$\\;\\;\\;\\;\\;$} & \\colhead{$\\;\\;\\;\\MBH(0) (\\Msun)\\;\\;\\;$} } \\startdata cs05bh0 & 5 & 0 \\\\ cs05bh0t\\tablenotemark{a} & 5 & 0 \\\\ cs10bh0 & 10 & 0 \\\\ cs15bh0 & 15 & 0 \\\\ cs20bh0 & 20 & 0 \\\\ cs20bh0t\\tablenotemark{a} & 20 & 0 \\\\ \\hline cs05bh7 & 5 & 4$\\times10^7$ \\\\ cs10bh7 & 10 & 4$\\times10^7$ \\\\ cs15bh7 & 15 & 4$\\times10^7$ \\\\ cs20bh7 & 20 & 4$\\times10^7$ \\\\ cs20bh7t\\tablenotemark{a} & 20 & 4$\\times10^7$ \\\\ \\hline cs05bh8 & 5 & 4$\\times10^8$ \\\\ cs10bh8 & 10 & 4$\\times10^8$ \\\\ cs15bh8 & 15 & 4$\\times10^8$ \\\\ cs20bh8 & 20 & 4$\\times10^8$ \\tablenotetext{a}{ The BH mass is varied with time as $\\MBH(t) = \\MBH(0) + \\int_0^t \\Mdot(t^\\prime)dt^\\prime$ assuming that all the inflowing mass is added to the central BH.} \\end{deluxetable} ", "conclusions": "\\subsection{Summary} We have presented detailed numerical models that explore the formation of substructures produced by the gas flow in barred galaxies. Previous models based on particle simulations (e.g., \\citealt{eng97,ann05,tha09}) did not have sufficient resolution to resolve nuclear spirals. On the other hand, studies that used the grid-based code CMHOG (e.g., \\citealt{pin95,mac04b}) unknowingly made mistakes in the force evaluation for the bar potential, so that the results needed to be recomputed. \\begin{figure} \\epsscale{1.2}\\plotone{fig19.eps} \\caption{ Temporal evolution of the mass inflow rates (top) and the BH mass (bottom) in Models cs05bh0t, cs20bh0t, and cs20bh7t where $\\MBH$ is allowed to vary with time. The mass inflow rates from the fixed-$\\MBH$ counterparts are compared as dotted lines. In all models, the total increase in the BH mass over 500 Myr is not large enough to cause significant changes in $\\Mdot$. \\label{fig:mdot_MBHt}} \\end{figure} In this paper we have corrected the errors in the original CMHOG code and run high-resolution hydrodynamical simulations. To resolve the nuclear regions, we employ a logarithmically-spaced cylindrical grid, with a zone size of $\\Delta R \\leq 6\\pc$ at $R\\leq1\\kpc$ where nuclear rings and spirals form. We have included the potential from a central BH, and studied the flow properties as the mass of the BH $\\MBH$ and the sound speed $\\cs$ in the gas are varied. For simplicity, the effects of gaseous self-gravity and magnetic fields are not included. The main results of the present work are summarized as follows: 1. \\emph{Off-axis Shocks} -- The imposed non-axisymmetric bar potential provides gravitational torques to the otherwise circular-rotating gas, perturbing its orbit. The perturbed orbits crowd at the downstream sides of the bar major axis and produce overdense ridges that eventually develop into off-axis shocks. At a quasi-steady state, the off-axis shocks are overall almost parallel to $\\xone$-orbits: they start from the bar major axis at the outer ends, are gradually displaced downstream as they move inward, and connect to the nuclear ring at the inner ends. While the positions of the off-axis shocks are almost independent of $\\MBH$, since the effect of a BH is negligible at large radii, they depend on $\\cs$ in such a way that the shocks are, on average, located closer to the bar major axis as $\\cs$ increases. This is primarily because gas with larger $\\cs$ should be more strongly perturbed to induce shocks, which occurs deeper in the potential well and thus results in shocks on lower $\\xone$-orbits \\citep{eng97}. The off-axis shocks are in general curved. Flow streamlines are complicated near the shocks in that they diverge before the shocks and are promptly swept inwards by inflowing gas right after the shocks. Therefore, the usual Ranking-Hugoniot jump conditions for planar, one-dimensional shocks are not applicable to the off-axis shocks. In fact, the shock strength, as measured by the peak density $\\Sigma_{\\rm max}$ at $R\\sim(1.5-1.8)\\kpc$ from the center, is $\\Sigma_{\\rm max}/\\Sigma_0\\sim3-6$ for models with no BH and do not sensitively depend on $\\cs$. The compression factor of the off-axis shocks are insensitive to the BH mass and depends on $\\cs$ roughly as $\\alpha_{\\rm max}\\sim 7.7 (\\cs/5\\kms)^{0.92}$. The off-axis shocks have very strong velocity shear amounting to $\\sim(1-3)\\times10^3\\kms\\kpc^{-1}$. This strong shear may suppress the growth of gravitational instability in the high-density off-axis shocks when self-gravity is included. 2. \\emph{Nuclear Rings} -- When gas passes through the off-axis shocks, it loses angular momentum, flows inward, and forms a nuclear ring where the centrifugal force balances the external gravity. The ring is attached to the inner ends of the off-axis shocks and thus becomes smaller in size as $\\cs$ increases. The rings that form in our models are all located inside the (outer) ILR, but this does not imply that the ring formation is related to the ILRs. When $\\cs$ is small, the pressure perturbations on the gas orbits are so weak that the rings are quite narrow and their shape is well described by $\\xtwo$-orbits. The mean radius is $\\Rring\\sim(0.8-0.9)\\kpc$ when $\\cs=5\\kms$ and $\\Rring\\sim(0.5-0.6)\\kpc$ when $\\cs=10\\kms$, independent of the BH mass. This suggests that the ring position is not determined by the $\\Omega-\\kappa/2$ curve, hence by the ILRs. When $\\cs \\geq 15\\kms$, on the other hand, large thermal pressure strongly affects the gas orbits in the nuclear rings. For example, some gas near the contact points between the ring and the off-axis shocks is forced out to follow relatively round orbits, while other gas is pulled in radially to make very eccentric orbits. These diverse gas orbits near the contact points tend to spread out the ring material, making the rings much broader than in models with smaller $\\cs$. 3. \\emph{Nuclear Spirals} -- Since even the non-axisymmetric bar potential is nearly axisymmetric in the central parts, the gaseous responses inside a nuclear ring are not as dramatic as in off-axis shocks. Nevertheless, non-axisymmetric $m=2$ perturbations are able to grow inside a ring and develop into nuclear spirals that persist for a long period of time, provided that $\\cs$ is small or $\\MBH$ is large. Although all models have weak spiral structures at early time, in models with large $\\cs$ they are soon destroyed by eccentric gas orbits as well as perturbations induced by the large pressure in the rings unless the BH mass is very large. When $\\MBH=4\\times10^8\\Msun$, the disruptive pressure perturbations from the rings cannot penetrate the very central parts where the gas has extremely large initial angular momentum. In this case, the nuclear spirals are protected from the surrounding and thus are long-lived. The shape of nuclear spirals is determined by the sign of $d(\\Omega-\\kappa/2)/dR$ such that spirals that form in the regions where $d(\\Omega-\\kappa/2)/dR$ is positive (negative) are leading (trailing), confirming the theoretical expectations of \\citet{but96}. With no BH, only the model with $\\cs=5\\kms$ has persistent leading spirals in the regions where the $\\Omega-\\kappa/2$ curve is an increasing function of $R$. The leading spirals in this model are quite strong and develop into shocks with the peak density $\\pSig/\\bSig\\sim7.9$ and the compression factor $\\alpha\\sim 3.4$ at $R=0.25\\kpc$ at the end of the run. Models with $\\MBH=4\\times10^7\\Msun$ initially have hybrid features comprising of trailing spirals at $R<\\Rmin$ and leading spirals at $\\Rmin 0.01\\Aunit$ in models with $\\MBH\\leq 4\\times10^7\\Msun$ and $\\cs\\geq15\\kms$ or $\\MBH=4\\times10^8\\Msun$ and $\\cs=20\\kms$ for different reasons depending on the BH mass. When $\\MBH\\leq 4\\times10^7\\Msun$, the gas orbits are affected by thermal pressure and some gas in the ring can take highly eccentric orbits, directly falling into the inner boundary. In models with $\\MBH= 4\\times10^8\\Msun$, on the other hand, the gas orbits near the center are more or less circular, but the density in the nuclear regions is enhanced greatly because of strong nuclear spirals, increasing $\\Mdot$. \\subsection{Discussion} Nuclear rings play an important role in evolution of barred galaxies by providing sites of active star formation near the centers (e.g., \\citealt{but86,gar91,bar95,mao01,maz08}). Rings certainly consist of gas that migrates from outer parts inward by losing angular momentum at off-axis shocks, but what stops further migration to form a ring remains a matter of debate. As a trapping mechanism of the ring material, \\citet{com96} proposed the non-axisymmetric bar toque that forces gas to accumulate between two ILRs or at a single ILR depending on the shape of the gravitational potential (see also \\citealt{but96}), while \\citet{reg03,reg04} favored the gas transitions from $\\xone$- to $\\xtwo$-orbits rather than orbital resonances. However, our numerical results show that a ring is formed at early time because of the centrifugal barrier that the migrating material feels. Later on, nuclear rings slowly shrink in size as gas with lower angular momentum gas is continuously added. Although nuclear rings are located in between two ILRs in models with no BH, this is a coincidence. Models with a central BH show that the specific ring positions are insensitive to the shape of the $\\Omega-\\kappa/2$ curves, suggesting that nuclear ring formation is not a consequence of the orbital resonances. In fact, the gas flows that produce the ring are not in force balance and have a large radial velocity, so that the concept of resonances and ILRs is not applicable to nuclear rings (e.g., \\citealt{reg03}). In addition, the notion of $\\xtwo$-orbits as the gas trapping locations is meaningful only for small $\\cs$.\\footnote{The models considered by \\citet{reg03,reg04} had the sound speed fixed to $\\cs=5\\kms$.} When the sound speed is large, thermal pressure at the contact points between the off-axis shocks and nuclear ring causes the gas orbits to deviate from $\\xtwo$-orbits considerably. The results of our simulations suggest that not all barred-galaxies possess nuclear spirals at their centers. Long-lasing nuclear spirals exist only when either the sound speed is small or the BH mass is large: they do not survive in models with \\emph{both} large $\\cs$ \\emph{and} small $\\MBH$. Two common views regarding the nature of nuclear spirals are low-amplitude density waves and strong gaseous shocks (see e.g., \\citealt{eng00,mac02,mac04a,mac04b,ann05,tha09}). And, our simulations indeed show that they are either tightly-wound density waves or shocks when the pith angle is relatively large ($\\ip>6^\\circ$). One may speculate that nuclear spirals are more likely to be shocks rather than density waves when $\\cs$ is small. In contrast to this prediction, however, nuclear spirals in models with $\\MBH=4\\times10^8\\Msun$ are density waves when $\\cs$ is small and shocks when $\\cs$ is large. This is of course because as $\\cs$ increases, waves in nuclear regions tend to be more open (with smaller $|k|$) in the beginning, and they are subsequently supplied with more gas from the rings as they grow. This is entirely consistent with the results of \\citet{ann05} who used SPH simulations to show that nuclear spirals are supported by shocks when $\\cs\\simgt15\\kms$ in models with a massive BH. We note however that weak trailing spirals seen in our Model cs10bh8 are absent in Model M2 ($\\cs=10\\kms$ and $\\MBH=4\\times10^8\\Msun$) of \\citet{ann05}, which is presumably due to an insufficient number of particles to resolve nuclear spirals in their SPH simulations. Of 12 models with differing $\\cs$ and $\\MBH(0)$ that we have considered, only 1 model possesses leading spirals, suggesting that galaxies with leading nuclear spirals would be very uncommon in nature. To our knowledge, only two galaxies, NGC 1241 and NGC 6902, are known to possess leading features in the nuclear regions \\citep{dia03,gro03}\\footnote{Leading arms in NGC 1241 are detected by Pa$\\alpha$ emissions tracing young stars, while those in NGC 6902 are observed in the K$^\\prime$ band tracing old populations. It is uncertain how these stellar features are related to gaseous nuclear spirals studied in this paper.}. Based on our simulations, the existence of leading spirals at centers requires two stringent conditions: (1) the gas should be dynamically cold enough to protect nuclear spirals from nuclear rings and (2) there should be a wide range of radii with $d(\\Omega-\\kappa/2)/dR<0$ in the central parts, which can be easily accomplished when there are two ILRs (or without a strong central mass concentration). The second condition is consistent with the linear theory that predicts short leading waves propagating outward from the inner ILR \\citep{mac04a}. The facts that NGC 6902 is a barred-spiral galaxy \\citep{lau04} and does not show significant $X$-ray emissions indicative of AGN activities \\citep{des09} are not inconsistent with the second requirement for the existence of leading nuclear spirals. Since NGC 1241 is a Seyfert 2 galaxy with an estimated BH mass of $\\log (\\MBH/\\Msun)=7.46$ \\citep{bia07}, however, the nuclear star-forming regions in this galaxy are unlikely to be associated with gaseous nuclear spirals studied in this work. Finally, we discuss the mass inflow rates derived in our models in regard to powering AGNs in Seyfert galaxies. The mass accretion rate is often measured by the Eddington ratio defined by $\\lambda \\equiv L_{\\rm bol}/L_{\\rm Edd} = 4.5\\times10^{-2} (\\epsilon/0.1) (\\Mdot/10^{-2}\\Aunit)(\\MBH/10^7\\Msun)^{-1}$, where $L_{\\rm bol}$ and $L_{\\rm Edd}$ denote the bolometric and Eddington luminosities of an AGN and $\\epsilon$ is the mass-to-energy conversion efficiency of the accreted material. Observations indicate that $\\lambda\\simlt 0.1$ for classical Seyfert 1 galaxies with broad iron emission lines (e.g., \\citealt{mey11}; see also \\citealt{pet97}), while $\\lambda\\sim 10^{-3}$ for low-luminosity Seyfert 1 AGNs (e.g., \\citealt{ho08}). In our numerical models, the mass inflow rates are larger for models with smaller $\\MBH$ and larger $\\cs$. Taking $\\epsilon\\approx0.1$ (e.g., \\citealt{yu02}), the mass inflow rates amount to $\\lambda \\simlt 10^{-4}$ for $\\MBH=4\\times10^8\\Msun$ and $\\cs\\simlt15\\kms$, $\\lambda \\sim 10^{-3}$ for $\\MBH=4\\times10^8\\Msun$ and $\\cs=20\\kms$ or for $\\MBH=4\\times10^7\\Msun$ and $\\cs=10\\kms$, and $\\lambda \\sim 0.02-0.4$ for $\\MBH=4\\times10^7\\Msun$ and $\\cs\\simgt 15\\kms$. For classical Seyfert galaxies with strong bars, this suggests that the masses of central BHs are likely to be less than $10^8\\Msun$, which appears to be consistent with the measured values from the relation between the BH masses and the velocity dispersions of stellar bulges (e.g., \\citealt{wat08}) and reverberation mapping techniques (e.g., \\citealt{gul09,den10}). Of course, this result may depend on many factors such as the axis ratio and strength of the bar, presence of self-gravity and magnetic fields, gas cooling and heating, turbulence, etc., all of which would affect gas dynamic significantly. Extending the present work to include these physical ingredients would be an important direction of future research." }, "1112/1112.3736_arXiv.txt": { "abstract": "{Circumstellar disks and envelopes of low-mass young stellar objects (YSOs) contain significant amounts of ice. Such icy material will evolve to become volatile components of planetary systems, such as comets in our solar system.} {To investigate the composition and evolution of circumstellar ice around low-mass YSOs, we observed ice absorption bands in the near infrared (NIR) towards eight YSOs ranging from class 0 to class II, among which seven are associated with edge-on disks.} {We performed slit-less spectroscopic observations using the grism mode of the Infrared Camera (IRC) on board AKARI, which enables us to obtain full NIR spectra from 2.5 $\\mu$m to 5 $\\mu$m, including the CO$_2$ band and the blue wing of the H$_2$O band, which are inaccessible from the ground. We developed procedures to carefully process the spectra of targets with nebulosity. The spectra were fitted with polynomial baselines to derive the absorption spectra. The molecular absorption bands were then fitted with the laboratory database of ice absorption bands, considering the instrumental line profile and the spectral resolution of the grism dispersion element.} {Towards the class 0-I sources (L1527, IRC-L1041-2, and IRAS04302), absorption bands of H$_2$O, CO$_2$, CO, and XCN are clearly detected. Column density ratios of CO$_2$ ice and CO ice relative to H$_2$O ice are $21-28$ \\% and $13-46$ \\%, respectively. If XCN is OCN$^-$, its column density is as high as $2-6$ \\% relative to H$_2$O ice. The HDO ice feature at 4.1 $\\mu$m is tentatively detected towards the class 0-I sources and HV Tau. Non-detections of the CH-stretching mode features around 3.5 $\\mu$m provide upper limits to the CH$_3$OH abundance of 26 \\% (L1527) and 42 \\% (IRAS04302) relative to H$_2$O. We tentatively detect OCS ice absorption towards IRC-L1041-2. Towards class 0-I sources, the detected features should mostly originate in the cold envelope, while CO gas and OCN$^-$ could originate in the region close to the protostar, where there are warm temperatures and UV radiation. We detect H$_2$O ice band towards ASR41 and 2MASSJ1628137-243139, which are edge-on class II disks. We also detect H$_2$O ice and CO$_2$ ice towards HV Tau, HK Tau, and UY Aur, and tentatively detect CO gas features towards HK Tau and UY Aur.} {} ", "introduction": "In molecular clouds, protostellar envelopes, and protoplanetary disks, significant amounts of oxygen and carbon are in the form of molecules in ice mantles, such as H$_2$O, CO, CO$_2$, and CH$_3$OH (e.g. Whittet \\cite{whittet93}, Murakawa et al. \\cite{murakawa00}, Gibb et al. \\cite{gibb04}). These interstellar ices are formed by the adsorption of gas-phase molecules onto grain surfaces and/or the grain-surface reactions of the adsorbed species (e.g. Aikawa et al. \\cite{aikawa05}). Ices in circumstellar disks of low-mass YSOs are of special interest as they contribute to the raw material of planetary formation. Observations of disk ices, however, are not straightforward for the following two reasons, and thus remain limited in number. Firstly, ices exist mostly near the midplane at $r\\ge$ several AU. Observations of disk ices are basically restricted to edge-on disks, where the ice bands absorb stellar light, scattered stellar light, and/or thermal emission of warm dust in the disk inner radius (Pontoppidan et al \\cite{crbr05}). The targets are naturally heavily extincted by dust, and thus are faint. Secondly, it is unclear whether the ice absorption bands, if detected, originate in the disk or other foreground components (i.e. ambient clouds). The source CRBR 2422.8-3423 was the first edge-on object towards which detailed ice observations were performed (Thi et al. \\cite{thi02}). Pontoppidan et al. (\\cite{crbr05}) concluded, via detailed analysis and modeling of the object, that only a limited amount ($< 20$ \\%) of detected CO ice may exist in the disk, while up to 50 \\% of water and CO$_2$ may originate in the disk. They also found that the 6.85 $\\mu$m band, which is tentatively attributed to NH$_4^+$, has a prominent red wing. Since this wing is reproduced by the thermal processing in the laboratory, it indicates that along the line of sight towards CRBR 2422.8-3423 there is warm ice in the disk. Honda et al. (\\cite{honda09}) succeeded in detecting water ice in a disk around a Herbig Ae star, which is not edge-on. They observed scattered light from a disk around HD142527 at multiple wavelength bands around $\\sim 3$ $\\mu$m. This method can be a powerful tool for investigating the spatial distribution of ices in disks. However, it would currently be difficult to detect ice features other than the 3 $\\mu$m water band via scattered light. In the present paper, we report near infrared (NIR) ($2.5-5$ $\\mu$m) observations of edge-on disks using the grism mode of the IRC on board AKARI. We selected seven edge-on young stellar objects (YSOs); three are considered to be in a transient state of class 0-I, and four are class II (see \\S 2 for more detailed descriptions). Although the absorption by the protostellar envelope may overwhelm the absorption by the disk in class 0-I YSOs, envelope material along our line of sight (perpendicular to the rotation axis and outflow) is likely to end up in the midplane region of the forming disk, and ices with a sublimation temperature of $> 50$ K (i.e. CO$_2$ and H$_2$O) would survive the accretion shock onto the disk (Visser et al. \\cite{visser09}). Ice observations towards low-mass YSOs have been performed using {\\it Spitzer Space Telescope (SST)} and various ground-based telescopes such as the Very Large Telescope (VLT) and Subaru. Pontoppidan et al. (\\cite{pontoppidan2003}) observed the 4.67 $\\mu$m CO ice band, which is observable from the ground, towards 39 YSOs. Using {\\it SST}, Furlan et al. (\\cite{furlan08}) and Pontoppidan et al. (\\cite{pontoppidan2008}) observed the strong CO$_2$ ice band at $15.3$ $\\mu$m towards class I sources. Observations with {\\it SST} also detected absorption bands of CH$_3$OH, HCOOH, H$_2$CO, and CH$_4$ (Boogert et al. \\cite{boogert08}, Zasowski et al. \\cite{zaowski09}, {\\rm \\\"{O}}berg et al. \\cite{oberg_ch4}). Since ground-based observations are restricted to atmospheric windows, and since {\\it SST} is restricted to the mid infrared ($5-36$ $\\mu$m), NIR observations via AKARI are important to comprehend the {\\it full} spectra of ices in circumstellar material. AKARI has a high enough sensitivity for the spectroscopic observation of low-mass YSOs; the sensitivity of the AKARI grism mode at 3 $\\mu$m is 120 $\\mu$Jy (1 $\\sigma$, 10min), which is almost comparable to the K-band sensitivity of the spectroscopic observation at 8-m ground-based telescopes (VLT and Subaru). While the L-band sensitivity on the ground is $\\sim 5$ mag or 100 times worse than at K-band, the sensitivity of the AKARI grism mode does not vary significantly at $2.5- 4$ $\\mu$m, and only gradually worsens at longer wavelengths, by up to a factor of 3-4 at 5 $\\mu$m. ", "conclusions": "We have observed ice absorption bands at $2.5-5$ $\\mu$m towards eight low-mass YSOs: three class 0-I protostellar cores with edge-on geometry, two edge-on class II objects, two multiple systems with edge-on class II, and one not-edge-on class II object. Towards the class 0-I objects, L1527, IRC-L1041-2, and IRAS04302, we have detected abundant H$_2$O, CO$_2$, and CO ice in the envelope. The column density ratio of CO$_2$ to H$_2$O ice is $21-28$ \\%, which coincides with the ratio observed by {\\it SST} towards YSOs with various inclinations. The weak absorption at $\\sim 4.1$ $\\mu$m can be fitted by HDO ice; the HDO/H$_2$O ratio ranges from 2 \\% to 22 \\%. The absorption in the vicinity of the CO band (4.76 $\\mu$m) is double-peaked and fitted by combining CO ice, OCN$^-$, and CO gas. The large column density of CO ice suggests that the envelope is still very dense and cold, while OCN$^-$ and CO gas features would originate in the region close to the protostar. The column density of OCN$^-$ is as high as $2-6$ \\% relative to H$_2$O, which is much higher than previous observations. Our lines of sight (high inclinations from the rotation axis) may preferentially trace the regions with high UV irradiation, such as the surface of a forming disk and/or torus envelope. The spectrum of IRAS04302 includes the 3.5 $\\mu$m absorption band, but the feature does not match either CH$_3$OH or CH$_4$. An OCS absorption band is tentatively detected towards IRC-L1041-2. Towards the edge-on class II objects, ASR41 and 2MASS J1628137-243139, we have detected the H$_2$O band. The low optical depth of the water feature is due to geometrical effects (Pontoppidan et al. \\cite{crbr05}). The detected water ice mainly originates in the disk. Both HK Tau B and HV Tau C are edge-on class II objects in multiple systems. In our spectra, which are dominated by the primaries, we have detected the absorption of H$_2$O ice and CO$_2$ ice. Ices in both the edge-on disks and foreground clouds would contribute to the absorption, although H$_2$O absorption towards HK tau could originate solely in the disk. Even if the observed features are due to the foreground clouds, the H$_2$O ice column densities in the clouds are much smaller than those observed towards HK Tau B and HV Tau C with Subaru (Terada et al. \\cite{terada07}), which confirms that the ice columns of the latter must originate in disks. The foreground H$_2$O ice and CO$_2$ ice are also detected towards UY Aur, which is not an edge-on system. We have tentatively detected CO gas towards HK Tau and UY Aur." }, "1112/1112.1505_arXiv.txt": { "abstract": "We study the implications of asymmetric dark matter on neutron stars. we construct a \"mixed neutron star\" model composed of ordinary baryons and of asymmetric dark matter baryons. We derive the general relativistic structure equations for each specie, the equation for the mass within a given radius, and the redshift as function of radius. We present one specific numerical model as an illustrative example. In this example, the mass of the dark neutron equals half that of the ordinary neutron. The main results are: a total mass of $3.74 M_{\\odot}$, a total mass within the neutron-sphere equaling $1.56 M_{\\odot}$, the neutrons mass is \\ $1.34 M_{\\odot}$, the star radius is \\ 31.9 km, the neutron-sphere radius is \\ 11.1 km, and the redshifts from the neutron-sphere and from the star surface are \\ 0.72, \\ 0.25, respectively. We comment briefly on possible astrophysical implications. ", "introduction": "\\subsection{Background} Cold dark matter which is favored by most astrophysical and cosmological observations can be realized in symmetric and (or) asymmetric scenarios. In the first class of models, dark matter is made of stable $X$ particles and an equal amount of $\\bar X$ antiparticles of mass $m_X$. In the early universe, these were in thermal equilibrium and their residual abundance $\\Omega_X$ is fixed at the \"freeze-out\" value when the rate of the Hubble expansion overcomes that of $\\bar X-X $ annihilation rate. A prototypical example which have been most extensively studied is SUSY(super-symmetry). Unfortunately, the searches for SUSY partners in the new large hadron collider (LHC) have failed to detect them. More specifically, insofar as SUSY dark matter models are concerned searches for electrons, positrons or photons in clumped dark matter in and around our Galaxy, and for energetic neutrinos resulting from annihilations of Xs do not provide solid \"indirect\" evidence for dark matter. Moreover, the ongoing direct underground searches put very strong bounds on the scattering crossections of massive Xs on nuclei. Additional constraints are related to accretion of galactic SUSY WIMPS onto the sun. The increased density of the captured WIMPS, accelerates the rate of particle-antiparticle annihilation. The resulting photons and electrons are trapped in the sun but the resulting ultra high energy neutrinos are not. Data from the ICE-CUBE Cerenkov radiation detector near the south pole, severely constrain such models \\cite{abbasi 09, abbasi 10} . In the case of asymmetric dark matter, there will be no such annihilation in the sun. Moreover, once the fraction of dark matter particles in the sun exceeds the ratio of $\\sigma_{Xn}/\\sigma_{XX}$, scattering on the already captured X particles in the sun dominates over scattering on the baryons. Finally, the observed number of satellite mini-halos around the milky way galaxy is two orders of magnitude smaller than predicted within the symmetric dark matter framework \\cite{Klypin, Moore}. Consequently there has been, in the very recent years, a renewed interest in asymmetric dark matter (ADM) which just like ordinary baryonic matter, is charge non-symmetric with say only the dark baryon (or generally only the particle) excess remains after the annihilation of most antiparticles .While there is no single clear-cut explanation for the ordinary baryonic asymmetry, the required dark matter density is readily achieved if the mass ratio of the X particles and baryons $m_X/m_b$ is tuned inversely with the corresponding ratio of asymmetries. An early example of such a model has been proposed in \\cite{nuss85}. For recent studies of asymmetric dark matter see \\cite{An etal 2010,mrm11} and references therein. \\subsection{Cosmological and astrophysical considerations} The idea of asymmetric dark matter should reconcile with astrophysical and cosmological data. An obvious constraint is imposed by big-bang nucleosynthesis, which implies that the mirror neutrinos and photons do not contribute to the rate of the cosmic expansion at that era. Another constraint is that the mirror large scale structure-formation should precede the recombination of ordinary matter, in order to serve as seeds gravitational potential wells for the latter. In the model described in \\cite{An etal 2010} the mirror neutrinos and the mirror photons are massive enough for these constraints to be satisfied. We note that even if the mirror neutrinos and photons are massless the above two constraints can be satisfied provided that the dark cosmic-temperature is lower than the ordinary one. The details of large scale structure formation will depend on the specific model for the asymmetrical dark matter. Generally, being self interacting one may expect it to form mirror structures on all astrophysical scales. Also, asymmetric dark matter will no longer provide collisionless dark halos as in the symmetric cold dark matter case, so that the flatness of galactic rotation curves will have to be readdressed in this new context. ", "conclusions": "We have demonstrated that a mixed neutron star can, as expected, have a mass higher than ordinary neutron stars. At the same time the physical radius, as probed by ordinary massless and massive particles, is the neutron-sphere radius which is similar to the radius of ordinary neutron stars. An important question, not addressed here, is that of stability. Since the models form a two-parameters family (the central densities) the question of stability is more complex than in the one parameter family of ordinary neutron star models. There are a number of quite interesting astrophysical implications with regard to phenomenology of compact x-ray sources. Can some of the stellar mass binary black holes be actually mixed neutron stars? The neutron-sphere redshift is about 50\\% higher than in the ordinary neutron star case, which may have interesting results for the temperature, radius and luminosity measured by a distant observer. The larger neutrons binding energy would lead to a smaller value of the maximal neutrons mass, compared to an ordinary neutron star. I wish to thank my collaborators, R. Mohapatra, S. Nussinov, D. Teplitz and V.~Teplitz as well as W. Kluzniak and L. Zdunik for interesting discussions. Thanks are due to the Afeka College Research Committee for financial support." }, "1112/1112.3706.txt": { "abstract": "We investigate gas accretion flow onto a circumplanetary disk from a protoplanetary disk in detail by using high-resolution three-dimensional nested-grid hydrodynamic simulations, in order to provide a basis of formation processes of satellites around giant planets. % Based on detailed analyses of gas accretion flow, we find that most of gas accretion onto circumplanetary disks occurs nearly vertically toward the disk surface from high altitude, % which generates a shock surface at several scale heights of the circumplanetary disk. % The gas that has passed through the shock surface moves inward because its specific angular momentum is smaller than that of the local Keplerian rotation, while gas near the midplane in the protoplanetary disk cannot accrete to the circumplanetary disk. % Gas near the midplane within the planet's Hill sphere spirals outward and escapes from the Hill sphere through the two Lagrangian points L$_1$ and L$_2$. % % We also analyze fluxes of accreting mass and angular momentum in detail and find that the distributions of the fluxes onto the disk surface are well described by power-law functions and that a large fraction of gas accretion occurs at the outer region of the disk, i.e., at about 0.1 times the Hill radius. The nature of power-law functions indicates that, other than the outer edge, there is no specific radius where gas accretion is concentrated. % These source functions of mass and angular momentum in the circumplanetary disk would provide us with useful constraints on the structure and evolution of the circumplanetary disk, which is important for satellite formation. ", "introduction": "Satellite systems around the giant planets in our solar system are commonly seen. They are thought to have formed in circumplanetary disks, which are believed to have existed around giant planets during their gas capturing growing stage. % MMSN model In earlier works, formation process of satellite systems have been considered based on a minimum mass subnebula (MMSN) model, in which satellites form from a disk that contains sufficient solid mass with solar composition for reproducing the current satellite systems (e.g., \\citet{Lunine82}), as an analog of the minimum mass {\\it solar} nebula model \\citep{Hayashi81}. However, it was suggested that the MMSN model has difficulty in reproducing current satellite systems around Jupiter and Saturn \\citep{Canup02}. One of severe problems is that the model leads to much higher temperature than that of H$_2$O ice sublimation at the current regular satellite region, which means that ice, which is the main component of the satellites, cannot be used as building material of the satellites. % Recent models (starved disk model) In order to overcome the difficulties of the MMSN-type models which assumes a closed and static disk, alternative models have been developed. % \\citet{Canup02} proposed a model in which an accretion disk with continuous supply of gas and solid is considered as a proto-satellite disk. % This model is based on results of hydrodynamic simulations of gas capturing process of giant planets \\citep[e.g.,][]{Lubow99, DAngelo02}, which demonstrated that gas accretion from the protoplanetary disk toward the parent planets occurs through a circumplanetary disk. % In this model, the surface density and temperature of the circumplanetary disk is kept lower than assumed in the MMSN model, and thus H$_2$O ice can be used as a solid building material of satellites. % Based on comparison among time scales of different processes, they concluded that formation of the Galilean satellites can be best explained if the circumplanetary disk had one order of magnitude lower gas surface density than the MMSN disk, with slow gas accretion rate corresponding to Jupiter's growth time scale longer than a few $\\times 10^6$ yr (see also \\citet{Sasaki10}). % On the other hand, \\citet{Mosqueira03a} proposed another disk model which consists of two components, i.e., inner MMSN-type massive disk and outer low-density extended disk. The model reproduced, for example, three inner Galilean satellites as well as only partially differentiated Callisto. % Although these models seem to reproduce the current satellite systems of the giant planets in our solar system, they needed to assume parameters for the disk structure such as surface density profile as a basis of physical processes of satellite formation. % Hydrodynamic simulations % The structure of a circumplanetary disk is closely related to the gas accretion process of giant planets. Gas flow around proto-giant-planets in protoplanetary disks has been studied using hydrodynamic simulations. % % 2D Earlier two-dimensional simulations showed that the gas in the Hill sphere rotates in the prograde direction \\citep{Miki82, Sekiya87, Korycansky96} and a pair of spiral shocks stands in the circumplanetary disks \\citep[e.g.,][]{Kley99, Lubow99, Tanigawa02, DAngelo02}. However, the scale height of the proto-planetary disk for Jupiter-sized planets is comparable to the Hill radius, thus simulations with two-dimension approximation cannot capture the feature of the accretion flow. % % 3D Recent three-dimensional hydrodynamic simulations revealed that the two-dimensional picture for circumplanetary disks is not appropriate for the flow in the Hill sphere \\citep[e.g.,][]{DAngelo03,Bate03,Machida08,Paardekooper08,Ayliffe09b,Coradini10}. % One of the most important features newly found in three-dimensional calculations is that, in the region of circumplanetary disks, the gas accretes nearly vertically downward toward the midplane from high altitude. % % Although some studies using hydrodynamic simulations were able to obtain the structures of circumplanetary disks, the direct use of results of hydrodynamic simulations for the structure, such as surface density, is problematic. % One of the reasons is that disk-like shear-dominant flow is susceptible to numerical viscosity, which is intrinsic nature of numerical hydrodynamic calculation, no matter if SPH method or mesh based method is used, thus it is difficult to obtain reliable structure of the disk directly from such simulations. % In particular, orbital radius of the rotating gas in disks basically changes only slightly with time, thus numerical error tends to accumulate easily in simulation of long-term evolution. % Also timescale required for low viscous disks such as circumplanetary disks to reach steady state is much longer than the typical dynamical time of the fluid. While high-resolution hydrodynamic simulations is needed to resolve the structure of circumplanetary disks, their long-term evolution is difficult to follow with such time-consuming simulation. % In addition, physical (non-numerical) viscosity in the disk is not well understood, and we need to assume specific viscosity models that are hard to justify. % Purpose of this study % As an alternative approach, in the present work, we examine gas accretion flow onto circumplanetary disks from proto-planetary disks, in order to determine gas accretion rate as a function of distance from parent planets. % Unlike the structure in a rotation disk, the accretion flow onto circumplanetary disks is not susceptible to numerical viscosity. % Also, because the circumplanetary disks are located at the downstream of super-sonic accretion flow, the accretion flow itself is hardly affected by the circumplanetary disk structure, which depends on poorly-known effective viscosity. % In \\S \\ref{sec_methods}, we describe our settings of hydrodynamic simulations. In \\S \\ref{sec_results}, results and analyses of the simulations are shown. We discuss implication of our analyses of the accretion flow in \\S \\ref{sec_discussion}. We summarize our results in \\S \\ref{sec_summary}. %\\clearpage %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "\\label{sec_discussion} %======================================================================= \\subsection{Size of Circumplanetary Disks} \\label{sec_disk_size} % The size of a circumplanetary disk is closely related to the location of satellite formation. However, it is not easy to define the outer edge of the disk, and there have been several attempts. % One natural way would be to define the disk edge based on its density distribution. But the density is monotonically and smoothly decreasing with increasing radius toward the Hill sphere, so it is difficult to define the edge from density distribution. % % \\citet{Ayliffe09b}, who performed three-dimensional hydrodynamic simulations, suggested a criterion for the disk edge based on angular momentum distribution. Their simulations show that specific angular momentum of the disk gas has a peak at a certain radial location, whereas specific angular momentum of a Keplerian disk increases monotonically. Their Figure 2 suggests that the turnover point is about 1/3 of the Hill radius, and they defined the disk edge by the radial location of this peak. % \\citet{Martin11} examined periodic orbits of a particle around a planet under the influence of the gravitational force from a central star and the planet, and found that, as the size of the orbit is increased, the orbits start crossing with each other at $r \\sim 0.4r_{\\rm H}$. They inferred that this corresponds to the location of the disk's outer edge where tidal torque of central star's gravity becomes strong, and called it tidal truncation radius ($r_{\\rm trunc}$). % The above value of $r_{\\rm trunc}$ is in agreement with the point of turnover of specific angular momentum found in their two-dimensional SPH simulation, as well as in the result of \\citet{Ayliffe09b}. % Our Figure \\ref{Fig_AM_vs_r} shows that the turnover point is $\\tilde{R} \\sim 0.3 \\tilde{r}_{\\rm H}$, which is also roughly in agreement with $r_{\\rm trunc}$.\\footnote{In this comparison, we assume that specific angular momentum in these works were not measured in inertial frame, but in the rotating frame.} %%% We here consider an alternative criterion by examining ${\\cal V}_R$ defined by Eq.~(\\ref{vr_per_vKep}) to define the position of the outer edge. % We observed that azimuthally-averaged radial velocity in the midplane is positive (outward) in almost all the region, although the value is very small for $\\tilde{R} \\lesssim 0.2$ (Fig.~\\ref{Fig_vr-per-vKep}), and the outward velocity significantly increases at $\\tilde{R} \\gtrsim 0.2$ (see also Fig.~\\ref{Fig_streamlines_outward}). % Since gas at $\\tilde{R} \\gtrsim 0.2$ moves outward and escapes from the Hill sphere quickly, the region $\\tilde{R} \\gtrsim 0.2$ can be regarded as outside of the circumplanetary disk. % We thus define the disk edge as the radial location where ${\\cal V}_R$ starts increasing significantly and has non-negligible positive value; $\\tilde{R} \\sim 0.2$ in our case shown in Fig.~\\ref{Fig_vr-per-vKep}. % % modified 2011/11/04 Note that a two-dimensional SPH simulation shows that negative torque is exerted on the outer disk \\citep{Martin11}, which gives rise to inward flow and is contrary to our results. This might be due to the fact that two-dimensional simulations, which create clearer spiral structure than three-dimensional ones, tend to enhance torque density. % modified 2011/12/12 In particular, the disk thickness in the outer region ($\\tilde{R} \\gtrsim 0.2$), at which outflow is observed, is very thick (thickness is comparable to radius). Thus, it would be unlikely that our three-dimensional calculation is significantly affected by the negative torque suggested by \\citet{Martin11}. % replaced by the above at 2011/12/12 %Thus, it would be unlikely that our three-dimensional calculation is %significantly affected by the negative torque suggested by %\\citet{Martin11}. The outer part of the circumplanetary disk is indeed %very thick (thickness is comparable to radius), thus two-dimensional %approximation is not valid. % % (connected the two paragraph at 2011/12/12) % % removed at 2011/12/12 %Similar outflow was reported by \\citet{Paardekooper08} based on %three-dimensional hydrodynamic simulations of gas flow around Earth size %planets. They explained that accreting material from the top pushes the %envelope around the planet and the envelope material was expelled into %the equatorial plane. Although the outflow in our simulation is much %slower than theirs, it seems to be explained by a similar mechanism, %i.e., it is caused by pushing disk material instead of an envelope. % The distribution of accreting angular momentum also seems to indicate a similar radius for the edge. As we see in Fig.~\\ref{Fig_j_zs}, $\\bar{\\tilde{j}}_{z,\\rm s}$ and $\\tilde{j}_{\\rm Kep}$ have different dependence on $\\tilde{R}$. If we fit each profile as a single power-law function, they cross each other at $\\tilde{R}\\sim 0.3$ in the present case. % (D).comment0705 This roughly agrees with the location of the disk edge defined above based on the significant increase of outward velocity. % The gas exterior to this radius has too large angular momentum to achieve Keplerian rotation, thus moves radially outward. % % removed at 2011/12/12 %This interpretation would be consistent with the explanation for the %outflow described in \\citet{Paardekooper08}, if we would consider %angular momentum instead of dynamical pressure. % ell % Once angular momentum of accreting gas is obtained, one can calculate mean specific angular momentum of the accretion flow, which is sometimes used in calculating the the so-called centrifugal radius $\\tilde{r}_{\\rm c} \\equiv \\tilde{\\ell}^2/(3\\tilde{r}_{\\rm H}^3)$ to infer the disk size \\citep[e.g.,][]{Mosqueira03a,Ward10}. % Mean specific angular momentum of the accretion flow within a radius $\\tilde{R}$ is given by \\begin{equation} \\tilde{\\ell}(\\tilde{R}) = \\frac{\\dot{\\tilde{J}}_{\\rm s}(\\tilde{R})} {\\dot{\\tilde{M}}_{\\rm s}(\\tilde{R})}, \\end{equation} where \\begin{equation} \\dot{\\tilde{J}}_{\\rm s}(\\tilde{R}) = \\int_0^{\\tilde{R}} d\\tilde{R}' \\int_0^{2\\pi} \\tilde{R}'d\\phi \\tilde{f}_{\\rm s}(\\tilde{R}',\\phi) \\tilde{j}_{\\rm s}(\\tilde{R}',\\phi). \\end{equation} We can see from Figure \\ref{Fig_ell} that $\\tilde{\\ell}(\\tilde{R})$ increases nearly linearly with radius and levels around $\\tilde{R} \\sim 0.2$, where $\\tilde{\\ell} \\sim 0.2$. This corresponds to $\\tilde{r}_{\\rm c} \\sim 0.013$, which might give us a rough estimation of the disk size. % % However, it is practically difficult to determine the upper bound of the integral range with respect to $\\tilde{R}$, which affects the value of $\\tilde{\\ell}$, because the circumplanetary disks is smoothly connected to the proto-planetary disks. % Also, unlike the case of a particle, angular momentum of a gas element is not a conserved quantity along a flow in principle, because gas is continuum medium which can transfer angular momentum through waves. This also makes it difficult to determine $\\tilde{\\ell}$ precisely, and thus $\\tilde{r}_{\\rm c}$ as well. % %This point also suffers us when $\\bar{\\tilde{f}}_{\\rm Kep}$ and %$\\dot{\\tilde{M}}_{\\rm Kep}$ is estimated, even though, from the point of %view of total budget of angular momentum, we expect that this estimation %should not be far from real one. %% %Therefore, in order to determine the disk size or edge, we would not %have to pay attention to $\\tilde{\\ell}$ (see also \\citet{Martin11}). %In addition, from the point of view of accreting mass or angular %momentum, typical lengths, such as centrifugal radius, should not exist %except for the outer edge of the disk because the rates of accreting %mass and angular momentum are well described by power-law functions (see %Figs.~\\ref{Fig_f_s_f_Kep} and \\ref{Fig_M_s_M_Kep}), %% %even though we do not rule out the possibility that surface density has %peaks somewhere in the circumplanetary disk because of disk internal %evolution, such as turbulent viscosity driven by MRI. %======================================================================= \\subsection{Picture of Gas Accretion Flow onto Circumplanetary Disks} \\label{sec_picture} % Figure \\ref{Fig_CP-disk} shows a schematic picture of gas accretion flow onto a circumplanetary disk based on the results obtained in the present work. Gas accretion occurs mostly downward from high altitude with high incident angle (see Fig.~\\ref{Fig_diskstructure}). The accreting gas is accelerated by the planet gravity to have almost free-fall velocity of the planet (Fig.~\\ref{Fig_flux_vs_theta}b). The value of angular momentum of the accreting gas normalized by the local Keplerian angular momentum is lower in the inner region of the disk and higher in the outer region (Fig.~\\ref{Fig_j_zs}). The falling gas reaches the shock surface formed on the top of the circumplanetary disk. %%% Gas near the midplane (especially where $\\tilde{R} \\lesssim 0.1$) is almost in Keplerian rotation and hydrostatic equilibrium in the $z$-direction. In this region, radial velocity is very small and it is difficult for the gas to accrete inward through the disk midplane, as mentioned above. % On the other hand, the gas at $\\tilde{R} \\gtrsim 0.2$ in the midplane shows significant outflow, which eventually escapes from the Hill sphere (see Fig.~\\ref{Fig_streamlines_outward}). % Thus inflow from high altitude and outflow near the midplane co-exist, and they do not interfere with each other. This means that there is a circulation across the Hill sphere and fresh (protoplanetary) gas is always supplied in the $\\tilde{R} \\gtrsim 0.2$ region. % In addition, the two Lagrangian points L$_1$ and L$_2$, which are often thought to be the most likely points of inflow, are actually the points of outflow, even in the gas accretion stage (Fig.~\\ref{Fig_flux_at_spheres}). % % added in response to the referee's comment 2). This seems consistent with \\citet{Klahr06}, who also performed three-dimensional hydrodynamic simulation with no explicit physical viscosity, but with lower resolution and with radiation. They showed that accretion mainly occurs via the poles of the planet and no inflow along the equatorial plane, which is quite similar to ours. But they explained that the circulation in their simulation is driven by accretion heating, which we do not consider. Thus it is not clear if the driving mechanisms of the circulation are the same. % % added in response to the referee's comment 2). Note that \\citet{Ayliffe09b} shows that the vast majority of the mass flows into the Hill sphere near the equator, which seems to be inconsistent with our result. There are several differences in setting between theirs and ours, so it is not easy to judge which factor causes the difference. Disk thickness would be one of the reasons for that, even though our disk thickness is not very different from theirs. Another possibility is gap formation. % When a deep gap is formed, the contribution of the vertically accreting gas would become less significant, which reduces the difference. % However, here we think that viscosity is the main reason for the difference. When there is explicit viscosity, the circumplanetary disk gas should transfer its angular momentum outward and most of gas would move inward accordingly, but this is not necessarily true in the inviscid limit. Actually, inviscid simulations by \\citet{Klahr06} showed similar results with ours. Also, inviscid limit might not be bad because circumplanetary disks are likely to be MRI inactive in most cases \\citep{Fujii11}. %%% % new paragraph for the new Fig.17 (2011/11/04) Since dust particles tend to settle down toward the midplane, gas accretion flow from high altitude is likely dust-poor gas, which diminishes dust-to-gas ratio in the circumplanetary disk. % On the other hand, since gas in the outflow region ($\\tilde{R} \\gtrsim 0.2$) rotates significantly slower than Keplerian velocity around the planet, dust particles would migrate inward quickly. Thus the outflow in the midplane is also likely dust-poor gas, which would be the source of solid material in the circumplanetary disk and would enhance dust-to-gas ratio in the disk. % Dust-to-gas ratio in the circumplanetary disk is one of the most important factors for satellite formation processes \\citep{CW06}, thus these two filtering effects would become important for satellite formation. % In addition, studies on size evolution of solid material in protoplanetary disks, such as \\citet{Kobayashi10}, would also be important for the filtering effects. % Further studies are needed on this issue. %%%% %Between the shock surface and the hydrostatic region with nearly %Keplerian rotation, there is a transition layer. %% %The gas that passed through the shock surface tries to merge into the %circumplanetary disk, but its angular momentum at the shock surface is %smaller than that for Keplerian rotation at the point. The gas %accordingly moves inward for adjustment. %% %As we see in Fig.~\\ref{Fig_j_zs}, specific angular momentum before the %shock normalized by Keplerian angular momentum decreases with decreasing %radius in $\\tilde{R} \\lesssim 0.1$, so inner region have relatively %larger inward movement in the transition region. %% %The inward stream in the transition region might play an important role %for the net mass accretion toward the planet in circumplanetary disks. %% %This is a kind of layered accretion. %% %This accretion picture should be explored more quantitatively by further %high-resolution simulations, using a code with lower artificial %viscosity such as the one with polar coordinates for numerical grids. %%% % modified 2011.7.21 As we see in Figure \\ref{Fig_flux_vs_theta}, there is the layer just under the shock at the disk surface where gas moves inward. This {\\it layered accretion} is explained by the fact that angular momentum of the accreting gas onto the disk surface is lower than that of the Keplerian rotation at the radius (Fig.~\\ref{Fig_j_zs}). % The inward stream in the layer might play an important role for the net mass accretion toward the planet in circumplanetary disks. % This accretion picture should be explored more quantitatively by further high-resolution simulations, using a code with lower artificial viscosity such as the one with polar coordinates for numerical grids. %%======================================================================= %\\subsection{Difficulties in Obtaining Surface Density of Circumplanetary % Disks} %% %% %Surface density of circumplanetary disks would be more directly needed %when one would like to consider satellite formation process, rather than %gas accretion rate onto circumplanetary disks as a function of radius. %% %Nevertheless we focused on the accretion rate because of several %reasons. % %% %First of all, effective viscosity, which is the key quantity for %determining surface density in quasi-steady state, is poorly fixed. %This is serious and essential because even the origin of viscosity is %not well understood, although turbulent viscosity driven by %magneto-rotational instability is the most likely source. %% %Even if viscosity can be fixed well, it is hard to determine surface %density by hydrodynamic simulations. One of the reasons is that low %viscosity, which would be realized in curcim-planetary disks, is %difficult to handle precisely. %% %Also, calculation time would be too long to reach a quasi steady state %for very high-resolution hydrodynamic simulation (see the description of %surface density in \\S \\ref{sec_diskstructure}) % % %On the other hand, gas accretion profile we obtained is a more robust %quantity in the sense that this quantity basically does not chenges even %if surface density does because the accretion flow onto the surface of %circumplanetary disks is mostly supersonic, in other words, no %disturbance in the circumplanetary disks can affect the supersonic %accretion flow, even though the height of the shock surface can change %to some degree. % % %Once the gas accretion rate as a function of distance from parent %planets is obtained, one can apply it to, for example, one-dimensional %circumplanetary disk evolution models. \\citet{Ward10} and %\\citet{Martin11} actually modeled one-dimensional circumplanetary disk %evolution by assuming distribution of the gas accretion rate, which is %the quantity we obtained in this paper. %% %Our accretion rate can basically apply those kinds of one-dimensional %models directly, even though our result only gives snapshots. % % %%% %Although specific angular momentum of the accretion flow $\\tilde{\\ell}$ %may not be important for determine the disk structure \\citep{Martin11}, %accretion rate $\\tilde{\\dot{M}}$ and evolution are definitely important. %In particular, hysteresis of the accretion rate is important for %satellite formation process via disk temperature and surface density %\\citep{Canup02}. %% %\\citet{Tanigawa07} and \\citet{Ward10} modeled the long term evolution of %the accretion rate onto giant planets by simple one-dimensional models. %However, the important period for satellite formation is the last minute %of the giant planet formation when the gas accretion is significantly %waned and about to cease. Precise treatment of disk physical processes, %such as dissipation of protoplanetary disks by photoevapolation, viscous %evolution, and gap formation is necessary, and thus is desired to be %examined in future works. %%======================================================================= %\\subsection{Dust to Gas Ratio of the Accreting Flow} %We have considered gas accretion flow, then we next discuss supply of %solid material into circumplanetary disks. %% %In this giant planet forming stage, large fraction of solid material in %protoplanetary disks should have been consumed to build protoplanets, %but still some dust particles or planetesimals should remain. %% %Supply of solid material to giant-planet systems is important for %satellite formation process and the fraction of heavy elements of the %giant planets in our solar system \\citep[e.g.,][]{Wuchterl00, Saumon04} %as well. % % %What we have found in this paper is that the width of gas accretion band %varies with height (see Fig.~\\ref{Fig_fate} and the corresponding text %in \\S 3.1). %% %If dust is well mixed and the ratio to gas is homogeneous in %protoplanetary disks, net dust to gas ratio of the accreting material is %the same as that of protoplanetary disks. %% %But if the ratio is not homogeneous, the net ratio differs from the %averaged one, and in reality, dust tends to settle down to the disk %midplane and thus the ratio would be enhanced around the midplane %\\citep[e.g.][]{Nakagawa86,Cuzzi93}. %% %According to Fig.~\\ref{Fig_fate}, the width is the narrowest at the %midplane and increases with height until $\\tilde{z} \\sim 0.5$, and keeps %nearly constant above the height. %% %This immediately implies that dust would be depleted in the accretion %flow if the ratio is enhanced around the midplane or the protoplanetary %disk, and that the tendency of the depletion is more noticeable for %larger size dust because sedimentation is more effective for larger %one. %% %This is the opposite sense for the larger content of heavy elements in %the current Jovian planets. % % % %Next we consider planetesimals accretion into circumplanetary disks. %% %Since motion of large planetesimals are usually decoupled from that of %gas, the motion is hardly affected by the gas and thus most of %planetesimals can pass through circumplanetary disks. %% %Even when planetesimals themselves pass through the circumplanetary %disks, dissipation due to the friction between planetesimals and %circumplanetary disk gas would give rise to ablation of the %planetesimals \\citep{Mosqueira03a}, which might be a important source %of solid in circumplanetary disks. %% %If the size of planetesimals is smaller, the planetesimals can be %directly captured by the gas drag with circumplanetary disks. %% %Even though there are some studies on the capturing process of %planetesimals during gas capturing growth phase of giant planets in the %context of accounting for composition of heavy elements in giant planets %\\citep{Zhou07, Shiraishi08}, no quantitative study have been reported %for the last stage of giant planet formation when circumplanetary disks %exist. %======================================================================= \\subsection{Effects of Gap Formation} % One important issue to be addressed is the effect of gap, which is a lower density annulus region near the planet orbit in the protoplanetary disk. % We observed the gap as a slightly lower density band formed around $\\tilde{x}=0$ in our simulations. % However, in the last stage of giant planet formation, a giant planet would become massive enough to create a deep gap, which would be able to truncate its growth. % Since the deep gap is associated with steep density gradient at the edge, gap formation may affect accretion flow and circumplanetary disk formation. % However, at the edge of such a deep gap, the gas density changes over a radial distance comparable to the disk scale height, while the width of the accretion band onto the circumplanetary disk is much narrower than the scale height (Fig.~\\ref{Fig_fate}). % Therefore, we think that the gap formation would not affect the qualitative feature of the accretion flow. % % Another possible effect of gap formation is the disk size. As described in \\S \\ref{sec_disk_size}, several mechanisms are proposed to explain the disk size. If the disk size is determined by the tidal effect as suggested by \\citet{Martin11}, it should not be affected by gap formation, except when the Bondi radius is smaller than the Hill radius. % replaced by the above at 2011/12/12 %Another possible effect of gap formation is the disk size. As described %in \\S \\ref{sec_disk_size}, several mechanisms are proposed to explain %the disk size. If the disk size is determined by the tidal effect as %suggested by \\citet{Martin11}, it should not be affected by gap %formation. %% On the other hand, if the disk size is determined by the radially outward velocity of the gas as described in \\S \\ref{sec_disk_size}, lower density around the Hill sphere and thus stronger radial pressure gradient near the disk edge may enhance the outflow from the circumplanetary disk, and the disk size may become smaller. % In any case, effects of gap formation should be examined in future works to check the validity of the results shown in this paper. % %If a deep gap, which should be expected to form along the orbits of %Jupiter mass planets, is developed in our simulation, resultant steeper %pressure gradient in radial direction would enhance more outward flow in %the disk. In this sense, $\\tilde{R}_+$ should not be larger than we %obtained even when deep gap is formed. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In order to understand the structure of circumplanetary disks, we performed high-resolution hydrodynamic simulation and analyzed gas accretion flow in detail. % We confirmed that gas accretion onto circumplanetary disks occurs in a manner that the gas is accreted from high altitude toward the disk surface downward with large incident angle, which was suggested by previous studies \\citep[e.g.,][]{DAngelo03,Bate03}, whereas \\citet{Ayliffe09b} showed that, in terms of mass flux, the downward accretion flow is not significant, which is inconsistent with our results. % replaced by the above at 2011/12/12 %We confirmed that gas accretion onto circumplanetary disks occurs in a %manner that the gas is accreted from high altitude toward the disk %surface downward with large incident angle, which was suggested by %previous studies \\citep[e.g.,][]{DAngelo03,Bate03}. %% We found that the gas that has passed through the shock surface moves inward because its specific angular momentum is smaller than that of Keplerian rotation, whereas gas accretion through the midplane does not occur. % While the net gas accretion across the Hill sphere is inward, outflow through near the two Lagrangian points is observed, although the regions around the two Lagrangian points have been, from the point of view of the potential energy, thought as a main accretion channel from protoplanetary disks. % This outflow was not observed in previous hydrodynamic simulations for Jupiter-sized planets \\citep[e.g.,][]{Bate03,Ayliffe09b}. % Outward radial velocity of the gas near the midplane significantly increases at a radial distance of about 0.2 times the Hill radius from the planet, and the gas can escape from the Hill sphere within a short period of time. %%% We also obtained the distribution of mass and angular momentum of accreting gas onto the surface of circumplanetary disks. We found that the accretion rates of mass and angular momentum can be well described by power-law functions. % This distribution would be useful in the study of satellite formation, for example, a radially one-dimensional viscous-evolution model for circumplanetary disks, such as \\citet{Ward10} and \\citet{Martin11}. % Recent development of viscous modeling for MRI turbulence in protoplanetary and circumplanetary disks \\citep{Okuzumi11, Fujii11} would also contribute to construct more realistic models for the circumplanetary disks. % However, in order to understand satellite formation processes, long term evolution of circumplanetary disks, which is determined by the evolution of accretion rate from protoplanetary disks to circumplanetary disks, is necessary and thus global evolution of protoplanetary disks with embedded giant planets until the complete dissipation of protoplanetary disks needs to be examined. % Together with such global models \\citep[e.g.,][]{Tanigawa07,Ward10}, we will be able to obtain long-term evolution of circumplanetary disks, which would provide better understanding of satellite formation processes. %%% Added by Ohtsuki-san 0705 Our results demonstrate that gas accretion toward and within circumplanetary disks has a complicated vertical structure. The width of the accretion band toward a circumplanetary disk depends on the initial height of gas elements (Fig.~\\ref{Fig_fate}). Also, after accretion onto the disk, the gas just below the shock surface migrates inward, while the gas near the midplane moves radially outward (Fig.~\\ref{Fig_CP-disk}). Such vertical heterogeneity of the flow may have significant influence on the dynamical evolution of solid bodies in the circumplanetary disk, which we will examine in our future work. % added in response to the referee's comment ). We also need to address the effect of viscosity and local calculation, which would affect the flow quantitatively. %% If you wish to include an acknowledgments section in your paper, %% separate it off from the body of the text using the" }, "1112/1112.0503_arXiv.txt": { "abstract": "Tracing the cosmic evolution of the Baryonic Acoustic Oscillation (BAO) scale with galaxy two point correlation functions is currently the most promising approach to detect dark energy at early times. A number of ongoing and future experiments will measure the BAO peak with unprecedented accuracy. We show based on a set of N-Body simulations that the matter distribution is anisotropic out to $\\sim150\\hMpc$, far beyond the BAO scale of $\\sim100\\hMpc$, and discuss implications for the measurement of the BAO. To that purpose we use alignment correlation functions, i.e., cross correlation functions between high density peaks and the overall matter distribution measured along the orientation of the peaks and perpendicular to it. The correlation function measured along (perpendicular to) the orientation of high density peaks is enhanced (reduced) by a factor of $~2$ compared to the conventional correlation function and the location of the BAO peak shifts towards smaller (larger) scales if measured along (perpendicular to) the orientation of the high density peaks. Similar effects are expected to shape observed galaxy correlation functions at BAO scales. ", "introduction": "\\label{sec:intro} The Baryonic Acoustic Oscillations (BAO) constitute a characteristic feature within the large scale structure of the Universe and can serve as standard ruler for constraining the properties of dark energy \\citep[e.g.,][]{Blake-Glazebrook-03, Linder-03, Seo-Eisenstein-03, Wang-06, McDonald-Eisenstein-07, Seo-Eisenstein-07, Seo-08, Seo-09, Kazin-Sanchez-Blanton-12}. The BAO in the baryon-photon fluid of the pre-recombination era imprint the sound horizon distance at decoupling as a typical scale in the matter correlation function or power spectrum \\citep{Peebles-Yu-70, Sunyaev-Zeldovich-70, Eisenstein-Hu-99, Bashinsky-Bertschinger-02}. These oscillations were detected in the cosmic microwave background \\citep[e.g.,][]{Page-03} and in the spatial distribution of galaxies \\citep{Eisenstein-05, Cole-05} and have been confirmed by a number of subsequent studies \\citep[e.g.,][]{Percival-07, Cabre-Gaztanaga-09, Sanchez-09, Percival-10, Reid-10, Kazin-10}. The next generation of large galaxy surveys, like the Panoramic Survey Telescope \\& Rapid Response System (Pan-STARRS, \\citealt{Kaiser-02}), the Dark Energy Survey (DES, \\citealt{DES-05}), the Baryonic Oscillation Spectroscopic Survey (BOSS, \\citealt{Schlegel-White-Eisenstein-09}), BigBOSS \\citep{Schlegel-11}, the Hobby Eberly Telescope Dark Energy Experiment (HETDEX, \\citealt{Hill-04}) and the space based Euclid mission \\citep{Cimatti-09}, will cover volumes much larger than current datasets, allowing for much more accurate determinations of the BAO. In the two-point correlation function the BAO are visible as a unique broad and quasi Gaussian peak \\citep{Matsubara-04}. However, the determination of the shape and location of the peak may be affected by sample variance \\citep{Cabre-Gaztanaga-09, Martinez-09, Kazin-10, Cabre-Gaztanaga-11}, non linear effects and the bias of the tracer galaxy population \\citep{Smith-03, Crocce-Scoccimarro-06a, Angulo-08, Crocce-Scoccimarro-08, Seo-08}. These difficulties lead \\cite{Prada-11} to suggest to use the zero-crossing of the two point correlation located at $\\sim 130\\hMpc$ as standard ruler instead of the peak location. Yet, several observations do not show the theoretically predicted zero-crossing at all \\citep{ Martinez-09, Kazin-10}. At this stage it is unclear whether this discrepancy is caused by systematic effects or cosmic variance or whether it represents a challenge for the concordance $\\Lambda$CDM model \\citep{Labini-09, Sanchez-09, Kazin-10}. One basic assumption for interpretation of the BAO measurements is that the matter distribution is isotropic at the relevant scales ($\\sim100\\hMpc$). In this work we use alignment correlation functions \\citep{Paz-Stasyszyn-Padilla-08, Faltenbacher-09} to show that the amplitudes of the two-point correlation function measured along the orientations of the high density peaks are larger than those derived from spherically averaged (conventional) clustering analysis out to scales of $\\gtrsim 150\\hMpc$ and discuss possible effects on measurements of the BAO. \\begin{figure*}[t] \\epsfig{file=fig1.eps,width=0.95\\hsize } \\caption{\\label{fig:z0} {\\it Upper panels:} 3 dimensional two point cross correlation functions between FoF groups of the indicated mass and the overall matter density field represented by 10\\% random subset of all simulation particles. Solid lines show the {\\em conventional} correlation functions. Dotted and dashed lines show the {\\em alignment} correlation functions along and perpendicular to the orientations of the FoF groups, respectively. {\\it Lower panels:} Difference between conventional and alignment correlation functions above. The shaded areas in the upper and lower panels give the cosmic variance. Poisson errors are negligible and are not shown here.} \\end{figure*} ", "conclusions": "The findings presented here are based on N-Body simulations however qualitative similar effects are expected to shape the observed galaxy correlation functions at scales up to $150\\hMpc$. Potential observational implications are:\\\\ $\\bullet$ {\\em Large-scale anisotropy and structure of two point correlation functions:} Our results indicate that matter is distributed anisotropically out to separations far larger then the BAO scale. At any given pair separation the conventional correlation function is the average of the alignment correlation function over the whole range of $\\theta$. We find that the alignment correlation function measured along high density peaks does not fall below zero (for separations $\\le 150\\hMpc$) which is counterbalanced by negative amplitudes at much smaller separations for measurements perpendicular to it. At the BAO scale galaxy two point correlation functions can be interpreted as the average of a more highly clustered component along the direction of high density peaks and a less clustered component perpendicular to it.\\\\ $\\bullet$ {\\em Shape and location of the BAO peak:} At BAO scales the amplitudes of the correlation functions between high density peaks and the overall matter distribution are significantly higher if measured along the orientation of the peaks. In this case the BAO peak is composed of a hump on top of the declining but still positive correlation function. The signal perpendicular to the orientations is dominated by the BAO hump itself with negligible underlying clustering signal. The location of the BAO peak is found at somewhat smaller (larger) scales if measured parallel (perpendicular) to the orientation of high density peaks. For the measurements along the density peak orientations the BAO hump is transformed into a plateau-like feature. If the survey volume is dominated by large filamentary structures the shape of the conventional correlation function should be close to that found for the parallel signal shown here. This may relate to a study by \\cite{Kazin-10} who find no apparent peak in 6\\% of their mock samples.\\\\ $\\bullet$ {\\em Zero crossing and large scale power:} Several publications report the non-detection of zero-crossing out to scales of $250\\hMpc$ which indicates (unexpected) large scale power \\citep[e.g.,][]{Martinez-09,Kazin-10,Labini-09}. We find similar results for the correlation functions measured along the orientations of high density peaks. Furthermore, the alignment cross correlation function measured along the orientations of high density peaks at $z=0.6$ shows zero-crossing in contrast to the behavior at $z=0$. The WiggleZ redshift-space correlation function at $z=0.6$ \\citep{Blake-11} shows a crossover as well. Whether these analogies are coincidental remains to be explored in future work.\\\\ $\\bullet$ {\\em Direct measurement of anisotropy:} With the advent of enormous cluster catalogs \\citep[e.g.,][]{Hao-10, Gilbank-11} it should in principle be possible to directly measure the large scale anisotropies with alignment correlation functions if cluster orientations can be determined with sufficient accuracy. An anisotropy signal may even be extracted simply by using orientations of the cluster central luminous red galaxies because the orientations of central galaxies and host systems are correlated \\citep{Faltenbacher-09, Okumura-Jing-Li-09, Schneider-11}. \\\\ $\\bullet$ {\\em Improvement of BAO measurements:} If the directional effects reported here are observable it would be worthwhile considering to measure the BAO peak perpendicular to the orientations of galaxy clusters (or luminous red galaxies) since in this direction the BAO peak is better confined." }, "1112/1112.0029_arXiv.txt": { "abstract": "We study the evolution of the Star Formation Rate Function (SFRF) of massive ($M_\\star>10^{10} M_\\odot$) galaxies over the $0.410^{10} M_\\odot$ range and determine the SFRF using the $1/V_{\\rm max}$ algorithm. We thus define simulated galaxy catalogues based on the predictions of three different state-of-the-art semi-analytical models of galaxy formation and evolution, and compare them with the observed SFRF. We show that the theoretical SFRFs are well described by a double power law functional form and its redshift evolution is approximated with high accuracy by a pure evolution of the typical SFR (SFR$^\\star$). We find good agreement between model predictions and the high-SFR end of the SFRF, when the observational errors on the SFR are taken into account. However, the observational SFRF is characterised by a double peaked structure, which is absent in its theoretical counterparts. At $z>1.0$ the observed SFRF shows a relevant density evolution, which is not reproduced by SAMs, due to the well known overprediction of intermediate mass galaxies at $z\\sim2$. Semi-analytical models are thus able to reproduce the most intense SFR events observed in the GOODS-MUSIC sample and their redshift distribution. At the same time, the agreement at the low-SFR end is poor: all models overpredict the space density of SFR $\\sim 1 M_\\odot/yr$ and no model reproduces the double peaked shape of the observational SFRF. If confirmed by deeper IR observations, this discrepancy will provide a key constraint on theoretical modelling of star formation and stellar feedback. ", "introduction": "The evolution of the star formation rate (SFR) over the cosmic time is a fundamental constraint for every theory of galaxy formation and evolution (see e.g. \\citealt{Hopkins04, HopkinsBeacom06}). Estimating SFRs for individual galaxies is a complex task, owing to the uncertainties involved in the reconstruction of this quantity from observational data. It is widely accepted that dusty molecular clouds are the main sites for star formation: this implies that newly born stars are subject to significant dust attenuation, until they are able to escape or disrupt their parent cloud. Young OB stars emit a considerable amount of energy in the restframe ultraviolet (UV) band, which has thus been considered a key waveband for recovering the instantaneous SFR in external galaxies. However, dust absorbs UV photons, heats up and re-emits this energy as thermal radiation in the Infrared (IR) bands. For this reason, SFR estimates based on UV luminosity include correction factors to account for dust attenuation and re-emission \\citep{Calzetti94, Kennicutt98,Bell03}. Dust attenuation is particularly relevant for star forming galaxies, where the dust emission peak is the dominant component of the galactic spectral energy distribution (SED, see e.g. \\citealt{Calzetti00}), not to mention the extreme cases of sub-mm galaxies (see e.g. \\citealt{Chapman04}). The study of the cosmic IR background shows that the global energy emitted by galaxies in the IR is comparable to the direct starlight emission, detectable in the optical (see e.g. \\citealt{Lagache05}), clearly showing that a significant fraction of star formation activity is expected to be heavily extinguished and detectable only in the IR. These uncertainties become more relevant, when SFR estimates for galaxies covering a wide redshift range are considered, since both galaxy physical properties and dust properties are expected to evolve with cosmic time \\citep{Maiolino04, Fontanot09a, Gallerani10, Fontanot11c}. Combining UV information with supplementary information from direct observations in the IR region has thus been proposed as the best tool to account for the total SFR \\citep{Kennicutt98,Bell07}. In particular, observations at $24 \\mu$m have shown to be extremely useful for estimating the global IR luminosity and the instantaneous SFR \\citep{Papovich07,Santini09}. Thanks to the advent of the {\\it Herschel} Space Observatory, we will be able to constrain total IR luminosities, by directly sampling the peak of the thermal emission. In fact, it has been recently shown by \\citet{Rodighiero10b}, that the combination of $24 \\mu$m data with {\\it Herschel} observations at longer wavelengths represents a very promising tool for a more accurate determination of SFRs of individual galaxies up to $z\\sim3$. Nowadays, our view of the evolution of galaxy properties has substantially changed thanks to the advent of multiwavelength surveys. If the spectral sampling is fine enough, these catalogues can be used to infer the physical properties also in the absence of spectroscopic information. In recent years several groups (see e.g. \\citealt{Fontana04,Panter07}) developed a number of ``SED fitting'' algorithms: these codes compare the available photometry for individual sources with synthetic SED libraries and the best-fit model template is chosen by means of a $\\chi^2$ minimisation. The resulting estimate for redshift (the so-called photometric redshift), stellar mass ($M_\\star$), dust extinction and SFR are the most widely used results of this procedure and have been of fundamental importance for our understanding of galaxy evolution. The common interpretation of these results is connected to the so-called ``downsizing'' scenario, in which the star formation shifts from high mass to low mass galaxies as redshift decreases (first introduced by \\citealt{Cowie96}). This picture has been recently revised by \\citet{Fontanot09b} by showing that the typical errors associated with the estimate of the physical quantities has to be taken into account, when comparing the observed evolutionary trends with the predictions of theoretical models. In particular, \\citet{Fontanot09b} concluded that the discrepancies seen between models and data for massive galaxies are not significant, if model predictions are convolved with typical observational errors, while the strongest discrepancies between model and data are seen for low-to-intermediate mass galaxies. The importance of the study of the redshift evolution of the cosmic SFR for galaxies at a given mass range has been widely recognised by a number of authors \\citep{Noeske07, Elbaz07, Zheng07, Drory08, Dunne08, Santini09, Gilbank11}. These results and the evolution of the stellar mass density provide fundamental information about the evolution of the global process of galaxy formation. The latter approach is best complemented by the analysis of the stellar mass function (i.e. the volume density of galaxies as a function of stellar mass), which allows us to characterise galaxy evolution as a function of both redshift and stellar mass. Despite the wealth of information already available in the literature about the stellar mass function and its redshift evolution, very little is known about the corresponding Star Formation Rate Function (SFRF hereafter) and its evolution: \\citet{Bell07} studied its evolution in the $0.2 10^{10} M_\\odot$, since at these stellar masses the GOODS-MUSIC is expected to be a complete sample. We thus compute the SFRF using SFR estimates coming from both SED fitting and/or a combination of UV light with $24 \\mu$m observations. We compare the resulting SFRF with the predictions of 3 independent semi-analytical models of galaxy formation and evolution (\\citealt{Wang08,Somerville08} and \\morgana) and we obtain the following results: \\begin{itemize} \\item{The agreement between the observed and predicted high-SFR end (SFR$>10 M_\\odot/yr$) of the SFRF is good over the whole redshift range (once the observational errors are taken into account). This result implies that theoretical models are able to reproduce the space density evolution of the strongest starbursts. It is worth noting that the predictions of the three models are remarkably similar in this SFR range.} \\item{The SFRF of massive galaxies, as seen by the GOODS-MUSIC is characterised by a complex evolution both in the typical SFR and in the normalisation. Theoretical models are able to cope with the SFR$^\\star$ evolution, but they predict a negligible density evolution. In particular, all models overpredict the space density of SFR $ \\sim 1 M_\\odot/yr$ galaxies. We interpret this behaviour as due to the well documented excess of intermediate mass galaxies ($10^{10} M_\\odot < M_\\star < 10^{11} M_\\odot$) in SAMs \\citep{Fontanot09b}.} \\item{The observed low-SFR end of SFRF is characterised by a double peaked shape. Despite the uncertainties in the determination of such low SFRs from SED fitting methods, it is worth noting that none of the theoretical model we consider is able to reproduce this feature, and all of them predict a smooth decrease of the space density of low-SFR galaxies, with a well defined power-law slope $\\alpha$. Since this is also the SFR range where models differ most, each of them predicting a different value for $\\alpha$, we stress that future observations providing stronger constraints on the low-SFR end of the SFRF will be of fundamental importance to understand the physical mechanisms responsible for the decline of SFR since $z\\sim2$ in massive galaxies.} \\end{itemize} The main result of this paper lies in the analysis at the high-SFR end, since this region is completely sampled by $24 \\mu$m observations, which are critical for the recovery of the SFR levels. Also the evolution of the typical SFR$^\\star$ is well sampled by means of IR data out to $z \\sim 1.0$, and allow us to put strong constraint on this relevant discrepancy seen between data and models. Unfortunately, for lower SFR levels this information is not currently available, and we have to rely entirely on the results of SED fitting. Our tests suggest that the use of SED fitting results does not introduce any obvious distortion in the SFR distribution at intermediate SFR values, but we could not completely exclude the double peaked feature in the SFRF to be due to systematics in the SED fitting algorithm. From optical photometry alone is not possible to uniquely associate SFR $ \\sim 1 M_\\odot/yr$ galaxies (the region were models and data differ most and roughly corresponding to the region between the two peaks of the SFRF) with known populations in the colour-magnitude diagram, i.e. the green valley or the blue cloud. The red $U-V$ colours, with a relatively small scatter, of galaxies belonging to the putative secondary peak provide only a qualitative support to a scenario where star formation in massive galaxies decrease at a slower pace with respect to theoretical expectations. If confirmed, this double peaked shape of the SFRF would thus represent a critical discrepancy between models and observations, since all SAMs we consider predict a smooth, featureless, single slope power-law shape for the SFRF. Modifications in the assumed star formation law, taking into account the relation between star formation and the atomic and molecular gas content of galaxies may alleviate this tension, since the properties of model galaxies with low SFR levels are very sensitive to this modeling layer \\citep[see e.g.]{Lagos11}. On the other hand, thanks to the new facilities (e.g. {\\it Herschel}), future observations in the IR region, able to characterise the properties of the IR peak, will finally determine the SFR levels of these objects and the relevance of this feature. This is a fundamental task, since the low-SFR end of the SFRF is the region where SAM predictions differ most and therefore provides potentially strong constraints to the different approaches to star formation and stellar feedback." }, "1112/1112.0359_arXiv.txt": { "abstract": "We present the results of the SuperLupus Survey for transiting hot Jupiter planets, which monitored a single Galactic disk field spanning 0.66 deg${^2}$ for 108 nights over three years. Ten candidates were detected: one is a transiting planet, two remain candidates, and seven have been subsequently identified as false positives. We construct a new image quality metric, $S_j$, based on the behaviour of 26,859 light curves, which allows us to discard poor images in an objective and quantitative manner. Furthermore, in some cases we are able to identify statistical false positives by analysing temporal correlations between $S_j$ and transit signatures. We use Monte Carlo simulations to measure our detection efficiency by injecting artificial transits onto real light curves and applying identical selection criteria as used in our survey. We find at 90\\% confidence level that 0.10$^{+0.27}_{-0.08}$\\% of dwarf stars host a hot Jupiter with a period of 1-10~days. Our results are consistent with other transit surveys, but appear consistently lower than the hot Jupiter frequencies reported from radial velocity surveys, a difference we attribute, at least in part, to the difference in stellar populations probed. In light of our determination of the frequency of hot Jupiters in Galactic field stars, previous null results for transiting planets in open cluster and globular cluster surveys no longer appear anomalously low. ", "introduction": "\\label{sec:intro} The discovery of short-period, giant extrasolar planets \\citep{1995Natur.378..355M, 1996ApJ...464L.147M} provided the exciting potential for large numbers of planets to be discovered by the transit method, as these ``hot Jupiters'' have a $\\sim$10\\% geometric probability of transiting, and do so every few days. However early predictions greatly overestimated the actual discovery rate \\citep{2003ASPC..294..361H}. This discrepancy resulted from simplistic assumptions and a misunderstanding of the effects that systematic noise would play in lowering detection efficiency \\citep{2006MNRAS.373..231P}. It also resulted from an over-estimation of the frequency of hot Jupiters. More accurate, and lower, predictions were provided in \\citet{2008ApJ...686.1302B}, where it was noted that objective and quantifiable detection criteria were required for more robust inferences of planet frequencies. By adopting such objective and quantifiable detection criteria, we propose that the frequency of hot Jupiters in the field has been overestimated by a factor of three \\citep{2011OHP...11.01008}. The SuperLupus Survey was established to detect hot Jupiters in a field positioned just above the Galactic plane ($b=11^{\\circ}$), and also to determine the fraction of stars that host hot Jupiters in this typical Galactic field. The survey design and data analysis was constructed so as to fulfill both of these objectives. In Section~\\ref{observations} of this paper, we describe the SuperLupus observations and data reduction. The photometry is described in Section~\\ref{photometry}. The criteria for candidate selection are detailed in Section~\\ref{sec:detection}, and the 10 identified candidates are described and analyzed in Section~\\ref{sec:candidates}. In Section~\\ref{mc_simulatinos}, we set out the details of the Monte Carlo simulations performed to calculate the detection efficiency and the effective number of stars probed for planets in the SuperLupus Survey. In Section~\\ref{frequency}, we apply this efficiency to the actual results of the survey to determine the fraction of stars in the field that host a hot Jupiter. Finally, in Section~\\ref{discussion} we summarize and discuss the implications of our results. ", "conclusions": "\\label{discussion} The frequency of hot Jupiters in the Galaxy is an important quantity that will ultimately provide a constraint on models of planet formation and migration. It has been suggested that the frequency of hot Jupiters in globular cluster environments is lower than that of field stars, and that this may be due to crowding or low metallicity affecting planet formation, migration, or survival \\citep{2000ApJ...545L..47G,2005ApJ...620.1043W}. Our results indicate, however, that there is little statistical disagreement between hot Jupiter frequencies in cluster and non-cluster environments, even if one of the remaining candidates turns out to be a genuine planet. This is consistent with the work presented in \\citep{2011...arXiv:1009.3013v1} which concludes that there is no evidence to support that open clusters have a lower frequency of hot Jupiters. Initial estimates for planet yields from transit surveys turned out to be far in excess of the actual discovery rate \\citep{2003ASPC..294..361H}. One of the many factors that led to this over-estimation was the adoption of the hot Jupiter frequency derived from early radial velocity surveys, which as we have shown is higher than is found from transit surveys. The {\\it Kepler\\/} result of 0.37$\\%$ \\citep{2011arXiv1103.2541H} obviously provides a robust statistic for hot Jupiter frequencies due to the high detection efficiency of that survey. However the result should be adopted cautiously when calculating expected yields from typical transit surveys, as it is based on a sample of ``solar type'' stars drawn from Kepler target stars, rather than the ensemble field stars monitored by most blind transit surveys." }, "1112/1112.4670_arXiv.txt": { "abstract": "We constrain properties of cluster haloes by performing likelihood analysis using lensing shear and flexion data. We test our analysis using two mock cluster haloes: an isothermal ellipsoid (SIE) model and a more realistic elliptical Navarro-Frenk-White (eNFW) model. For both haloes, we find that flexion is more sensitive to the halo ellipticity than shear. The introduction of flexion information significantly improves the constraints on halo ellipticity, orientation and mass. We also point out that there is a degeneracy between the mass and the ellipticity of SIE models in the lensing signal. ", "introduction": "The properties of galaxy and cluster haloes are of great interest in cosmology, and can be powerful tests of the cosmological paradigm and the nature of dark matter. Two important parameters that describe a dark matter halo are its mass and shape, which are related to many physical processes, such as the growth and merging history \\citep{1993MNRAS.264..201K,2005Natur.435..629S}. Models of dark matter haloes beyond the spherical approximation are favored by many numerical simulations \\citep{2002ApJ...574..538J, 2004IAUS..220..421S,2004ApJ...611L..73K, 2006MNRAS.367.1781A} and observations \\citep{2000A&A...364..377R, 2004ApJ...601..599L, 2005ApJ...625..108D, 2006ApJ...645..170S, 2010ApJ...718..762W}. Furthermore, numerical simulations with different assumptions predict different properties of dark matter haloes \\citep[e.g.][]{2002sgdh.conf..109B, 2005ApJ...627..647B,2007MNRAS.380...93W}. Current models based on N-body simulations, semi-analytic models or hydrodynamic simulations can predict several halo properties, but several ingredients of these models remain uncertain. A precise understanding of halo properties such as mass and ellipticity is important to confirm and improve the existing models of galaxy formation and probe the physical nature of dark matter. Gravitational lensing is a powerful tool to study mass distributions, independent of the nature or dynamical state of the matter \\citep[see][for reviews]{1999ARA&A..37..127M,2006glsw.book..269S, 2008PhR...462...67M}. Galaxy-Galaxy Lensing (GGL) is concerned with the mass associated with galaxies and dark matter haloes in which galaxies reside \\citep{1984ApJ...281L..59T, 1996ApJ...466..623B, 1998ApJ...503..531H, 2004ApJ...606...67H, 2004AJ....127.2544S, 2006MNRAS.368..715M}. The distortion caused by a single galaxy cannot be detected, but the statistics of many foreground-background pairs yield a detectable signal for a population of galaxies. \\citet{1996ApJ...466..623B} discovered a significant GGL shear signal. \\citet{1997ApJ...474...25S} developed a maximum likelihood analysis that can constrain the halo properties of the lens galaxy populations through GGL, allowing to estimate the mean velocity dispersion and the characteristic scale for a non-singular isothermal sphere halo model. Flexion as the gradient of the projected mass density, is sensitive to the small-scale variations of mass distributions \\citep{2002ApJ...564...65G,2005ApJ...619..741G,2006MNRAS.365..414B}. Different techniques have been developed to measure flexion \\citep[see][for examples]{2006ApJ...645...17I, 2007ApJ...660..995O, 2008A&A...485..363S,2011MNRAS.416.1616F}. Recently, \\citet{2011MNRAS.412.2665V} applied the shapelets technique on the COSMOS survey, and \\citet{2011ApJ...736...43C} introduced a new method (so-called analytic image model) to analyse lensing flexion images. It has been noted that flexion can contribute to cosmology in several aspects, such as exploring the mass distribution of dark matter haloes of galaxies and clusters, especially substructures \\citep{2009MNRAS.395.1438L, 2010MNRAS.409..389B, 2010arXiv1008.3088E}. \\citet{2011MNRAS.412.1023H} also propose to reduce the distance measurement errors of standard candles using lensing shear and flexion maps. In \\citet{2009MNRAS.400.1132H, 2011A&A...528A..52E, 2011MNRAS.417.2197E}, the ellipticity of a galaxy halo has been studied with flexion. It was found that the constraints from flexion are tighter than those from shear. In \\citet{2005ApJ...619..741G}, Galaxy-Galaxy lensing Flexion (GGF) has been studied using the distribution function of orientations between the line connecting foreground and background pairs and the flexion of the background galaxies. A GGF signal has been detected by \\citet{2007ApJ...666...51L} using images taken by HST ACS in the cluster Abell 1689. Moreover, combining shear and flexion information provides tighter constraints on halo properties by studying mass distribution on different scales \\citep{2010arXiv1008.3088E,2010MNRAS.404..858S,2011MNRAS.412.1023H}. In this paper, we combine shear and flexion data to constrain the properties of dark matter haloes. The tangential shear is mainly sensitive to the mass of the halo, whereas flexion is sensitive to the halo ellipticity. Moreover the usable number density of flexion data is relatively low, since the flexion signal drops faster than the shear signal with the angular distance to the centre of the halo. It is thus not sufficient to constrain the halo properties with shear or flexion alone. Therefore, we propose to take advantage of both shear and flexion in our analysis. In particular, we use the angular positions, tangential shear and flexion of galaxies in our likelihood functions. A singular isothermal ellipsoid model and an elliptical NFW model for a galaxy or cluster halo with 3 parameters (mass, ellipticity and orientation) are adopted in this paper. In Sec. 2, we recall the basic lensing equations. Our likelihood function is introduced in Sec. 3. We perform numerical tests of our method and results in Sec. 4 and present our conclusions in Sec. 5. Throughout this paper, we adopt a $\\Lambda$CDM model with $\\Omega_{\\Lambda}=0.75$, $\\Omega_{\\rm m}=0.25$, and a Hubble constant $H_0 = 73$ km\\,s$^{-1}$\\,Mpc$^{-1}$. ", "conclusions": "In this paper, we study the potential of weak lensing flexion in the study of galaxy cluster haloes. We use mock data including shear, $\\cal F$ flexion, $\\cal G$ flexion and redshift information. We find that the inclusion of flexion significantly improves the estimate of foreground haloes parameters, although the details are model-dependent. In particular, in the case of a SIE halo, the presence of a mass-ellipticity anti-correlation implies that analyses where the halo is incorrectly assumed to be spherical will overestimate the halo mass. On the other hand, we did not find significant correlation between the halo mass and ellipticity in the eNFW model. The noise in the mock data is determined by the dispersion of the intrinsic shear and flexion distributions, and by the density of background galaxies. After applying stringent cuts in the data, we are left with a galaxy density of roughly 10 arcmin$^{-2}$. Our approach assumes that the flexion estimators are linear in the flexion observables and the point spread function can be removed without producing a bias. In reality, the noise of flexion estimators can be complicated and non-Gaussian. An accurate study of flexion noise is important to evaluate the estimated error. The analysis considers a single cluster halo. This approach is not possible if the number density of background galaxies is low. In that case, stacking of several halo fields can be used to increase the number of background images and thus the signal-to-noise. That approach requires the alignment of the major axis of various foreground galaxies and selecting haloes with similar properties, for example similar shapes of their central galaxies. Such stacking analysis can constrain the halo shapes more tightly, and as a function of other halo properties, e.g. mass. Our results emphasize that a combined weak lensing analysis will be a useful technique for precise measurements of the properties of galaxy or cluster haloes from future weak lensing surveys, such as EUCLID." }, "1112/1112.3116_arXiv.txt": { "abstract": "Recent advances in N-body simulations of dark matter haloes have shown that three-parameter models, in particular the Einasto profile characterized by $d\\ln\\rho(r)/d\\ln r\\propto r^\\alpha$ with a shape parameter $\\alpha \\lesssim 0.3$, are able to produce better fits to the 3D spatial density profiles than two-parameter models like the Navarro, Frenk and White (NFW), and Moore et al. profiles. In this paper, we present for the first time an analytically motivated form for the 2D surface mass density of the Einasto family of dark matter haloes, in terms of the 3D spatial density parameters for a wide range of the shape parameter $0.1$ $\\leq$ $\\alpha$ $\\leq$ $1$. Our model describes a projected (2D) Einasto profile remarkably well between $0$ and $(3-5)$ $r_{200}$, with errors less than $0.3$ per cent for $\\alpha$ $\\leq 0.3$ and less than $2$ per cent for $\\alpha$ as large as 1. This model (in 2D) can thus be used to fit strong and weak lensing observations of galaxies and clusters whose total spatial(3D) density distributions are believed to be Einasto-like. Further, given the dependence of our model on the 3D parameters, one can reliably estimate structural parameters of the spatial (3D) density from 2D observations. We also consider a Sersic-like parametrization for the above family of projected Einasto profiles and observe that fits with a Sersic profile are sensitive to whether one fits the projected density in linear scale or logarithmic scale and yield widely varying results. Structural parameters of Einasto-like systems, inferred from fits with a Sersic profile, should be used with caution. ", "introduction": "Gravitational lensing signatures are a response to the projected surface mass column density of matter $\\Sigma(\\vec R)$ along the line of sight in galaxies and clusters. Upon a suitable deprojection and circular averaging, an estimate of $\\Sigma (R)$ can, in principle, be used to trace the spherically averaged 3D density profile $\\rho(r)$. In the past few years, N-body simulations have shown [\\cite{P03}, \\cite{Nav04} (Nav04), \\cite{M06} (M06), \\cite{S09}(S09)] that three-parameter models, especially the Einasto \\citep{Einasto} profile and the \\cite{PS97} de-projected Sersic profiles (PS97), are able to produce better fits to the 3D density profiles of galaxy and cluster-sized dark matter haloes than two-parameter models (\\cite{NFW} $\\&$ (1996) (NFW), \\cite{Moore99}). While the PS97 profile has a well known 2D sky projected form - the Sersic \\citep{Sersic} profile, there has been no such analytical counterpart for the Einasto profile. In this paper, we present a very good approximation for the 2D projection of the Einasto family of 3D profiles. Thus, if the 3D total mass density is believed to be Einasto-like, our model can be used to parametrically describe the projected 2D surface mass densities of galaxies and clusters in the weak and strong lensing regimes. However, note that even upon radial averaging to smooth out the substructure, not all haloes subscribe to an Einasto profile. For the rest of this paper, we will limit the discussion to the 2D projection of the special case where the 3D profile is Einasto-like. In 3D, the functional form of the Einasto \\citep{Einasto} profile is given by: \\begin{align}\\label{einasto} \\ln[\\frac{\\rho(r)}{\\rho_s}]= -b[(\\frac{r}{r_s})^{\\frac{1}{n}} -1] \\end{align} where, $\\rho(r)$ is the 3D (spatial) density at $r$, $n$ (or $\\alpha=\\frac{1}{n}$) is the shape parameter, $b$ is a function of $n$, $\\rho_s$ the spatial density at a scale radius $r_s$ and $\\rho(0) = \\rho_s e^b$. In 2D, the Sersic \\citep{Sersic} profile, which has been used to describe the projected surface brightness profiles of galaxies, is similar in form to the Einasto profile in 3D \\eqref{einasto} and is given by: \\begin{align}\\label{sersic} \\ln [\\frac{\\Sigma_S(R)}{\\Sigma_{R_E}}] = -q[(\\frac{R}{R_E})^{\\frac{1}{m}} -1] \\end{align} such that $\\Sigma_S(0)$$=$$\\Sigma_{R_E}~e^q$ where, $R$ is the projected distance in the plane of the sky, $\\Sigma_{R_E}$ is the line of sight projected surface brightness at a projected scale radius $R_E$, which can be defined to be the half-light radius of a Sersic profile under the condition: $q=2m-0.3333+0.009876/m$ (PS97) with $m$ (or $\\lambda$$=$$\\frac{1}{m}$) characterizing the shape of the Sersic profile. The parameter $\\alpha$$=$$\\frac{1}{n}$ in \\eqref{einasto} defines the shape of the Einasto profile. In the first ever fits to N-body haloes with a Einasto profile, Nav04 found an average value of $0.172$$\\pm$$0.032$ for a wide range of halo masses from dwarfs to clusters. For galaxy-sized haloes \\cite{Pr06} found $0.133$$\\leq$$\\alpha$$\\leq$$0.167$. \\cite{HW08} observed an evolution of $\\alpha$ with mass and redshift ($z$) in the Millennium Simulation ($MS$) of \\cite{Spring05} and found $\\alpha$$\\sim$$0.17$ for galaxy and $\\sim$ $0.23$ for cluster-sized haloes. A similar trend is supported by \\cite{G08} where $\\alpha$$\\sim$$0.3$ for the most massive clusters in $MS$. Hence, although in this paper we discuss our approximation to a projected Einasto profile for $0.1$$\\leq$$\\alpha$$\\leq$$1$, particular attention is drawn to the domain $\\alpha$$\\lesssim$$0.3$; where as we shall show in \\S 3, the errors due to our approximation out to $30$$r_{-2}$ ($\\sim$$(3-5)$$r_{200}$) are $<$ $0.3\\%$. The parameters $r_s$, $b$ and $n$ in the Einasto profile are not independent. It can be seen, that in terms of a dimensionless length $X=\\frac{r}{r_s}$, the logarithmic slope of the density profile is given by: \\begin{align} \\beta = \\frac{d \\ln (\\rho/\\rho_s)}{d \\ln X}= -\\frac{b}{n} X^{1/n} \\end{align} Nav04 chooses to define $r_s$ such that $b/n=2$ (the isothermal value of $\\beta=-2$) at $r=r_s$ and hence label $r_s$ as $r_{-2}$ and $\\rho_s$ as $\\rho_{-2}$. Another approach is to use the convention of M06, requiring $r_s$ to include half the total mass. They quote a numerical estimate of $b$$=$$3n-0.333+0.0079/n$. In this paper we will follow the Nav04 parametrization. Since the Sersic profile describes the 2D surface brightness profiles of galaxies reasonably well, a natural question is: does the 2D surface mass density of Einasto-like 3D dark matter haloes also follow a Sersic-like description? For the Nav04 simulations, \\cite{M05} (M05) have shown that a Sersic function does produce fits with acceptable errors in the range of $r_{conv}$ to $r_{200}$, where $r_{conv}$ is the minimum radius of convergence and $r_{200}$ is the virial radius in the N-body simulations of P03 and Nav04. In this paper, we focus on the analytical description of the projected Einasto profile. In order to see how well a projected Einasto profile is described by a Sersic profile, we numerically project \\eqref{einasto} and find that a Sersic profile produces acceptable fits only in a limited range of the projected radius $R$. We shall show in \\S 3 that this range also depends on whether one fits the numerically projected Einasto profile in linear scale (large errors with increasing R) or log scale (large errors with decreasing R) of density. It is thus evident that the Sersic profile does not give an adequate description of a projected Einasto profile for all $R$. The choice of Sersic fit (log or linear) may, for example, have possibly strong implications in the strong and weak lensing regimes respectively, yielding incorrect results. It is with this perspective, and the observation that the projection of an Einasto profile (i.e. integral of a Sersic-like function) is not a Sersic-like function but rather a Gamma-like function, we present a derivation of the surface mass density (\\S 2.1) for the Einasto profile. Throughout this paper we will describe 3D spatial parameters as $r$, $r_s$ (or $r_{-2}$), $\\rho_s$ (or $\\rho_{-2}$), $b$ and $n$ (or $\\alpha = 1/n$))and 2D projected parameters as $R$, $R_E$, $q$ and $m$ (or $\\lambda = 1/m$). We will also refer to the Sersic profile as $\\Sigma_S$, our approximation to a projected 2D Einasto profile as $\\Sigma_E$ and a numerically projected Einasto profile as $\\Sigma_N$. Further, for this paper, $r_{conv}$ and $r_{200}$ have no physical meaning as such. We note from the Nav04 simulations, that on an average $r_{conv} \\sim 5$ per cent of $r_{-2}$, and $r_{200} \\sim 8-10$ per cent of $r_{-2}$ for galaxies and $\\sim 4$ per cent of $r_{-2}$ for clusters. In this paper we will, for the sake of discussion, refer to $r_{conv} \\sim$ $0.05$ $r_{-2}$ and $r_{200}$ as $\\sim$ $6$ $r_{-2}$. In \\S 2 we derive a semi-analytical form \\eqref{sigmae} for $\\Sigma_E$, which for the Nav04 parametrization of $b=2n$, simplifies to \\eqref{sigmae2}. In \\S 3 we discuss details of estimating the fit parameters in our expression for $\\Sigma_E$, followed by a discussion of errors in our approximation. We then present a comparison between $\\Sigma_N$ and the best-fitting Sersic profile $\\Sigma_S$. In \\S 3.4, we note the remarkable accuracy with which \\eqref{sigmae2}, along with \\eqref{zeta2param} and \\eqref{muparam}, can be used to extract the 3D parameters of Einasto-like systems from its 2D form. ", "conclusions": "Non-parameteric estimates of density profiles in N-body simulations (Nav04,M06) favour Einasto-like profiles, since they provide better fits than the two-parameter NFW and Moore profiles. \\cite{M06} have also shown that a de-projected Sersic profile fits the 3D halo mass distribution almost as well as the Einasto profile, and a Sersic profile provides good fits to non-parametric estimates of surface mass densities (M05) of the Nav04 N-body haloes. We have observed that fits with a Sersic function ($\\Sigma_S$) to a numerically projected Einasto profile ($\\Sigma_N$) are sensitive to whether one fits using linear density $\\Sigma_S$ (errors increasing for large R) or log density $ln(\\Sigma_S)$ (errors increasing for small R) yielding widely varying results. Consequently, the Sersic profile does not give an adequate description of the projected Einasto profile. Sersic profile fits to the surface mass density of N-body haloes (M05), whose 3D spatial densities are well fit by Einasto profiles with $0.12$$\\leq$$\\alpha$$\\leq$$0.22$, have been obtained from the limited radial range of $r_{conv}$ to $r_{200}$. For the haloes in M05 and Nav04, this range is generally less than two decades in radius. Hence, if the 3D distribution is indeed Einasto-like, interpreting structural properties from fits with a Sersic profile, especially $m$ in \\eqref{sersic} as the shape parameter, $\\Sigma_S(0)$ as the central density and $R_E$ as the half-mass radius, can be misleading. In this paper, we have provided an analytical approximation \\eqref{sigmae2} to the projected surface mass density of Einasto-like 3D density distributions. The fit errors are well contained to $<$$2\\%$ for the projected radial range $0$$\\leq$$R$$\\leq (10-30)$ $r_{-2}$ equivalent to $(3-5)$$r_{200}$ and shape parameter, $1$$\\leq$$n$$\\leq$$10$, or $0.1$$\\leq$$\\alpha$$\\leq$$1$. This model can therefore be used both as a fitting function for 2D observations and also to extract the 3D parameters of Einasto-like profiles. Since $\\Sigma_E$ fits a projected Einasto profile in a wide radial range, it can be used for fitting strong and weak lensing observations in systems whose total 3D density distribution is believed to be Einasto-like. One can also numerically integrate \\eqref{sigmae2} to get reliable estimates of the mass enclosed. Finally, we note that the form similarity of \\eqref{einasto} and \\eqref{sersic}, i.e. fitting functions that describe the 3D mass density of dark matter haloes and the 2D light distributions of galaxies, respectively, could be largely coincidental and should be used with caution when drawing conclusions about the similarity of dynamical evolution that lead to the formation of the stellar components of ellipticals and dark matter haloes." }, "1112/1112.1922_arXiv.txt": { "abstract": "*{} \\abstract{Long tails and streams of stars are the most noticeable upshots of galaxy collisions. Their origin as gravitational, tidal, disturbances has however been recognized only less than fifty years ago and more than ten years after their first observations. This Review describes how the idea of galactic tides emerged, in particular thanks to the advances in numerical simulations, from the first ones that included tens of particles to the most sophisticated ones with tens of millions of them and state-of-the-art hydrodynamical prescriptions. Theoretical aspects pertaining to the formation of tidal tails are then presented. The third part of the review turns to observations and underlines the need for collecting deep multi-wavelength data to tackle the variety of physical processes exhibited by collisional debris. Tidal tails are not just stellar structures, but turn out to contain all the components usually found in galactic disks, in particular atomic / molecular gas and dust. They host star-forming complexes and are able to form star-clusters or even second-generation dwarf galaxies. The final part of the review discusses what tidal tails can tell us (or not) about the structure and content of present-day galaxies, including their dark components, and explains how tidal tails may be used to probe the past evolution of galaxies and their mass assembly history. On-going deep wide-field surveys disclose many new low-surface brightness structures in the nearby Universe, offering great opportunities for attempting galactic archeology with tidal tails.} ", "introduction": " ", "conclusions": "It is an undeniable fact that tidal forces and the formation of tidal tails are overall a second order process in galaxy evolution. The fraction of stars expelled in the intergalactic medium is low, at most a few percent in major mergers. The fraction of gas is more important, but the bulk of the gaseous reservoir is funneled into the central regions. Collisional debris may host star-forming regions, but their contribution to the total star formation rate is minimum. Clearly, most of the activity occurs in the more central and nuclear regions where starbursts and/or AGN fueling is triggered. However, one of the aims of the present Review is to emphasize the idea that tidal debris can provide insightful information about the properties of galaxies, the same way as garbage in trash cans tells us much about the way of life of their owners. The presence of tidal features is an unambiguous proof that a major/minor merger occurred in the recent past, and that at least one of the colliding galaxies had a stellar and/or gaseous disk. The converse is not true though, as not all collisions produce prominent tidal features. Determining when the merger took place is less strait-forward. However, numerical simulations done in cosmological context will soon be able to constrain the survival time of collisional debris and thus give predictions on their age. Comparisons between observations and simulations should then allow us to reconstruct the mass assembly of galaxies. Current generation of wide-field-of-view cameras and the on-going extremely deep surveys of the nearby Universe detect numerous new tidal features of very low surface brightness, offering interesting prospects to galactic archeology. At high redshift, the census of tidal perturbations is much more complex, not only because of dimming and band shifting issues, but also because distant galaxies are much more gas--rich and therefore are intrinsically irregular. This makes the separation between secular and external effects rather ambiguous. Multi-wavelength surveys have revealed the presence in collisional debris of all the constituents of regular galaxies though with different proportions: young and old stars, atomic gas, molecular gas, even possibly dark gas, heavy elements and dust. Star formation seems to proceed there in a similar way as in isolated spiral disks, despite the very different environment at large scale. Tidal tails may in principle even be used to probe some fundamental aspects of physics, including, of course, the properties of tidal forces but also the laws of gravitation, as shown by recent experiments with modified gravity. The fact that tidal forces can be compressive and for instance contribute to the stability of star clusters whereas they are usually associated with destruction processes has only recently been understood. The shape of the tidal tensor explains why massive tidal dwarf galaxies may only be formed within an extended dark matter halo. A theoretical study on the nature and the role of tidal forces in galaxies remains largely to be done and might provide further surprises." }, "1112/1112.4500_arXiv.txt": { "abstract": "The electromagnetic modes possibly unstable in strongly magnetized plasmas are identified. The regime where this instability might stand out compared to the incoherent electron-cyclotron radiation is explored. These modes are relevant to the inertial confinement fusion and the gamma ray burst. ", "introduction": "The magnetic field is ubiquitous in plasmas. The plasma interaction with the magnetic field often determines its dynamical behavior~\\citep{Stone, Biskamp, Fishman, Nakar, Innes}. In particular, the magnetic field could convert the electron kinetic energy into the collective photons, as in the magnetron or the gyrotron~\\citep{gyrotron3, tgyro} and the free electron laser (FEL)~\\citep{Fisch, sonlandau, sonbackward}. In the FEL, a relativistic electron beam gets periodically accelerated amplifying coherent electromagnetic (E\\&M) waves. The resulting lasers have various applications. One natural question would be whether the electron thermal gyro-motion in the magnetic field, not the specifically designed population-inverted plasma state, can excite an analogous process. In this letter, we examine the instability of collective E\\&M waves arising from the electron thermal gyro-motion. The analysis of the electron motion in the presence of a strong magnetic field leads to a theoretical framework similar to that of the Landau damping~\\citep{sonlandau, songamma}, which reveals that there could exist numerous unstable E\\&M modes in relativistic plasmas. Our theory is similar to the well-known electron maser theory~\\citep{chu, Treumann}, but our theory deals with the long time limit while the previous researches mainly focus on the short time limit, which will be discussed in detail at the end of Sec.~(4). We identify the regime where the coherent radiation from this instability might be more intense than the incoherent cyclotron radiation. Various implications of our study on the astrophysical and laboratory plasmas are discussed, including the short gamma ray burst~\\citep{Nakar}, the non-inductive current drive~\\citep{Fisch, sonprl} and the soft x-ray generation in the inertial confinement fusion~\\citep{Tabak}. This paper is organized as follows. In Sec.~(2), the theory of the instability is developed based on the Landau damping theory for the non-relativistic plasmas. In Sec.~(3), the fully relativistic theory is developed. This section is the major result of this paper. In Sec.~(4), we identify the instability inherent in the relativistic plasmas. In Sec.~(5), we discuss the implications of our results. ", "conclusions": "To summarize, the coherent E\\&M instabilities of strongly magnetized relativistic electrons are analyzed in the context of the Landau damping theory. Our main results are Eqs.~(\\ref{eq:landau2}) and (\\ref{eq:landau3}). Our theory based on the long-time scale is relevant with the hard x-ray and the gamma ray generation in the strongly magnetized dense plasmas. In an astrophysical plasma, a strong and large spatial scale magnetic field in the range of $10^{10} - 10^{15}$ gauss is often encountered. Its radiation may have to be re-examined in terms of the possible unstable modes identified in our analysis. More specifically, the theory would be relevant with the generation of the soft and the hard x-ray, and the gamma ray~\\citep{gamma}. The complications when applying the above theory to astrophysical plasmas are the relativistic effect and the electron quantum diffraction~\\citep{sonpla, sonprl, sonbackward, sonlandau}. One more relevant phenomenon is the non-inductive current drive~\\citep{Fisch, sonprl}. During the E\\&M wave amplification, the electrons giving the energy to the E\\&M wave lose their momentum at an increased rate, as the collision frequency increases compared to the case of no interaction with the E\\&M wave. This leads to the well-known non-inductive current drive~\\citep{Fisch, sonprl}. The authors are thankful to Dr.~I.~Dodin and Prof.~N.~J.~Fisch for many useful discussions and advice." }, "1112/1112.2921_arXiv.txt": { "abstract": "We measured the magnetic field of the bright, evolved hot subdwarf HD\\,76431 by means of high-resolution spectropolarimetry. In contrast to previous measurements we were not able to detect a significant magnetic field. We discuss the possibility that this field may be variable. Our search for a possible companion star to HD\\,76431 led to inconclusive results. ", "introduction": "HD\\,76431 (PG\\,0853+019, $m_{\\rm B}=8.9\\,{\\rm mag}$) is a bright B-type star located at high Galactic latitude. According to its atmospheric parameters \\citep[$T_{\\rm eff}=31\\,000\\,{\\rm K}$, $\\log{g}=4.51; $][]{ramspeck01} it is located below the main sequence in the $T_{\\rm eff}$-$\\log{g}$-diagram (see Fig.~\\ref{fig:tefflogg}). \\citet{ramspeck01} measured a subsolar He abundance ($\\log{y}=-1.51$) and concluded that HD\\,76431 is most likely a post-EHB object. \\begin{figure} \\begin{center} \\includegraphics[width=10cm]{tefflogg_hd76431.eps} \\label{fig:tefflogg} \\caption{$T_{\\rm eff}-\\log{g}$ diagram. The grey circles mark sdBs from the SPY project \\citep{lisker05}. EHB evolutionary tracks are taken from \\citet{dorman93}. The location of the main sequence is indicated by the long-dashed horizontal lines. The atmospheric parameters of HD\\,76431 are taken from Ramspeck et al. (2001).} \\end{center} \\end{figure} The role of magnetic fields in stellar evolution remains unclear. One way to address this issue is the analysis of the magnetic properties of evolved stars. While the strong MG-fields of magnetic white dwarfs are easily detectable due to the prominent Zeeman splitting of their spectral lines, weaker fields can only be detected with spectropolarimetry. Such fields of the order of a few kG have been detected in white dwarfs and central stars of planetary nebulae. A first attempt to determine the magnetic fields of bright hot subdwarf stars using this method was made by \\citet{elkin98}, who also targeted HD\\,76431, but did not detect a magnetic field ($B_{\\rm e}=-50\\pm130\\,{\\rm G}$). \\citet{otoole05} on the other hand reported the detection of a magnetic field in HD\\,76431 with a mean longitudinal field strength of $-1096\\pm91\\,{\\rm G}$ using the medium resolution instrument FORS1 mounted at the ESO-VLT. Here we present the results of a spectropolarimetric monitoring campaign of this star using high-resolution data. ", "conclusions": "In contrast to the detection reported by \\citet{otoole05} no significant magnetic field could be found in HD\\,76431. This result is consistent with the non-detection by \\citep{elkin98}. Given that the measurement obtained by \\citet{otoole05} is correct, the magnetic field of this star may be variable on a timescale of years. The reason for this variation may be slow rotation of the star. The sharp metal lines indicate a projected rotational velocity $<10\\,{\\rm km\\,s^{-1}}$. However, Petit et al. (these proceedings) also report the non-detection of a magnetic field in HD\\,76431 based on high-resolution spectropolarimetry. Furthermore, the authors reanalysed the FORS1-data on which the results of \\citet{otoole05} are based and detected no significant magnetic field either. They conclude that the detection of \\citet{otoole05} was spurious and caused by instrumental effects. The question, whether the subdwarf has a companion, remains unclear." }, "1112/1112.0115_arXiv.txt": { "abstract": "{ Physical and phenomenological properties (radius, luminosity, shape of the light curve, etc.) of Cepheids strongly depend on the pulsation period, with the exception of the pulsation amplitude. A possible factor causing a wide range of pulsation amplitudes might be the different atmospheric metallicities of individual Cepheids. } { We studied the influence exerted by the atmospheric iron content, [Fe/H], on the pulsational amplitude of Galactic Cepheids.} { We searched for correlations between the [Fe/H] value and both the observed amplitudes and amplitude related parameters. } { The amplitude of the Cepheid pulsation slightly decreases with increasing iron abundance. This effect is more pronounced for the radial velocity variations and for the shorter pulsation periods. The wavelength dependence of photometric amplitudes is also found to be sensitive to the metallicity. Some of these effects are not consequences of differential line blanketing. Based on the calibrations of the metallicity sensitivity relationships, we derived photometric iron abundance for 21 Galactic Cepheids. The dichotomic behaviour dividing Galactic Cepheids that pulsate in the fundamental mode into short- and long-period groups at the period of 10\\fd47 can be noticed in some diagrams that show metallicity-related dependences. } { We confirm that variety in atmospheric metallicity in Cepheids contributes to the finite range of pulsation amplitudes at a given period. Effects of metallicity on the amplitudes revealed from observational data and the occurrence of the dichotomy also derived from phenomenological data have to be confirmed by appropriate theoretical models of stellar structure and pulsation.} {} ", "introduction": "\\label{sect_1} Classical Cepheids are radially pulsating supergiant stars. This type of pulsation, maintained by the $\\kappa$-mechanism in the outer layers of the star, is stable in a narrow range of the effective temperature, in the nearly vertical {\\em instability strip} in the Hertzsprung-Russell diagram. The fact that this pulsation is a free radial oscillation of the star results in a relationship between the period, $P$, and luminosity, $L$, of Cepheids. Owing to the existence of the period-luminosity ($P$-$L$) relationship, Cepheids are primary distance indicators in astronomy. The precise calibration of this relationship has been in the forefront of Cepheid-related studies for decades. A recently emerged aspect of these studies is the role of metallicity in shaping the $P$-$L$ relationship. In spite of thorough theoretical and observational investigations, the results on the metallicity dependence are controversial (see summaries by \\cite{Retal08}; \\cite{M09}; \\cite{FM10}). Along with other Cepheid-related relationships, the $P$-$L$ relationship is roughly linear if plotted against $\\log P$. The only exception is the period-amplitude ($P$-$A$) relationship. The wide range of possible photometric and radial velocity amplitudes at a given pulsation period may be caused by various effects (pulsation mode and energy, companion star(s), metallicity, etc.). Motivated by the fact that the dependence of the pulsation amplitude on $\\log P$ is neither linear nor unique, we studied the $P$-$A$ relationship of Galactic Cepheids in Paper~I (\\cite{KSz09}). We revised the $P$-$A$ graphs for the $U$, $B$, $V$, $R_{\\rm C}$, and $I_{\\rm C}$ photometric bands. In addition, a $P$-$A$ graph was constructed using the observed peak-to-peak amplitudes of the pulsational radial velocity variations updating the much earlier diagrams (including that by \\cite{CS84} to which we inadvertently missed to refer in Paper~I). One purpose of constructing new $P$-$A$ diagrams has been to study the possible effect of metallicity on the pulsational amplitude, which may contribute to the observed finite range of actual amplitudes at a given pulsation period. \\begin{table*}[t] \\caption{Average [Fe/H] values for various groups of Cepheids} \\begin{tabular}{lcccccccccccccc} \\hline \\noalign{\\smallskip} Sample && \\multicolumn{3}{c}{Solitary Cepheids} && &\\multicolumn{3}{c}{Binary Cepheids}& &&\\multicolumn{3}{c}{Total sample}\\\\ \\noalign{\\smallskip} \\cline{3-5} \\cline{8-10} \\cline{13-15} && [Fe/H] & $\\sigma$ & $N$ &&& [Fe/H] & $\\sigma$ & $N$ &&&[Fe/H] & $\\sigma$ & $N$ \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} Fundamental mode (F) Cepheids, $\\log P > 1.02$ && 0.167 & 0.164 & 49 &&& 0.184 & 0.154 & 34 &&& 0.174 & 0.160 & 83\\\\ Fundamental mode (F) Cepheids, $\\log P < 1.02$ && 0.091 & 0.158 & 114 &&& 0.103 & 0.100 & 78 &&& 0.096 & 0.137 & 192\\\\ First overtone (1OT) Cepheids && 0.108 & 0.142 & 28 &&& 0.068 & 0.118 & 24 &&& 0.089 & 0.133 & 52\\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\label{tab_aveFe} \\end{table*} An influence of the chemical composition on the amplitude of pulsation is expected based on different patterns of the $P$-$A$ plots for Cepheids in our Galaxy and the Magellanic Clouds. van Genderen (\\cite{vG78}) was the first who attributed the differences in the largest pulsation amplitudes of Cepheids to the different average metallicities of the respective host galaxies. If atmospheric metallicity has an influence on largest possible amplitude at a given pulsation period, then it should exert influence on the actual amplitude of individual Cepheids. Here we study the effect of the atmospheric iron content, [Fe/H], on peak-to-peak amplitudes and amplitude-related parameters introduced in Paper~I for classical Cepheids of our Galaxy. Input values have been taken from Paper~I (where the sources of the amplitude and spectroscopic [Fe/H] data as well as the method of their homogenization is also described); in addition, [Fe/H] values published recently by Luck \\& Lambert (\\cite{LL11}) and Luck et~al. (\\cite{Letal11}) have also been taken into account (after applying the same correction as for the [Fe/H] data tabulated in Paper~I). The present sample consists of 329 Galactic Cepheids with known spectroscopic [Fe/H] value. This paper is organized as follows. Section~\\ref{sect_per} deals with the period dependence of the iron content of Galactic Cepheids. Correlations between the [Fe/H] value and various amplitudes and amplitude-related parameters are presented in Sect.~\\ref{sect_amp}. Consequences of our results are discussed in Sect.~\\ref{sect_disc}, while the conclusions are drawn in Sect.~\\ref{sect_concl}. ", "conclusions": "\\label{sect_concl} We confirmed that the observed variety in the atmospheric iron content of Cepheids contributes to the wide range of the actual pulsation amplitudes: more metal-deficient Cepheids tend to pulsate with larger amplitudes. We were able to reveal the metallicity dependence of several parameters ($q$, $k$, $m$, $\\Delta k$, and $\\Delta m$) of Galactic Cepheids. In physical terms, these relationships mean that the amplitude of the radial velocity variations, the ratio of photometric to radial velocity amplitudes, and the wavelength dependence of the photometric amplitudes depend on the iron abundance of the pulsating stellar atmosphere. In some cases, the metallicity sensitivity of the amplitude parameters is not the consequence of the differential line blanketing with the result that stars of higher metallicity are fainter towards shorter wavelengths. These correlations, although not strong, can still be used for deriving the [Fe/H] value of individual Cepheids from photometric observational data. Improvement of these correlations is mainly hindered by the current uncertainties in the spectroscopic abundance determination: individual [Fe/H] values have an error of $0.05 - 0.10$. From multicolour photometric data, we derived a photometric [Fe/H] value for 21 Galactic long-period Cepheids, one of which (GY~Sge) is lacking prior spectroscopic abundance determination. Availability of direct spectroscopic abundance values and multicolour light curves of Cepheids in the Magellanic Clouds would extend the newly found correlations to more negative [Fe/H] values because the interval of atmospheric iron content of the Galactic Cepheids (typically $-0.25 < {\\rm [Fe/H]} < +0.25$) is not particularly wide. Quite a few Cepheids more deficient in iron are known in the outer spiral arm of the Milky Way system but these Cepheids are so distant that their available spectroscopic [Fe/H] values are too uncertain for calibration purposes. Magellanic Cepheids are much more metal-deficient: the mean [Fe/H] is $-0.34 \\pm 0.03$ for the LMC and $-0.64 \\pm 0.04$ for the SMC (Keller \\& Wood \\cite{KW06}). The available observational data on the abundance of individual Magellanic Cepheids are scanty: photometric metallicity values determined by Harris (\\cite{H81}) using the Washington system resulted in metallicities of limited accuracy, and spectroscopic metallicities for a sample consisting of 12-12 Cepheids in each cloud (\\cite{Retal05}). The finite interval of actual [Fe/H] values of Cepheids situated in different parts of the same galaxy is a drawback when attempting to decrease the scatter of the $P$-$L$ relationship for Cepheids in individual galaxies and in settling the question of the metallicity sensitivity of the $P$-$L$ relationship. The wide variety of iron abundance values of Cepheids in the same host galaxy is a reason to be cautious when assigning an `average' metallicity to all Cepheids in a given galaxy. As a rule, the determination of the heavy element abundance of external galaxies (except the nearest ones) is based on the chemical properties of their bright HII regions (\\cite{Zetal94}). Metallicity of these regions, however, corresponds to youngest population of stars involving long-period Cepheids. Short-period Cepheids are somewhat older (and evolve slower) and may have formed from a less processed interstellar matter. Both the effect of metallicity (of non-blanketing origin) on the amplitudes revealed from observational data and the occurrence of the dichotomy at the limit of 10\\fd47 (including its relation to the $2P_{\\rm 2OT} = P_{\\rm F}$ resonance phenomenon) have to be investigated through appropriate theoretical models of stellar pulsation." }, "1112/1112.5730_arXiv.txt": { "abstract": "The differential intensities of Cosmic Rays at Earth were calculated using a 2D stochastic Montecarlo diffusion code and compared with observation data. We evaluated the effect of stretched and compressed heliospheres on the Cosmic Ray intensities at the Earth. This was studied introducing a dependence of the diffusion parameter on the heliospherical size. Then, we found that the optimum value of the heliospherical radius better accounting for experimental data. We also found that the obtained values depends on solar activity. Our results are compatible with Voyager observations and with models of heliospherical size modulation. ", "introduction": "HelMod Code\\cite{HelMod2011} solves the bi-dimensional Parker's particle transport equation\\cite{parker1965}\\!. A Monte Carlo technique is applied on a set of Stochastic Differential Equations (SDEs) fully equivalent to the Parker's equation\\cite{AstraArticle2011}\\!. The model takes into account particle drift effects and latitudinal dependence of the solar wind speed and of the Interplanetary Magnetic Field (IMF). It is described in details in Ref.~\\refcite{HelMod2011}. In the model, the IMF from Parker\\cite{parker1958} is modified introducing a small latitudinal components as described in Ref.~\\refcite{langner2004}. For periods of low solar activity, we take a solar wind speed gradually increasing from the Earth position up to a maximum value near the heliospherical poles ($\\simeq 760$\\,km/s)~\\cite{McComas2000}\\!. For periods approaching the solar maximum we assume a solar wind speed independent on the latitude. The symmetric part of the diffusion tensor, in a reference frame with one axis aligned with the Parker's magnetic-field, is purely diagonal containing transverse ($K_{\\perp \\theta}$ and $K_{\\perp r}$) and parallel ($K_{||}$) components~\\cite{potgieter2000}\\!. The diffusion coefficients are given by~\\cite{potgieter1994} \\begin{eqnarray}\\label{duffusion_comp} \\begin{array}{l} K_{||} = \\beta \\, K_0(t) \\,K_{P}(P)\\,\\left[\\frac{B_{\\oplus}}{3B} \\right] \\ , \\\\ K_{\\perp r} = \\rho_k \\ K_{||} \\ , \\\\ K_{\\perp \\theta} = \\iota(\\theta) \\ \\rho_k \\ K_{||} \\ , \\end{array} \\end{eqnarray} \\noindent where $\\beta = v/c$, $v$ the particle velocity and $c$ the speed of light; the diffusion parameter $K_0$ accounts for the dependence on the solar activity; $ B_{\\oplus}$ is the measured value of IMF at the Earth position - typically $ \\approx 5$\\,nT, but changing with time - obtained from Ref.~\\refcite{SW_web}; $B$ is the magnitude of the large scale IMF as a function of heliocentric coordinates; finally, the term $K_{P}$ takes into account the dependence on the rigidity $P$ of the GCR particle usually expressed in GV.~In the present model $K_P \\approx P$ (e.g., see Ref.\\cite{perko1987}). Furthermore \\ $\\rho_k = 0.05$ \\ and, as described in Ref.~\\refcite{HelMod2011}, \\begin{eqnarray}\\label{enhaced_K} \\iota(\\theta) = \\left\\{ \\begin{array}{rll} 10 \\ , & \\textrm{ in the polar regions},\\\\ 1 \\ , & \\textrm{ in the equatorial region}. \\end{array}\\right. \\end{eqnarray} After the transformations from 3D field-aligned into 2D heliospherical coordinates~\\cite{burg2008}\\!, the symmetric components of the diffusion tensor contains both diagonal ($K_{r r}$ and $K_{\\theta \\theta}$) and off-diagonal terms ($K_{r \\theta}$ and $K_{\\theta r}$), resulting by a proper combination of $K_{\\perp \\theta}$, $K_{\\perp r}$ and $K_{||}$~\\cite{HelMod2011}\\!. ", "conclusions": "We presented the HelMod 2D Monte Carlo code for the study of Cosmic Rays propagation in the inner heliosphere. Both heliospherical shape and size are supposed to be relevant for the modulation process. We introduced a dependence of the diffusion parameter on the heliospherical size, which accounts for the variation with time and solar activity. We compare modulated spectra with experimental data covering the solar cycle 23. Then we found, for our 2D model, the best value of the heliospherical radius, which changes with time. Most of the solar modulation occurs in the inner heliosphere and differences in the heliospherical radius are effective only at energy below a few hundred MeV. Our results are not in contradiction with Voyager observations and models of TS distance as a function of solar activity. We found that LIS form Ref.~\\refcite{Burger2000} fits better observation data at low energy." }, "1112/1112.0609_arXiv.txt": { "abstract": "A quantitative study of the observable radio signatures of the sausage, kink, and torsional MHD oscillation modes in flaring coronal loops is performed. Considering first non-zero order effect of these various MHD oscillation modes on the radio source parameters such as magnetic field, line of sight, plasma density and temperature, electron distribution function, and the source dimensions, we compute time dependent radio emission (spectra and light curves). The radio light curves (of both flux density and degree of polarization) at all considered radio frequencies are than quantified in both time domain (via computation of the full modulation amplitude as a function of frequency) and in Fourier domain (oscillation spectra, phases, and partial modulation amplitude) to form the signatures specific to a particular oscillation mode and/or source parameter regime. We found that the parameter regime and the involved MHD mode can indeed be distinguished using the quantitative measures derived in the modeling. We apply the developed approach to analyze radio burst recorded by Owens Valley Solar Array and report possible detection of the sausage mode oscillation in one (partly occulted) flare and kink or torsional oscillations in another flare. ", "introduction": "Oscillations and quasi-periodic pulsations (QPPs) in radio emission from solar flares have been observed for half a century \\citep{Young61}. Pulsation phenomena can be formally categorized by wavelength of the emission and by frequency/period of the QPPs. For example, at meter and decimeter wavelengths, rapid QPPs occur with periods of tens to hundreds of milliseconds. Typically, these QPPs have rapidly varying amplitudes, variable periods, and narrow bandwidths $\\delta f/f\\lesssim 1$, \\citep[e.g.][]{Young61, Zlobec_etal_1987, Stepanov_Yurovsky_1990, Kurths_etal_1991, Yurovsky_1991, Fleishman94, Aschwanden_etal_1995, Fl_etal_2002, PRL, Magdalenic_etal_2002, Benz_etal_2005, Meszarosova_etal_2005, Chen_Yan_2007, Benz_etal_2011, Chen_etal_2011}. Longer-period QPPs, with periods $\\tau\\gtrsim10$~s are observed in microwaves; such QPPs are often accompanied by QPPs in hard X-rays \\citep[HXR; e.g.,][] {Parks_Winckler_1969, Janssens_etal_1973, Zaitsev_Stepanov_1982b, Kane_etal_1983, Nakajima_etal_1983, Asai_etal_2001, Grechnev_etal_2003, Nakariakov_etal_2003, Stepanov_etal_2004, Melnikov_etal_2005, Fl_etal_2008}. Numerous models have been proposed to account for the various types of oscillations, QPPs, and related phenomena observed at radio wavelengths \\citep[e.g.,][and references therein]{Aschwanden87, Nakariakov_etal_2003, Fl_etal_2008}. Generally speaking, for radio emissions produced by coherent radiation processes (such as fast millisecond pulsations), QPPs are believed to result from nonlinear, self-organizing wave-wave or wave-particle interactions \\citep{Zaitsev_Stepanov_1975b, Meerson_etal_1978, Bardakov_Stepanov_1979, Aschwanden88,Fleishman94,Korsakov98}. In contrast, for radio emission produced by incoherent gyrosynchrotron (GS) radiation from energetic electrons, QPPs are believed to result from modulation of the source parameters (such as the energetic electron distribution, the magnetic field strength, the line of sight, etc.) via magnetohydrodynamic (MHD) oscillations (kink, sausage, or torsional modes) \\citep{Aschwanden04} and/or modulation of electron distribution function via time-dependent acceleration and injection. Most of the reported so far QPPs of GS microwave emission \\citep[e.g.,][]{Kane_etal_1983, Nakajima_etal_1983, Fl_etal_2008} are consistent with modulation of fast electron acceleration/injection, while there is no compelling case in which modulation of GS emission by an MHD oscillation mode would be firmly confirmed. As has been shown by \\citet{Fl_etal_2008}, a well spectrally resolved radio spectrum is crucial to distinguish between MHD oscillations and modulation of fast electron injection/acceleration. Another component needed to firmly detect a particular MHD oscillation mode from the observed radio data is detailed quantitative modeling of observable measures, such as oscillation phase or modulation power, of the GS flaring emission modulated by one or another MHD mode, which is yet unavailable. This paper attempts to partly remedy the situation by analyzing dependence of GS radiation on modulation of the source parameters by various MHD modes, namely, large-scale kink, sausage, and torsional oscillation modes, while we do not consider a special case of the longitudinal slow mode (essentially, the sound wave) where only the plasma density oscillates since this mode can only affect low-frequency radiation from a dense plasma, which has already been studied by \\citet{Nakariakov_Melnikov_2006}. For each mode, time-domain and Fourier analyses are performed on the radio emission and polarization to characterize the properties of the pulsations, and the results are evaluated to identity a signature by which each oscillation mode can be distinguished in observational data. In contrast to approximate sausage-mode modulation modeling performed by \\citet{Fl_etal_2008} based on relatively crude Dulk-Marsh approximation, in this study the microwave (GS plus free-free) emission is modeled using precise fast GS codes recently developed by \\cite{Fl_Kuzn_2010}. The results of the modeling are then used to search for the corresponding QPP signatures in the total power microwave database available from the Owens Valley Solar Array (OVSA). Specifically, we looked through well-calibrated data recorded during 2001-2002 when microwave emission from 412 solar flares was recorded \\citep{Nita_etal_2004}. We found the the vast majority of the radio bursts does not display unambiguous QPPs in the dynamic spectra; only about 10\\% of all events show time variations at the light curves that can be classified as QPPs. From them we selected a few events in which the presence of QPPs is apparent from their dynamic spectra; for those selected events we computed all possible quantitative measures to be compared with the corresponding measures derived from the modeling in the attempt to detect an indication of one or another MHD loop oscillation mode. In most of the case we were unable to firmly identify the genuine cause of the QPPs; however, in two best cases, presented in the paper, the observed QPPs indeed could be produced by an MHD mode: the sausage mode in an (occulted) flare of 28 August 2002 21:41 UT and kink or torsional mode in a flare of 20 July 2002 21:27 UT. % We note that some MHD oscillation modes may or may not produce corresponding oscillations in HXR and/or soft X-rays (SXR), which we also tested for the mentioned events using HXR data available from the Reuven Ramaty High Energy Solar Spectroscopic Imager (RHESSI) and SXR data from the Geostationary Operational Environmental Satellites (GOES). % ", "conclusions": "We have described a detailed modeling of the MHD loop oscillation effect on the GS microwave emission from solar flares. Unlike most of other studies concentrating on observed QPP periods and their possible correspondence to one or another MHD mode, we address the question how the oscillation phase and corresponding modulation amplitude behave as a function of the radio frequency, given that these QPP measures were shown to be highly efficient to identify the true cause of QPPs \\citep{Fl_etal_2008}. We found that the QPP relative phase behavior is distinctly different for the cases of relatively strong magnetic field when the phase repeatedly changes at each of the local peaks at low gyroharmonics and of either dense or tenuous flaring loops with a lower field, displaying no or $\\pi$ phase shift around the (single) spectral peak, respectively. Furthermore, the sausage mode oscillations are distinguishable from the kink and torsional mode oscillations using the frequency behavior of the modulation amplitude. The kink and torsional modes are distinguishable from each other in the case of tenuous source, when oscillations of the source visible area dominate the kink mode effect on the radio QPPs in the optically thick region. However, the kink and torsional mode oscillations in a dense loop are indistinguishable from each other as far as linear order effect (which is identical for these modes) dominates the radio emission response on these MHD oscillations. We have used the modeling results to analyze the OVSA microwave burst database and search for the expected MHD oscillation signatures in some bursts displaying QPPs. Although we did not find many promising events-candidates, we did identified two reasonably good examples in which detection of one or another MHD oscillation mode is likely. Specifically, in one of them the radio QPPs might be caused by the sausage mode, while in the other one by either kink or torsional mode. Comparison of the radio QPPs with SXR and HXR QPPs favors the MHD loop oscillations for these two events compared with the fast electron injection/acceleration oscillations detected in another event using a similar analysis \\citep{Fl_etal_2008}. However, this conclusion must be taken with some caution (given incompleteness and insufficient accuracy of the data) and must be further evaluated and tested on other events with radio QPPs; we plan to perform a more comprehensive statistical study based on full available OVSA database elsewhere including analysis of the polarization data when available, because as evident from the above modeling, such a study would benefit greatly from accurate polarization data. We anticipate that similar data analysis will be even more important for soon-to-be-available imaging spectroscopy data after on-going OVSA expansion (EOVSA) has been completed. In particular, all EOVSA antennas will be equipped with dual-polarization feeds for simultaneous right-hand and left-hand circular polarization measurements. Furthermore, the expansion will greatly improve the image quality and the frequency resolution, which are all needed to perform a more meaningful analysis and yield more reliable conclusions on the cause of radio QPPs in each particular case." }, "1112/1112.0323_arXiv.txt": { "abstract": "I summarise recent results on multi-wavelength properties of distant lensed galaxies, with a particular focus on {\\it Herschel}. Submm surveys have already resulted in a breakthrough discovery of an extremely efficient selection technique for strong gravitational lenses. Benefitting from the gravitational magnification boost, blind mm-wave redshifts have been demonstrated on IRAM, SMA and GBT, and follow-up emission line detections have been made of water, [O{\\sc iii}], [C{\\sc ii}] and other species, revealing the PDR/XDR/CRDR conditions. I also discuss HST imaging of submm lenses, lensed galaxy reconstruction, the prospects for ALMA and e-Merlin and the effects of differential magnification. Many emission line diagnostics are relatively unaffected by differential magnification, but SED-based estimates of bolometric fractions in lensed infrared galaxies are so unreliable as to be useless, unless a lens mass model is available to correct for differential amplification. ", "introduction": "Gravitational lens magnification is a fast route to faint object and faint population studies. This technique has been exploited to reveal or constrain the populations responsible for integrated far-IR and submm extragalactic backgrounds (e.g. \\cite[Smail et al. 1997]{Smail97}, \\cite[Knudsen et al. 2008]{Knudsen08}). Lensing is also one of the few available means of inferring the distribution of dark matter in galaxies (e.g. \\cite[Gavazzi et al. 2007]{Gavazzi07}). Rare lens configurations even have the promise of yielding $\\Omega_{\\rm M}$ and $w$ to $10\\%$; this requires $50$ double Einstein rings, requiring in turn a parent sample of thousands of lenses (\\cite[Gavazzi et al. 2008]{Gavazzi08}). Indeed many current and future applications of lensing are limited by sample size (\\cite[Treu 2010]{Treu10}). Furthermore, most lens surveys have selected strongly for foreground lenses at relatively low redshifts, e.g. $z<0.2$, restricting studies of the evolving dark matter distribution in galaxies. These are strong drivers for both larger and higher redshift samples of lenses. The past year has seen an explosion of interest in submm-selected strong gravitational lenses, mainly driven by {\\it Herschel} open and guaranteed time surveys and their follow-ups. New techniques pioneered with {\\it Herschel} are capable of detecting lenses in much higher numbers and to much higher redshifts than previous optical/near-IR surveys such as SLACS or SLQS (\\cite[Bolton et al. 2006]{Bolton06}, \\cite[Oguri et al. 2006]{Oguri06}). This paper will briefly review some recent key infrared lensing results with a particular focus on the multiwavelength spectral energy distributions and on new data from {\\it Herschel}. ", "conclusions": "Galaxy cluster lenses are a rapid route to probing the populations dominating the cosmic infrared backgrounds, both through direct detections and stacking analyses. The bulk of the observed or predicted cosmic infrared backgrounds at $15\\,\\mu$m, $100\\,\\mu$m, $160\\,\\mu$m and $850\\,\\mu$m have now been resolved into individual sources through ultra-deep imaging in the fields of foreground gravitationally-lensing galaxy clusters. Meanwhile, there is a clear need for larger numbers of strong galaxy-galaxy lens systems, particularly at higher lens redshifts. Following the spectacular confirmation of the $\\sim100\\%$ gravitational lens selection efficiency from submm-selected lenses, there is now a tremendous scope for such lens discovery in the submm with {\\it Herschel}. A further breakthrough has been ``blind'' submm/mm-wave redshift determination, i.e. without recourse to the torturous multi-stage multi-wavelength cross-identification and long-term 8m/10m-class optical/near-IR spectroscopy. Emission line diagnostics already yielding constraints on the physical conditions in these lensed galaxies. However, differential magnification effects cannot be neglected. In particular, the observed CO SLED in any lensed system can easily be un-representative of the SLED in the underlying source. Furthermore, the uncorrected bolometric fractions of strongly-lensed galaixes (magnifications $\\mu>10$) inferred from broad-band SEDs are so unreliable as to be useless. SS thanks friends and colleagues the H-ATLAS (\\cite{Eales10}) and HerMES consortia, whose results are quoted in this paper, STFC (grant ST/G002533/1) for financial support, and the conference organisers for the kind invitation." }, "1112/1112.0053_arXiv.txt": { "abstract": "Be stars possess gaseous circumstellar decretion disks, which are well described using standard $\\alpha$-disk theory. The Be star 28\\,CMa recently underwent a long outburst followed by a long period of quiescence, during which the disk dissipated. Here we present the first time-dependent models of the dissipation of a viscous decretion disk. By modeling the rate of decline of the $V$-band excess, we determine that the viscosity parameter $\\alpha=1.0\\pm0.2$, corresponding to a mass injection rate $\\dot{M}=(3.5\\pm 1.3) \\times 10^{-8}\\ M_\\sun\\,\\mathrm{yr}^{-1}$. Such a large value of $\\alpha$ suggests that the origin of the turbulent viscosity is an instability in the disk whose growth is limited by shock dissipation. The mass injection rate is more than an order of magnitude larger than the wind mass loss rate inferred from UV observations, implying that the mass injection mechanism most likely is not the stellar wind, but some other mechanism. ", "introduction": "\\label{intro} It is now generally accepted that Classical Be stars possess gaseous circumstellar disks that are responsible for their excess emission \\citep[for reviews, see][]{por03,baa01}. Furthermore spectroastrometry and optical interferometic measurements indicate that this material is in Keplerian rotation \\citep{oud11,kra11}. Consequently, as detailed in the review of \\citet{car11}, the steady-state viscous decretion disk model \\citep{lee91} is capable of explaining many of the (time-averaged) observational properties of Be star disks. Most likely these disks are fed by mass-loss from their rapidly rotating central stars. Both observational \\citep{riv98,nei02} and theoretical \\citep{and86,cra09} studies suggest that non-radial pulsations may play an important role. However, to create a viscous decretion disk, this material must be placed into orbit. Since the stars most likely are rotating at less than their critical speed \\citep[e.g.,][]{cra05}, the mass loss mechanism must add some angular momentum before injecting the material into to the disk. From a theoretical point of view, this requirement is problematic, so, as yet, no entirely satisfactory mechanism has been proposed \\citep[see][for a detailed assessment]{owo06}. Regardless of how material is injected into the disk, if this material is in orbit, turbulent viscosity will redistribute it on a viscous diffusion timescale, producing an outflowing decretion disk, as long as the disk is fed continually. However, Be stars exhibit a wide range of variability, ranging from emission line profile variations \\citep[e.g, $V/R$ variability, indicative of density waves within the disk; ][]{ste09,car09}, to photometric and polarimetric variability, indicating that frequently the disks are not fed at a constant rate. In some cases, the disk can disappear for decades before being rebuilt \\citep[see review by][]{und82,wis10}. Oftentimes, this rebuilding phase occurs in a series of outbursts, during which the disk apparently is fed for a short period of time followed by a period of quiescence, during which the nascent disk can partially reaccrete onto the star. Similar outbursts can occur while the disk is disappearing. Once such example is the Be star 28\\,CMa. Over the past 40 years, 28\\,CMa has exhibited quasi-regular outbursts, every 8 years or so, when the star brightens by about half a magnitude in the $V$-band. This usually occurs via multiple individual injections (e.g., the rapid rises seen in Fig. 1), during which the star brightens at a rate of $0.02\\pm0.01\\,\\mathrm{mag\\,d}^{-1}$, due to free-bound and free-free radiation from the disk. The outbursts last for about two years \\citep[see][for a detailed discussion of the long-term photometric variations of 28\\,CMa]{ste03}. A particularly well-monitored outburst occurred from 2001 to 2004, and this outburst was followed by a long quiescent phase from 2004 to 2009. Furthermore it appears that during this quiescent phase, very little additional material was injected into the disk (only a few minor events). Thus nature has provided us the perfect experiment to study how Be star disks dissipate. If we accept as a premise that the viscous decretion disk model is correct (and that no other mechanisms participate in the disk mass budget), then the timescale of the post-outburst photometric decline is controlled solely by the disk viscosity. In this paper, we model the $V$-band light curve with the goals of: 1) directly measuring the disk viscosity parameter, $\\alpha$, and 2) further testing the viscous decretion disk model by comparing its time-dependent predictions of the disk dissipation against the detailed shape of the photometric light curve. \\begin{figure*} \\centerline{\\includegraphics[width=0.85\\linewidth]{f1.eps}} \\caption{$V$-band Light Curve. Visual observations of 28\\,CMa (grey triangles) are shown in comparison to our model fits for different values of $\\alpha$, as indicated. Phase~I (MJD = 52900--53070) is the initial decline, which was used to determine the value of $\\alpha$. Phase~II (MJD = 53070--54670) is the slow disk-draining phase. The inset shows the reduced chi-squared of the Phase~I fit for different values of $\\alpha$. The horizontal dotted line indicates the 90\\% confidence level. } \\label{fig:LightCurve} \\end{figure*} ", "conclusions": "Light curve fitting to the dissipation of a disk can be a powerful tool for measuring $\\alpha$. Further it provides additional quantitative tests of the viscous decretion disk model. Here we have shown that the viscous decretion disk model can reproduce the detailed time-dependence of the dissipation of a Be star disk. Using the $V$-band excess to measure how the disk density decreases with time, we have been able for the first time to directly measure the value of the disk viscosity parameter. We find $\\alpha = 1.0\\pm0.2$, which provides an important clue about the origin of the turbulent viscosity, suggesting that it likely is produced by an instability in the disk whose growth is limited by shock dissipation. Finally, knowing the value of $\\alpha$, we % calculate that the mass decretion rate of 28\\,CMa was ${\\dot M} = (3.5\\pm1.3) \\times 10^{-8}\\ M_\\sun\\ \\mathrm{yr}^{-1}$. Such a large required mass injection rate places strong constraints on any proposed stellar mass loss mechanism." }, "1112/1112.0862_arXiv.txt": { "abstract": "{} {\\zp, one of the closest and brightest massive stars, was the first early-type object observed by the current generation of X-ray observatories. These data provided some surprising results, confirming partly the theoretical predictions while simultaneously unveiling some problematic mismatches with expectations. In this series of papers, we perform a thorough study of \\zp\\ in X-rays, using a decade of \\xmm\\ observations.} {\\zp\\ was observed 18 times by \\xmm, totaling 1Ms in exposure. This provides the highest-quality high-resolution X-ray spectrum of a massive star to date, as well as a perfect dataset for studying X-ray variability in an ``archetype'' object. } {This first paper reports on the data reduction of this unique dataset and provides a few preliminary results. On the one hand, the analysis of EPIC low-resolution spectra shows the star to have a remarkably stable X-ray emission from one observation to the next. On the other hand, the fitting by a wind model of individual line profiles recorded by RGS confirms the wavelength dependence of the line morphology.} {} ", "introduction": "With its very early spectral type (O4Infp, \\citealt{wal72}) and a distance of only 335\\,pc \\citep{van07,mai08}, the star Naos, better known as \\zp\\ (or HD\\,66811), is one of the closest and brightest massive stars. It is therefore one of the most studied objects amongst the O-star population. However, despite the intense work, many open questions remain on its nature. Indeed, \\zp\\ displays several intriguing properties. First, its visible spectrum shows clear signs of helium overabundance and chemical enrichment by CNO-processed material (e.g. \\citealt{pau01}) as well as fast rotation (more than 200\\kms\\ for $v \\sin(i)$, \\citealt{pen96,how97}). Second, it is a known runaway (e.g. from Hipparcos data, \\citealt{mof98}). These properties have led to speculations on its evolutionary status. On the one hand, the chemical enrichment and fast rotation could result from mass and angular momentum exchange through Roche lobe overflow in a binary. \\zp\\ could therefore have been the secondary component of such a system, the supernova explosion of its companion having ejected it from its birth place a few millions years ago \\citep{van96}. On the other hand, \\zp\\ {displays a similar Hipparcos parallax} as stars of the Vela R2 association\\footnote{ This conclusion was based on the original release of the Hipparcos catalog, hence the use of the `old' distance of 430\\,pc in the Schaerer et al. paper, but the parallax similarity remains when using the new reduction of Van Leeuwen (and thus the `new' distance of 335\\,pc).} \\citep{sch97}, and dynamical interactions within this association could have led to the ejection of the (single) O-star \\citep{van96}. In this scenario, the chemical enrichment of \\zp\\ would be explained by the intense rotational mixing occurring in the fast-rotating main-sequence progenitor \\citep{mey00}. In addition, \\zp\\ displays double-peaked emission lines, suggested to arise in a rotating wind \\citep{con74,pet96}, and a compression of the wind in the equatorial plane was detected by \\citet{har96}. Due to its brightness, \\zp\\ was one of the first massive stars observed with high-resolution in X-rays \\citep{kah01,cas01}. At first, its X-ray lines appeared to match expectations as they did show the broad, blueward-skewed profiles expected for the wind embedded shock model \\citep{owo01}. However, the devil was in the details. When quantitatively fitting the line profiles, \\citet{kra03} found a much lower wind attenuation than expected on the basis of the mass-loss rate determined from optical and UV observations (see also \\citealt{osk06}). They also found that the typical optical depths $\\tau_*$, used in the wind-shock models, seemed independent of wavelength, which can only be explained by invoking porosity \\citep{fel03, osk06}. To improve the fitting of the X-ray line profiles, \\citet{leu07} included the effect of resonance scattering: better fits were indeed obtained, without the need of a large reduction in the mass-loss rate, but they also showed that some unexplained discrepancies remain. Re-analyzing the {\\it Chandra} data of \\zp, \\citet{coh10} argue in favor of a reduced mass-loss rate, without the need of any porosity as their new derivation of the optical depths implies an increase with wavelength, as expected from the bound-free absorption opacity of the (cool) wind. Except for \\citet{leu07}, all above studies relied on a single 68\\,ks {\\it Chandra} observation or a 57\\,ks \\xmm\\ exposure taken in 2000. Both facilities have their advantages: while \\xmm\\ globally has a higher sensitivity, {\\it Chandra} has a lower background, and a higher spectral resolution and sensivity at short wavelengths for its grating spectra. Today, however, much more data are available (see below). Considering the uncertainties in the line profile results and the lack of new variability studies, we decided to re-investigate \\zp\\ using the best dataset available at the present time: 18 \\xmm\\ exposures, corresponding to an exposure of 1Ms totaling $>$700\\,ks of useful time (i.e. an improvement by an order of magnitude compared to most previous studies). This dataset thus provides the most detailed X-ray view of an O star to date. The results that we obtained will be presented in a series of papers. The first one will present the data, their reduction, and a few first results; the second one will focus on the X-ray variations of \\zp, using EPIC and RGS data; the last one will present a global analysis of the high-resolution X-ray spectrum, using the merged high-resolution data. This first paper is organized as follows. The dataset and its reduction are presented in Sect. 2, the spectral fits are presented in Sect. 3, the individual line profile fitting in Sect. 4, and the results are summarized and discussed in Sect. 5. ", "conclusions": "In the past decade, about 1Ms of data were obtained by \\xmm\\ on \\zp. Of these, about 30\\% is strongly affected by flares, reducing the useful exposures to 579ks for EPIC-MOS, 477ks for EPIC-pn, and 751ks for RGS. A variety of modes was used, the most reliable being the SW+Thick Filter mode; the use of the medium filter yielded piled-up data, while the data taken in LW+Thick Filter mode do not appear significantly different from those obtained with SW+Thick Filter mode. Attention was paid to this problem, notably by choosing similar extraction regions for all datasets and using only single events. Broad-band EPIC data taken with the Thick filter were analyzed using absorbed optically-thin thermal emissions. Four temperatures are needed to reproduce in a reasonable way the data. Note that the fits are not formally acceptable, but that (1) instruments do not always agree with one another and (2) the reduced noise amplifies the limitations due to the imperfect atomic parameters and standard line profiles. Nevertheless, the EPIC spectra appear remarkably stable over the decade of observations, with only 3\\% dispersions around the average fluxes. A detailed variability study, based on lightcurves, will be presented in Paper II. The combined high-resolution RGS spectrum confirms that the X-ray line profiles vary with wavelength. Fitting individual line profiles using a wind model yields similar onset radius for the X-ray emission, but wind continuum opacities depending on wavelength. This is simply due to the fact that the cool absorbing clumps in the wind are not fully optically thick at all wavelengths, though further modelling is needed in order to adequately reproduce the opacity variations." }, "1112/1112.0786_arXiv.txt": { "abstract": "Astrophysical sources of TeV gamma rays are usually established by Cherenkov telescope observations. These counting type instruments have a field of view of few degrees in diameter and record large numbers of particle air showers via their Cherenkov radiation in the atmosphere. The showers are either induced by gamma rays or diffuse cosmic ray background. The commonly used test statistic to evaluate a possible gamma-ray excess is Li and Ma (1983), Eq.\\ 17, which can be applied to independent on- and off-source observations, or scenarios that can be approximated as such. This formula however is unsuitable if the data are taken in so-called \"wobble\" mode (pointing to several offset positions around the source), if at the same time the acceptance shape is irregular or even depends on \\bref{operating parameters such as the pointing direction or telescope multiplicity}. To provide a robust test statistic in such cases, this paper explores a possible generalization of the likelihood ratio concept on which the formula of Li and Ma is based. In doing so, \\bref{the multi-pointing nature of the data and the typically known instrument point spread function are fully exploited} to derive a new, semi-numerical test statistic. Due to its flexibility and robustness against systematic uncertainties, it is not only useful for detection purposes, but also for skymapping and source shape fitting. \\bref{Simplified Monte Carlo simulations are presented} to verify the results, and \\bref{several applications and further generalizations of the concept are discussed}. ", "introduction": "\\label{sec:intro} The field of very-high-energy (VHE, $>100\\gev$) gamma-ray astronomy is currently in its third generation of instruments, with an advanced future project, CTA \\citep{cta}, already being in its preparatory phase. The currently operated systems H.E.S.S. \\citep{hess}, MAGIC \\citep{magicstereoperformance} and VERITAS \\citep{veritas} have established more than 100 VHE sources in the sky\\footnote{http://tevcat.uchicago.edu/}. The acceptence of these telescope systems covers few degrees in diameter, and declines smoothly with increasing distance from the pointing position. Establishing a gamma-ray signal from an astrophysical source requires to significantly \\bref{prove} a gamma-ray excess over background events that typically remain dominant even after selection cuts. This background is mostly composed of diffuse hadrons, part of which appears almost identical \\bref{to} electromagnetic showers and has to be considered to be irreducible \\citep{irreducible}. Besides these \"gamma-like hadrons\", the irreducible background also contains smaller fractions of diffuse electrons and gamma rays. \\bref{This irreducible background, and the statistical and systematic uncertainties that come with it, are one of the main limiting factors of TeV astronomy; usually, an observational campaign for a given source either reveals one source or none, and only in few cases, or if the effort of a large scan is undertaken, additional unexpected sources are detected. Therefore, a statistical source detection technique that is both sensitive to weak sources and stable against systematic effects is of crucial importance to the field.} The standard test statistic to evaluate an excess of gamma rays from a given sky direction is Li and Ma \\cite{lima}, Eq.\\ 17, which hereafter will be refered to as $S\\tin{LM}$. It was established among several alternatives they evaluated in their paper, based on the fact that this likelihood ratio test statistic was the only one that yielded a satisfyingly Gaussian null-hypothesis distribution. The formula they presented was designed to compare the event numbers of an on- and off-source observation ($N\\tin{on}$, $N\\tin{off}$), and allows for a scaling factor $\\alpha$ between the effective observation times ($t\\tin{on} = \\alpha\\, t\\tin{off}$) to account for unequal exposures. In modern observation practice, most Cherenkov telescope data are \\textit{not} taken in On/Off-mode (see Fig.~\\ref{fig:on_off_wobble}, left), because this strategy implies a lot of observation time dedicated to empty sky regions. Also, it is prone to systematic differences between the on- and off-data caused by instabilities in electronic or atmospheric \\bref{operating} conditions, especially if the off-data \\bref{could not be scheduled contemporaneously enough with the on-source observation.} Therefore, usually the \"wobble\" technique \\citep{wobble} is applied, in which the data are taken at two or more observation positions offset in different directions from the main target coordinates (see Fig.~\\ref{fig:on_off_wobble}, right). In this way, each wobble set provides both on- and off-data at the same time. In its original idea, the exposure shape is considered to have some circular symmetry, and for each wobble set the off-data can be taken from the same observation, at sky positions of similar distance to the telescope pointing position (\"reflected regions\" \\citep{reflectedregion}). In that case, the off-exposure ratio $\\alpha$ is the same for all wobble data sets, and $S\\tin{LM}$ can be applied after summing up the on- and off-events of all wobble sets. This constant $\\alpha$ can also be achieved approximately if the off-data are taken from \\textit{hadron-like} background events (\"template background\", \\citep{templatebackground}), or a ring area around the source position (\"ring background\" \\citep{hessskymapping}), both of which require \\bref{also symmetry assumptions or} an efficiency correction through Monte Carlo simulations of the isotropic background. \\begin{figure} \\centering \\includegraphics[width=13cm]{on_off_wobble.eps} \\caption{ Observation schemes with asymmetric acceptance shapes (green areas). The sky position to be evaluated for a signal is marked with a yellow star, the corresponding off-regions with grey stars. Left: Original On/Off scheme. The off-data provides $N\\tin{off}$, the on-data $N\\tin{on}$, and it needs one exposure ratio $\\alpha=t\\tin{on}/t\\tin{off}$ between those. Right: Wobble scheme with four observation positions. The off data of each wobble set can be taken from all other wobble sets, resulting in four $\\alpha$ parameters and their corresponding $N\\tin{on}$, $N\\tin{off}$. } \\label{fig:on_off_wobble} \\end{figure} In the general case, though, the acceptance symmetry might not hold and a Monte Carlo correction may involve too high uncertainties. This occurs for instance in very \\bref{low-energy observations}, where camera-based acceptance inhomogeneities (dead pixels, trigger fluctuations) are both difficult to model in simulations and furthermore lead to \\bref{features in the acceptance shape that can depend on the Alt/Az pointing direction or other operation parameters of the system}. This is particulary troublesome in two-telescope systems like MAGIC \\citep{magicstereoperformance}, where already the basic geometrical overlap of the two fields of view implies an elongated, Alt/Az-dependent exposure \\bref{shape}. Under these conditions, the \"reflected regions\" approach does not hold, and the off-data for each wobble set have to be taken from the other wobble sets (see Fig.~\\ref{fig:on_off_wobble}, right). This approach can provide \\bref{a sensitive measurement}, because with several wobble sets, the off regions are numerous and well-populated, but it results in a different $\\alpha$ for each wobble set, which is not supported by $S\\tin{LM}$. On top of that, if this procedure has to be done \\bref{separately for different types of data (be it for instance different Azimuth angles or telescope multiplicities)}, it leads to many more $\\alpha$-parameters, and in general, some off-events may happen to be oversampled if they lie in more than one off-region, which is also not considered in $S\\tin{LM}$. As a consequence, while the background density can still be modeled under certain assumptions and some numerical effort \\citep{icrcadvanced}, the test statistic $S\\tin{LM}$, if still applied in some way, becomes very approximative. This may be dealt with in practice by making additional high requirements to the signal-to-background ratio of a detection \\citep{magicstereoperformance}, which however \\bref{limits the effective sensitivity of the instrument}. Therefore, unlike previous efforts \\citep{hessskymapping,icrcadvanced}, this work will not pursue to extract the variables needed for $S\\tin{LM}$ through the complex task of explicit background modeling. Instead, \\bref{the likelihood ratio concept behind that test statistic will be generalized, and new formulae will be derived} that can directly be applied to multi-wobble Cherenkov telescope data. To do so, \\breftwo{no} Monte Carlo simulations or exposure symmetry assumptions will be required. It will only be assumed that the telescope acceptance shape is the same for different wobble data sets if they are taken \\bref{under similar operating conditions}\\footnote{Note that if this basic assumption does not hold, any other symmetry assumption also breaks.}. Besides that, \\bref{this work will also address the disadvantage of $S\\tin{LM}$ that it} depends on the size of the signal region in which $N\\tin{on}$ is calculated. This area is usually defined through an integration radius (\"$\\theta^2$-cut\", with $\\theta$ being the angular distance between reconstructed gamma direction and source position). Its optimization either depends on the background density and an assumption of the source strength, or may involve several trials. \\bref{In this paper,} these assumptions \\bref{are reduced} by accomodating the knowledge of the point spread function (PSF) in the formulae, which makes them independent of the source strength or background density. Likelihood ratios are frequently used to convert a complex likelihood maximization problem to a test statistic that follows a $\\chi^2$-distribution. This possibility was first proven in \\cite{llhratio}, and was suggested for astronomical purposes in \\citep{llhratioastronomy}. The technique is now widely used in counting type experiments, mainly in X-ray and gamma-ray astronomy, and can be applied both for detection and optimization purposes \\citep{llhrvschi2}. \\bref{It should be pointed out} that criticism and potentially more accurate or more general alternatives to the likelihood ratio concept exist \\citep{bayesianlima,limacritics,fisherexactlima}, but are not subject of this work. The structure of the paper is to first define the coordinate systems and naming conventions needed for the calculations. Then, a likelihood function is set up and maximized to gain all relevant parameters \\breftwo{(Sec.~\\ref{sec:llh})}. Based on that, the test statistic is formulated in Sec.~\\ref{sec:ts} and its intrinsic inclusion of $S\\tin{LM}$ is demonstrated. In Sec.~\\ref{sec:generalizations}, some further possible generalizations and applications of the formulae are discussed and a recipe for \"Likelihood Ratio Skymapping\" is suggested. Finally, some example toy simulations are shown in Sec.~\\ref{sec:mc} to verify the method. ", "conclusions": "This paper derives a new generalized test statistic (\\eref{eq:significance}) that can be used for significance calculation in VHE astronomy. The advantages over the existing test statistics are that it flexibly takes into account any number of data subsets from different wobble coordinates and \\bref{operating conditions of the system}, even if the acceptance shape is very irregular \\bref{and different between these operating conditions}. Also, it \\bref{takes} advantage of a known gamma-ray PSF while not requiring the optimization of an integration radius (\"$\\theta^2$-cut\"). The test statistic can be applied to any position in the field of view, so it is very suitable for skymapping purposes. The advantages of this approach is that the test statistic only makes minimal assumptions on the acceptance field of view and does not require any exposure symmetry or Monte Carlo simulations. It is hence unaffected by many systematic uncertainties. \\bref{A} \"Likelihood Ratio Skymapping\" procedure \\bref{is suggested} in Sec.~\\ref{sec:procedure}. The log-likelihood function (\\eref{eq:lllhfphi}) can also be applied to fit the shape and position parameters of the source. \\bref{The} formulae are furthermore extendable to accomodate established sources in the field of view, a non-homogeneous PSF shape or gamma-to-hadron acceptance ratio, an unbinned analysis approach or the \"orbit\" observation mode. \\bref{If the background event density is sufficiently high, the relative} \\breftwo{excess} parameter $\\phi$ is well-suited to calculate a gamma-ray flux map of the field of view. \\bref{In several simulated scenarios it is verified} that the test statistic can not only handle the difficult situations it is designed for, but \\bref{also seems to be systematically higher in the signal case, and therefore more sensitive,} than the commonly used test statistic after Li and Ma \\cite{lima}, Eq.\\ 17." }, "1112/1112.2260_arXiv.txt": { "abstract": "In order to understand environmental effects on star formation in high-redshift galaxies, we investigate the physical relationships between the star formation activity, stellar mass, and environment for $z\\simeq1.2$ galaxies in the 2 deg$^2$ COSMOS field$^*$. We estimate star formation using the \\OII$\\lambda 3727$ emission line and environment from the local galaxy density. Our analysis shows that for massive galaxies ($M_* \\gtrsim 10^{10} M_{\\sun}$), the fraction of \\OII\\ emitters in high-density environments ($\\Sigma_{\\rm 10th}\\gtrsim 3.9\\ {\\rm Mpc^{-2}}$) is $1.7\\pm 0.4$ times higher than in low-density environments ($\\Sigma_{\\rm 10th} \\lesssim 1.5\\ {\\rm Mpc^{-2}}$), while the \\OII\\ emitter fraction does not depend on environment for low-mass $M_* \\lesssim 10^{10} M_{\\sun}$ galaxies. In order to understand what drives these trends, we investigate the role of companion galaxies in our sample. We find that the fraction of \\OII\\ emitters in galaxies with companions is $2.4\\pm0.5$ times as high as that in galaxies without companions at $M_*\\gtrsim 10^{10} M_{\\sun}$. In addition, massive galaxies are more likely to have companions in high-density environments. However, although the {\\it number} of star forming galaxies increases for massive galaxies with close companions and in dense environments, the {\\it average} star formation rate of star forming galaxies at a given mass is independent of environment and the presence/absence of a close companion. These results suggest that interactions and/or mergers in high-density environment could induce star formation in massive galaxies at $z\\sim1.2$, increasing the fraction of star-forming galaxies with $M_* \\gtrsim 10^{10} M_{\\sun}$. ", "introduction": "A key question in understanding the formation and evolution of galaxies is to find what physical parameters are most sensitive to the star formation process in galaxies; e.g., how the star formation activity depends on the environment, and how the relation between star formation and environment changes during the course of galaxy evolution over 10 Gyrs. Observational properties of galaxies in the local universe have been extensively studied in the past. Typically, actively star-forming galaxies in the local universe have lower masses than passive galaxies and the most massive galaxies tend to be inactive for star formation (so-called mass-downsizing; Cowie et al. 1996). In addition, the star formation activity strongly depends on environment: star formation rate (SFR) decreases with increasing galaxy density (e.g., Lewis et al. 2002; Gomez et al.2003; Kauffmann et al. 2004; Mahajan et al. 2010) and the fraction of star-forming galaxies also decreases with increasing galaxy density (e.g., Carter et al. 2001; Balogh et al. 2004; Mateus \\& Sodr\\'e 2004). The fraction of early-type (passive) galaxies is higher in higher-density regions while the fraction of late-type (star-forming) ones is lower in such environment, called the ``morphology-density relation'' and consistent with the above results (e.g., Dressler 1980; Dressler et al. 1997; Goto et al. 2003; Capak et al. 2007). These findings indicate that the star formation activity in galaxies is strongly related to both the stellar mass and their environment. On the other hand, it is well known that the star formation rate density (SFRD) steeply increases in the first $\\sim 2$ Gyrs, peaks at $z\\sim 1-3$, and then decreases by an order of magnitude toward the present day (e.g., Madau et al.1996; Lilly et al. 1996; Shioya et al. 2008; Bouwens et al. 2009). These results suggest that the redshift range of $z \\sim 1-3$ is the key epoch for the most active cosmic star formation in the history of universe. Therefore, studies of galaxies at $1 < z < 3$ are important for understanding galaxy evolution. In recent years, wide and deep surveys have allowed us to study star formation activity in $z \\gtrsim 1$ galaxies. For example, Elbaz et al. (2007) and Cooper et al. (2008) studied the relation between the star formation rate (SFR) in galaxies and the environment at $z\\sim 1$, in the Great Observatories Origins Deep Survey (GOODS; Giavalisco et al. 2004) and the DEEP2 Galaxy Redshift Survey (DEEP2; Davis et al. 2003). These studies showed that the average star formation rate increases with increasing galaxy density (Elbaz et al. 2007; Cooper et al. 2008). We also investigated the relation between the fraction of star-forming galaxies with \\OII\\ $\\lambda 3727$ emission (hereafter ``\\OII\\ emitters'') and the local galaxy density at $z\\simeq 1.2$ in the Cosmic Evolution Survey (COSMOS; Scoville et al. 2007a; Koekemoer et al. 2007) and found that the fraction of star-forming galaxies in high-density regions is as high as that in low-density regions (Ideue et al. 2009; hereafter ``Paper {\\sc i}''). Similar trends were also reported by other observational studies in groups of galaxies (Iovino et al. 2010; Li et al. 2011; Sobral et al. 2011) and cluster environments (Hayashi et al. 2010; Hilton et al. 2010) at $z \\gtrsim 1$. Since the fraction of star-forming galaxies decreases with increasing galaxy density in the local universe as mentioned above, these results suggest that the relationship between the star formation activity in galaxies and their environment dramatically changed from $z\\sim1$ to the present day. On the other hand, many studies suggest that the star formation activity or the fraction of star-forming galaxies also strongly depends on stellar mass even at $z\\gtrsim 1$ (e.g., Noeske et al. 2007; Elbaz et al. 2007; Ilbert et al. 2010). Therefore, it is important to investigate the relationships between the star formation activity, the stellar mass of galaxies, and the environment at $z\\sim 1$, in order to reveal the origin of the observed difference in the environmental dependence of the star formation activity between $z\\sim0$ and $z\\sim1$. In this paper, we focus on the relationships between the star formation, stellar mass, and environment of galaxies at $z \\simeq 1.2$ in the COSMOS field. We have already carried out Subaru imaging observations of the COSMOS field (Taniguchi et al. 2007). We can obtain the sample of \\OII\\ emitters in the COSMOS field by using a narrowband filter NB816 ($\\lambda_c = 8150 $ \\AA\\ and $\\Delta \\lambda {\\rm (FWHM)} = 120 $ \\AA, see Takahashi et al. 2007) and estimate their SFR by using the \\OII\\ luminosity. The \\OII\\ emission line provides a good estimator of the star formation in galaxies at intermediate redshift (e.g., Kennicutt 1998; Jansen et al. 2001). The stellar mass in galaxies in the COSMOS field was obtained from the fitting of their spectral energy distributions (SEDs; Ilbert et al. 2010). Thanks to the very large survey area ($\\sim$ 2 deg$^2$) of the COSMOS field, we are able to obtain an unbiased picture of the star formation activity in galaxies, avoiding the cosmic variance effect. Indeed, the COSMOS field includes various regions with a wide range of galaxy density (e.g., Scoville et al. 2007a, 2007b; Takahashi et al. 2007), which enables us to investigate systematically the environmental effects on galaxy properties. This paper is organized as follows. In section 2, we describe the sample selection and possible contamination of the \\OII\\ emitter sample. Descriptions of the measurements of local galaxy density, stellar mass, and SFR are presented in section 3. We investigate the dependence of star formation on environment and stellar mass, and the effect of close companions on the star formation activity, for our sample at $z\\simeq 1.2$ in section 4. In section 5, we summarize our findings and discuss implications of our results on the evolution of galaxies from $z\\sim 1.2$ to $z\\sim 0$. Throughout this paper, magnitudes are given in the AB system. We adopt a flat universe with the following cosmological parameters; $\\Omega_{\\rm matter} = 0.3$, $\\Omega_{\\Lambda} = 0.7$, and $H_0 = 70\\; {\\rm km\\;s^{-1}\\;Mpc^{-1}}$. ", "conclusions": "In this paper, we have investigated the relations between the star formation, stellar mass, and environment at $z\\simeq 1.2$ in the COSMOS field. As shown in section 4, our results are different from the observational properties of galaxies in the local universe. Based on our observational results, we discuss possible origins of the difference between $z\\sim 0$ and $z\\sim 1$ from a viewpoint of the evolution of galaxies. \\subsection{The Observational Properties of Galaxies at $z \\simeq 1.2$ in the COSMOS Field} The findings of this study are summarized by the following: \\begin{enumerate} \\item At $M_* \\gtrsim 10^{10} M_{\\sun}$, the fraction of \\OII\\ emitters in high-density environments is $1.7\\pm0.4$ times as high as in low- and intermediate-density environments. The fraction of \\OII\\ emitters does not depend on environment at $M_* \\lesssim 10^{10} M_{\\sun}$ (Figure \\ref{fracoii-mass-env}). \\item The fraction of \\OII\\ emitters in galaxies with a companion (likely interacting galaxies) is $2.4\\pm0.5$ times higher than that in those without a companion (likely isolated galaxies) over $M_* \\sim 10^{10}-10^{11.5} M_{\\sun}$ (Figure \\ref{fracoii-mass-compnoncomp}). \\item The fraction of galaxies with a companion for the \\OII\\ emitters at high mass ($M_{*}=10^{10}-10^{11.5}M_{\\sun}$) is $1.5\\pm0.3$ times higher than that at low mass ($M_{*}<10^{10}M_{\\sun}$), while that for the non-\\OII\\ emitters at high mass is $0.7\\pm0.2$ times that at low mass (Figure \\ref{fraccomp-mass-oiinonoii}). \\item The fraction of \\OII\\ emitters with a companion is higher in the higher-density environments (Figure \\ref{fraccomp-mass-env}). The fractions in the low-, intermediate-, and high-density environment at high mass ($M_{*}=10^{10}-10^{11.5}M_{\\sun}$) are $0.2\\pm0.1$, $0.3\\pm0.1$, and $0.4\\pm0.1$, respectively. \\item The average SFR of \\OII\\ emitters strongly correlates with stellar mass in all the environments. For example, the linear fit gives the relation in the high-density environment, $\\log {\\rm }=(0.59\\pm0.08)\\times\\log M_{*}-(4.56\\pm0.78)$. The SFR at a given mass is independent of the environment and the presence of a close companion (Figures \\ref{sfr-mass-env} and \\ref{sfr-mass-compnoncomp}). \\end{enumerate} From the first item described above, it is expected that high-mass \\OII\\ emitters contribute significantly to the active star formation in high-density regions at $z\\simeq 1.2$. Moreover, items 2, 3, and 4 suggest that interactions and/or mergers could trigger star formation in massive galaxies preferentially in high-density environments, while the star formation in low-mass galaxies appears to be independent of environment and does not seem to be affected by the interactions and/or mergers. In other words, $EW$(\\OII) of some inactive high-mass galaxies may become higher due to the active star formation induced by galaxy interactions, and thus we observe such high-mass galaxies as \\OII\\ emitters. On the other hand, considering item 5, it appears that the strength of star formation activity in the star-forming galaxies is not influenced by environmental effects, although the environment and interactions affect the fraction of star-forming galaxies. The environmental dependence of the \\OII\\ fraction is seen only at high mass. It is interesting to note that the influence of interactions and/or mergers on star formation depends both on environment and stellar mass. This can be interpreted such that massive galaxies need external triggers (i.e., interactions and/or mergers) for active star formation, while the less massive galaxies are form stars regardless of the presence of a close companion. One possible triggering mechanism for less massive galaxies is minor merger events (e.g., Taniguchi \\& Wada 1996). However, such satellite galaxies would be too faint to be detected in our imaging survey. As an another possible interpretation, it is expected that less massive galaxies contain much gas and can naturally form stars since evolution of those galaxies is slow compared to massive galaxies, which is expected from mass-downsizing (e.g., Cowie et al. 1996). Therefore, we conclude that the environmental dependence of the \\OII\\ fraction (i.e., the fraction of star-forming galaxies) is seen only at high mass, because 1) interactions and/or mergers induce the star formation only in massive galaxies and 2) the probability of interactions and/or mergers depends on local galaxy density. Peng et al. (2010, 2011) recently found that the environment quenching (or the satellite quenching) depends on the local galaxy density but not on the stellar mass of galaxies. We observe a different environmental effect from that found by Peng et al. (2010, 2011), since our results suggest that the influence of interaction and/or mergers depend on both environment and stellar mass. \\subsection{Implications to the evolution of galaxies from $z\\sim 1$ to $z\\sim 0$} In the local universe, it is observed that the fraction of red galaxies in high-density environments is higher than that in low-density environments, and this difference becomes larger at lower mass (e.g., Baldry et al. 2006). Namely, the fraction of blue galaxies in high-density environments is lower than that in low-density environments, especially at low mass. Iovino et al (2010) also found a similar trend at $z<0.5$. Furthermore, they found that the fraction of blue galaxies at $z=0.6-0.8$ does not depend on the environment in $M_* > 10^{10}$. These are in contrast to the environmental dependence of the \\OII\\ fraction in massive galaxies at $z\\simeq 1.2$ mentioned above. Now a question arises: what is the origin of the difference between $z\\sim 1$ and $z\\sim 0$? Our results suggest that interactions and mergers are likely to induce star formation in massive galaxies in high-density environments at $z\\simeq 1.2$, resulting in the observed high \\OII\\ fraction of massive galaxies in high-density environments. These inactive massive galaxies where star formation would be induced by interactions are likely to have some cold gas for star formation. We have also found that the average SFR at a given mass is similar among the different environments and is independent of the presence of a close companion (Figures \\ref{sfr-mass-env} and \\ref{sfr-mass-compnoncomp}), the strength of the star formation induced by interactions in massive galaxies does not seem to be significantly different from other massive star-forming galaxies on average. Therefore we expect that there are already some inactive high-mass galaxies with sufficient cold gas for SFR at a given mass at $z\\simeq 1.2$. On the other hand, the fraction of gas-poor galaxies is higher in higher density regions at low-$z$ (e.g., Giovanelli \\& Haynes 1985; Bertram et al. 2006). Therefore it is unlikely that a significant starburst occurs in interactions or mergers between gas-poor galaxies (i.e., dry mergers; e.g., Tran et al. 2005; Cattaneo et al. 2008), even though interactions and/or mergers still occur in high-density regions at low-$z$ (e.g., Patton et al. 2011). In this context, one important difference in galaxies between $z\\sim 1.2$ and low $z$ is the cold gaseous content. Since galaxies at $z\\sim 1.2$ are in early stages of their evolution, it is likely that their cold gaseous content is, on average, larger than that in galaxies at low $z$ even if such galaxies are located in relatively high-density regions. It is thus expected that interactions and mergers at $z\\simeq 1.2$ could cause intense star formation more frequently since it is known that interactions and mergers between two gas-rich galaxies (wet mergers) can induce active star formation and rapidly consume the cold gas (e.g., Barton et al. 2000; Woods et al. 2006). In fact, Lin et al. (2008, 2010) suggest that the fraction of dry mergers gradually increases from $z\\sim 1$ to $z\\sim 0$, while the fraction of wet mergers decreases at the redshift range. If there are many inactive high-mass galaxies with sufficient cold gas for star formation at $z\\simeq 1.2$ as mentioned above, there may be some mechanisms responsible for quenching star formation in galaxies while retaining their gas. For example, it might be expected that high-mass galaxies have experienced starbursts in the past, and then their gas becomes hot and their star formation stops in spite of the existence of the (hot) gas. Later the hot gas cools down into cold gas, and then high-mass galaxies actively form stars when interactions and/or mergers occur. As another scenario, high-mass galaxies might have consumed most of gas in the past, which is expected from the mass-downsizing of galaxy formation (e.g., Cowie et al. 1996), and the gas density was already too low to cause intense star formation (Kennicutt 1989). However, if interactions can sufficiently perturb the remaining gas, the star formation can be triggered and they can be observed as high-mass star-forming galaxies in high-density environment. In this case, such galaxies immediately consume their gas in the star formation and these massive galaxies in high-density environment are observed to be passive at $z \\lesssim 1$. From these considerations, we suggest the following scenario; The high-mass star-forming galaxies for which star formation are induced by interactions and/or mergers contribute significantly to the active star formation in high-density regions at $z\\simeq 1.2$. These massive star-forming galaxies in high-density regions at $z\\simeq 1.2$ could quickly consume most of the accreted cold gas. If this is the case, the star formation activity may not be enhanced when interactions and/or mergers occurred in high-density environments at lower redshift. In this context, the quenching of star formation in massive galaxies in high density environments is expected to lead to the shift of major star formation in the universe from high-density regions to low-density ones at $z\\lesssim 1$. We would like to thank the anonymous referee for her/his very useful comments and suggestions. We also thank all members of the COSMOS team. This work was financially supported in part by the Japan Society for the Promotion of Science (Nos. 17253001, 19340046, 23244031, and 23654068). Y. I. is financially supported by the Japan Society for the Promotion of science (JSPS) through JSPS Research Fellowship for Young Scientists." }, "1112/1112.0337_arXiv.txt": { "abstract": "We investigate whether any multi-planet systems among {\\em Kepler} candidates (2011 February release) can harbor additional terrestrial-mass planets or smaller bodies. We apply the ``packed planetary systems'' hypothesis that suggests all planetary systems are filled to capacity, and use a Hill stability criterion to identify eight 2-planet systems with significant gaps between the innermost and outermost planets. For each of these systems, we perform long-term numerical integrations of 10$^7$ years to investigate the stability of 4000$-$8000 test particles injected into the gaps. We map out stability regions in orbital parameter space, and therefore quantify the ranges of semi-major axes and eccentricities of stable particles. Strong mean-motion resonances can add additional regions of stability in otherwise unstable parameter space. We derive simple expressions for the extent of the stability regions, which is related to quantities such as the dynamical spacing $\\Delta$, the separation between two planets in units of their mutual Hill radii. Our results suggest that planets with separation $\\Delta < 10$ are unlikely to host extensive stability regions, and that about 95 out of a total of 115 two-planet systems in the {\\em Kepler} sample may have sizeable stability regions. We predict that {\\em Kepler} candidate systems including KOI 433, KOI 72/Kepler-10, KOI 555, KOI 1596, KOI 904, KOI 223, KOI 1590, and KOI 139 can harbor additional planets or low-mass bodies between the inner and outer detected planets. These predicted planets may be detected by future observations. ", "introduction": "Early studies of extrasolar planetary systems showed residual velocity trends in Keplerian orbit fits to radial velocity data \\citep[e.g.,][]{marc98,butl98,marc99,vogt00,fisc01}, suggesting that these systems may host additional, undetected planets. \\citet{fisc01} noted that about half of the stars in their sample of 12 systems showed residual trends greater than the expected scatter due to measurement uncertainties and stellar noise. Most of these systems were later confirmed to harbor additional planet(s). In more recent years, the study and prediction of undiscovered planets have been aided by long-term N-body simulations. These numerical investigations searched for stability zones in multi-planet systems by integrating hundreds to thousands of test bodies, which were injected into empty regions between known planets \\citep[e.g.,][]{meno03,barn04,raym05,ji05,rive07,raym08}. For example, a putative Saturn-mass planet in HD~74156 was first predicted by \\citet{raym05} through numerical simulations that showed a stable region between planets b and c. The planet was later discovered by \\citet{bean08}, although there have been questions about the validity of the detection~\\citep{witt09,mesc11}. This prediction was motivated by the ``packed planetary systems'' (PPS) hypothesis. The PPS hypothesis is the idea that planetary systems are formed ``dynamically full'' and filled to capacity, and any additional planets will cause the systems to be unstable \\citep[e.g.,][]{barn04, raym05, raym06, barn07}. Consequently, planetary systems with stable stability zones between the innermost and outermost planets are suggestive of additional, yet-undetected planets. Reasons for the non-detections of hypothetical planets include lack of sufficient data, such as non-transiting planets that require more data to detect them via transit timing variations, or planetary masses/radii that are below detection limits. The orbital properties of predicted planets can be identified through long-term numerical simulations. Support for the PPS hypothesis comes from early observations of packed multi-planet systems that led to this hypothesis \\citep[e.g.,][]{butl99,marc01b,marc01a,fisc02,mayo04}, apparent consistency between the planet-planet scattering model and packed systems \\citep{raym09}, the remarkably dense and packed Kepler-11 system \\citep{liss11a}, theoretical work \\citep[e.g.,][]{cham96,smit09}, and other investigations \\citep[e.g.,][]{rive00,gozd06}. In the present study, we apply the PPS hypothesis to multi-planet candidate systems discovered by the {\\em Kepler} team during the mission's first four and a half months of data \\citep{boru11}. The {\\em Kepler} mission is a transit survey designed to search for Earth-sized planets \\citep{boru10,koch10,jenk10,cald10}, and is sensitive to terrestrial-class and larger planets located at a large range of separations from their host star. {\\em Kepler} can detect multiple transiting systems for densely-packed planets with nearly coplanar configurations or with serendipitous geometric alignment, and the dynamics and statistics of {\\em Kepler} multi-planet systems are providing a wealth of information about planetary systems \\citep[e.g.,][]{stef10,lath11,liss11a,liss11b}. Given that planetary systems have been discovered with densely packed planets, we seek to test the PPS hypothesis and predict additional planets in {\\em Kepler} candidate multi-planet systems. In Section \\ref{datamethods}, we discuss {\\em Kepler}'s sample of multi-planet systems as well as our methodology for evaluating each planetary system's level of dynamical packing. We also explain our methods for running numerical simulations and our choice of initial conditions. In Section \\ref{results}, we present the results from long-term numerical integrations and illustrate them using stability maps. Section \\ref{mountain} discusses the dynamical interpretation of our work, in particular the relationship between planetary spacing and the extent of an inter-planet stability region. We then summarize the restrictions and scope of our study (Section \\ref{scope}) and state our conclusions (Section \\ref{conclusion}). ", "conclusions": "\\label{conclusion} The ``packed planetary systems'' model advocates the idea that all planetary systems are formed to capacity. To test this hypothesis, we investigated the packing density of {\\em Kepler} candidate two-planet systems from the first four and a half months of the mission. Through numerical calculations, we determined whether regions of stability exist between known planets with wide separations, i.e.,~in systems that seemed the most unpacked based on how well they satisfy Hill stability. Discovery of a stable region suggests that a low-mass body could be present in the gap, which would then bring the system to a more ``packed'' state. With time, such predictions will be shown to be correct or incorrect, allowing us to gauge the success of this model. We performed detailed numerical simulations of eight, two-planet {\\em Kepler} systems, selected using an analytical $\\beta/\\beta_{\\rm crit}$ stability criterion. In addition to the known planets, we included 4000$-$8000 test particles per planetary system, allowing both circular and non-circular, and coplanar and non-coplanar orbits. These test particles are good proxies for low-mass bodies such as terrestrial planets as well as small bodies such as asteroids or dwarf planets. We integrated all bodies for 10$^7$ years; we defined stable particles as those that remained stable during the length of the integration and unstable particles as particles that experienced a collision or ejection. Our results (Figures \\ref{maps1} to \\ref{maps3}) indicated that all of the planetary systems investigated here (KOIs 433, 72, 555, 1596, 904, 223, 1590, and 139) can pack additional, yet-undetected bodies in the identified stable locations. We also discussed relationships relating dynamical spacing between known planets and the extent of the inter-planet stability region. We derived an analytical relationship relating the largest possible eccentricity of a stable test particle to the semi-major axes and Hill radii of the two planets surrounding the particle. We also demonstrated that $\\Delta$, the separation between two planets in units of their mutual Hill radii, can be a reasonable predictor of whether or not stability regions can exist between planets. The cut-off occurs at a critical $\\Delta$ between 10 and 15. We suggest that planets with separation $\\Delta < 10$ are unlikely to host extensive stability regions. Based on this $\\Delta=10-15$ cut-off, we suggest that about 95 out of a total of 115 two-planet systems in the February 2011 {\\em Kepler} sample may have sizeable stability regions." }, "1112/1112.1107_arXiv.txt": { "abstract": "We build a hydrodynamical model for computing and understanding the Sun's large-scale high latitude flows, including Coriolis forces, turbulent diffusion of momentum and gyroscopic pumping. Side boundaries of the spherical 'polar cap', our computational domain, are located at latitudes $\\geq 60^{\\circ}$. Implementing observed low latitude flows as side boundary conditions, we solve the flow equations for a cartesian analog of the polar cap. The key parameter that determines whether there are nodes in the high latitude meridional flow is $\\epsilon=2 \\Omega n \\pi H^2/\\nu$, in which $\\Omega$ is the interior rotation rate, n the radial wavenumber of the meridional flow, $H$ the depth of the convection zone and $\\nu$ the turbulent viscosity. The smaller the $\\epsilon$ (larger turbulent viscosity), the fewer the number of nodes in high latitudes. For all latitudes within the polar cap, we find three nodes for $\\nu=10^{12}{\\rm cm}^2{\\rm s}^{-1}$, two for $10^{13}$, and one or none for $10^{15}$ or higher. For $\\nu$ near $10^{14}$ our model exhibits 'node merging': as the meridional flow speed is increased, two nodes cancel each other, leaving no nodes. On the other hand, for fixed flow speed at the boundary, as $\\nu$ is increased the poleward most node migrates to the pole and disappears, ultimately for high enough $\\nu$ leaving no nodes. These results suggest that primary poleward surface meridional flow can extend from $60^{\\circ}$ to the pole either by node-merging or by node migration and disappearance. ", "introduction": "It has been demonstrated both observationally and theoretically that meridional circulation plays a crucial role in the workings of the solar cycle. In flux-transport dynamos, the meridional circulation is responsible for determining the cycle period \\citep{ws91,dc99,krs01} and also plays an important role in setting the cycle 'shape', that is, its rise and fall patterns. It follows that it is very important for understanding and predicting solar cycles that we have the best possible information about meridional circulation on the Sun. The accurate measurements of surface flow-patterns as a function of latitude as well as time would be very useful for simulating the evolutionary pattern of the Sun's large-scale fields, particularly the polar fields \\citep{bsss04,wls05,ddg08}. \\begin{figure}[hbt] \\epsscale{1.0} \\plotone{f1lores.eps} \\caption{The Sun's characteristic conveyor-belts during cycles 22 (left frame) and 23 (middle frame), derived from surface observations and mass-conservation, are shown. The speed associated with the colors is shown in the colorbar (right frame). The maximum surface flow-speeds at $25^{\\circ}$ latitude were almost the same in both the conveyor-belts, but the flow turning down towards the equator around $60^{\\circ}$ latitude during cycle 22 made the length of the primary conveyor-belt shorter than that during cycle 23, in which flow went all the way to the pole before turning equatorward.} \\label{mc2223} \\end{figure} While meridional flow at all latitudes is important for the solar dynamo, its pattern near the poles of the Sun may be particularly significant. Recently \\citet{dgdu10} showed the variation in solar cycle duration could be explained by variations in the latitudinal extent of the Sun's primary 'conveyor belt' or meridional circulation. Observations of the Sun's surface Doppler plasma flow from Mount Wilson Observatory data indicate that in both cycles 22 and 23, the maximum poleward surface flow-speed in the primary belt was the same (\\citet{ub05}; also see the figure 1 of \\citet{dgdu10}). But in cycle 22, the primary circulation cell flowed poleward only to about $60^{\\circ}$ latitude, thus making a shorter path for the magnetic flux transport via the conveyor belt and resulting in a cycle duration of $\\sim 10.5$ years (see the left frame of Figure 1). On the other hand, in cycle 23, the poleward surface flow went all the way to the poles (see the middle frame in Figure 1), leading to a longer path via the conveyor-belt cycle of $\\sim 12.5$ years. In the past, some flux-transport dynamo calculations \\citep{bebr05} and surface-transport calculations \\citep{jcss08} have dealt with possible multi-cell meridional flow scenarios, in the context of understanding the role of these flows in solar cycle features. We now see that these studies are not just the merely playing with models; such scenarios could happen in reality. The change in the surface poleward flow-pattern -- its reversing around $60^{\\circ}$ as it did in cycle 22 and maintaining poleward flow all the way to the pole as in cycle 23 -- have significantly impacted the duration of cycle 23 and the length of the minimum that followed it, compared to that in cycle 22. The plasma velocity can also be determined by helioseismic analysis, either from ring diagrams or from time-distance diagrams \\citep{gdsb97, hetal02,zk04,gkhhk08,gbs10}. Alternatively, features seen on the images such as magnetic structures and supergranule cells, can be tracked with cross-correlation analysis to yield a drift velocity for that feature \\citep{khh93,sd96,sks06,szk07,sksb08}. \\citet{u10} reanalyzed the Mount Wilson surface Doppler data for cycles 22 and 23, from 1986 through 2009. \\citet{u10} computed meridional flow profiles up to at least $80^{\\circ}$ latitude, found smooth evolution of the signal from one year to the next, and confirmed the difference in high latitude flow patterns between cycles 22 and 23, in particular, the existence of a second reversed cell poleward of about $60^{\\circ}$ during most of cycle 22, and a single cell with poleward surface flow all the way to the poles during the major part of cycle 23. Differential rotation throughout the solar convection zone is fairly well known from helioseismic measurements \\citep{t04}. It is difficult to measure at high latitudes because of foreshortening and other effects. Most methods measure the linear rotational velocity rather than the angular rotation rate, so it is particularly difficult to calculate the angular measure with the short moment-arm near the poles. It appears that on average the angular measure of differential rotation at the surface declines monotonically to the poles \\citep{b00}, though the existence of a 'polar vortex' has been suggested by theory \\citep{g79}. Although our focus on this paper is more on meridional flow at high latitudes than on differential rotation there, our model will calculate both quantities, so we will need to compare results from the model with both flows. Given the important consequences of a second, reversed meridional cell in high latitudes, it is important to develop hydrodynamical theories that could indicate what to expect to occur in the Sun. What physics determines the presence or absense of a second cell? Can we develop a simple physical argument that we should expect to see a single primary cell, or two or more cells? This is the question we attempt to answer in this paper. The theory we develop filters out all convective instability and concentrates on forcing high latitude meridional circulation mechanically by the primary meridional flow, and possibly by differential rotation, from lower latitudes. We recognize that in the Sun there could be both this mechanical forcing and axisymmetric convection in high latitudes. Meridional circulation is produced in virtually all fluid dynamical models used to simulate and understand the origins of the differential rotation of the Sun. These fluid dynamical models fall generally into two classes: mean field models that are axisymmetric and include parameterizations of turbulent transport of momentum \\citep{drsc89,r05}, and global 3D numerical models that simulate global convection in a deep rotating spherical shell \\citep{mbdt08}. For both types of models there are numerous results that show a wide variety of meridional flow patterns, but recently both approaches have been converging toward a common result of a dominant meridional flow cell that has poleward flow near the outer boundary, and equatorward return flow near the bottom \\citep{r05,mbdt08}. For some parameter choices the mean field models give a second, reversed cell at high latitudes \\citep{r05} and the 3D global convection models can give multiple high latitude cells \\citep{mbdt08}. Both classes of models also show that the percentage fluctuations in the amplitudes of meridional flow with time are much larger than for the differential rotation. This appears to be due to time-fluctuations in the Reynolds stresses and the fact that the meridional flow is a result of a slight imbalance between large forces, while the forces responsible for differential rotation are small enough to make the differential rotation change only very slowly. These models are all global, and focused primarily on the problem of understanding the details of differential rotation with depth and latitude in the Sun, as well as such features as torsional oscillations. To focus primarily on the fluid dynamics of high latitudes, it is possible to build simpler models, particularly at first, and then build up to more realistic models more comparable to the global models just described. This is the approach we take below. ", "conclusions": "We have developed a relatively simple hydrodynamical model of the circulation at high latitudes in the solar convection zone and photosphere that contains only three forces: pressure gradients, viscous and Coriolis forces. The model equations are solved in a cartesian 'analog' of a spherical polar cap that leaves out curvature effects as well as the large density increase with depth. This system is assumed to be stress-free at the top and bottom, corresponding to the top and bottom of the solar convection zone. It is forced with meridional flow and differential rotation, guided by observations, imposed at the low latitude boundary of the cap, placed at latitudes between $60^{\\circ}$ and $70^{\\circ}$ latitude. While the inclusion of such a boundary is artificial, it is intended to separate the physics of low and mid-latitudes, responsible for the primary meridional circulation cell that has poleward flow at the top, from the physics active at high latitudes, which should be the primary determinant of the circulation found there. Our general results are that, as the turbulent viscosity is increased, the number of nodes decreases. As the meridional flow is increased for a given turbulent viscosity, the latitude of a node increases; or, in other words, the node migrates poleward. The first of these results is explained by the viscous forces increasingly overpowering the deflecting effect of the coriolis force to allow the poleward flow of the primary cell to reach a higher latitude. The second is due to the increased poleward momentum of the poleward moving particles that allows them to reach a higher latitude before sinking down to feed the return flow. Overall, we find that our general results are not particularly sensitive to the differential rotation imposed at the boundary, provided it is plausible for the Sun. Unlike for the meridional circulation, changes in the differential rotation of the Sun at all latitudes are very small percentages of the mean differential rotation. Most interestingly, we find that the decrease in the number of nodes as the turbulent viscosity is increased is not monotonic. In particular, in the neighborhood of $\\nu=10^{14} {\\rm cm}^2\\,{\\rm s}^{-1}$ in this model, we find that two nodes merge as the meridional flow is increased, leaving no nodes for higher meridional flow at the boundary. This is true even though, for still higher viscosity, there remains one node. Two nodes implies the presence of both a reversed (clockwise) cell and, on its poleward side, a second, weak, counterclockwise cell that has a poleward flow near the top, just as in the primary cell. With the merger of the two nodes, the reverse cell is squeezed out, merging the primary cell with its polar counterpart. We speculate that it is this phenomenon that could be responsible for the primary cell reaching all the way to the poles during much of cycle 23, whereas there was a node near $65^{\\circ}$ in both cycles 22 and the early stages of cycle 24. At present, observations can not tell us whether there is a second node near $80^{\\circ}$ as our model predicts. Better observations in the future of velocities at the highest latitudes would obviously be valuable for testing this theory and that of the spherical polar cap that will be developed later. Better information about variations of meridional flow with depth at high latitudes would also be very useful. Our solutions are for steady flow, but we know that the meridional flow at high latitudes on the Sun changes with time. Given our results, we can expect that a time dependent theory could determine whether changes in meridional flow at the boundary would lead to changes in meridional flow at higher latitudes that agree with observed changes in flow at the highest latitudes. If the Sun merges two nodes with a meridional flow increase, does the model do the same? Does the model predict a merger as a result of an increase in meridional flow that is not observed on the Sun? Questions such as these can only be answered with a time dependent model. Our results suggest that it may not be necessary to invoke any additional physics to explain changes in the meridional flow cells with time in high latitudes. That does not mean, however, that the presence or importance of such physics can be ruled out at this time. There are several additional effects that should be included in the high latitude meridional circulation model to make it more realistic. Even keeping the same physics, results from this model using spherical geometry could change significantly. With spherical geometry the meridians converge to the pole, making it harder for as much mass flux to reach the pole as does in the straight channel. In addition, the spherical problem does not separate easily in radius and colatitude due to the Coriolis forces. This has the effect of linking different latitudes and different depths of the flow in ways not present in the cartesian analog. All of these spherical effects should influence the structure of the flow, including the location of nodes in the streamfunction, and how many nodes there are for a given turbulent viscosity. Within spherical geometry, the flow patterns will change substantially when the density increase with depth through the convection zone is included; the patterns of mass flux, or $\\rho v$, could look somewhat like those without the density variation, but the velocity itself should decline substantially with depth. This effect could also change the location and number of nodes present. In this version of the model, effects of the variation of the turbulent viscosity with depth should also be studie; we have already seen that the number and latitude of nodes in the meridional flow is sensitive to the turbulent viscosity used. Allowing this quantity to vary with radius in the spherical case could change the results substantially, as could allowing for the density increase with depth. In its present form, the model is for axisymmetric motions, and, beyond the turbulent diffusion, our model contains no effect of organized global scale Reynolds stresses. Yet these stresses are surely important in driving the global differential rotation as well as playing a role in the maintenance of the primary meridional cell. We do not know to what latitude these Reynolds stresses reach, but they could be active within the polar cap. Their possible effects on the polar meridional flows should also be considered. Our model currently also does not include explicitly any thermodynamics. Allowing for departures of the temperature from the adiabatic gradient could be important, particularly at the top and the bottom of the convection zone. Finally, our model is hydrodynamic, so no effects of magnetic fields are included. But the Sun is a dynamo which generates fields, some quite strong, throughout the convection zone, so the effects of these should also be taken into account. One possible effect is that different amplitudes of polar fields might influence the amplitude of meridional flow in high latitudes, as well as the locations of its nodes." }, "1112/1112.3102_arXiv.txt": { "abstract": "Disk accretion onto weakly magnetized astrophysical objects often proceeds via a boundary layer that forms near the object's surface, in which the rotation speed of the accreted gas changes rapidly. Here we study the initial stages of formation for such a boundary layer around a white dwarf or a young star by examining the hydrodynamical shear instabilities that may initiate mixing and momentum transport between the two fluids of different densities moving supersonically with respect to each other. We find that an initially laminar boundary layer is unstable to two different kinds of instabilities. One is an instability of a supersonic vortex sheet (implying a discontinuous initial profile of the angular speed of the gas) in the presence of gravity, which we find to have a growth rate of order (but less than) the orbital frequency. The other is a sonic instability of a finite width, supersonic shear layer, which is similar to the Papaloizou-Pringle instability. It has a growth rate proportional to the shear inside the transition layer, which is of order the orbital frequency times the ratio of stellar radius to the boundary layer thickness. For a boundary layer that is thin compared to the radius of the star, the shear rate is much larger than the orbital frequency. Thus, we conclude that sonic instabilities play a dominant role in the initial stages of nonmagnetic boundary layer formation and give rise to very fast mixing between disk gas and stellar fluid in the supersonic regime. ", "introduction": "Accretion onto astrophysical objects possessing a material surface (as opposed to accretion onto black holes) always involves a non-trivial interaction of the incoming gas with the object's outer layers. Examples of such objects include white dwarfs in cataclysmic variables (CVs), young stars gaining material from a protoplanetary disk, and accreting neutron stars. In all these systems, one must understand how the incoming gas shares its angular momentum with the accreting object and how it mixes with the previously accreted material. If the magnetic field of the central object is strong enough, it can disrupt the disk at some distance above the surface. Inside this region, accretion proceeds along field lines \\citep{GhoshLamb,Koldobaetal}. The critical value of the magnetic field at the surface of the central object for magnetic disruption is given by \\ba B_{*,\\text{crit}} = 5 \\times 10^4 \\left(\\frac{\\beta}{.5} \\right)^{-7/4} \\left(\\frac{M_*}{.6 M_\\odot}\\right)^{1/4} \\left(\\frac{\\dot{M}}{10^{-8} M_\\odot \\ \\text{yr}^{-1}}\\right)^{1/2}\\left(\\frac{R_*}{9 \\times 10^3 \\ \\text{km}}\\right)^{-5/4} \\text{G}. \\label{Bdisrupt} \\ea Here, we have used parameters for the mass and radius that are typical for CVs in outburst \\citep{Bergeronetal,Hachisu,Patterson}. The parameter $\\beta$ is a dimensionless factor of order unity that depends on the model adopted for the disruption of the disk by the stellar magnetic field \\citep{GhoshLamb}. For $B_* < B_{*,\\text{crit}}$, the accretion disk will extend all the way to the surface of the star, for which there is good observational evidence in neutron star low mass X-ray binaries \\citep{GilfanovRevnivtsevMolkov} and in dwarf nova systems in outburst \\citep{WheatleyMaucheMattei}. Since the rotation rate of the star, $\\Omega_*$, is slower than the Keplerian rotation rate at the stellar surface, $\\Omega_K(R_*)$, a boundary layer (BL) will exist inside of which $d \\Omega / d R > 0$ and the rotation profile of the star attaches smoothly to that of the disk. If the accretion rate is high enough, it is also possible for material to spread meridionally to high latitudes forming a spreading layer (SL) \\citep{InogamovSunyaev, PiroBildsten}. For both a BL and a SL, the intense energy release in a localized region near the stellar surface leads to easily observable signatures such as hard spectral components, variability of emission, and so on. A considerable amount of effort has been previously devoted to understanding the structure of well-developed, steady-state BLs, which have been evolving for long enough to establish a smooth rotation profile in their interior. An outstanding question in the study of steady-state BLs is identifying the mechanism of angular momentum transport in the layer. Several potential mechanisms have been explored over the years, among them shear instabilities \\citep{KippenhahnThomas}, baroclinic instabilities \\citep{Fujimoto, Hanawa}, and the Tayler-Spruit dynamo \\citep{Spruit, PiroBildsten1}. Despite this effort, no clear answer exists at present regarding the nature of the angular momentum transport and mixing in well-developed BLs. \\subsection{Initiation of Boundary Layers} An interesting aspect of the BL problem that has received less attention in the past is the issue of BL {\\it initiation}, i.e. the initial stage of BL formation which must occur when the accreted material first touches the surface of the star. In this case the physical setup is going to be quite different from the steady-state BL primarily because of the much larger velocity shear, justifying the study of BL initiation in the limit of an essentially discontinuous rotation profile at the interface between the star and accreting material. Even though the initiation stage is just a transient phase in the BL evolution, one still needs to understand it to get a full picture of the BL phenomenon. Also, one need not think that the BL initiation is a unique event for every accreting object --- it may, in fact, be repetitive. The most common situation in which this BL initiation is recurrent involves the dramatic increase of the mass accretion rate (due to some sort of outburst triggered by an instability) through the disk which is magnetically disrupted outside the star under normal circumstances. According to Equation (\\ref{Bdisrupt}), a sudden increase of $\\dot M$ by several orders of magnitude as typical for some accreting systems (e.g. dwarf novae, FU Ori outbursts, etc.) can rapidly compress the stellar magnetosphere to the point at which disk material starts touching the stellar surface and a BL starts to form. There is good observational evidence for this kind of recurrent behavior. For instance, \\citet{LivioPringle} have argued that the observed lag in the rise time of the UV emission relative to the optical in a CV system transitioning to outburst can be explained if the disk is magnetically disrupted in quiescence but not in outburst. Thus, the hottest, innermost part of the disk is evacuated in quiescence, and UV radiation is not emitted immediately in the transition to outburst, since the empty region requires time to become filled. FU Ori stars are another type of system in which the disk can extend all the way to the stellar surface, due to the high accretion rate - $\\dot{M} \\sim 10^{-4} M_\\sun \\text{yr}^{-1}$ \\citep{KHH}. In these systems, the BL can puff up to become of order the stellar radius and the distinction between the boundary layer, the disk, and the star becomes blurred \\citep{PNFuOri}. The goal of our present work is to make a first step towards understanding the {\\it formation} of the BL. As in the case of a well-developed BL, the major issue for this initial stage lies in details of the angular momentum transport and mixing. However, because of the enormous shear present at the star-disk interface at this stage, it is highly likely that purely hydrodynamical shear instabilities would dominate the transport of mass and momentum rather than anything else. For that reason, in this work we primarily focus on exploring the properties of various shear-driven instabilities under the conditions typical at the BL formation stage. {First, we seek to identify a particular variety of the shear instability that most efficiently initiates mixing between the two fluids, i.e. has the fastest growth rate. Second, we examine the conditions needed for its operation, such as the density contrast between fluids, initial velocity profile, etc.} There are several physical ingredients which can potentially affect operation of shear instabilities: stratification, rotation, magnetic fields, turbulence, radiative transfer, and the supersonic nature of the flow. The latter aspect of the problem is very important and is inevitable during the BL initiation phase, when disk material rotating at the Keplerian speed comes into contact with the more slowly spinning stellar surface. The differential azimuthal velocity between the two interacting flows is then bound to be a significant fraction of the Keplerian velocity at $R_*$ and should highly exceed the sound speed both in the disk and in the outer layers of the star. \\subsection{Shear Instabilities in Compressible Fluids} The study of shear instabilities in compressible fluids is of fundamental physical significance and has received much attention in the past. \\citet{Landau}, \\citet{Hatanaka}, \\citet{Pai}, \\citet{MilesKH}, and \\citet{Gerwin} have all studied the problem in the vortex sheet approximation, where one half plane of compressible fluid moves at constant velocity over another. When the two fluids move at a low Mach number relative to each other, one finds that infinitesimal perturbations are governed by the classical Kelvin-Helmholtz (KH) dispersion relation for two incompressible fluids. However, \\citet{MilesKH} showed that above a critical Mach number, the vortex sheet becomes marginally stable to two-dimensional disturbances along the direction of the flow. This is a surprising result considering that one might have expected increasing the shear to lead to an increased growth rate for the instability rather than stabilization. Understanding the implications of this result for the momentum and mass transport in the astrophysical BLs is one of the goals of our study. Later, \\citet{Blumenetal}, \\citet{Ray}, \\citet{ChoudhuryLovelace} and \\citet{Glatzel} studied the stability of a fluid with a continuous and monotonically varying velocity profile. They found that unlike the vortex sheet, continuous velocity profiles with a supersonic velocity difference across them were unstable even at high Mach number. \\citet{Glatzel} showed that the instability was similar to the Papaloizou-Pringle instability which operates in hydrodynamical disks with radial boundaries \\citep{PP}. Thus, for high Mach number flow in a compressible fluid, a finite thickness velocity profile exhibits fundamentally different behavior from the vortex sheet, since it can support unstable modes. Moreover, the growth rate of the unstable modes scales as $\\propto 1/L$, where $L$ is the thickness of the shear layer. Thus, the thinner the shear layer, the faster the instability operates! This is exactly the opposite of what one might have expected from the vortex sheet stability criterion and we will provide an explanation for this in \\S \\ref{finite_width_SL}. Previous studies of possible instabilities inside astrophysical BLs have been primarily concerned with the sub-sonic regime when the flow can be considered as almost incompressible \\citep{KippenhahnThomas, Fujimoto1}. This regime may indeed apply in well-developed BLs with smooth shear profiles, even though at the present level of knowledge one can hardly exclude the possibility of the existence of localized regions in steady-state BLs where compressibility is still important. Our primary goal here is to extend these studies into the regime of highly compressible, supersonic flows and to explore the implications for astrophysical objects. In the course of studying supersonic shear instabilities, we will sometimes also incorporate stratification in our calculations and explore the role of rotation. The paper is organized as follows. In \\S \\ref{goveqn} we present our governing set of equations and describe the formalism we use to study the stability problem. Then in \\S \\ref{compress_KH} we study the case of the compressible vortex sheet with no gravity and show that our formalism reproduces the dispersion relation obtained by previous authors \\citep{Landau, Hatanaka, Pai, MilesKH, Gerwin}. In \\S \\ref{compress_grav}, we introduce gravity as a small parameter and perform a perturbation expansion to study what influence this has on the stability of the vortex sheet. Finally, in \\S \\ref{finite_width_SL} we study the stability of a finite width shear layer with a linear velocity profile. We derive new dispersion relations for this case and test the growth rate of the fastest growing mode using the Godunov code Athena \\citep{Stoneetal}. ", "conclusions": "\\label{discon} We have studied supersonic shear instabilities that could drive the turbulence in the BLs of stars for which the disk is undisrupted by a magnetic field. Our study is aimed mainly at identifying the instabilities that lead to the formation of the BL when the disk just touches the surface of the star. The main result of our work is the identification of two types of instabilities that could operate in the BLs of such systems and had not been previously discussed in this context. The first is an instability of a vortex sheet at high Mach number caused by gravity. Although the vortex sheet is stable above a critical Mach number, the addition of a small amount of gravity destabilizes it. We have found that the eigenfrequencies of the dispersion relation in the limit $G \\rightarrow 0$ acquire a purely imaginary term, which is proportional to the small parameter $G$. This has the effect of making the upper branch unstable. We now consider whether the instability of the upper branch in the limit $G \\rightarrow 0$ is likely to be relevant during the initiation of the BL, when one may expect the vortex sheet approximation to be valid. Redimensionalizing Equation (\\ref{uvp1}), we obtain \\ba \\label{gfreq} \\omega_1 \\sim \\Omega_K \\eps^{1/2} i, \\ea where $\\Omega_K$ is the Keplerian velocity at the surface of the star. During the initiation of the BL, we expect disk material to be less dense than stellar material, which means that $\\eps \\lesssim 1$. It then follows from Equation (\\ref{gfreq}) that the characteristic growth rate of the instability is $\\lesssim \\Omega_K$ and is independent of the wavenumber. If the BL is radially thin, as would be expected during the initiation phase, then the growth rate given by Equation (\\ref{gfreq}) is small compared to the shear rate $S \\sim \\Omega_K R_*/\\delta_{BL} \\gg \\Omega_K$. Considering that $S$ is the characteristic growth rate for shear instabilities, if there are other mechanisms for instability with growth rates proportional to $S$, they will quickly become dominant. In particular, we have demonstrated that the growth rate of the sonic instability of a finite width shear layer at high Mach number is proportional to $S$. The sonic instability is similar in nature to the Papaloizou-Pringle instability in that both are global instabilities and cannot be derived from a local analysis. There are two destabilizing mechanisms for the sonic instability. The first corresponds to emission of radiation and gives instability over a broad range of wavenumbers. The second corresponds to over-reflection of trapped modes and results in sharply peaked, disconnected regions of instability in $k$-space. Because sonic instabilities operate on a much faster timescale than the gravity mechanism for a vortex sheet, this makes them an appealing candidate for the initial stages of boundary layer formation, when one might expect large shears to be present. We mention that \\citet{AlexakisMiles} have considered the Miles instability \\citep{Mileswave} in the context of boundary layer formation. Specifically, they invoked the Miles instability to generate mixing of WD and stellar material and explain the enrichment in heavy elements observed in nova explosions. The Miles instability was proposed as a mechanism for generating waves over water at low wind speeds and operates through a resonant interaction between the wind and the water wave. Due to this interaction, a component of the pressure perturbation is created that is in phase with the velocity of the air-water interface, and much like pumping a swing, swings up the interface to large amplitudes. However, the Miles instability was initially introduced as a way of explaining the formation of waves on water at weak wind speeds, for which the air-water interface is stable to the KH instability. During boundary layer formation, however, large shears are generated, so we are not in the weak wind regime, and the Miles instability is likely to be swamped by sonic instabilities. There are two main astrophysical implications of our findings. First, we demonstrate that the initiation of the BL is likely to take only very short amount time after the first material from the disk arrives to the stellar surface. This is because the sonic instabilities that we explored in this paper have an extremely short growth rate, which is very weakly dependent on the density contrast between the disk and stellar material.\\footnote{This statement is not true for the supersonic KH instability with gravity investigated in \\S \\ref{compress_grav}, see Equation (\\ref{uvp1}).} Thus, mixing of the two fluids starts almost immediately after they come into contact. Second, given the efficiency with which the sonic instability operates, it is likely that it may play important role also in more developed phases of the BL evolution. As long as the BL possesses some effective \"boundaries\" (e.g. sharp changes in the velocity of density behavior) and the gas flow within it is supersonic the purely hydrodynamic sonic instabilities are going to operate in it potentially providing means for continuing mixing and angular momentum transport inside the layer. Future numerical calculations capable of following the nonlinear development of sonic instabilities should be able to address this issue." }, "1112/1112.4514_arXiv.txt": { "abstract": "We present the discovery of the \\starname\\ planetary system, which we initially identified through the detection of five distinct periodic transit signals in the \\kepler\\ light curve of the host star \\tmid. From high-resolution spectroscopy of the star, we find a stellar effective temperature $\\teff=5455\\pm100$~K, a metallicity of \\feh$=0.01\\pm0.04$, and a surface gravity of $\\logg=4.4\\pm0.1$. We combine these estimates with an estimate of the stellar density derived from the transit light curves to deduce a stellar mass of $\\mstar=0.912\\pm0.034~\\msun$ and a stellar radius of $\\rstar=0.944^{+0.060}_{-0.095}~\\rsun$. For three of the transit signals, we demonstrate that our results strongly disfavor the possibility that these result from astrophysical false positives. We accomplish this by first identifying the subset of stellar blends that reproduce the precise shape of the light curve and then using the constraints on the presence of additional stars from high-angular resolution imaging, photometric colors, and the absence of a secondary component in our spectroscopic observations. We conclude that the planetary scenario is more likely than that of an astrophysical false positive by a factor of $2\\times10^5$ (\\planetb), $1\\times10^5$ (\\planetc), and $1.1\\times10^3$ (\\planetd), sufficient to validate these objects as planetary companions. For \\planetc\\ and \\planetd, the blend scenario is independently disfavored by the achromaticity of the transit: From \\spitzer\\ data gathered at 4.5~\\micron, we infer a ratio of the planetary to stellar radii of $0.075\\pm0.015$ (\\planetc) and $0.065\\pm0.011$ (\\planetd), consistent with each of the depths measured in the \\kepler\\ optical bandpass. We determine the orbital periods and physical radii of the three confirmed planets to be $3.70$~d and $1.91^{+0.12}_{-0.21}~R_{\\Earth}$ for \\planetb, $10.85$~d and $3.07^{+0.20}_{-0.31}~R_{\\Earth}$ for \\planetc, and $77.61$~d and $2.75^{+0.17}_{-0.30}~R_{\\Earth}$ for \\planetd. From multi-epoch radial velocities, we determine the masses of \\planetb\\ and \\planetc\\ to be $8.7\\pm2.2~M_{\\Earth}$ and $16.1\\pm3.5~M_{\\Earth}$, respectively, and we place an upper limit on the mass of \\planetd\\ of $20.1~M_{\\Earth}~(2~\\sigma)$. ", "introduction": "Systems with multiple exoplanets, and transiting exoplanets, each bolster confidence in the reality of the planetary interpretation of the signals and offer distinct constraints on models of planet formation. The first extrasolar planets were found around a pulsar \\citep{wols92}, and it was the multi-planetary nature --- in particular the gravitational perturbations between the planets \\citep{rasio92, wols94} --- which solidified this outlandish claim. Around Sun-like stars as well, the origin of radial velocity signals continued to be questioned by some, at the time multiple planets were found around ups Andromeda \\citep{butler99}. The orbital configuration of planets relative to each other has shed light on a host of physical processes, from smooth radial migration into resonant orbits \\citep{lee02} to chaotic scattering into secular eccentricity cycles \\citep{malh02,ford05}. Now with ever-growing statistics of ever-smaller Doppler-detected planets in multiple systems \\citep{mayor11}, the formation and early history of planetary systems continues to come into sharper focus. Concurrently, transiting exoplanets have paid burgeoning dividends, starting with the definitive proof that Doppler signals were truly due to gas-giant planets orbiting in close-in orbits \\citep{charbonneau00, henry00}. Transit lightcurves offer precise geometrical constraints on the orbit of the planet \\citep{winn10}, such that radial velocity and photometric measurements yield the density of the planet and hence point to its composition \\citep{adams08, miller11}. Transiting configurations also enable follow-up measurements \\citep{charbonneau02, knutson07, triaud10} which inform on the mechanisms of planetary formation, evolution, and even weather. These two research streams, multiplanets and transiting planets, came together for the first time with the discovery of Kepler-9 \\citep{holman10, torres11}. This discovery was enabled by data from the \\kepler\\ Mission \\citep{borucki10, koch10a}, which is uniquely suited for such detections as it offers near-continuous high-precision photometric monitoring of target stars. Based on the first 4 months of \\kepler\\ data, \\citet{borucki11} announced the detection of 170 stars each with 2 or more candidate transiting planets; \\citet{steffen10} discussed in detail 5 systems each possessing multiple candidate transiting planets. A comparative analysis of the population of candidates with multiple planets and single planets was published by \\citet{latham11}, and \\citet{lissauer11a} discussed the architecture and dynamics of the ensemble of candidate multi-planet systems. The path to confirming the planetary nature of such \\kepler\\ candidates is arduous. At present, three stars (in addition to Kepler-9) hosting multiple transiting candidates have been presented in detail and the planetary nature of each of the candidates has been established: These systems are Kepler-10 \\citep{batalha11, fressin11a}, Kepler-11 \\citep{lissauer11b}, and Kepler-18 \\citep{cochran11}. Transiting planets are most profitable when their masses can be determined directly from observation, either through radial velocity (RV) monitoring of the host star or by transit timing variations (TTVs), as was done for Kepler-9bc, Kepler-10b, Kepler-11bcdef, and Kepler-18bcd. When neither the RV or TTV signals is detected, statistical arguments can be employed to show that the planetary hypothesis is far more likely than alternate scenarios (namely blends of several stars containing an eclipsing component), and this was the means by which Kepler-9d, Kepler-10c, and Kepler-11g were all validated. While such work proves the existence of a planet and determines its radius, the mass and hence composition remain unknown save for speculation from theoretical considerations. This paper presents the discovery of a new system, \\starname, with five candidate transiting planets. We validate three of these by statistical argument; we then proceed to use RV measurements to determine the masses of two of these, and we place an upper limit on the mass of the third. We do not validate in this paper the remaining two signals (and hence remain only candidates, albeit very interesting ones, owing to their diminutive sizes), rather the validation of these two remaining signals is addressed in a separate effort \\citep{fressin12}. The paper is structured as follows: In \\S2, we present our extraction of the \\kepler\\ light curve (\\S2.1), our modeling of these data (and RVs) to estimate the orbital and physical parameters of the planets and star (\\S2.2), as well as limits on the motion of the photocentroid during transit (\\S2.3) and a study of the long-term astrophysical variability of the star from the \\kepler\\ light curve (\\S2.4). In \\S3 we present follow-up observations that we use to argue for the planetary interpretation, including high-resolution imaging (\\S3.1) and \\spitzer\\ photometry (\\S3.2), and the spectroscopy we use to characterize the star and determine the radial velocity signal (\\S3.3). In \\S4, we present our statistical analysis that validates the planetary nature of the three largest candidate planets in the system. In \\S5 we consider the dynamics of the system, and in \\S6 we discuss the constraints on the composition and formation history of the three planets. \\subsection{Nomenclature} Throughout the course of the \\kepler\\ Mission, a given star is known by many different names \\citep[see][for an explanation of \\kepler\\ naming conventions]{borucki11}, and we pause here to explain the relationship of these names in the current context. The star that is the subject of this paper is located at $\\alpha=$\\kicra, $\\delta=$\\kicdec\\ (J2000). It was already known as \\tmid, and in the \\kepler\\ input catalog it was designated \\kicid. After the identification of candidate transiting planets it became a \\kepler\\ Object of Interest (KOI) and was further dubbed \\koi, and it appeared as such in the list of candidates published by \\citet{borucki11}. Some authors have elected to denote KOIs using a different nomenclature, in which case \\koi\\ would be identified as KOI-70. After the confirmation of the planetary nature of three of these candidates it was given its final moniker \\starname. This paper describes that process of confirmation, but for simplicity we refer to the star as \\starname\\ throughout. The three confirmed exoplanets were initially assigned KOI designations representing the chronological order in which the transiting signals were identified, but to avoid confusion we will refer to them henceforth by their \\starname\\ designations in which they are ordered by increasing orbital period $P$; \\planetb\\ (\\koib, $P=3.70$~d), \\planetc\\ (\\koic, $P=10.85$~d), and \\planetd\\ (\\koid, $P=77.61$~d). We will refer to the two remaining candidates as \\koie\\ and \\koif, but note (as described below) that the period of \\koie\\ ($P=6.10$~d) is intermediate between those of \\planetb\\ and \\planetc, and the period of \\koif\\ ($P=19.58$~d) is intermediate between those of \\planetc\\ and \\planetd. ", "conclusions": "" }, "1112/1112.3875_arXiv.txt": { "abstract": "Thanks to {\\tt SAURON\\/} integral-field observations we uncovered the Planetary Nebulae (PNe) populations inhabiting the central and nuclear regions of our galactic neighbours M32 and M31, respectively, and discuss the significant differences between their corresponding PNe luminosity functions in light of the properties of their parent stellar populations. In particular, we conclude that the lack of bright PNe in the nuclear regions of M31 is likely linked to the nearly Solar value for the stellar metallicity, consistent with previous suggestions that a larger metallicity would bias the Horizontal-Branch (HB) populations toward bluer colors, with fewer red HB stars capable of producing PNe and more blue HB stars that instead could contribute to the far-UV flux that is observed in metal-rich early-type galaxies and, incidentally, also in the nucleus of M31. ", "introduction": "Planetary Nebulae (PNe) in external galaxies are mostly regarded either as tracers of the gravitational potential (e.g., Romanowsky et al. 2003) or as indicators for the distance of their galactic hosts (e.g., Ciardullo et al. 1989), with the latter advantage owing to the nearly universal -- though not fully understood -- shape of the PNe luminosity function (PNLF, generally in the \\Oiii$\\lambda5007$ emission). Yet extra-galactic PNe can also be used as probes of their parent stellar population (see, e.g., Ciardullo 2006) and understanding in particular the origin of the PNLF is a puzzle that, once solved, promises to reveal new clues on the late stages of stellar evolution and on the formation of PNe themselves (e.g., Ciardullo et al. 2005; Buzzoni, Arnaboldi \\& Corradi 2006). PNe originates from horizontal-branch (HB) stars that climb back the asymptotic giant branch (AGB) at the end of their helium-burning phase, when these stars leave the AGB and quickly cross the Hertzprung-Russell diagram on their way towards the cooling track of white dwarves (WD). For a population with a given age and metallicity, HB stars have nearly the same helium core mass ($\\sim 0.5\\,M_{\\odot}$) but a range of hydrogen shell mass ($\\sim 0.001 - 0.3 \\,M_{\\odot}$), with the reddest stars having also the largest H-shells and originating from the most massive main-sequence progenitors. Only HB stars with a considerable H-shell ascend toward the AGB and eventually lead to the formation of a PN, whereas the bluest HB stars with little envelope mass head straight toward the WD cooling curve by evolving first to higher luminosities and effective temperatures (the so-called AGB-manqu\\'e phase). According to this simple picture, galaxies with on-going star formation should show brighter PNe than quiescent systems where massive stars have long disappeared (e.g., Marigo et al. 2004), but in fact the PNLF of young and old galaxies are relatively similar. In particular, all extra-galactic PNe surveys indicate a common and bright cut-off for the PNLF, which led Ciardullo et al. (2005) to suggest a binary evolution for the progenitors of the brightest PNe that would be common to different kind of galaxies. If galaxies seem to invariably host very bright PNe, their specific content of PNe - that is the number of PNe normalised by a galaxy bolometric luminosity - appears to vary with the metallicity of the stellar population. More specifically, Buzzoni, Arnaboldi \\& Corradi (2006) found that more metal rich galaxies show comparably fewer PNe, which also corresponds to larger far-UV fluxes. Interestingly, this may indicate that at a given mean stellar age, a larger metallicity would bias the HB population towards fewer stars with massive H-shell capable to lead to the formation of PNe, with a larger fraction of blue HB stars that contribute instead to the overall far-UV flux of their host galaxy (i.e. the so-called UV-upturn, Burnstein et al. 1988) as they follow their AGB-manqu\\'e evolution. Within this context, we note that whereas our knowledge of both the shape and normalisation of the PNLF comes chiefly from the peripheral PNe populations of galaxies, both measurements for the stellar metallity and the UV spectral shape of galaxies pertain to their optical regions. This is because narrow-band imaging or slitless spectroscopy - the most common techniques employed to find extragalactic PNe - find it hard to detect PNe against a strong stellar background, whereas measuring the strength of stellar absorption lines or imaging the far-UV flux of galactic halos is prohibitively expensive in terms of telescope time. Such a dramatic spatial inconsistency needs to be resolved if we ought to understand the link between PNe and the properties of their parent stellar populations, in particular if we consider that such a connection may already not be entirely within our grasp, as Hubble Space Telescope (HST) observations for the UV color-magnitude diagram of M32 suggests (Brown et al. 2008). ", "conclusions": "" }, "1112/1112.4387_arXiv.txt": { "abstract": "We explore a scenario in the Next-to-Minimal-Supersymmetric-Standard-Model (NMSSM) with both a light ${\\cal O}(10)$ GeV neutralino and a CP-odd Higgs boson with significant coupling to down-type fermions, evading all current B physics, LEP and WMAP bounds. Motivated by a possible slight lepton universality breaking hinted in $\\Upsilon$ decays, we consider the effect of the mixing of $\\eta_b$ resonances with the pseudoscalar Higgs on the spin-dependent scattering neutralino cross section off nucleons. We conclude that this mechanism could be relevant provided that non-perturbative effects enhance the effective $\\eta_b$-nucleon coupling, taking over velocity/$q^2$ suppression factors, perhaps giving a new insight into the current controversial situation concerning direct search experiments of dark matter. ", "introduction": "Evidence has been accumulated both from astrophysics and cosmology that about 1/4 of the energy budget of the present universe consists of the so-called (cold) dark matter (DM), namely, a component which is non-relativistic and neither feels the electromagnetic nor the the strong interaction. It is fair to say that the most popular DM candidate for a WIMP (Weakly Interacting Massive Particle) is the lightest supersymmetric particle (LSP) in supersymmetric models with $R$-parity conservation. Leaving aside the axion and the axino, the superpartners with the right properties for playing the role of a WIMP in the universe are the gravitino and the lightest neutralino ($\\chi$) - by far, the most discussed case in the literature. Although the LHC is running smoothly and collecting large amounts of data useful to look for physics beyond the Standard Model (SM), other complementary facilities are certainly needed, especially concerning DM detection. In fact, DAMA/LIBRA, CoGeNT, and more recently CRESST experiments have reported the observation of events in excess of the expected background, hinting at the existence of a light WIMP~\\cite{Bernabei:2010mq,Aalseth:2011wp,Angloher:2011uu}. However, exclusion limits set by other direct searches, such as Xenon10~\\cite{arXiv:1104.3088} and Xenon100~\\cite{arXiv:1104.2549}, are in tension with the above claims. In the NMSSM, a light neutralino (as a DM candidate) can efficiently annihilate through the resonant s-channel via a light pseudoscalar Higgs mediator satisfying the requirements from the relic density~\\cite{Gunion:2005rw}. However, following a scan of the NMSSM parameter space, the authors of \\cite{Das:2010ww} obtained upper limits on the spin-independent (SI) $\\chi$-nucleon cross section which are substantially below the requirements of DAMA and CoGeNT. The spin-dependent (SD) cross section (via $Z$-exchange) was also found several orders of magnitude below current experimental bounds. On the other hand, the authors of \\cite{Gunion:2010dy} were able to achieve a somewhat larger SI cross section in a similar scenario. Admittedly, such cross sections can be further enhanced by increasing the $s$-quark content of the nucleon, but the agreement with the low range of DAMA results turns out to be only marginally acceptable. In this work we revisit SD $\\chi$-nucleon scattering via pseudocalar-exchange in the NMSSM, usually neglected in most analyses \\cite{Das:2010ww}, which however might be enhanced due to a non-perturbative mechanism as later argued. \\begin{table*}[hbt] \\caption{Phase-space corrected leptonic branching fractions, $\\hat{\\cal B}\\left(\\Upsilon(nS) \\to \\ell \\ell\\right)$ (in \\%), and error bars (summed in quadrature) of $\\Upsilon(1S)$, $\\Upsilon(2S)$, and $\\Upsilon(3S)$ resonances \\cite{pdg}. Error bars of the ratios $\\hat{R}_{\\tau/\\ell}(nS)$ are likely overestimated because of expected correlations between the numerator and denominator experimental uncertainties.} \\label{FACTORES} \\begin{center} \\begin{tabular}{|c|c|c|c|c|c|} \\hline & $\\hat{\\cal B}\\left(e^+e^-\\right)$ & $\\hat{\\cal B}\\left(\\mu^+\\mu^-\\right)$ & $\\hat{\\cal B}\\left(\\tau^+\\tau^-\\right)$ & $\\hat{R}_{\\tau/e}(nS)$ & $\\hat{R}_{\\tau/\\mu}(nS)$\\\\ \\hline $\\Upsilon(1S)$ & $2.48 \\pm 0.07$ & $2.48 \\pm 0.05$ & $2.62 \\pm 0.10$ & ${\\bf 0.057 \\pm 0.050}$ & ${\\bf 0.057 \\pm 0.046}$ \\\\ \\hline $\\Upsilon(2S)$ & $1.91 \\pm 0.16$ & $1.93 \\pm 0.17$ & $2.01 \\pm 0.21$ & ${\\bf 0.052 \\pm 0.141}$ & ${\\bf 0.041 \\pm 0.141}$ \\\\ \\hline $\\Upsilon(3S)$ & $2.18 \\pm 0.21$ & $2.18 \\pm 0.21$ & $2.30 \\pm 0.30$ & ${\\bf 0.056 \\pm 0.171}$ & ${\\bf 0.056 \\pm 0.171}$ \\\\ \\hline \\hline $\\psi(2S)$ & $0.773 \\pm 0.017$ & $0.77 \\pm 0.08$ & $0.772 \\pm 0.100$ & ${\\bf -0.001 \\pm 0.100}$ & ${\\bf 0.002 \\pm 0.100}$ \\\\ \\hline \\end{tabular} \\end{center} \\end{table*} ", "conclusions": "In this paper, we have considered a particular scenario within the NMSSM with both a light neutralino and a light CP-odd Higgs boson, the latter sizably mixing with pseudoscalar $\\eta_b$ resonances. Implicit in our work is the idea that non-perturbative effects (e.g. instanton-induced interaction) may lead to a non-negligible pseudoscalar contribution to the $\\chi$-nucleus scattering, thereby introducing a momentum-dependent form factor in the cross section which might be helpful (see e.g. Ref.~\\cite{Feldstein:2009tr}) to interpret the results of direct DM search experiments, with variable sensitivity along the nuclear recoil energy range. To conclude we stress that an accurate experimental test of lepton universality in $\\Upsilon$ decays, the discovery of the $\\eta_b(2S)$ resonance together with the measurements of ${\\cal B}[\\eta_b(nS) \\to p\\bar{p}]$ at a (Super) B factory \\cite{Bona:2007qt} could be relevant for a better understanding of DM searches and related astrophysical questions. \\subsection*" }, "1112/1112.0498_arXiv.txt": { "abstract": "{We describe the growth of gas giant planets in the core accretion scenario.} { The core growth is not modeled as a gradual accretion of planetesimals but as episodic impacts of large mass ratios, i.e.\\ we study impacts of 0.02 - 1 \\mearth onto cores of 1-15 \\mearth. Such impacts could deliver the majority of solid matter in the giant impact regime. We focus on the thermal response of the envelope to the energy delivery. Previous studies have shown that sudden shut off of core accretion can dramatically speed up gas accretion. We therefore expect that giant impacts followed by periods of very low core accretion will result in a net increase in gas accretion rate. This study aims at modelling such a sequence of events and to understand the reaction of the envelope to giant impacts in more detail. } {To model this scenario, we spread the impact energy deposition over a time that is long compared to the sound crossing time, but very short compared to the Kelvin-Helmholtz time. The simulations are done in spherical symmetry and assume quasi-hydrostatic equilibrium.} {Results confirm what could be inferred from previous studies: gas can be accreted faster onto the core for the same net core growth speed while at the same time rapid gas accretion can occur for smaller cores -- significantly smaller than the usual critical core mass. Furthermore our simulations show, that significant mass fractions of the envelope can be ejected by such an impact. } { Large impacts are an efficient process to remove the accretion energy by envelope ejection. In the time between impacts, very fast gas accretion can take place. This process could significantly shorten the formation time of gas giant planets. As an important side-effect, the episodic ejection of the envelope will reset the envelope composition to nebula conditions. } ", "introduction": "\\label{sec:introduction} We study the formation of gas giant planets in the core accretion scenario \\citep{mizuno1980,1996Icar..124...62P,1986Icar...67..391B}. In this scenario, a planetary embryo grows by accreting from a swarm of planetesimals. At some point the embryo becomes massive enough to gravitationally attract a gaseous envelope. The growth process, both in terms of solid and gas accretion, is controlled by the planetesimal accretion, which is typically modeled as a gradual accretion of small planetesimals. However, in the giant collision phase under certain conditions, the accretion process could be dominated by relatively large impacts \\citep{1969Icar...10..109S}. This is confirmed by Monte-Carlo planet formation models (T. Schr\\\"oter et al., in preparation) and recent results from N-body simulations \\citep{Raymond:2005p21625,Nimmo:2006p21624}. In such cases, while the collisions are less frequent, each one increases the mass of the protoplanet by a significant amount (typically of order 10\\%). This possibility led us to investigate the importance of the nature of the solid accretion process in the overall growth of giant planets. In particular, we want to investigate if episodic but large impacts result in changes in the mass and structure of the envelope when compared to gradual core growth? A body of work investigates the importance of impacts and/or core luminosity on the evolution of the envelope of giant planets. One such study concerned itself with the possibility of stripping the envelope of Uranus by an impact induced shockwave \\citep{1990Icar...84..528K}. However, the authors did not follow the long-term evolution of the post impact planet and did not consider the possibility of subsequent re-accretion of gas. \\citet{2006ApJ...650.1150I} study the collision of two giant planets to explain the low envelope mass of HD~149026b. Another study \\citep{2007A&A...466..717A} tried to assess the effect of a large impact on the long-term luminosity evolution of a giant planet with an eye on its potential detection. Other studies investigated the influence of the thermal energy content of the solid core as well as the energy provided by its contraction on the overall evolution of the luminosity \\citep{2008A&A...482..315B}. Further studies \\citep{2007MNRAS.381..640P} concern the dynamic response of proto-planetary envelopes to a perturbation within an ideal gas approach. Recently, \\citet{2010ApJ...720.1161L} have studied the merger of planetary embryos focussing on the re-distribution of heavy elements following the merger. Most closely related to the problem at hand is the sudden core luminosity shut-off scenario: the evolution of the planet when the core luminosity is suddenly shut off. This has been studied in detail by \\citet{2000ApJ...537.1013I} and \\citet{Hubickyj2005415}. It is thus expected, that the periods of low core accretion in-between impacts will lead to massive gas accretion. The effect of sporadic, relatively massive impacts during the growth phase of the core on the gas accretion has, however, not been studied in detail. Here we attempt to determine the thermal response of a gaseous envelope upon a sudden energy input delivered by a large impact to the core, and how such episodic events could modify the build-up of the envelope when compared to the nominal case of gradual accretion. ", "conclusions": "% \\label{sec:conclusion} We have analyzed the effect of relatively massive impacts onto the cores of giant planets in the growth phase. Previous studies \\citep{2000ApJ...537.1013I,Hubickyj2005415} have shown how the envelope reacts to a shut-off of core luminosity with rapid envelope accretion. Therefore, intermittent giant impacts with long periods of very low core accretion in-between are expected speed up gas accretion. The aim of this study was to study the reaction of the envelope to the giant impact and the subsequent period of low core luminosity in detail. Wether or not a net acceleration of gas accretion takes place and how strong this effect is, depends on the timescale of the reaction to the impact in relation to the time in-between impacts. There are two major effects: 1) Due to a very large core luminosity after the impact, a large fraction of the envelope is ejected. The remaining tiny envelope allows a very fast energy transport and in consequence the huge impact energy is used up or radiated away very fast. The huge luminosity in combination with a small envelope leads to a small Kelvin-Helmholtz time. Therefore the envelope 'forgets' that the impact has taken place in very short time. Afterwards, very little solid accretion is necessary for a given net core growth speed and the subsequent evolution up to the next impact can be understood by the well-studied sudden shut-off of core luminosity. However, due to the episodic impacts, further core growth is still possible. 2) This effect is enhanced by a second: The very large luminosity during the impact reconfigures the envelope structure so that it is in hydrostatic equilibrium for very high energy fluxes. Shutting off core accretion after this reconfiguration triggers higher gas accretion rates than the shut-off without prior impact. This can also be understood in terms of the negative gravothermal specific heat of self-gravitating non-degenerate gases. In this state the impacts actually lower the central temperature which explains why subsequent gas accretion can be faster after the impact. This second effect helps reduce the time it takes to accrete once more the envelope gas that has been ejected by the impact. Once the planet has again the same envelope mass as before the impact but a much lower core luminosity, the evolution follows the shutdown scenario. Together, the alternation between very high and low energy input allows more gas to be accreted even though the impact initially ejects some (or all) of the envelope gas. The subsequent high gas accretion rate quickly reforms the original envelope and continues to accrete faster than the gradually growing case. In this way, even a rapid succession of impacts can lead to a faster envelope growth. A further interesting consequence of the envelope loss caused by the impact relates to the dust opacity: every time the envelope is ejected, new fresh gas is accreted from the nebula, therefore resetting any former modifications to the opacity caused by e.g. dust growth and settling. In summary, we find that episodic large impacts significantly speed up gas accretion as was expected from shut-off calculations. The new result of this study is the fact that 1) almost the entire envelope of the planet is ejected as a consequence of the impact energy and 2) it takes only a very short time to accrete the lost gas after the ejection. In fact, this is so fast that the ejection has practically no effect on the long-term evolution - envelope masses are equivalent with the shutdown-case using the increased post-impact core mass. We can therefore conclude the following for the formation of giant planets: If planetesimals are accreted in a regime where mass ratios are large, i.e. most mass is delivered by massive impacts, this will accelerate envelope build-up. Furthermore, the ratio of envelope to core mass will be significantly enhanced and smaller cores can begin rapid gas accretion while the core is growing by large impacts. The speedup with consecutive impacts, can be understood in principle from these calculations. We will study a full evolution calculation based on episodic core growth in a future publication." }, "1112/1112.1065.txt": { "abstract": "{Interacting galaxies are well-known for their high star formation rates and rich star cluster populations, but it is also recognized that the rapidly changing tidal field can efficiently destroy clusters. We use numerical simulations of merging disc galaxies to investigate which mechanism dominates. The simulations include a model for the formation and evolution of the entire star cluster population, accounting for the evaporation of clusters due to two-body relaxation and tidal shocks. We find that the dynamical heating of stellar clusters by tidal shocks is about an order of magnitude higher in interacting galaxies than in isolated galaxies. This is driven by the increased gas density, and is sufficient to destroy star clusters at a higher rate than new clusters are formed: the total number of stellar clusters in the merger remnant is 2--50\\% of the amount in the progenitor discs, with low-mass clusters being disrupted preferentially. By adopting observationally motivated selection criteria, we find that the observed surplus of star clusters in nearby merging galaxies with respect to isolated systems is caused by the observational bias to detect young, massive clusters, and marks a transient phase in galaxy evolution. We provide a general expression for the survival fraction of clusters, which increases with the gas depletion time-scale, reflecting that both the formation and the destruction of clusters are driven by the growth of the gas density. Due to the preferential disruption of low-mass clusters, the mass distribution of the surviving star clusters in a merger remnant develops a peak at a mass of about $10^{3}~\\msun$, which evolves to higher masses at a rate of 0.3--0.4~dex per Gyr. Briefly after a merger, the peak mass depends weakly on the galactocentric radius, but this correlation disappears as the system ages due to the destruction of clusters on eccentric orbits. We discuss the similarities between the cluster populations of the simulated merger remnants and (young) globular cluster systems. Our results suggest that the combination of cluster formation and destruction should be widespread in the dense star-forming environments at high redshifts, which could provide a natural origin to present-day globular cluster systems.} ", "introduction": "\\label{sec:intro} Merging and interacting galaxies host huge starbursts and large populations of young massive stellar clusters \\citep[e.g.][]{holtzman92,schweizer96,whitmore99}. A galaxy interaction triggers inflows of interstellar gas towards the galaxy centres, where it fuels a burst of star formation \\citep{hernquist89,mihos96,barnes96}. Merger-induced starbursts play a central role in the history of the universe, as galaxies are thought to have formed through hierarchical merging \\citep[e.g.][]{white78,white91,cole00}. Some fraction of this star formation takes place in compact stellar clusters \\citep{elmegreen83,whitmore99,bressert10} with masses in the range $10^2$--$10^8~\\msun$ \\citep{portegieszwart10}. The clusters that remain after a merger are often used as fossils to trace the formation history of the galaxy \\citep{larsen01}. During the past two decades, observations with the Hubble Space Telescope have revealed that many nearby ongoing galaxy mergers host exceptionally rich star cluster populations with cluster masses exceeding $10^7~\\msun$ \\citep{schweizer82,holtzman92,miller97,schweizer98,bastian06}, which are formed due to the perturbation of the interstellar medium (ISM) \\citep{schweizer87,ashman92}. The multitude of star clusters suggests that they are useful tracers of past galaxy mergers, especially because they are easily observed up to distances of several tens of megaparsecs. The observed clusters ($>10^4~\\msun$) are distributed according to a power law with index $-2$ down to the detection limit \\citep{zhang99}. These clusters are thought to be just the `tip of the iceberg', since the initial cluster mass function (ICMF) appears to continue beyond the detection limit and down to {a certain} physical lower mass limit \\citep[see e.g.][]{portegieszwart10}. However, high gas densities and tidal shocks, both of which are prevalent in coalescing galaxies, are known to have a disruptive effect on star clusters \\citep{spitzer58,weinberg94b,gieles06}. The destruction rate of star clusters decreases with increasing cluster mass and density \\citep{spitzer87,lamers05}.\\footnote{Unless the environment in which they reside is so disruptive that it can efficiently destroy a cluster regardless of its mass. In that case, the recently argued scenario in which cluster disruption is mass-independent \\citep{whitmore07} can arise \\citep{elmegreen10b,kruijssen11}. This would then not be universal, but depends on the environmental conditions.} This indicates that the effects of star cluster disruption could be masked by observational selection effects and go unnoticed in observations, i.e. the brightest and therefore most massive clusters are easiest to detect but also least affected by disruption. If the ICMF is universal, i.e. all stellar clusters are formed according to a power law with index $-2$ throughout space and time \\citep[e.g.][]{kruijssen11d}, then the important role of cluster disruption is supported by the old (`globular') star cluster systems that are observed in nearby spiral and giant elliptical galaxies, which are strongly lacking low-mass clusters with respect to the young populations in presently merging galaxies \\citep{vesperini01,fall01,elmegreen10}. For a power law ICMF, the size-of-sample effect would also require that the most massive clusters are formed in the largest bursts of star formation, implying that globular clusters originate from starburst environments. The question thus arises whether or not the disruption of star clusters dominates over their formation in starburst galaxies. This is not easily determined on analytical grounds. Globular cluster systems are present over most of the galaxy mass range \\citep[e.g.][]{peng08}, and as such it is evident that they were not only formed in interactions between massive spiral galaxies; the presence of globular clusters in dwarf galaxies suggests that these also endured starbursts during their early evolution. While the globular clusters of dwarf galaxies are generally metal-poor, the colour distribution of globular clusters is often bimodal in massive spiral galaxies and giant ellipticals \\citep{searle78,forbes97,kundu01,peng06}. This colour bimodality may translate into a metallicity bimodality, although it has recently been suggested that it is a relic of a non-linear relation between colour and metallicity \\citep{yoon06,chies11,yoon11}. {Regardless of whether the metallicity distribution is bimodal}, a popular explanation for the broad range in metallicities \\citep{muratov10} is that the metal-poor clusters preferentially originate from accreted dwarf galaxies \\citep{prieto08}, while the metal-rich population was mainly formed in-situ, either by disc instabilities \\citep{shapiro10} or in galaxy mergers \\citep{ashman92}. The formation of globular clusters has been investigated in several theoretical and numerical studies \\citep{harris94,elmegreen97,bekki02,li04,bournaud08}. As expected from {power law statistics}, these studies all point to dense, gas-rich environments, which are typically correlated with high star formation rate densities. However, the present-day population of globular clusters is not recovered in these studies, because they only concern cluster formation and contain either no description for the further evolution of clusters or a very simplified one. {Separate studies, both analytical and numerical, have shown that} the evolution of (globular) clusters is related to the galactic environment \\citep{spitzer87,baumgardt03,lamers05a,gieles06,elmegreen10b,kruijssen11}. To obtain a more complete, quantitative understanding of the origin of present-day globular clusters, it is necessary to consider their formation and further evolution simultaneously. At present, the most commonly used method to model the evolution of star clusters is through $N$-body simulations \\citep{vesperini97b,portegieszwart98,baumgardt03,gieles08,praagman10,renaud11}. However, this method is computationally too expensive to follow the formation and evolution of the entire cluster population. \\citet{kruijssen11} therefore introduced a method in which numerical simulations of galaxies are supplemented with a semi-analytic model for the formation and evolution of star clusters, of which the results are consistent with (observed and simulated) formation and destruction rates from the literature. {This model enables us to track the formation and evolution of the entire star cluster population throughout the assembly histories of galaxies.} As a first effort to understand the (im)balance between the formation and destruction of star clusters in starburst environments, we use the method from \\citet{kruijssen11} to model the star cluster populations of galaxy mergers. This allows us to quantify the net effect of a galaxy merger on its cluster population. With this setup, we aim to investigate: \\begin{itemize} \\item[(1)] {the relative importance of} cluster formation and destruction in interacting galaxies; \\item[(2)] whether galaxy mergers can produce the progenitors of present-day metal-rich globular clusters. \\end{itemize} In Sects.~\\ref{sec:clform} and~\\ref{sec:clevo} we summarise our model, while the initial conditions of the simulations are presented in Sect.~\\ref{sec:init}. The evolution of the star cluster population in galaxy mergers is assessed in Sect.~\\ref{sec:evo}, where we also address their sensitivity to model parameters. We end this paper with a summary of our conclusions. ", "conclusions": "\\label{sec:concl} We have performed a numerical study of major mergers of comparable-mass disc galaxies, complemented with a sub-grid model for the ongoing formation and evolution of their star cluster populations. The simulations have been used to address the relative contributions of cluster formation and disruption over the course of a galaxy merger, and to investigate the potential formation of the metal-rich part of a globular cluster system. The main results from our model are as follows. \\begin{itemize} \\item[(i)] During a galaxy merger, the total number of star clusters decreases. The increase of the star formation rate during merger-induced starbursts is compensated by a stronger increase of the cluster disruption due to tidal shock heating by dense gas. \\item[(ii)] Although during certain episodes the destruction rate is high enough to disrupt clusters independently of their mass, over the entire course of a merger low-mass clusters are most strongly affected by the destruction. When considering increasingly massive clusters, their number decreases by a smaller amount during a merger. If the cluster sample is limited to massive and young clusters to mimic observational selection effects, the net destruction cannot be detected and changes to a transient increase of the number of clusters during the starbursts, in agreement with observational results. \\item[(iii)] The relative decrease of the number of clusters is stronger for higher peak star formation rates, because the enhanced formation and destruction of clusters are both caused by the high gas density. This trend is weaker for higher masses and may be reversed above $M\\sim 1$--$3\\times10^5~\\msun$, where a stronger starburst may produce {\\it more} clusters than a weak starburst. In Eq.~\\ref{eq:tdepl}, we provide a generalised expression for the survival fraction of clusters as a function of the gas depletion time-scale, which reflects the intensity of the starburst. \\item[(iv)] {The peaks in the cluster age distribution and star formation history can be offset with respect to each other due to the elevated cluster disruption rate at the height of a starburst. This offset can be as large as 200~Myr \\citep{kruijssen11}, which implies that while the cluster age distribution can be used to reveal the occurrence of a starburst, it cannot necessarily be used to determine its time or duration.} \\item[(v)] The orbital kinematics of the star clusters in a merger remnant are isotropic within galactocentric radii of $\\sim 50$--60~kpc due to the destruction of clusters on highly eccentric orbits. This value is similar to the result for the accretion of globular clusters from satellite dwarf galaxies \\citep{prieto08}, which shows that it may not be possible to distinguish between in-situ and ex-situ cluster formation based on solely the orbital (an)isotropy of the cluster population. \\item[(vi)] The preferential destruction of low-mass clusters causes the power law initial cluster mass function to develop a peak at a mass of about $10^{2.5}~\\msun$ during the final coalescence of the galaxies. This is a lower limit, as the precise value depends on the relation between cluster mass and radius, with the post-merger peak mass potentially reaching up to $10^4~\\msun$ if the cluster radii are completely unrelated to their masses. The peak mass only weakly correlates with galactocentric radius due to the destruction of clusters on radially anisotropic orbits, and (for the adopted mass-radius relation) increases by about 0.3--0.4~dex per Gyr after the completion of a merger. Young to intermediate-age ($\\sim 2$~Gyr old) merger remnants should display a peak in the star cluster mass distribution at about $10^3~\\msun$ due to the destruction of low-mass clusters (see Figs.~\\ref{fig:histgc} and~\\ref{fig:gcmfsens}). \\item[(vii)] After a merger is completed, the star cluster population is similar to what a young globular cluster system would look like. Firstly, the ejection of clusters from star-forming regions into the stellar halo produces a spatial distribution that is comparable to that of globular clusters. Secondly, the peaked cluster mass distribution is intermediate to that of young massive clusters and old globular clusters. Thirdly, the high star formation rate during a merger is capable of producing clusters that are massive enough to survive for a Hubble time. \\end{itemize} Interestingly, the high disruption rate after (globular) cluster formation could lead to a mass distribution with a peak mass of $10^3$--$10^4~\\msun$ on such a short time-scale that only little further disruption is required to obtain the current peak mass of the globular cluster mass distribution. This would imply that even the subset of clusters on the widest orbits around their host galaxies would be able to reach it before the present day. If the ICMF of globular clusters had a Schechter-type truncation at the high-mass end \\citep{kruijssen11d}, any further disruption would not yield an additional increase of the peak mass because it then saturates at about 10\\% of the truncation mass\\footnote{This percentage applies if the ICMF of globular clusters followed a power law with index $-2$ below the truncation, and the mass dependence of the disruption time-scale is $\\gamma\\sim0.7$ \\citep{gieles09,kruijssen09b}.} \\citep{gieles09}. This would thus lead to a `universal' globular cluster mass function, independent of galactocentric radius and current galactic environment. The current peak mass of globular cluster systems throughout the universe indeed happens to be $\\sim 2\\times10^5~\\msun$ \\citep[e.g.][]{jordan07}, around 10\\% lower than the estimated truncation mass of their ICMF \\citep{kruijssen09b}. The increased cluster disruption rate in galaxy mergers is driven by the high gas densities that also cause the burst of star formation. This indicates that the mechanism of enhanced disruption is not necessarily constrained to major mergers, and can be generalised to any environment with a high gas density and a correspondingly high SFR. While major mergers may provide an efficient formation channel for globular cluster populations, they are not a prerequisite. Any extremely high-density, dynamically active, star-forming environment -- be it in a starburst dwarf galaxy, during bulge assembly, in an unstable high-redshift disc or in a major merger -- would cause the enhanced disruption of clusters at young ages. The clusters that eventually survive are characterised by a more quiescent evolution due to cluster migration and natural selection \\citep{kruijssen09b,elmegreen10,elmegreen10b,kruijssen11}. Indeed, the wide variety of galaxy types with remarkably similar globular cluster mass distributions is hard to explain if cluster disruption is governed by the present-day environment, and suggests that the bulk of the disruption occurred at the epoch of globular cluster formation, when the host galaxies were likely more similar. A generalisation to all dense environments is supported by dwarf galaxies like Fornax, which has not undergone a major merger and yet harbours a handful of globular clusters \\citep{shapley39,hodge61} {that presumably formed in a starburst during the early formation of the galaxy}. If such a generalisation to all dense environments indeed holds, it would suggest that globular cluster populations may be the inevitable outcomes of the large starbursts occurring in the early universe." }, "1112/1112.4933_arXiv.txt": { "abstract": "We report on our second-year campaign of X-ray follow-up observations of unidentified \\F $\\gamma$-ray sources at high Galactic latitudes ($|b|>10^{\\circ}$) using the X-ray Imaging Spectrometer onboard the \\SU X-ray Observatory. In this second year of the project, seven new targets were selected from the First \\F Catalog, and studied with $20-40$\\,ks effective \\SU exposures. We detected an X-ray point source coincident with the position of the recently discovered millisecond pulsar PSR~J2302+4442 within the $95\\%$ confidence error circle of 1FGL~J2302.8+4443. The X-ray spectrum of the detected counterpart was well fit by a blackbody model with temperature of $kT \\simeq 0.3$\\,keV, consistent with an origin of the observed X-ray photons from the surface of a rotating magnetized neutron star. For four other targets which were also recently identified with a normal pulsar (1FGL~J0106.7+4853) and millisecond pulsars (1FGL~J1312.6+0048, J1902.0$-$5110, and J2043.2+1709), only upper limits in the $0.5-10$\\,keV band were obtained at the flux levels of $\\simeq 10^{-14}$\\,erg\\,cm$^{-2}$\\,s$^{-1}$. A weak X-ray source was found in the field of 1FGL~J1739.4+8717, but its association with the variable $\\gamma$-ray emitter could not be confirmed with the available \\SU data alone. For the remaining \\F object 1FGL~J1743.8$-$7620 no X-ray source was detected within the LAT $95 \\%$ error ellipse. We briefly discuss the general properties of the observed high Galactic-latitude \\F objects by comparing their multiwavelength properties with those of known blazars and millisecond pulsars. ", "introduction": "Since its successful launch in 2008 June, the Large Area Telescope (LAT) onboard the {\\it Fermi} Gamma-ray Space Telescope \\citep{atw09} has enabled many important breakthroughs in the understanding of the origin of high energy $\\gamma$-ray emissions of various classes of astrophysical objects. The number of detected $\\gamma$-ray sources increased dramatically, from 271 objects listed in the 3rd EGRET Catalog \\citep[3EG;][]{har99}\\footnote{See also \\citet{cas08} for the revised catalog of EGRET $\\gamma$-ray sources.} to 1873 in the Second \\F Catalog \\citep[2FGL;][]{2FGL}. About 800 $\\gamma$-ray sources included in 2FGL have been identified as blazars \\citep{2LAC}, i.e., jetted active galactic nuclei (AGN) characterized by strong relativistic beaming. Other associations included pulsars \\citep[e.g.,][]{LATPSR}, high-mass X-ray binaries \\citep[e.g.,][]{HMXB}, radio galaxies \\citep[e.g.,][]{MAGN}, pulsar wind nebulae \\citep[e.g.,][]{pwn}, supernova remnants \\citep[e.g.,][]{snr}, globular clusters \\citep[e.g.,][]{glbcls}, starburst galaxies \\citep[e.g.,][]{sburst}, and distinct objects like the Large Magellanic Cloud \\citep{lmc}. However, no obvious counterparts at longer wavelengths have been found for as much as $38\\%$ of \\F objects so that several hundreds of GeV sources currently remain \\emph{unassociated} with any known astrophysical systems. Fortunately, an improved localization error for the \\F \\citep[typical $95\\%$ confidence radii $r_{\\rm 95} \\sim 0^{\\circ}.1-0^{\\circ}.2$, and even $0^{\\circ}.005-0^{\\circ}.01$ for the brightest sources;][]{2FGL}, when compared to that of EGRET (typical $r_{\\rm 95} \\simeq 0^{\\circ}.4-0^{\\circ}.7$), allows for much more effective follow-up studies at radio, optical, and X-ray frequencies, which can help to unravel the nature of the unidentified $\\gamma$-ray emitters. In this context, X-ray follow-up observations of unidentified \\F objects are of particular importance, since some classes of astrophysical sources of $\\gamma$-rays such as AGN are strong X-ray emitters as well, while the others like most of $\\gamma$-ray emitting pulsars are faint X-ray sources. Note that assuming the keV--to--GeV emission continuum in the form of a broad-band power-law ($F_{\\nu} \\propto \\nu^{-\\alpha_{{\\rm x} \\gamma}}$), which could be a relatively good zero-order approximation in the case of blazar sources but not necessarily in the case of other classes of $\\gamma$-ray emitters, the monochromatic X-ray flux energy density scales as $[\\nu F_{\\nu}]_{\\rm 1\\,keV} = \\left({\\rm 1\\,keV}/{\\rm 0.1\\,GeV}\\right)^{1-\\alpha_{{\\rm x} \\gamma}} \\times [\\nu F_{\\nu}]_{\\rm 0.1\\,GeV} \\simeq 3 \\times 10^{-3} \\times [\\nu F_{\\nu}]_{\\rm 0.1\\,GeV}$ for a relatively flat spectral index of $\\alpha_{{\\rm x} \\gamma} \\simeq 0.5$. Hence, if an X-ray counterpart of a bright \\F source is characterized by, e.g., $[\\nu F_{\\nu}]_{\\rm 0.1\\,GeV} \\simeq 10^{-11}$\\,erg\\,cm$^{-2}$\\,s$^{-1}$ and the X-ray--to--$\\gamma$-ray power-law emission continuum with the slope $\\alpha_{{\\rm x} \\gamma} \\geq 0.5$, such source can be expected to be detectable with modern X-ray instruments such as {\\it Chandra}, XMM-{\\it Newton}, {\\it Swift}, and \\SU within reasonable exposure times. In particular, a point source search to the level of $[\\nu F_{\\nu}]_{\\rm 1\\,keV} \\sim (10^{-14} - 10^{-13})$\\,erg\\,cm$^{-2}$\\,s$^{-1}$ is easily attainable with the X-ray Imaging Spectrometer \\citep[XIS;][]{koy07} onboard \\SU \\citep{mit07} with relatively short exposures of few tens of ksec \\citep[e.g.,][]{aka11}. In the case of a positive detection, correlated flux changes at X-ray and $\\gamma$-ray frequencies provide an identification. The lack of correlated variability, or non-detection of an X-ray counterpart, provide on the other hand only circumstantial evidence regarding the nature of a studied target. Yet in many cases such evidence may be crucial, since the non-detection of an X-ray counterpart despite a long, dedicated observation of a bright \\F object may disprove a potential association with given classes of astrophysical sources. That is because, as mentioned above, only a few established high-energy emitters are that bright in $\\gamma$-rays but very faint in X-rays \\citep[e.g., Geminga pulsar; see the discussion in][]{tho04,mat07}. Thus motivated we started a project to investigate the nature of unidentified high Galactic-latitude \\F objects through deep X-ray follow-up observations with \\SU XIS. The results of the first-year campaign conducted over the span of \\SU AO4 were presented in \\citet{mae11}. The AO4 program included four steady/weakly variable \\F sources from the initial \\F Bright Source List \\citep[0FGL;][]{0FGL} and can be summarized as follows. The X-ray counterpart for one of the brightest unassociated \\F objects, 1FGL~J1231.1$-$1410 (also detected by EGRET as 3EG~J1234$-$1318 and EGR~J1231$-$1412), was found. The X-ray spectrum of the counterpart was well fit by a blackbody model with a temperature of $kT \\simeq 0.16$\\,keV plus an additional power-law component dominating above 2\\,keV photon energies. This power-law component was confirmed in subsequent {\\it Swift} and XMM-{\\it Newton} exposures. Considering a recent identification of 1FGL~J1231.1$-$1410 with the millisecond pulsar (MSP) PSR~J1231$-$1411 \\citep{ran11}, in \\citeauthor{mae11} we concluded that the detected thermal X-ray photons originate from the surface of a rotating magnetized neutron star, while the non-thermal X-ray component is most likely produced within the pulsar magnetosphere. In the case of 1FGL~J1311.7$-$3429, two possibly associated X-ray point sources were discovered, one of which is now excluded from the smaller error ellipse of the GeV emitter as catalogued in the 2FGL \\citep{2FGL}. The identification of the remaining X-ray counterparts with the respective $\\gamma$-ray objects remain uncertain despite a robust determination of the the spectral and variability properties of the X-ray sources. In the case of 1FGL~J1333.2+5056, we found several weak X-ray sources within the \\F error circle, and speculated on the AGN nature of the target. Finally, one X-ray point source was detected at the edge of the error ellipse of 1FGL~J2017.3+0603. The physical connection was however viewed as unlikely, since the X-ray source did not coincide with the location of the MSP PSR~J2017+0603 discovered by the Nan\\c{c}ay radio telescope which constituted a more highly probable association with the \\F object \\citep{cog11}. The MSP identification was later indeed confirmed by the detection of the pulsed emission in the \\F data, with the same period as the radio pulsations. In this paper, we report the results of our second-year campaign, conducted over the span of \\SU AO5 (2010 April to 2011 March) which included observations of seven \\F sources located at high Galactic latitudes ($|b|>10^{\\circ}$). The targets were selected from the First \\F Catalog of point sources \\citep[1FGL;][]{1FGL} as objects \\emph{unidentified} at the time of writing of the \\SU AO5 proposal. Since then however, four of the selected targets have been associated with MSPs: 1FGL~J1902.0$-$5110 with PSR~J1902$-$5105 \\citep{cam11}, 1FGL~J2043.2+1709 with PSR~J2043+1711 \\citep{gui11}, 1FGL~J2302.8+4443 with PSR~J2302+4442 \\citep{cog11}, and 1FGL~J1312.6+0048 with PSR~J1312+00 \\citep{2FGL}. And moreover, 1FGL~J0106.7+4853 have very recently been associated with a normal pulsar PSR~J0106+4855 \\citep{ple11}. The XMM-{\\it Newton} and {\\it Swift} satellites detected the weak X-ray counterpart of the MSP in 1FGL~J2302.8+4443 \\citep{cog11}. In the following, our new \\SU observations and data reduction procedure are described in section~\\ref{sec:obs}. The analysis results are given in section~\\ref{sec:results}, and discussed further in section~\\ref{sec:disc_and_conc} in the context of multiwavelength studies of unidentified \\F objects. ", "conclusions": "\\label{sec:disc_and_conc} In this paper, we report on the results of X-ray follow-up observations of seven bright \\F sources at high Galactic latitudes ($|b|>10^{\\circ}$) using \\SU XIS. We discovered the X-ray counterpart of 1FGL~J2302.8+4443 coinciding with the position of the MSP PSR~J2302+4442 recently claimed to be associated with the $\\gamma$-ray emitter. We did not however, detect X-ray counterparts for the other four \\F objects similarly identified with a normal pulsar (1FGL~J0106.7+4853) and MSPs, namely for 1FGL~J1312.6+0048, 1FGL~J1902.0$-$5110 and 1FGL~J2043.2+1709. (In a few cases the X-rays sources have been detected within the 2FGL error ellipses, but none at the positions of the pulsars.) A relatively weak X-ray source was found inside the 2FGL error region of 1FGL~J1739.4+8717. Finally, no candidate for the X-ray counterpart was detected for the remaining object 1FGL~J1743.8$-$7620. Including our previous observations of 1FGL~J1231.1$-$1410, 1FGL~J1311.7$-$3429, 1FGL~J1333.2+5056, and 1FGL~J2017.3+0603 reported in \\citet{mae11}, our sample of high Galactic-latitude \\F objects initially selected as unidentified and studied with \\SU consists now of eleven targets. For eight of these, we have detected single or multiple X-ray sources within the LAT error ellipses. Over the time period when the \\SU observations were being obtained, six targets from the $\\gamma$-ray sample were found to be associated with MSPs, one target (1FGL~J0106.7+4853) has been associated with a normal pulsar, and one source (1FGL~J1333.2+5056) has been classified as an AGN candidate, all in agreement with the gathered X-ray data. Still, four objects from the list remain unidentified. As argued below, one of these four, 1FGL~J1739.4+8717, is quite likely a high-redshift blazar. The source 1FGL~J1739.4+8717 was characterized by an enhanced flux level within the LAT photon energy range during the the first seven months of \\F operation\\footnote{\\texttt{http://heasarc.gsfc.nasa.gov/FTP/fermi/data/lat/catalogs/source/lightcurves/2FGLJ1738.9+8716.png}}. After that time, the activity of the $\\gamma$-ray emitter decreased. The photon index in the LAT energy band is $\\Gamma_{\\gamma}$ = 2.1 $\\pm$ 0.1 \\citep[where dN/dE $\\propto$ E$^{-\\Gamma_{\\gamma}}$ is the differential photon flux; ][]{2FGL}, which is a typical value for the $\\gamma$-ray spectra of BL Lac type blazars \\citep[see][]{2LAC}. Importantly, as written in section~\\ref{sec:J1739} one relatively bright radio source, NVSS\\,J173722+871744, is located inside the 2FGL error region of 1FGL~J1739.4+8717 (see Figure\\,\\ref{fig:J1739_Ximage}). With a typical radio spectral index for blazar sources and radio-to-optical and optical-to-X-ray spectral indices that are consistent with blazar broadband spectrum, it is quite likely that 1FGL~J1739.4+8717 is indeed associated with a distant blazar currently characterized by an activity level low enough so that its X-ray emission was below the detection limit of the XIS instrument ($\\sim 10^{-15}$\\,erg\\,cm$^{-2}$\\,s$^{-1}$) at the time of the performed \\SU observations. In general, unidentified sources constituted a large fraction of the population of $\\gamma$-ray emitters detected by EGRET ($\\sim 60\\%$ in 3EG), and at present about 31 $\\%$ of \\F sources in 2FGL catalog remain unassociated (specifically 273 sources at high Galactic latitudes ($| b | > 10^{\\circ}$) and 303 sources at low Galactic latitudes $| b | < 10^{\\circ}$). Those located at the lowest Galactic latitudes ($|b| < 5^{\\circ}$) are most widely expected to be associated with local systems such as molecular clouds, supernova remnants, massive stars, high-mass X-ray binaries, radio quiet pulsars, and pulsar wind nebulae \\citep[e.g.,][]{kaa96,yad97,rom99}. In particular, half a dozen of the brightest 3EG sources in the Galactic plane were identified as young pulsars \\citep{tho99}, despite the relatively poor localization of the EGRET sources and the source confusion complicated substantially the identification procedure. On the other hand, most of the unassociated 3EG sources at high Galactic latitudes ($|b| > 10^{\\circ}$) were later identified as blazars \\citep{sow03,sow04}. Pulsars were therefore expected to be found mainly among GeV emitters at low Galactic latitudes, while blazars were supposed to constitute the main population of GeV emitters at high Galactic latitudes. But the $\\gamma$-ray--bright pulsars were also expected to be found at intermediate Galactic latitudes \\citep[$5^{\\circ} < |b| < 73^{\\circ}$;][]{cra06}. The identification of a number of \\F objects located above the Galactic plane ($| b | > 10^{\\circ}$) with such systems (predominantly with MSPs) confirmed these expectations \\citep[see, e.g.,][and the discussion in \\citealt{mae11}]{ran10}. In Figure\\,\\ref{fig:ratio} we plot the X-ray--to--$\\gamma$-ray energy flux density ratios ($F_{2-10\\,{\\rm keV}}/F_{0.1-100\\,{\\rm GeV}}$) versus radio--to--$\\gamma$-ray energy flux density ratios ($F_{1.4\\,{\\rm GHz}}/F_{0.1-100\\,{\\rm GeV}}$) for the \\F objects from our \\SU sample discussed here (blue circles) with the radio data available in the literature \\citep[see Table\\,\\ref{tab:fits}]{1FGL,con98}. These can be compared with the analogous ratios evaluated for bright \\F objects identified with blazars and MSPs (denoted in the figure by red crosses and squares, respectively). We remind the reader that the blazar class includes flat spectrum radio quasars (FSRQs) and BL Lacertae objects (BL Lacs). In addition, in the figure we plot the two targets discussed in \\citet{mae11}, namely 1FGL~J1333.2+5056 most likely associated with an AGN, and a peculiar object 1FGL~J1311.7$-$3429 (pink stars). As shown, the blazar and MSP populations are clearly separated in the constructed flux ratio plane. Also, four objects from our sample which have recently been associated with a normal pulsar (1FGL~J0106.7+4853) and MSPs (1FGL~J1312.6+0048, 1FGL~J2043.2+1709, and 1FGL~J2302.8+4443) occupy the same region in the analyzed parameter space as the previously known MSPs detected in the GeV range\\footnote{In the case of 1FGL~J1312.6+0048, the X-ray--to--$\\gamma$-ray energy flux density ratio shown in Figure\\,\\ref{fig:ratio} is evaluated assuming the association of the $\\gamma$-ray sources with the MSP PSR~J1312+00. That is, the \\SU XIS upper limit derived at the position of the pulsar is considered, and not the X-ray flux of the \\SU source detected within 2FGL error region.}. On the other hand, in the case of 1FGL~J1739.4+8717 the evaluated energy flux density ratios --- which are very similar to those characterizing 1FGL~J1333.2+5056 --- are consistent with the blazar identification proposed above, if only NVSS\\,J173722+871744 is considered as the true counterpart of the $\\gamma$-ray emitter. In all, we conclude that the gathered \\SU XIS data together with the broad-band properties of the analyzed \\F objects are in agreement with the identification of most of them as MSPs. Yet a few cases in the analyzed sample (1FGL~J1739.4+8717, 1FGL~J1333.2+5056) constitute quite probable associations with AGN (high-redshift blazars). Finally, the nature of the remaining few targets (like 1FGL~J1311.7$-$3429) is still an open question, although, as inferred from Figure \\ref{fig:ratio}, the MSP identification seems more viable than the blazar one. In the near future, we are further continuing our X-ray studies during the \\SU AO6 cycle, focusing on both new targets (1FGL~J0103.1+4840, 1FGL~J1946.7$-$5404, and 1FGL~J2339.7$-$0531), but also performing ultra-deep exposures on particularly intriguing sources like 1FGL~J1311.7-3429. Finally, let us comment in more detail on the case of 1FGL~J2302.8+4443. This object, as already mentioned above, has recently been associated with the millisecond pulsar PSR~J2302+4442 discovered by the Nan\\c{c}ay radio telescope \\citep{cog11}. The rotation period of the pulsar is $P \\simeq 5.19$\\,ms, the spin-down luminosity is $\\dot{E} \\simeq 3.74 \\times 10^{33}$\\,erg\\,s$^{-1}$, and the characteristic age can be evaluated as $\\tau \\simeq 6.2$\\,Gyr. \\citet{cog11} reported also on the detection of the X-ray counterpart of the pulsar with XMM-{\\it Newton}, with the unabsorbed $0.5-3$\\,keV flux of $3.1^{+0.4}_{-0.4} \\times 10^{-14}$\\,erg\\,cm$^{-2}$\\,s$^{-1}$. This is consistent with our \\SU detection (the re-calculated flux in the same photon energy range $2.9^{+1.1}_{-1.2} \\times 10^{-14}$\\,erg\\,cm$^{-2}$\\,s$^{-1}$). Anticipating the distance of the pulsar $d \\simeq 1.18$\\,kpc which was inferred from the NE2001 model of Galactic free electron density \\citep{cor01}, the X-ray luminosity of PSR~J2302+4442 can therefore be evaluated roughly as $L_{\\rm x} \\sim 3 \\times 10^{30}$\\,erg\\,s$^{-1}$. This, together with $\\dot{E}$ provided above, is then in good agreement with the scaling relation between the X-ray and spin-down luminosities $L_{\\rm x} \\sim 10^{-3} \\times \\dot{E}$ established for MSPs \\citep{bec97,gae06,zha07}, although the inferred distance of PSR~J2302+4442 is indicated to be smaller by a factor of four considering an unphysically high $\\gamma$-ray efficiency and, instead, assuming average efficiency of $\\gamma$-ray MSPs $\\sim$ 10 \\% \\citep{cog11}. The spectrum of the X-ray counterpart of PSR~J2302+4442/1FGL~J2302.8+4443 was well fit by a blackbody model, and this is again in agreement with the idea that the observed X-ray photons originate from thermal emission from the surface of a rotating magnetized neutron star. Because of the limited photon statistics in the higher energy range of XIS, we could not however confirm the presence of a non-thermal component above 2\\,keV \\citep[clearly detected in the case of 1FGL~J1231.1$-$1410 by][]{mae11}. Interestingly, for the isolated pulsars as old as PSR~J2302+4442 (characteristic age of about 6\\,Gyr), the surface temperature of the neutron star is expected to be $T \\lesssim 10^{5}$\\,K \\citep{nam87,pag92}. Our detection of the X-ray counterpart indicated $kT \\simeq 0.31 \\pm 0.03$\\,keV, i.e. $T \\simeq 3.6 \\times 10^{6}$\\,K instead. Some reheating process is therefore required, possibly related to the impact of relativistic particles on polar caps \\citep[and references therein]{bec97}. C.C. Cheung's work at NRL is sponsored by NASA DPR S-15633-Y. \\L .S. is grateful for the support from Polish MNiSW through the grant N-N203-380336. We thank the anonymous referee for a careful reading of the manuscript and useful suggestion which helped to improve the paper." }, "1112/1112.4334_arXiv.txt": { "abstract": "We study the effect of the non-Gaussianity induced by gravitational evolution upon the statistical properties of absorption in quasar (QSO) spectra. Using the generic hierarchical ansatz and the lognormal approximation we derive the analytical expressions for the one-point PDF as well as for the joint two-point probability distribution (2PDF) of transmitted fluxes in two neighbouring QSOs. These flux PDFs are constructed in 3D as well as in projection (i.e. in 2D). The PDFs are constructed by relating the lower-order moments, i.e. cumulants and cumulant correlators, of the fluxes to the 3D neutral hydrogen distribution which is, in turn, expressed as a function of the underlying dark matter distribution. The lower-order moments are next modelled using a generating function formalism in the context of a {\\em minimal tree-model} for the higher-order correlation hierarchy. These different approximations give nearly identical results for the range of redshifts probed, and we also find a very good agreement between our predictions and outputs of hydrodynamical simulations. The formalism developed here for the joint statistics of flux-decrements concerning two lines of sight can be extended to multiple lines of sight, which could be particularly important for the 3D reconstruction of the cosmic web from QSO spectra (e.g. in the BOSS survey). These statistics probe the underlying projected neutral hydrogen field and are thus linked to ``hot-spots'' of absorption. The results for the PDF and the bias presented here use the same functional forms of scaling functions that have previously been employed for the modelling of other cosmological observation such as the Sunyaev-Zel'dovich effect. ", "introduction": "Ongoing Cosmic Microwave Background (CMB) experiments such as Planck\\footnote{http://www.rssd.esa.int/Planck}, the Atcama-Cosmology Telescope\\footnote{http://www.physics.princeton.edu/act/} (ACT) and the South Pole Telescope\\footnote{http://pole.uchicago.edu/} (SPT) will pinpoint the cosmological parameters that describe the background geometry and dynamics of the Universe in an unprecedented detail. Along with very precise constraints on the structure of the Universe on the largest scales, smaller-scales observables will be crucial in order to further constrain the cosmological concordance model or find possible deviations from it. In particular, galaxy clustering (BOSS, 6dF, etc.) surveys and future weak lensing and clustering observations (e.g. EUCLID) could probe smaller scales and new redshif regimes. Spectroscopic surveys such as BOSS\\footnote{http://cosmology.lbl.gov/BOSS/} (and BIG BOSS) will also trace the large scale distribution of the baryonic matter in the Universe through the study of the flux distribution of the Lyman-${\\alpha}$ absorption systems in a very large number of quasar (QSO) spectra. The Lyman-$\\alpha$ ``forest'', the many absorption features in QSO spectra, produced by intervening neutral hydrogen in the intergalactic medium (IGM) along the line-of-sight, is well known to be an important cosmological probe (for a recent review see Meiksin 2009). In the standard cosmological paradigm, the IGM consists of mildly non-linear gas, making up the cosmic web, that traces the dark matter and is photo-heated by a Ultra Violet (UV)-background. The Lyman-$\\alpha$ forest is thus the main probe of the IGM and it has been shown to arise naturally in hierarchical structure formation scenarios. Astrophysical effects produced by feedback from galaxies and/or AGNs do not seem to strongly affect the vast majority of the baryons in the cosmic web \\citep{Mc05,T02}, thereby this can be used as a dark matter tracer. The relation between the Lyman-${\\alpha}$ forest flux and the underlying matter field is a nonlinear one and it is generally expected that statistics of Lyman-$\\alpha$ are biased relative to the underlying dark matter distribution. The Lyman-${\\alpha}$ forest has been studied using variety of analytical techniques such as the Zel'dovich approximation, which is valid in the quasilinear regime and often used in modelling of the nonlinear gravitational clustering \\citep{DS97, HGZ97,McGill90,MM02}. In addition, the lognormal approximation \\citep{CJ91} is frequently used to model the statistics of Lyman-$\\alpha$ forest \\citep{Bi93,GH96,BD97,RPT01,Viel02}. Models based on the hierarchical or scaling {\\em ansatz} \\citep{BS89,BeS92,BS99} for higher-order correlation functions have also been investigated in order to model the statistics of Lyman-$\\alpha$ forest \\citep{VSS99}. In addition to analytical modelling, hydrodynamical simulations have also played a very important role in this field (e.g. \\cite{Cen94,GH98,Cr98,Cr99,MW01}) and support the simple analytical picture. Thus, analytical schemes, including the ones that we develop here, can be calibrated using numerical simulations and this in turn allows to explore a large parameter space efficiently. However, numerical simulations are required to resolve the Jeans scale of the photo-ionized warm IGM, and this requirement typically means small box sizes that sample larger scales modes rather poorly. Indeed several semi-numerical prescriptions, which are not entirely based on hydrodynamical simulations, have also been developed to model Lyman-$\\alpha$ flux and recover the correlation function from observed data sets \\citep{Sls11}. \\begin{figure} \\begin{center} {\\epsfxsize=10 cm \\epsfysize=5.1 cm {\\epsfbox[29 436 581 714]{flux.eps }}} \\end{center} \\caption{The PDF $p(F_{\\alpha})$ and bias $b(F_{\\alpha})$ of the flux $F_{\\alpha}$ is plotted as a function of the flux $F_{\\alpha}$. The PDF (left-panel) and the bias (right-panel) both are constructed using a lognormal model for the underlying mass distribution. The resulting PDFs for $\\tilde\\delta_{}$ are next transformed into the flux PDFs using the fluctuating Gunn-Peterson approximation Eq.(\\ref{eq:gp}). We compare the results from the lognormal approximation (solid lines) against the one based on the Gaussian approximation (dashed lines). The values of $A(z)$ and $\\beta$ are constructed using the functional fit in Eq.(\\ref{eq:fit}) given by \\citep{Kim07} see text for more details. The bias changes signs at an intermediate flux value which depends on cosmology and hydrodynamical parameters that define the equation of state of the photo-ionized medium $A(z)$ and $\\beta$. A fiducial value for the variance $\\sigma=3$ was assumed for this plot. } \\label{fig:demo} \\end{figure} The two most commonly used approaches in Lyman-${\\alpha}$ studies are based either on decomposing the information encoded in the transmitted flux via Voigt profile fitting or treating the flux as a continuous field. In the first approach, the shapes and clustering properties of absorption lines fitted by Voigt profiles have been investigated in variety of studies involving the temperature of the IGM \\citep{STLE99,MMR01}, in order to constrain the reionization history \\citep{TSS02,HH03}, to measure the matter power spectrum and cosmological parameters \\citep{Croft99,VHS04,MSB06,MM99,Rol03,Copp06,Gui07,VH06}. By using the second set of methods, statistical properties of the flux such as the mean flux level, flux PDF, flux power spectrum \\citep{VHS04,Viel08,SSM06} and flux bi-spectrum are typically employed to explore flux statistics. For example, it has been shown that the mean-flux level can be used to constrain the amplitude of intergalactic UV background \\citep{T04,Bolt05} while the flux PDF \\citep{Mc09, Bolt09} is sensitive to the thermal evolution of the IGM (see also \\cite{L06}). The flux power spectrum on the other hand can be used to constrain the cosmological parameters and nature of dark matter \\citep{Croft02}. The flux-bispectrum \\citep{Viel09} contains useful information about the primordial, as well as gravity-induced (i.e. secondary) non-Gaussianity. The data typically used in these investigations consists mainly of two different sets of QSO spectra: the SDSS low resolution low signal-to-noise spectra and UVES/VLT or HIRES/KECK high resolution spectra. The number of SDSS spectra is about a factor $\\sim 200$ larger than that of high resolution samples though the later probes the smaller scales with greater accuracy. More recently, it has been argued that BOSS-like QSO spectroscopic surveys could detect Baryon Acoustic Oscillations (BAO) signatures at high redshift \\citep{ME07} and a sample of QSO pairs can constrain the geometry of the high redshift universe \\citep{Mc03}. Furthermore, analysis of coincident absorption lines in QSO pairs can also allow departures from the Hubble flow and non-gravitational effects to be measured \\citep{Rauch05}. The SDSS-III/BOSS survey aims at identifying and observing more than 160,000 QSO over $\\sim 10,000$ square degrees within a redshift range $z=(2.15-3.5)$. This survey is primarily design to studying baryonic acoustic oscillations by performing a full 3D sampling of the matter density. Such studies will also provide an unprecedented opportunity to study the clustering statistics using projected Lyman-$\\alpha$ flux decrements of QSOs. In this paper, we will consider the Lyman-$\\alpha$ flux decrement in two dimensions (2D) which is related the projected density of neutral hydrogen. The statistical study of projected Lyman-$\\alpha$ flux decrement can be performed using one- or two-point PDF or their lower order moments. The PDFs contain information of cumulants or their correlators to an arbitrary order and can be constructed using well-established machinery of hierarchical ansatz \\citep{MCM1,MCM2,MCM3}. Several authors have recently studied the lower-order cumulants of Lyman-$\\alpha$ forests and cross-correlated them against weak lensing convergence as well as to the CMB sky \\citep{VDSS09,VVDS11}. The results presented here are complementary to such studies as we take into account the lower order moments to an arbitrary order not just for the one-point cumulants but also for their two-point counterparts or {\\em cumulant correlators}. The cumulant correlators are the two-point analogues of one-point cumulants and are already in use in different areas of cosmology, e.g. in analysis of galaxy surveys \\citep{SS97}. The lowest in the two-point hierarchy is the two-point correlation function. In the context of Lyman-$\\alpha$ studies, the two-point correlation function has already been introduced in studies involving two neighbouring line of sights \\citep{Viel02,Dod02}. Our study generalizes these results to probe non-Gaussian correlation functions involving multiple line of sight. We model the statistics of underlying neutral hydrogen distribution using lognormal distribution as well as an extension of perturbation theory approach \\citep{VaMu04} in 3D. Predictions from these models are then tested using hydrodynamical simulations at three different redshifts ($z=2,3,4$). Next we use these results to build and test statistical description of the projected flux distribution. The results presented here can be generalized to cross-correlation studies involving external data sets and Lyman-$\\alpha$ flux distribution. The correlation functions (CCs) are equivalent to their Fourier (harmonic) space counterpart, the multispectra, recently introduced by \\citep{MuHe09}. The plan of this paper is as follows. In \\textsection\\ref{sec:form} we introduce the notations and define the relevant quantities such as the transmitted flux and its relation to the underlying density contrast. In \\textsection\\ref{sec:sims} we give details of the simulations that were used in our study. In \\textsection\\ref{sec:3D} details of modelling of the 3D flux are presented. In \\textsection\\ref{sec:lower} we derive the lower order statistics for the flux in terms of that of the underlying density contrast. In \\textsection\\ref{sec:hier} we provide a very brief introduction to the hierarchical ansatz which we use to model the statistics of underlying density contrast. In \\textsection\\ref{sec:pdf} we provide derivation of the PDF and the bias associated with the flux distribution and finally \\textsection\\ref{sec:conclu} is left for discussion of results. We also provide a brief appendix introducing the lognormal approximation as well as the hierarchical ansatz. ", "conclusions": "\\label{sec:conclu} The diffuse intergalactic medium acts as a significant reservoir of baryons at low to intermediate redshift ($z<5$), which can be probed via the absorption lines in the spectrum of distant QSOs. The Lyman-$\\alpha$ absorption lines along the line of sights are due to a vast range of completely different classes of object. These objects include under-dense neutral hydrogen clouds, halos of large and over-dense systems which may have already reached virial equilibrium, as well as UV heated systems that are strongly coupled to their environments. Previous studies have mainly dealt with the problem of detailed modelling of number of these objects as a function of their clustering and internal properties. Using self-consistent scaling models, which have a long history and were initially employed in galaxy clustering statistics, several authors have computed the column density distribution of Lyman-$\\alpha$ absorption systems not only for the low column density Lyman-$\\alpha$ forest systems but also for Lyman-limit systems and the damped absorption systems. In addition to the hierarchical modelling, lognormal approximation is also quite successful in reproducing the clustering statistics of Lyman-{$\\alpha$} absorption lines. In this study we showed the approach taken by the lognormal approximation and the hierarchical ansatz generates near identical results. While previous studies focussed only on one-point statistics we extend these results to the two-point distributions and their lower order moments. In addition to the statistics of absorption lines, the statistics of the transmitted flux play an important complementary role in Lyman-$\\alpha$ studies. The flux can be treated as a continuous field. Various statistics which are often employed in analysing the flux include the LOS power spectrum measurements, estimation of bispectrum and more recently the entire PDF. The flux PDF contains the information regarding cumulants to an arbitrary order. Clearly the flux statistics and the column density distributions are related. One major goal in this study was to unify these two pictures in the context of the hierarchical ansatz or scaling models as well as the lognormal approximation. The hierarchical model is primarily valid at smaller scales where the correlation functions assumes a hierarchical form. Gravitational clustering is known to develop such a form of hierarchy both in the perturbative (quasi-linear) as well as in the highly non-linear regime. The hierarchical ansatz can be used to describe the mass functions and the bias associated with collapsed objects using the scaling function $h(x)$ and $b(x)$; the variable $x$ is a scaling variable. Previous studies that have employed the hierarchical ansatz have shown that the column density distribution of absorption systems can be described using the same functional form for $h(x)$ and $b(x)$ that are often used for wide ranging studies from the galaxy clustering to thermal Sunyaev-Zel'dovich effects or X luminosity of clusters of galaxies. We showed that the same analytical framework can also be used to understand the statistics of QSO flux measurements, thereby providing a unified statistical approach to what at first sight appear to be very different observables. We have approached the modelling of the flux PDF in two different ways. For the 3D analysis, the PDF of the neutral hydrogen density contrast $\\hat\\delta$ is modelled according to the hierarchical ansatz as well as lognormal distribution. We find very good agreement with the numerical simulations for the entire range of redshifts that we have considered. The predictions from both these models are almost identical and differ only marginally in the less interesting under-dense regions. We also employ a modified version of scale invariant hierarchical approximation which was developed recently \\citep{VaMu04}. This particular approximation provides a very accurate model for the clustering of $\\hat\\delta$. Using the fluctuating Gunn-Peterson approximation we map the $\\hat\\delta$ PDF to that of the transmitted flux $F_{\\alpha}$. We find reasonable approximation for the allowed range of parameters $A(z)$ and $\\beta$ that define the Gunn-Peterson approximation. Next we consider the projected or the 2D distribution of flux. For projected statistics, we start with linking the cumulants and cumulant correlators for the flux and the underlying neutral hydrogen density contrast $\\hat \\delta$. Then $\\hat \\delta$ is taken to be a tracer of the underlying density contrast $\\delta$, modelled statistically using the hierarchical ansatz. Next it is shown that, under certain simplifying assumptions, the PDF of $\\delta$ and PDF of a suitably defined reduced flux decrement are linked through a very simple relation. Tests against numerical simulations shows good agreement. Results were obtained for the entire bias functions that act as a generating function for the cumulant correlators. The scaling function $h(x)$ and the bias function $b(x)$ are also known to describe the number density and bias of over-dense objects respectively. The formalism developed here can also be used to probe the higher order cross-correlation statistics involving Lyman-$\\alpha$ flux decrement and weak lensing convergence of CMB maps or those extracted from convergence maps constructed using the weak gravitational lensing of optical galaxies." }, "1112/1112.1382_arXiv.txt": { "abstract": "{Radial velocity (RV) measurements from near-infrared spectra have become a potentially powerful tool to search for planets around cool stars and sub-stellar objects. As part of a large survey to characterize M-dwarfs using NIRSPEC at Keck~II, we obtained spectra of eight late M-dwarfs (spectral types M5.0-M8.0) during two or more observing epochs per target. These spectra were taken with intermediate spectral resolving powers ($R\\sim20,000$) in the $J$-band. } {We search for relative radial velocity variability in these late M-dwarfs and test the NIRSPEC capability of detecting short period brown dwarf and massive planetary companions around low-mass stars in the $J$-band ($\\approx 1.25~\\mu$m). Additionally, we reanalyzed the data of the M8-type star vB10 (one of our targets) presented in Zapatero Osorio et al. (2009), which were obtained with the same instrumentation as our data.} {To achieve a precise RV measurement stability, the NIRSPEC spectra are self-calibrated by making use of the telluric absorption lines, which are present in the observed spectra and used as a long-term stable reference. In the modeling process a multi-parameter $\\chi^2$-optimization is employed to generate an accurate description of the observation. The telluric lines allow us to model the instrumental profile of the spectrograph and the determination of the Doppler shift of the stellar absorption lines. } {For the entire M-dwarf sample, we do not find any evidence of relative RV variations induced by a short period brown dwarf or massive planetary companion. The typical RV precision of the measurements is between 180 and 300~m~s$^{-1}$, which is sufficient to detect hot Neptunes around M-dwarfs. Also, we find that the spurious RV shift in Zapatero et al.~(2009) of the star VB10 was caused by asymmetries in the instrumental profile between different observing epochs, which were not taken into account in their analysis.} {} ", "introduction": "The search for extrasolar planets has led to more than 700 confirmed discoveries\\footnote{The Extrasolar Planets Encyclopedia; http://www.exoplanet.eu; 2011-Nov-15} by using all detection techniques. Up to now, most of them have been detected by means of the radial velocity (RV) technique using high-resolution spectrographs ($R=\\lambda / \\Delta \\lambda \\ge 40,000$) at optical wavelengths. Most discoveries are giant gaseous planets (typically hot Neptunes and Jupiters) of short periods (of a few days) around stars of spectral types F, G and K. As potential hosts to rocky planetary companions, M-dwarfs have become increasingly popular as targets for RV searches (e.g. Endl et al. 2006, Charbonneau et al. 2009, Mayor et al. 2009, Zechmeister et al.~2009). Very cool stars such as M-dwarfs are the most abundant type ($\\sim70 \\%$) of stars in the solar neighborhood and the Milky Way in general (Henry at al.~1997). The effective temperatures and masses of M-dwarfs, respectively, are in the range 3700 to 2200~K and 0.5 to 0.07 solar masses for the M0 to M9.5 spectral types. They exhibit prominent absorption features corresponding to strong neutral atoms, H$_2$0, FeH, VO, CO, and TiO. Owing to the low masses of these objects, the reflex motion of the host star due to the gravitational pull of the extrasolar planet is higher and more easily detectable than for more massive host stars. Since M-dwarfs are very cool stars in comparison with solar-type stars, short period planets would more likely be situated in the habitable zone. M-dwarfs emit most of their energy around $1.1-1.3~\\mu {\\rm m}$, in the near-infrared (NIR), while they appear very faint at optical wavelengths. First attempts to measure RV variations among very cool M-dwarfs at NIR wavelengths were done by Mart\\'in et al.~(2006). They achieved a RV precision of around 300 m~s$^{-1}$ for the M9.5-dwarf LP944-20 by using the spectrograph NIRSPEC, which is mounted on the Keck II telescope in Hawaii (McLean et al.~1998). Recently, several research groups have reported high-precision RV measurements taken in the NIR with CRIRES (K\\\"aufl et al.~2004), mounted at the UT1/VLT in the Paranal Observatory of ESO in Chile. Bean et al.~(2010a) conducted high-resolution spectroscopic data of over 60 M-dwarfs (spectral types M4-M9) and used a NH$_3$ gas cell spectrum as a stable reference, and report an RV precision of better than $5~{\\rm m~s^{-1}}$. Figueira et al.~(2010) took observations of the planetary candidate TW~Hya and achieved a RV precision better than $10~{\\rm m~s^{-1}}$ by adopting telluric lines as a stable reference. Blake et al.~(2010) report RV measurements of 59 M- and L-dwarfs using the Keck/NIRSPEC spectrograph, with the aim to detect low-mass companions. They made use of strong CO absorption features around 2.3~$\\mu$m in M- and L-dwarfs and achieved RV precisions between 50 and 200~m~s$^{-1}$. Tanner et al.~(2010) report preliminary results of a late M-dwarf survey by using Keck/NIRSPEC with RV precisions between 150-300~m~s$^{-1}$. In 2009, Pravdo \\& Shaklan~(2009) announced a massive planet around the M-dwarf vB10 discovered by means of astrometrical data. Zapatero Osorio et al.~(2009; hereafter ZO09) made use of our NIRSPEC data set and found evidence for RV variations, which supported the planet hypothesis. They achieved a RV precision of about 300~m~s$^{-1}$. However, this planet was later refuted by different groups: Bean et al. (2010b), who took high-resolution spectra ($R=\\lambda/\\Delta\\lambda\\sim100,000$) with CRIRES and who achieved a RV precision of $\\sim 10$~m~s$^{-1}$, and by Anglada-Escud\\'e et al. (2010). Additionally, Lazorenko et al.~(2011) carried out an astrometric survey using the FORS2 camera of the ESO/VLT on Cerro Paranal, Chile, but found no evidence for the existence of a massive planet orbiting vB10. As part of this work, we aimed at finding out what had caused the spurious RV variations in the data analysis of ZO09. Here, we report relative RV measurements of 8 late M-dwarfs with NIRSPEC, and we support the capability of this instrument to detect giant planetary companions with short orbital periods. In Section 2 we describe our M-dwarf sample, our observations and data reduction. In Section 3 we outline the details of the data analysis, followed by the results and discussion (Section~4). ", "conclusions": "We analyzed the data sets with our relative radial velocity measurement approach and determined the relative RV measurements with respect to the selected reference epoch. For any of the eight M-dwarfs in our sample, we have not found significant evidence of relative RV variations at the level of 3$\\sigma$ (Table~\\ref{xoxo:T3}), where $\\sigma$ stands for the measurements uncertainty. The RV precisions are in the order of 180-300~m~s$^{-1}$, except for the observations in July 2008, which were taken at low SNR. We investigated the period and mass range of companions which could be detected with such RV precisions. We determined the minimum mass of the planet by employing a Monte-Carlo analysis, thereby probing planetary orbits with different parameters and investigating how many of these orbits could be recovered for the five measurements of vB10. We considered only the case of a circular orbit and the mass of vB10, which is $m_\\star=0.078~{\\rm M}_\\odot$. Fig.~\\ref{xoxo:F1} shows the 3$\\sigma$ detection limit. We find that for companions with only a few days period, even planets with minimum masses of $m_{\\rm{p}}\\sin i \\ge 0.3~M_{\\rm Jup}$ can be detected with a RV precision of $\\sim220$~m~s$^{-1}$. \\begin{figure} \\includegraphics[angle=270,scale=.4]{out.dat.ps} \\caption{Monte-Carlo analysis for the five vB10 measurements. For the mass of vB10, we adopted 0.078~M$_\\odot$ from Pravdo \\& Shaklan (2009). It is shown that with a RV precision of $\\sim220$~m~s$^{-1}$ even hot Jupiters with minimum masses $m_{\\rm p} \\sin i> 0.3~$M$_{\\rm Jup}$ could be detected around late M-dwarfs with 3$\\sigma$ confidence. We note that for a larger number of measurements the number of aliasing peaks can be significantly decreased.\\label{xoxo:F1}} \\end{figure} In Fig.~4 we show our relative RVs of vB10 and the measurements by ZO09. We note that for a proper comparison, we adopted the same reference epoch as in ZO09. The agreement between ZO09 and our measurements is within 1$\\sigma$ of the quoted uncertainties for all epochs except for the 2001 epoch (BJD = 2452076). We provide next an explanation for the discrepancy of this one measurement. Similar to our data analysis, ZO09 used the telluric lines present in the target spectra as a stable reference, but contrary to our analysis they did not account for any IP variations in their analysis, but calculated the RVs by cross correlation. In our analysis, we do not see any RV shift exceeding the RV-precision for any measurement. We get evidence that the different instrumental setting used on 2001-Jun-15 (0.576\" slit instead of the standard setting of 0.432\") produced an asymmetric instrumental profile (Fig.~\\ref{xoxo:F2}), which led to a significant RV shift when a simple cross correlation is adopted for the RV determination. Our results clearly demonstrate the importance of modeling the IP especially when observations are carried out with different instrumental settings. \\begin{figure} \\includegraphics[angle=270,scale=.355]{xoxo-fg2.ps} \\caption{Our RV measurements of vB10 (crosses) vs. the measurements of ZO09 (open circles). The RV precision given by our analysis is about $~220$~m~s$^{-1}$, except for the last epoch in 2008, which was hampered by bad weather. We prove that the RV shift in the work of ZO09 in the first 2001 epoch (BJD$=245~2076$) originates from unaccounted asymmetries in the IP rather than from a planetary companion. \\label{xoxo:F2A}} \\end{figure} \\begin{figure} \\includegraphics[angle=270,scale=.36]{new-ip-nogrid6.ps} \\caption{Instrumental profiles (IPs) of NIRSPEC for two different observing epochs of vB10. On 2001-06-15, a broader slit was used ($0.576\"$; solid line) than for the reference epoch ($0.432\"$; 3007-06-25; dashed line). To visualize the asymmetries between both IPs, we calculated the ratio between both IPs and scaled the resulting function for better visibility (dotted line). ZO09 did not account for these asymmetries between both IPs, which led to a spurious RV shift of about 1~km~s$^{-1}$ for the 2001-06-15 measurement in their data analysis.\\label{xoxo:F2}} \\end{figure} \\begin{table} \\begin{center} \\caption{Relative radial velocity measurements.} \\label{xoxo:T3} \\begin{tabular}{lcccc} \\hline\\hline Obs. Date & BJD & rel. RV \\\\ & 245~0000+ & (m~s$^{-1}$) \\\\ \\hline \\multicolumn{2}{l}{{\\bf 2MJ2331-2749}:} \\\\ 2007-Jun-24 & 4276.10874 & $194\\pm224$\\\\ 2007-Jun-25 & 4277.09424 & {\\it ref. epoch}\\\\ \\hline \\multicolumn{2}{l}{{\\bf GJ 406}:} \\\\ 2007-Apr-30 & 4220.79751 & {\\it reference epoch} \\\\ 2007-Dec-23 & 4458.14962 & $-77\\pm 238$ \\\\ \\hline \\multicolumn{2}{l}{{\\bf GJ 905}:} \\\\ 2007-Jun-25 & 4277.11249 & $-233\\pm 201$\\\\ 2007-Oct-27 & 4400.95657 & {\\it reference epoch} \\\\ \\hline \\multicolumn{2}{l}{{\\bf GJ 1156}:} \\\\ 2007-Apr-30 & 4220.81513 & $141\\pm185$\\\\ 2007-Dec-22 & 4457.16844 & {\\it reference epoch} \\\\ \\hline \\multicolumn{2}{l}{{\\bf LHS 1363}:} \\\\ 2007-Oct-26 & 4400.01234 & $-27\\pm196$\\\\ 2007-Oct-27 & 4401.01511 & {\\it reference epoch}\\\\ \\hline \\multicolumn{2}{l}{{\\bf LP 412-31}:} \\\\ 2007-Oct-26 & 4400.03654 & {\\it reference epoch} \\\\ 2007-Oct-27 & 4401.04491 & $298\\pm260$\\\\ \\hline \\multicolumn{2}{l}{{\\bf RXJ2208.2}:} \\\\ 2007-Jun-24 & 4276.07879 & $-189\\pm 307$\\\\ 2007-Jun-25 & 4277.07266 & {\\it reference epoch}\\\\ \\hline \\multicolumn{2}{l}{{\\bf vB10}:} \\\\ 2001-Jun-15 & 2076.08951 & $19\\pm230$ \\\\ 2001-Nov-02 & 2215.70560 & $69\\pm223$ \\\\ 2001-Nov-02 & 2215.74225 & $-3\\pm217$ \\\\ 2007-Jun-25 & 4277.05865 & {\\it reference epoch} \\\\ 2008-Jul-28 & 4675.75669 & $131\\pm497$ \\\\ \\hline \\end{tabular} \\end{center} \\end{table} We compare our results to the work of Blake et al.~(2010), who searched for companions to M- and L-dwarf wby using NIRSPEC at a spectral resolving power of $\\sim25,000$ in the $K$-band. They adopted one spectral order covering the wavelength range from 2.285 to 2.318~$\\mu$m to measure the dense and strong CO-absorption line pattern present in those dwarfs. As a stable wavelength reference, they made use of the CH$_4$ telluric absorption lines present in the observations. Similarly to us, they employed a self calibrating approach, with the difference that they adopted theoretical models for M- and L-dwarfs, which well-described the observations. Blake et al. obtained measurements with SNRs in the range of 50 to 100 in the pseudo stellar continua, and they report RV precisions of 100-300 m/s for slowly rotating late-M and L dwarfs. The uncertainty of Blake et al. in the K-band is in agreement with our derivation of 180-300 m/s in view of our SNRs. However, our wavelength coverage is about twice that in Blake et al. According to the relative RV precision formulae, we should have obtained better velocity precision in terms of wavelength coverage, which is not the case. We conclude that both the larger number of deep lines (more than 30 lines with a line depth larger than 50\\%) in the CO-band region as compared to the J-band (only a few lines with a depth larger than 25\\%) as well as the use of theoretical template spectra instead of deconvolved stellar spectra appear to compensate for the shorter wavelength coverage in a similar factor (c.f. Equation~6 in Butler et al. 1996). We note that Reiners et al.~(2010) and Rodler et al.~(2011) carried out theoretical RV precision studies of M- and L-dwarfs, by adopting theoretical models of M-dwarfs (e.g. del Burgo et al.~2009). As result, they find that the highest RV precision for M-dwarfs is attained in the $Y$ band around $1~\\mu$m, rather than in the $J-$ , $H-$ or $K$-band. For L-dwarfs, however, Rodler et al.~(2011) reported that the highest RV precision is attained in the $J$-band. We conclude that for an accurate relative RV determination with NIRSPEC, a self-calibrating approach, which accounts for changes in the instrumental setting, produces the best measurements in terms of RV precision. Although with our RV precision we would be able to detect massive hot Neptunes around late M-dwarfs, we have not found any brown dwarf or massive planetary companion in our survey. Additionally, the re-analysis of the data of the M8-dwarf vB10 presented in ZO09 now clearly confirms the non-existence of a massive planet orbiting that dwarf and agrees with the results by other research groups (e.g. Anglada-Escud\\'e et al.~2010; Bean et al.~2010b; Lazorenko et al.~2011)." }, "1112/1112.3542_arXiv.txt": { "abstract": "{Convergent point (CP) search methods are important tools for studying the kinematic properties of open clusters and young associations whose members share the same spatial motion.} {We present a new CP search strategy based on proper motion data. We test the new algorithm on synthetic data and compare it with previous versions of the CP search method. As an illustration and validation of the new method we also present an application to the Hyades open cluster and a comparison with independent results.} {The new algorithm rests on the idea of representing the stellar proper motions by great circles over the celestial sphere and visualizing their intersections as the CP of the moving group. The new strategy combines a maximum-likelihood analysis for simultaneously determining the CP and selecting the most likely group members and a minimization procedure that returns a refined CP position and its uncertainties. The method allows one to correct for internal motions within the group and takes into account that the stars in the group lie at different distances. } {Based on Monte Carlo simulations, we find that the new CP search method in many cases returns a more precise solution than its previous versions. The new method is able to find and eliminate more field stars in the sample and is not biased towards distant stars. The CP solution for the Hyades open cluster is in excellent agreement with previous determinations. } {} ", "introduction": "Ever since their discovery, the existence of stellar groups with common space motion in the solar neighborhood has been an intriguing issue whose understanding is still far from complete. The origin and evolution of these comoving groups of stars, usually referred to simply as moving groups, is explained by different scenarios including cluster disruption, dynamical effects, and accretion events \\citep{Eggen(1996),Dehnen(2000),Fux(2001),Navarro(2004)}. Moving groups, which are observed as overdensities in the velocity space and exhibit a low internal velocity dispersion, typically a few km/s or less \\citep{Mathieu1986}, allow study of the large-scale structure and dynamics of the Milky Way \\citep{Antoja2008}. Because of perspective effects, the proper motions of comoving stars\\footnote{In fact, the proper motions of a group of stars converge to a vertex either if their space motions are parallel or if they are expanding uniformly from a moving point. These two dynamical states are strictly equivalent as far as proper motions are concerned and radial velocity information is needed to distinguish between them \\citep{Blaauw1964}.} appear to converge to a single point in the celestial sphere referred to as the \\textit{convergent point} (hereafter CP) of the moving cluster. The CP is important not only for determining which stars are actual members of a moving cluster, but also for deriving individual kinematic distances of moving group members, provided that their radial velocities are known. This is very valuable when the trigonometric parallax from the ground is not accessible and \\rm{\\sc{Hipparcos}} parallaxes are not available \\citep[for recent applications of this strategy see][]{Mamajek2005, Bertout2006}. The first algorithms implementing a method for calculating the CP coordinates come from Charlier and Bohlin in 1916 \\citep[see][]{Smart1968}. Each of them derived an equation independently using the position angle of stars to determine the CP position on the celestial sphere. However, the constant least-square coefficients involved in solving these equations were subject to measurement errors, thus leading to systematic errors in the derived CP coordinates. Later, \\citet{Seares1945}, \\citet{Petrie1949} and \\citet{Roman1949} proposed different ways of correcting the Charlier and Bohlin equations. Their strategies represent, in a first approximation, different ways of using the position angle of stars to determine the CP solution. Another approach was proposed by \\citet{Brown1950}. He introduced a reference frame with the origin approximately at the center of the cluster and a fundamental plane defined by drawing a great circle through the origin in the direction of the average proper motion of the cluster. The proper motion of each star is then resolved into two components, one parallel to the reference plane and the other one perpendicular to the same plane, with the former expected to be much larger than the latter. The coordinates of the CP were then derived by applying the method of maximum likelihood. Based on this method, \\citet[][hereafter J71]{Jones1971} presented a twofold algorithm dedicated to simultaneously selecting the moving group members and calculating the CP position. This method was later improved and reformulated by \\citet[][hereafter B99]{deBruijne1999} to make full use of the \\textit{Hipparcos} data and allow for internal motions within the moving group. There are many methods of finding the CP coordinates of moving groups. Several of them use, as J71 and B99 do, the observed proper motions of stars and differ in the details of the search strategy \\citep[see for example][]{Makarov2001, Makarov2007a, Makarov2007b}. Other methods use different observed quantities, such as parallaxes and radial velocities \\citep[]{Chen1997, Chereul1999, Hoogerwerf1999,Asiain1999}. The choice of which method to use depends essentially on the observational information that is available. The new CP search method that we propose in this article uses the observed proper motions to select the stars that belong to the moving group and determine the CP position. It builds on the works of J71 and B99, so we give in Sect.~2 some details on these methods that will be useful in presenting our own work. In the rest of this paper we present and test a new version of the CP search method. Section~3 describes our new algorithm while Sect.~4 deals with the construction of synthetic data from moving group simulations dedicated to testing and investigating the performance of our algorithm in comparison with previous ones. Section~5 presents an application of our method to the Hyades open cluster using \\rm{\\sc{Hipparcos}} data, and finally Sect.~6 summarizes the results of this work. ", "conclusions": "We presented a new CPSM based on the idea of representing the stellar proper motions by great circles on the celestial sphere. The new CP method combines (i) maximum-likelihood analysis to simultaneously determine the CP and select moving group members with (ii) a direct minimization routine used to return a more refined CP position and its uncertainties. Our new method allows us to correct for internal motions within the group by applying an individual correction for each star that depends on its proper motion. This procedure takes into account that the stars in the group are not located at the same distance. We performed extensive Monte Carlo simulations to test and compare our new CPSM with the classic CPSM regarding the selection of moving members and the accuracy of the CP solution. We investigated the effects of (a) the velocity dispersion of the cluster, (b) observational errors on proper motion, (c) cluster distance, (d) number of group members, and (e) angular distance to the CP on the CP solution. Our new CPSM returned a more precise CP solution than the classic CPSM for 95\\% of the simulations. We verified that in the absence of velocity dispersion (ideal moving group) both methods exhibit the same performance. We also found that the new CPSM finds and eliminates more field stars than the classic CPSM. This situation is even more evident in the presence of background stars that generally have smaller proper motions. The new CPSM is able to retrieve more than 80\\% of all cluster members with a contamination around 20\\% of the total number of field stars at distances of 100~pc and 200~pc. At larger distances the efficiency of rejecting field stars decreases, but it is still higher than the efficiency of the classic CPSM. The definition of the $X^{2}$ function in the new CP method considers both the amplitude and direction of the stellar proper motion vector instead of only one directional component that removes the bias towards distant stars that is observed in the classic CPSM. Additional information (e.g. parallaxes and radial velocities) is required to eliminate the remaining field stars in the sample that were not rejected by the CPSM. The new CPSM is shown to work well when applied to the Hyades open cluster, and its results are agree well with previous determinations of the CP position. The individual parallaxes derived using our CP solution are fully consistent with the trigonometric parallaxes given in the \\rm{\\sc{Hipparcos}} catalog. Although our results agree well with data from both reductions, we find that our external precision is slightly better for HIP07. Our results are also in good agreement with the secular parallaxes found in the literature, which are more precise than the trigonometric parallaxes given in the \\rm{\\sc{Hipparcos}} catalog. The original implementation of the new CPSM was adapted to take the full covariance matrix of the \\rm{\\sc{Hipparcos}} catalog into account. The same procedure can also be applied to the future data of the \\textit{Gaia} mission, which will also be published with the full covariance matrix. The new CPSM will be used in forthcoming papers to investigate the kinematic properties of several young associations." }, "1112/1112.4154_arXiv.txt": { "abstract": "{We study the structure of neutron stars in $R+\\beta R^{\\mu \\nu} R_{\\mu \\nu}$ gravity model with perturbative method. We obtain mass--radius relations for six representative equations of state (EoSs). We find that, for $|\\beta| \\sim 10^{11}$ cm$^{2}$, the results differ substantially from the results of general relativity. Some of the soft EoSs that are excluded within the framework of general relativity can be reconciled for certain values of $\\beta$ of this order with the 2 solar mass neutron star recently observed. For values of $\\beta$ greater than a few $10^{11}$ cm$^2$ we find a new solution branch allowing highly massive neutron stars. By referring some recent observational constraints on the mass--radius relation we try to constrain the value of $\\beta$ for each EoS. The associated length scale $\\sqrt{\\beta}\\sim 10^6$ cm is of the order of the the typical radius of neutron stars implying that this is the smallest value we could find by using neutron stars as a probe. We thus conclude that the true value of $\\beta$ is most likely much smaller than $10^{11}$ cm$^{2}$.} ", "introduction": "Einstein's theory of gravity, general relativity, enjoyed impressive observational and experimental support in the past century. Main successes of this theory were explanations it brought to phenomena such as precession of the perihelion of Mercury, bending of light and the gravitational redshift of light near massive bodies. Although all of these impressive tests were done solely in the solar system, the standard model of cosmology assumes the validity of general relativity in all scales up to the very large scale structure of the universe. However, as the 20th century was ending, data from distant supernovae Ia \\cite{per99,rie98,rie04} were interpreted as evidence of late time acceleration in the expansion rate of the universe. If we continue to assume the validity of general relativity in all scales, then the best fit to observational data requires us to introduce a non--vanishing positive cosmological constant into the theory. This is the simplest way to follow without changing the basic paradigm. Yet there are several theoretical problems related with the existence of cosmological constant (see for example \\cite{Weinberg,Peebles,Nobbenhuis,Bousso}). One of the most important among these problems is the lack of a quantum theoretical method to calculate its inferred value from cosmological data \\cite{car01}. There are several proposals in order to avoid the problems of cosmological constant with alternative routes of explanations \\cite{uzan,tsu10}. These proposals can be collected into a few groups: to explain late time accelerated expansion one can either modify general relativity by modifying the Einstein--Hilbert action, or add new gravitational degrees of freedom other than the metric to the theory of gravity, or change how the matter fields and perhaps the cosmological constant gravitates \\cite{man05,clif11}. A modified gravity theory, which is proposed to solve late time cosmic acceleration problem, should also be able to pass several tests before it can be considered a viable theory of gravity. First of all, in the weak gravity regime, such a theory should be compatible with the solar system tests and table--top experiments. In cosmological scales, other than producing the late time accelerated expansion, it should be free of gravitational instabilities, and obey constraints of the standard model of cosmology. Such a theory is also expected to do well in strong gravity regime: for example it should have solutions for neutron stars with mass--radius relation inside the current observational bounds. There are alternatives to and generalizations of general relativity theory which have the same predictions in the weak-field regime as the general relativity, and also provide an explanation of the evolution of the universe in the large scale. Thus, the difference between general relativity and alternatives might become prominent in the strong-gravity regime \\cite{psa08}. One important family of modifications of Einstein--Hilbert action is the $f(R)$ theories of gravity (see reviews \\cite{Odintsov-rev,Capozziello-rev,Sotiriou-rev,deFelice-rev} and references therein). In such theories one uses a function of curvature scalar as the Lagrangian density. The $f(R)$ term must have a lower order expansion in Ricci scalar in order to include general relativity as, perhaps, a weak-field limit of it. Such models of gravity can be made to pass Solar System tests, explain the late-time accelerated expansion of the universe and also work well in the strong-field regime. In a previous work \\cite{ara11} a simple version of $f(R)$ gravity theory, in which $f(R) = R +\\alpha R^2$, is studied by two of the present authors. It is curious that predictions of such theories can be shown to be equivalent to scalar-tensor theories of gravity \\cite{mag94}, which have been analyzed throughly since the seminal paper of Brans and Dicke \\cite{bd}. However, if we modify the Einstein--Hilbert term by contractions of Ricci and Riemann tensors, $R_{\\mu \\nu }R^{\\mu \\nu }$ and $R_{\\mu \\nu \\rho \\sigma}R^{\\mu \\nu \\rho \\sigma}$, then we have an alternative gravity theory with independent predictions. Inspiration for such an alternative gravity theory may come from string theory. Absence of ghosts in low energy string theory in flat backgrounds requires the quadratic corrections to Einstein's gravity to be of the Gauss-Bonnet (GB) form \\cite{zwi85}: \\begin{equation} S=\\int d^{4}x\\sqrt{-g}\\left[ R+\\gamma ( R^2 -4 R_{\\mu \\nu }R^{\\mu \\nu } +R_{\\mu \\nu \\rho \\sigma}R^{\\mu \\nu \\rho \\sigma} )\\right]\\ . \\label{zwei} \\end{equation} However, the Gauss--Bonnet term will not contribute to the equations of motion, because it is equivalent to a total derivative in four dimensions. This means that variations of $R_{\\mu \\nu \\rho \\sigma}R^{\\mu \\nu \\rho \\sigma}$ can be expressed in terms of variations of $R^2$ and $R_{\\mu \\nu }R^{\\mu \\nu }$. Contraction of Riemann tensors, could have relevance only in the cases related to quantum gravity \\cite{psa09,par99}. Therefore for the classical physics applications one can take the action of our string-inspired gravity as \\begin{equation} S=\\int d^{4}x\\sqrt{-g}\\left[ R+\\alpha R^2 +\\beta R_{\\mu \\nu }R^{\\mu \\nu }\\right]\\ , \\label{rqugr} \\end{equation} where $\\alpha $ and $\\beta $ are free-parameters, taken independent of each other. However, special relations between $\\alpha $ and $\\beta $ exists and such cases correspond to some special theories: for example, $\\beta=-3\\alpha$ correspond to the Weyl tensor squared modification of general relativity \\cite{s08,fs09}, and $\\beta=-4\\alpha$ has unique energy properties \\cite{bay02}. As mentioned above, the absence of ghosts in the low energy string theory in flat backgrounds requires the quadratic corrections to Einstein's gravity to be of the form given in equation (\\ref{zwei}). This means that the alternative gravity theory defined by (\\ref{rqugr}) cannot be free of ghosts and this brings up the question of stability of the neutron star solutions discussed in this paper. However, all such modified gravity theories, bar the theory defined with (\\ref{zwei}), suffer from the same problem, the solution of which is beyond the scope of this paper. Our aim in this paper is to see the possibility of neutron star solutions in a specific modified gravity theory defined by (\\ref{rqugr}) with $\\alpha =0$, to analyze solutions with mass-radius relations obeying the observational constraints, and to see if there are problems in this gravity model other than the well known stability issue. When combined with the results of \\cite{ara11} this analysis will allow for a comparison of the effects of $R^2$ and $R_{\\mu \\nu }R^{\\mu \\nu }$ terms in (\\ref{rqugr}) on the structure of neutron stars. We further note that the mass dimensions of quadratic corrections to Einstein--Hilbert action in (\\ref{rqugr}) is order $[L]^{-4}$. It is possible to add terms with the same mass dimensions to the above action. In fact referring to the quantum gravity arguments in \\cite{par99}, we could also add the following dimension $[L]^{-4}$ terms to the action: \\begin{equation} \\square R\\quad \\mathrm{and}\\quad \\nabla ^{\\mu }\\nabla ^{\\nu }R_{\\mu \\nu }\\ . \\end{equation} However, as it is pointed out in \\cite{emi10} these terms do not contribute to the field equations and therefore in the context of the present paper they are also redundant. The effect of the value of $\\alpha $, while $\\beta =0$, on the mass-radius (M-R) relation of neutron stars has been studied, for a polytropic EoS in \\cite{coo10}, and for realistic EoS in \\cite{ara11}. The latter work constrained the value of $\\alpha $ to be $|\\alpha |\\lesssim 10^{10}$ cm$^{2}$. In an other study, Santos analyzes neutron stars in this model of gravity for a single EoS of ideal neutron gas \\cite{emi11} and for a specific choice of $\\beta=-2\\alpha$, $\\sqrt{\\alpha}=0.96$ km. In that work, the author shows, for this restricted choice of EoS and parameters, that stable configurations of neutron stars are possible even for arbitrarily large baryon numbers of the neutron star. Our approach in this work is different than \\cite{emi11} in two ways: (i) Rather than choosing a fixed specific value for $\\beta$, while $\\alpha =0$, we study its effect, as a free parameter, on the M-R relation and constrain its value by referring to observations, (ii) As the interaction between nucleons can not be neglected, the EoS of ideal neutron gas can not provide realistic M-R relations. We thus employ six different realistic EoSs corresponding to a variety of assumptions for the interaction between nucleons rather than the very restrictive EoS of ideal neutron gas. In the present work, we adopt an alternative theory of gravity, in which the Einstein-Hilbert action is modified with the term $R_{\\mu\\nu}R^{\\mu\\nu}$ only. In \\S 2, we assume a perturbative form of our alternative gravity model and obtain the field equations. The reason of perturbative approach is that the equations of motion derived from this alternative gravity theory are fourth order differential equations, and their treatment in four dimensions is problematic. To obtain the modified Tolman--Oppenheimer--Volkoff (TOV) equations from the field equations we also assumed perturbative forms of metric and hydrodynamical functions. In \\S 3, we solve the structure of neutron stars in this gravity model for six representative equations of state describing the dense matter of neutron stars. We plot the mass-radius relations of neutron stars for $\\beta$ changing in the range $\\sim \\pm 10^{11}$ cm$^2$. This way the value of the perturbative parameter $\\beta$ is constrained by the recent measurements of the mass-radius relation \\cite{oze10} and the observed 2 solar mass neutron star \\cite{dem10}. We identify that $\\beta \\sim \\pm 10^{11}$ cm$^2$ produces results that can have observational consequences. Lastly, the results of the numerical study and the significance of the scale of the perturbation parameter is discussed in the conclusions. ", "conclusions": "In this paper we studied the structure of neutron stars (NSs) in a generalized theory of gravity motivated by string theory. In the first section we discussed the motivations for modifying gravity. In the second section we derived the field equations of this gravity model from its action. In the third section we obtained the hydrostatic equilibrium equations in spherical symmetry from the field equations by using a perturbative approach in which general relativity (GR) stands for the zeroth order gravity model. In the fourth section we solved the hydrostatic equilibrium equations for NSs by using numerical methods. In order to solve the equations we have used realistic equations of state (EoSs) that describe the dense matter inside NSs and obtained the mass-radius (M-R) relations depending on $\\beta$, the free parameter of the generalized gravity model considered. These M-R relations are then compared with the recent observational measurements of mass and radius of NSs to constrain the value of $\\beta$. We have shown that observationally significant changes on the M-R relation are induced for $\\beta \\sim 10^{11}$ cm$^{2}$. An order of magnitude smaller values for $\\beta $ gives results that are not significantly different from what GR predicts. An order of magnitude greater values, on the other hand, leads to results that can not be associated with known properties of NSs. For such values, the perturbative approach breaks down too. For this selection of EoSs we see that none of them are consistent with the observations for $\\beta<-5\\times 10^{11}$ cm$^2$ and $\\beta>4\\times 10^{11}$ cm$^2$. The only exception is MPA1 which accepts large positive values like $\\beta_{11}\\sim 10$. We thus conclude from this analysis that $|\\beta |\\lesssim 5\\times 10^{11}$ cm$^{2}$ is an upper limit brought by observations of NSs \\cite{oze10,dem10}. We note that, for the the gravity model $R+\\alpha R^2$, the constraint on $\\alpha$ obtained by \\cite{ara11} by using NSs confronting the M-R relation with the same observations, is of the order $10^{10}$ cm$^2$, two orders of magnitude smaller than the constraint obtained on $\\beta$ in the present work, although both $\\alpha$ and $\\beta$ have the same dimensions ($[L]^2$). This indicates that the $R^2$ term in eq. (\\ref{rqugr}) is more effective on the structure of neutron stars than the $R_{\\mu \\nu }R^{\\mu \\nu }$ term. The last observation is very interesting from a theoretical point of view. As mentioned in the introduction, theory described by (\\ref{rqugr}) is equivalent to Weyl tensor squared modification of general relativity for $\\beta = -3\\alpha$ \\cite{s08,fs09}. That is, from the point of view of Weyl tensor squared modification, $\\alpha$ and $\\beta$ are in the same order. Our perturbative analysis in this paper also in some way forces us to make a similar statement: if $\\beta$ is taken in the same order as of $\\alpha$, then there would not be the problem of the break-down of the perturbative approach. What extra we learn from this analysis is that even though the $R^2$ and $R_{\\mu \\nu }R^{\\mu \\nu }$ terms seem to be on equal footing in the expansion of a Weyl tensor square term, \\begin{equation} \\label{Weyl} \\frac12 C_{\\mu \\nu \\rho \\sigma}C^{\\mu \\nu \\rho \\sigma} = -\\frac13 R^2 + R_{\\mu \\nu }R^{\\mu \\nu } + \\mathrm{topological\\, term}, \\end{equation} the contributions of them to neutron star physics are not equal. This is pointed out to us by the analysis in this paper and it is an important statement just by itself. It would be significant if this observation is confirmed and understood analytically. We plan to do this in a future publication. We find that, some of the EoSs, which do not give M-R relations consistent with the observations within the framework of GR, can be reconciled with these observations via the free parameter $\\beta $ in the generalized gravity model considered in this work. This then brings up the question of degeneracy between the EoSs and the free parameter $\\beta $. This degeneracy does not effect the constraint $|\\beta|\\lesssim 5\\times 10^{11}$ cm$^{2}$ which is a bound for all EoSs we considered except for MPA1. We finally comment that the constraint we obtained is actually the strongest constraint we could obtain by using NSs as the experimental apparatus. Another estimate of the nominal value $\\beta _{0}$ of the previous section is as follows: Typical radius of a NS is $R_{\\ast}\\sim 10$ km, the only length scale in the system. This corresponds to an estimate of curvature $R^{\\mu \\nu }R_{\\mu \\nu }\\sim R_{\\ast}^{-2}\\sim 10^{-12}$ cm$^{-2}$ and so the new perturbative term $\\beta R^{\\mu \\nu }R_{\\mu \\nu }$ will become of order $R_{\\ast}$ and lead to variations on the structure of neutron stars for $\\beta \\sim R_{\\ast}^{2}\\sim 10^{12}$ cm$^{2}$. As we obtain such variations in this limit, the value that we should obtain by using NSs, we infer that the actual value of $\\beta $ is likely much smaller than this. As we mentioned before, deviations from GR are not significant for NSs for values much less than this value." }, "1112/1112.4648_arXiv.txt": { "abstract": "{} {We intend to compile a new galaxy group and cluster sample of the latest available SDSS data, adding several parameter for the purpose of studying the supercluster network, galaxy and group evolution, and their connection to the surrounding environment.} {We used a modified friends-of-friends (FoF) method with a variable linking length in the transverse and radial directions to eliminate selection effects and to find reliably as many groups as possible. Using the galaxies as a basis, we calculated the luminosity density field.} {We create a new catalogue of groups and clusters for the SDSS data release~8 sample. We find and add environmental parameters to our catalogue, together with other galaxy parameters (e.g., morphology), missing from our previous catalogues. We take into account various selection effects caused by a magnitude limited galaxy sample. Our final sample contains 576493 galaxies and 77858 groups. The group catalogue is available at \\texttt{http://www.aai.ee/$\\sim$elmo/dr8groups/} and from the Strasbourg Astronomical Data Center (CDS).} {} ", "introduction": "Observations of the local Universe have shown that basically all galaxies are located in groups -- it is their natural environment. Groups and clusters of galaxies form the basic building blocks of the Universe. Therefore, it is essential to extract groups of galaxies from galaxy surveys, and their study can provide new understanding of the evolution of galaxies, of the large-scale structure, and of the underlying cosmological model. In our previous papers \\citep{Tago:08,Tago:10} we have extracted groups from the SDSS DR5 and DR7 samples, respectively. In these papers we have given an extensive review of papers dedicated to group search methods and of the published group catalogues. In this introduction we present only a short review of the studies of galaxy groups. During the last decade, several group catalogues that use spectroscopic redshifts have been published, either based on the 2dFGRS \\citep{Eke:04, Yang:05, Tago:06}, or on earlier releases of the SDSS \\citep{Einasto:03, Merchan:05, Zandivarez:06, Berlind:06, Berlind:09, Yang:07,Koester:07}. However, similar algorithms used to compile these catalogues have yielded groups of galaxies with rather different statistical properties. Several authors have recently compiled group catalogues up to the redshift 0.6: GAMA (Galaxy And Mass Assembly) by \\citet{Robotham:11} is a galaxy group catalogue based on the SDSS target catalogue; \\citet{Farrens:11} derived a catalogue, based on the LRGs and QSOs in the 2dF-SDSS surveys. Using photometric redshifts several group/cluster catalogues have been compiled \\citep[e.g.][]{Gal:09,Szabo:11}. \\citet{Hao:10} applied a Gaussian mixture BCG algorithm to the SDSS DR7 data and assembled a photometric group catalogue up to the redshift 0.55. Using spectroscopic redshifts, \\citet{Knobel:09} compiled the deepest group catalogue so far, reaching up to the redshift 1 in the zCOSMOS field. In addition, catalogues of rich galaxy clusters have been created by \\citet{Miller:05}, \\citet{Aguerri:07}, and \\citet{Popesso:07}. We discuss and compare some of these catalogues in a separate paper. The papers dedicated to group and cluster search use a wide range of both sample selection methods as well as cluster search methods and parameters. The choice of the methods and parameters depends on the goal of the catalogue. For example, while \\citet{Weinmann:06} searched for compact groups in the SDSS DR2 sample, applying strict criteria in the friend-of-friend (FoF) method, then \\citet{Berlind:06} applied the FoF method to the volume-limited samples of the SDSS with the goal to measure the group multiplicity function and to constrain dark matter haloes. Hence, the parameters of the algorithm depend on the goal of the study. Our goal is to generate an up-to-date catalogue of groups and clusters for large-scale structure studies. The catalogue is based on the SDSS data release~8 (DR8). Since the SDSS spectroscopic main sample basically did not change from DR7 to DR8, we use exactly the same group finding algorithm and parameters as described in \\citet{Tago:10}, yielding a group sample with similar properties. The photometry of the galaxies in DR8 has been reprocessed, yielding more accurate luminosities. This is important for detailed photometric galaxy modelling (Tempel et al. in prep). Compared to our previous catalogues, we added several additional descriptors, including environmental parameters (both local and global) and morphology. For the DR7, such data have been already used in several papers: the environment and morphology have been used to study the environmental effects on galaxy evolution \\citep{Tempel:11a}; global environments have been used to extract superclusters from the cosmic network \\citep{Liivamagi:10}. The present catalogue, based on the DR8, has already been used to compare the local and global environments of galaxies (Lietzen et al. in prep), to study the structure of rich groups \\citep{Einasto:12}, and to study the photometric structure of galaxies (Tempel et al. in prep). The paper is organised as follows. The data used are described in Sect.~\\ref{sect:data}. Section~\\ref{sect:cat} gives a brief overview of the group finding algorithm used, together with a short comparison with our DR7 catalogue. In Sect.~\\ref{sect:dens} we describe our method to calculate the luminosity density field. In Sect.~\\ref{sect:param} we describe the additional galaxy and group parameters. All the parameters in the resulting catalogue are described in Appendix~\\ref{app:cat}. The catalogue can be downloaded from \\texttt{http://www.aai.ee/$\\sim$elmo/dr8groups/} or from the Strasbourg Astronomical Data Center (CDS)\\footnote{Galaxy and group/cluster tables will be available at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/}. Throughout this paper we assume the following cosmology: the Hubble constant $H_0 = 100\\,h\\,\\mathrm{km\\,s^{-1}Mpc^{-1}}$, the matter density $\\Omega_\\mathrm{m}=0.27$ and the dark energy density $\\Omega_\\Lambda=0.73$. ", "conclusions": "We have constructed a group catalogue for the SDSS DR8 sample, following the same procedure that we used for the DR7 sample \\citep{Tago:10}. Since the spectroscopic data is the same in the DR7 and DR8, the new catalogue is similar to the previous DR7 group catalogue. The improvements are in the area of initial galaxy selections and from the SDSS side, the photometry of galaxies. In addition to the properties, presented in our previous catalogue, we added some new qualitative information. Most importantly, we tracked the missing galaxies in the SDSS (due to fibre collisions) and calculated the environmental density parameters for each galaxy and group. We also added our galaxy morphology as derived in \\citet{Tempel:11a}. In the study of galaxy groups, the most important problem at present is the dynamical status of groups of galaxies. Recently \\citet{Plionis:06} and \\citet{Tovmassian:09} studied shapes and virial properties of groups and found a strong dependence on richness and concluded that groups are not in dynamical equilibrium but rather are at various stages of their virialisation process. The dynamical status of groups is also characterised by the existence of subgroups and substructures seen in many studies \\citep{Burgett:04, Coziol:09, Einasto:10}. These studies also support the non-virialised nature of groups of galaxies. With both observational and simulated data \\citet{Niemi:07} showed that about 20\\% of nearby groups are not bound and are groups merely in a visual sense. The controversial results obtained by various authors may be an indication that our present knowledge of groups of galaxies is poor. An optimistic viewpoint is that the study of a broad and inhomogeneous class of galaxy systems -- groups of galaxies -- will help step by step to solve important problems of galaxy formation and evolution, and of the evolution of the large scale structure." }, "1112/1112.4012_arXiv.txt": { "abstract": "We have studied the Residual Blaze Functions (RBF) resulting from division of individual echelle orders by extracted flat-field in spectra obtained by slit-fed OES spectrograph of 2m telescope of Ond\\v{r}ejov observatory, Czech Republic. We have eliminated the dependence on target and observation conditions by semiautomatic fitting of global response function, thus getting the instrument-only dependent part, which may be easily incorporated into data reduction pipeline. The improvement of reliability of estimation of continuum on spectra of targets with wide and shallow lines is noticeable and the merging of all orders into the one long spectrum gives much more reliable results. ", "introduction": "The requirement of a modern astrophysical analysis, to study the wide range of the object spectrum in the greatest details, or to observe its variability in many spectral lines at the same time, lead naturally to the usage of cross-dispersed echelle spectrographs. Although the reduction of the echelle raw frames is quite straightforward and a number of automatic reduction pipelines has been in use for many instruments, there remains still unsolved the problem of the reliable normalization of individual extracted spectral orders before merging into one long continuous spectrum. Due to its inherent nature the echelle blaze function has to be removed from all orders with the immense precision, otherwise the strong ripple patterns appear on the merged spectrum. They make difficult the fitting of the continuum an hence influence the reliability of the measurement of equivalent widths of wide shallow lines, the placement of wings of hydrogen lines of hot rapidly rotating stars, or the studies of diffuse interstellar bands. ", "conclusions": "We have shown a one possible way how to tackle the problem of precise unblazing of echelle spectra using the separation of correction function to instrument-only dependent part (IRBF) changing on the scale of one echelle order and observation-dependent part (GSF) smoothly changing over the whole observed spectrum. Even if this is not the final solution, and still some quickly changing variations remain, the resulting corrected echelle orders may be merged in the one long spectrum much more easily avoiding strong ripples. If the estimate of GSF is known for given observation, the spectra may be normalized to continuum almost automatically. Our method is just a first simple approximation of the general correction procedure, it may be instrument dependent (better suited to slit-fed spectrographs) and so the future investigation of the problem is still highly desirable." }, "1112/1112.5201.txt": { "abstract": "%\\begin{abstract} In this work we present a model of dark matter based on scalar-tensor theory of gravity. With this scalar field dark matter model we study the non-linear evolution of the large scale structures in the universe. %using a dark matter %model steaming from a scalar-tensor theory of gravitation. The equations that govern the evolution of the scale factor of the universe are derived together with %and also the appropriate Newtonian equations to follow the non-linear evolution of the structures. Results are given in terms of the power spectrum that gives quantitative information on the large-scale structure formation. The initial conditions we have used are consistent with the so called concordance $\\Lambda$CDM model. %We also show how to estimate %the scalar field dark matter parameters using a sample of observed rotation curves. % ", "introduction": "\\label{sec:Intro} %The Big Bang theory The standard model of cosmology is supported by three main astronomical observations: the surveys of supernovae Ia, the cosmic microwave background radiation (CMB), and the primordial nucleosynthesis. These observations together with other % %M modern cosmological observations, % like galaxies surveys (SDSS, 2dF), galaxy rotation curves, the Bullet Cluster observation, studies of clusters of galaxies, % establish that the universe behaves as dominated by dark matter (DM) and dark energy. However, the direct evidence for the existence of these invisible components remains lacking. Several theories that would modify our understanding of gravity have been proposed in order to explain the large scale structure formation in the universe and the galactic dynamics. The best model we have to explain the observations is the $\\Lambda$CDM model, i.e., the model of cold dark matter (CDM) --non-relativistic particles %at the epoch of last scattering of unknown origin-- with cosmological constant ($\\Lambda $), in particular, this model explains very well the universe on scales of galaxy clusters and up \\citep{Breton2004}. The $\\Lambda$CDM model has become the theoretical paradigm leading the models of the universe to explain the large scale structure (LSS) formation and several other observations. Where ``large'' means scales larger than $1$ Mpc --about the size of the group of galaxies that our Milky Way belongs. Together with the cosmic inflation theory, this model makes a clear prediction about the necessary initial conditions that the universe has to have in order to have the structures we observe and that those structures build hierarchically due to a gravitational instability. % One of its main predictions is that the density profile of galaxies, clusters of galaxies, and so on, is of the form \\citep{Navarro96,Navarro97}, \\begin{displaymath} \\rho_{NFW}(r) = \\frac{\\rho_0}{(r/r_0)(1+r/r_0)^2}, \\end{displaymath} a density profile known as Navarro-Frenk-White profile (NFW). Parameters $\\rho_0$ and $r_0$ must be fitted, for example, using rotation curves of galaxies. The $\\Lambda$CDM model and its success in explaining several observations --this is why this model is also known as the concordance model--", "conclusions": "\\label{sec:Conclusions} We have used a general, static STT, that is compatible with local observations by the appropriate definition of the background field constant, i. e., $\\langle \\phi\\rangle = (1+\\alpha)/G_N$, to study the LSS formation process. The initial condition for the several cases (different values of parameter $\\alpha$) was built in such a way that the geometry of the model universe were flat. Quantitatively, this demands that our models have $\\Omega/(1+\\alpha)=1$ and this changes the amount of dark matter and energy of the models in order to have a flat cosmology. Using the resulting modified dynamical equations, we have studied the LSS formation process of a $\\Lambda$CDM universe. We varied the amplitude and sign of the strength of the SF (parameter $\\alpha$) in the interval $(-1,1)$ and performed several 3D-simulations with the same initial conditions. From our simulations we have found that the inclusion of the SF changes the local dynamical properties of the most massive groups, however, the overall structure is very similar, as can be seen in Fig. \\ref{fig:LSSF}. The general gravitational effect is that the interaction between dark matter particles given by the potential $\\Phi_N$ (see equation (\\ref{eq:pointMassPotential})) changes by a factor $F_{SF}$, equation (\\ref{eq:FSFFactor}), in comparison with the purely Newtonian case. Thus, for $\\alpha > 0$ the growth of structures speeds up in comparison with the Newtonian case (without SF). For the $\\alpha <0$ case the effect is to diminish the formation of structures. It is important to note that particles in our model are gravitating particles and that the SF acts as a mechanism that modifies gravity. The effective mass of the SF ($m_{SF}=1/\\lambda$) only sets an interaction length scale for the Yukawa term. In this work we only varied the amplitude of the SF --parameter $\\alpha$-- leaving the scale length, $\\lambda$, of the SF unchanged. However, in other studies we have done \\citep{mar2009b} we have found that increasing $\\lambda$ enhances the structure formation process for $\\alpha$ positive, and decreasing $\\lambda$ makes the structures grow at a slower rate. We have computed the mass power spectrum in order to study the LSS formation process. The theoretical scheme we have used is compatible with local observations because we have defined the background field constant $<\\phi> = G_{N}^{-1} (1+\\alpha)$ or equivalently that the local gravitational constant is given by $ (1 + \\alpha) \\langle \\phi \\rangle^{-1} $, instead of being given by $\\langle \\phi \\rangle^{-1}$. % A direct consequence of the approach is that the amount of matter (energy) has to be increased for positive values of $\\alpha$ and diminished for negative values of $\\alpha$ with respect to the standard $\\Lambda$CDM model in order to have a flat cosmological model. % Quantitatively, our model demands to have $\\Omega/ (1+\\alpha) =1$ and this changes the amount of dark matter and energy of the model for a flat cosmological model, as assumed. The general gravitational effect is that the interaction including the SF changes by a factor $F_{SF}(r,\\alpha,\\lambda) \\approx 1+\\alpha \\, \\left( 1+\\frac{r}{\\lambda} \\right)$ for $r<\\lambda$ in comparison with the Newtonian case. Thus, for $\\alpha >0$ the growth of structures speeds up in comparison with the Newtonian case. For the $\\alpha <0$ case the effect is to diminish the formation of structures. For $r> \\lambda$ the dynamics is essentially Newtonian. Comparison with the power spectrums from galaxies in the SDSS catalog and that inferred from Lyman-$\\alpha$ forest observations tell us that $\\Lambda$CDM predicts more structure formation in the regime of smaller scales ($k>0.4$ $h/$Mpc). Whereas, the model with SF with $\\alpha=-1/2$ and $\\lambda=5$ Mpc$/h$, follows the general trend of the observed power spectrum. In this way we are able to construct a model that predicts the observed structure formation in the regime of small scales, with lesser number of halo satellites than the $\\Lambda$CDM model. % %\\begin{acknowledgement} %This work received financial support by %CONACYT grant number \\ CB-2007-01-84133. %\\end{acknowledgement} %%%%%%%%%%%%%%%% REFERENCIAS %%%%%%%%%%%%%%%% %" }, "1112/1112.1073.txt": { "abstract": "The CO-\\htwo \\ conversion factor in galaxies is typically described as bimodal: one value for discs and quiescent regions, and another (lower) value for mergers and starbursts. In this proceeding, I will describe both empirical observational evidence that the conversion factor varies with physical environment, as well as a theoretical model which aims to understand the physical processes which drive these variations. I present a functional form for \\xco \\ which can be applied to observations ranging in scale from $\\sim 70$ pc to galaxy-wide scales, and show the consequences of the application of this model to the Kennicutt-Schmidt star formation law. ", "introduction": "Deriving an \\htwo \\ gas mass from a giant molecular cloud (GMC) or galaxy full of clouds involves the usage of tracer molecules such as $^{12}$CO (hereafter, CO). This is because \\htwo \\ has no permanent dipole moment, and its first quadrupole moment lies $\\sim 500$ K above ground, significantly warmer than the typical ISM temperature in a GMC. Converting from a CO emission line strength to an underlying \\htwo \\ gas mass involves the usage of a CO-\\htwo \\ conversion factor. The CO-\\htwo \\ conversion factor goes by two names in the literature: \\xco\\footnote{\\xco \\ is also referred to as the $X$-factor. I will use \\xco \\ or the $X$-factor arbitrarily in this proceeding.} \\ and \\alphaco. The former has units of \\xcounits, while the latter \\msun ${\\rm pc}^{-2}$(K-\\kmsend)$^{-1}$. The two are equivalent, and simply related via \\xco = $6.3 \\times 10^{19} \\alpha_{\\rm CO}$ in the aforementioned units. Since both \\xco \\ and \\alphaco \\ are used in the literature, I will present the results of this proceeding in terms of both units. Observationally determining an $X$-factor requires deriving an independent measurement of \\htwo \\ gas mass, and comparing this to an observed CO line flux. Via a variety of methods \\citep[some of these are summarised in the seminal conference proceeding by][]{sol91}, it has been found that the conversion factor is roughly constant in the Galaxy and nearby galaxies (when the metallicity is of order solar) with value \\xco= $2-4 \\times 10^{20} \\xcounits$, or \\alphaco $\\approx 3-6$ \\alphacounits \\ \\citep[for a reference list of these observations, please see the introduction of][]{nar11c}. However, despite the seeming constancy of the $X$-factor in relatively ``normal'' GMCs, observations of nearby ultraluminous infrared galaxies (ULIRGs) show that the usage of a Milky Way \\xco \\ in these environments would cause the inferred \\htwo \\ gas mass to exceed the measured dynamical mass \\citep{dow98}. Hence, in ULIRGs, the $X$-factor must be lower than within the Galaxy. In Figure~\\ref{figure:observational_xco}, we show the distribution of \\xco \\ and \\alphaco \\ values from the \\citet{dow98} survey, as well as a shaded region for the typical range of Galactic values. There is a dispersion amongst the ULIRG $X$-factors, though on average they are lower than the Galactic mean. Despite the observed dispersion in ULIRG conversion factors, the fact that they are lower on average than the Milky Way value has led the community to largely adopt a bimodal picture of \\xco: a value of $\\alphaco \\approx 4$ for disc galaxies, and $\\alphaco \\approx 0.8$ for starbursts and mergers. The ramifications of a bimodal form of the CO-\\htwo \\ conversion factor are significant. As an example, \\citet{dad10b} and \\citet{gen10} demonstrated the effects of assuming a bimodal \\xco \\ on the Kennicutt-Schmidt (KS) star formation rate-gas density relation. While in $\\Sigma_{\\rm SFR}-\\Sigma_{\\rm CO}$ space, galaxies at low and high-\\z \\ lie on an arguably unimodal relation, the introduction of a bimodal \\xco \\ to convert the CO luminosity to an \\htwo \\ gas mass causes the KS relation to become bimodal. That is, mergers and discs lie on different track on the KS relation when utilising a different \\xco \\ for each. This is shown explicitly in Figure~\\ref{figure:ks}, where I have compiled the galaxies presented in \\citet{dad10b} and \\citet{gen10}, and plotted (in the left panel) the $\\Sigma_{\\rm SFR}-\\Sigma_{\\rm CO}$ relation, and (in the middle panel), the $\\Sigma_{\\rm SFR}-\\Sigma_{\\rm H2}$ relation which utilises a bimodal $X$-factor. The mergers (squares) and discs (circles) occupy different tracks. This has given rise to a terminology in the literature, as well as at conferences (including this one!), that mergers and discs have different ``modes'' of star formation. Mergers, according to this relation, form stars more efficiently (i.e. needing less gas to sustain a given SFR) than discs. We will neglect the right panel of Figure~\\ref{figure:ks} for now, and will return to it shortly. ", "conclusions": "The principal result from this work is that the CO-\\htwo \\ conversion factor in galaxies varies with the physical conditions in the galaxy. It is not bimodal with the global morphology of the galaxy. Rather, the physical environment sets the CO-\\htwo \\ conversion factor. The dominant drivers in the conversion factor are the gas temperature, velocity dispersion and metallicity. While the former two are difficult to observe, they can be parameterised by the observable CO intensity. With this, the CO-\\htwo \\ conversion factor can be described as function of CO intensity and gas phase metallicity. We show that this function can be derived both from empirical observational results, as well as numerical modeling. %\\begin{figure} % \\centering % \\includegraphics[angle=90,scale=0.4]{observational_xco.ps} %\\caption{\\label{figure:observational_xco}} %\\end{figure} %%%%%%%" }, "1112/1112.1637_arXiv.txt": { "abstract": "We outline the scientific motivation behind a search for gravitational waves associated with short gamma ray bursts detected by the InterPlanetary Network (IPN) during LIGO's fifth science run and Virgo's first science run. The IPN localisation of short gamma ray bursts is limited to extended error boxes of different shapes and sizes and a search on these error boxes poses a series of challenges for data analysis. We will discuss these challenges and outline the methods to optimise the search over these error boxes. ", "introduction": "Short hard gamma ray bursts (short GRBs) are believed to be produced by mergers of either two neutron stars or a neutron star and a stellar mass black hole \\cite{Nakar:2007, ShibTan06, Berger:2010qx}. These events are ideal sources for strong gravitational wave emission \\cite{ACST94, Kiuchi:2010ze}. If an observation of both gamma rays and gravitational waves (GW) originating from the same event could be achieved, it will increase confidence and allow for better science output. It is thus of great importance to constantly monitor and record GRBs to allow a GW search to be performed around the burst times. Systematic analyses of GW data around short GRB times have been done in the past and the most recent publications from the LIGO-Virgo group contain results from the Swift-observed GRBs during LIGO's fifth science run (S5) and Virgo's first science run (VSR1) \\cite{Abadie:2010uf, Collaboration:2009kk}. This paper reports the methodology and motivation for a proposed GW search around the times of 20 additional GRB during S5/VSR1. These bursts were observed by the InterPlanetary Network (IPN) \\cite{Hurley:2002wv, HurleyHTML}, a group of satellites orbiting the Earth and Mars and operating, among other equipment on board, gamma ray detectors. These bursts were detected between 2006 and 2007 and have been localized, in both time and sky location, such that a targeted GW search is possible. The InterPlanetary Network \\cite{Hurley:2002wv, Hurley:1999ym} employs several space missions and synthesizes data obtained from the detection of the same burst by different spacecraft equipped with gamma ray detectors. The IPN has been operating for three successive generations; presently the third IPN (IPN3) began its operation in November 1990. Currently the spacecraft gathering data are Konus-WIND, Suzaku, INTEGRAL, RHESSI, Swift, Fermi/GBM (in Earth orbit), MESSENGER (in Mercury orbit) and Mars Odyssey (in Mars orbit) \\cite{HurleyHTML}. When the duty cycles and effective fields of view of all the missions in the network are considered, the IPN is an all-time, isotropic GRB monitor. In this paper, we first review the motivation for performing a joint GW-GRB search as an overview and then provide details of the IPN short GRBs that were identified during the LIGO-Virgo S5-VSR1 science runs. We discuss the necessary changes to the current ongoing analysis (for the S6/VSR2-3 science runs Swift and Fermi-observed GRBs, \\cite{lvc:s6grb, Harry:2010fr}) that are needed to implement a search on GRBs identified by the IPN network which may be less well localized in both sky position and time than the corresponding bursts identified by the Swift satellite and used in previous analyses \\cite{Abadie:2010uf, Collaboration:2009kk}. ", "conclusions": "We have presented the methodology for a search for gravitational waves around the times of short GRBs, detected by IPN during S5 and VSR1. This search has all the needed tools and will commence in a short time. The work in this paper was presented in the form of a poster at the 10th Amaldi Conference for Gravitational Waves organised in Cardiff, UK in July 2011. These types of searches are very promising for the future detection of gravitational waves and combined with the prospect of detecting other electromagnetic counterparts from GRBs, e.g. radio or optical pre- or afterglows (summarised in \\cite{Predoi:2009af, Coward:2011yr}), may open the doors for true multi-messenger astronomy with gravitational waves." }, "1112/1112.2772_arXiv.txt": { "abstract": "{ The stellar astronomy has always been considered the fundamental source of knowledge about the basic building blocks of the universe --- the stars. It has proved correctness of many physical theories --- like e.g. the idea of nuclear fusion in stellar cores, the exchange of mass in interacting binaries or models of stellar evolution towards white dwarfs or neutron stars. Despite its well acknowledged importance it seems to be loosing its interestingness for students, for telescope allocation committees at large observatories, as well as for granting agencies. In the domain of big telescopes it has been gradually overtaken by the extra-galactic research and cosmology, surviving however at smaller observatories and among most advanced amateur astronomers. We try to analyse the main obstacles lowering the efficiency of research in contemporary stellar astronomy. We will shortly tackle several problems induced by paradigmatic changes in handling the extraordinary amount of data provided by current instruments as well as by introduction of economical criteria and factory-like management into the modern astronomy. Finally we speculate the reasons of a marginal role of Virtual observatory in contemporary stellar research and give some ideas of possible improvements. } \\FullConference{Accelerating the Rate of Astronomical Discovery, sps5\\\\ August 11-14, 2009\\\\ Rio de Janeiro, Brazil} \\begin{document} ", "introduction": "The technological potential of the current astronomical research is enormous. The opening of almost all windows of electromagnetic spectrum facilitated by advances in detector technology together with specialized astronomical satellites as well as 8 to 10m class telescopes harnessed by extremely sensitive instrumentation provides the astronomical community with overwhelmingly massive amount of data about the very deep structure of our universe as well as detailed multi-wavelength information about the billions of even very distant objects. The most powerful GRIDs of computers as well as clever infrastructure of Virtual Observatory are supposed to easily handle the data avalanche and so amount of principal astronomical discoveries should follow the technological Moore's law. But is it really so ? Unfortunately, there is a danger of loosing the important dimension --- the time --- from observations of the most powerful telescopes and their archives. The astronomy of past was based on long-term monitoring of selected objects and most physical analysis was concerning their time-dependent nature (e.g. the search of periods, phase dependent changes of line profiles or changes of intensity or spectra during the outbursts of novae or cataclysmic variables). The largest amount of data coming from large facilities is currently only some kind of snapshots of the Universe, being forced by the way the telescope time is allocated and by pressure on rapid publication of results. All of this may lead to the paradoxical situation when the next generation of astronomers will have from the then available data the feeling of the universe like something static and unchangeable as it was in the ancient times before the Galileo's invention of telescope. ", "conclusions": "Stellar astronomy seems to be now in stagnation period. The interest of professional astronomers is turned towards very distant extragalactic objects. The still existing professional stellar astronomy has been driven towards producing only snapshots than systematic monitoring. The stars are interesting for many astronomers only thanks to their planets, whereas most of the physical problems of the stellar structure and stellar evolution as well as nature of some phenomena remain still unsolved. Unfortunately, the stellar research seems not to be very appealing for students either, despite the great importance of stellar astronomy for understanding of galactic structure and evolution. On the other hand there is a lot of enthusiastic people among amateur astronomers, who continue to monitor the stellar variability on the long time scales and report outburst of cataclysmic variables, but they can hardly take spectra of most interesting stars, mainly due to their limited capabilities. The solution of the problems described might be the Virtual observatory, allowing the integration of dispersed observations from different (even small) teams. The VO principles might attract the number of young people fascinated by new software technologies. Unfortunately, there is a lack of stellar spectra in VO archives as well as little support of specific techniques important in stellar astronomy in VO tools with respect to existing legacy applications. There is also very little VO awareness among stellar astronomers and their students. Thus the possible solution of the current stellar crisis could be expressed by several goals: \\begin{itemize} \\item Convince people to put more stellar spectra into VO archives \\item Circumvent the data jealousy by convincing people about advantage of openness \\item Get resources needed for development of new more versatile analysis tools with VO interface (or for adding the VO interface to legacy applications) \\item Rise the level of VO awareness by systematic VO education or rather \"VO evangelization\" \\end{itemize} We can hopefully still preserve the image of the dynamic eternally changing universe for our grandsons instead of hinting them the idea of the Immutable Heavens again." }, "1112/1112.0777_arXiv.txt": { "abstract": "We report the discovery of two consecutive supernovae (SNe), \\object{2010cu} and \\object{2011hi}, located at 0.37\\arcsec\\ (180 pc) and 0.79\\arcsec\\ (380 pc) projected distance respectively from the centre of the \\textit{K}-band nucleus of the luminous infrared galaxy \\object{IC 883}. The SNe were discovered in an ongoing near-infrared \\textit{K}-band search for core-collapse SNe in such galaxies using the ALTAIR/NIRI adaptive optics system with laser guide star at the Gemini-North Telescope. These are thus the closest SNe yet discovered to a LIRG nucleus in optical or near-infrared wavelengths. The near-infrared light curves and colours of both SNe are consistent with core-collapse events. Both SNe seem to suffer from relatively low host galaxy extinction suggesting that regardless of their low projected galactocentric distances, they are not deeply buried in the nuclear regions of the host galaxy. ", "introduction": "Luminous ($L_{{\\rm IR}} > 10^{11} L_{\\odot}$) and ultraluminous ($L_{{\\rm IR}} > 10^{12} L_{\\odot}$) infrared (IR) galaxies (LIRGs and ULIRGs, respectively), have high star formation (SF) rates. The fraction of SF in U/LIRGs in the local universe is small compared to that in normal spiral galaxies; however, at high redshift they become the dominant sources of SF \\citep[e.g.,][]{magnelli11}. Stars more massive than $\\sim8M_{\\sun}$ explode as core-collapse supernovae (CCSNe), and due to their short life cycles they are a very useful tool for tracing ongoing SF rates, independent of the conventional way of using the galaxy IR luminosity. Recently, \\citet{anderson11} have presented evidence that the SN population in the interacting LIRG system \\object{Arp 299} differs from those observed in normal spiral galaxies with the Ib and IIb SNe being more numerous and centrally concentrated than more common Type II SNe. Until now, this kind of study has been impossible due to insufficient numbers of SNe discovered in U/LIRGs, providing a strong motivation for SN searches in galaxies with high SF rates. The extinction-free searches at radio wavelengths using very long baseline interferometry (VLBI) have been successful in discovering SNe and SN remnants in U/LIRGs, e.g., in \\object{Arp 220} \\citep[][]{parra07, batejat11} and \\object{Arp 299} \\citep{perez-torres09, ulvestad09}. The reason for the low optical discovery rate of CCSNe in U/LIRGs is the high dust extinction combined with spatial concentration of SF in the crowded nuclear regions within the central few hundreds of parsecs \\citep[see e.g.,][]{soifer01}. Therefore imaging in the near-IR where the extinction is strongly reduced, and high spatial resolution is achievable with ground-based adaptive optics (AO) or space-based imaging is required. Recently such methods have been shown to be very efficient with the discovery of several SNe in the LIRG nuclear/circumnuclear regions. \\object{SN 2004ip} was discovered with the Very Large Telescope using the natural guide star NACO AO system \\citep{mattila07}. \\object{SNe 2008cs} and \\object{2004iq} reported in \\citet{kankare08} were discovered using the Gemini-North Telescope with the laser guide star (LGS) assisted ALTAIR AO system and the \\textit{Hubble Space Telescope} (\\textit{HST}) NICMOS archive data, respectively. In this Letter, we report the consecutive discovery of two SNe in the same host galaxy with the Gemini-North Telescope\\footnote{Based on observations obtained at the Gemini Observatory, 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 Science and Technology Facilities Council (United Kingdom), the National Research Council (Canada), CONICYT (Chile), the Australian Research Council (Australia), Minist\\'{e}rio da Ci\\^{e}ncia e Tecnologia (Brazil) and Ministerio de Ciencia, Tecnolog\\'{i}a e Innovaci\\'{o}n Productiva (Argentina).}, \\object{SN 2010cu} and \\object{SN 2011hi}. \\citet{sanders03} report the host LIRG \\object{IC 883} (also known as e.g., \\object{UGC 8387}, \\object{Arp 193}, and \\object{IRAS 13183+3423}) to have an IR luminosity of $L_{IR} = L[8-1000 \\mu m] = 4.7 \\times 10^{11} L_{\\odot}$ and a distance of 100 Mpc (H$_{0}$ = 75 km~s$^{-1}$Mpc$^{-1}$) giving a projected linear distance scale $1\\arcsec = 485$~pc. The IR luminosity of \\object{IC 883} indicates a CCSN rate of $\\sim1.3$ yr$^{-1}$ derived with the empirical relation of \\citet{mattila01}, consistent with the discovery of two SNe within two years. In \\citet{romero-canizales11b} we report the details of our radio follow-up of \\object{SNe 2010cu} and \\object{2011hi} and investigate the innermost nuclear regions of \\object{IC 883} through high-angular resolution, high-sensitivity radio observations. ", "conclusions": "The adaptive optics assisted discovery of two consecutive SNe located at $<400$~pc from the LIRG nucleus demonstrates the importance of high spatial resolution in searches for CCSNe in highly crowded starburst regions in LIRGs. We note that of the five SNe discovered using high resolution near-IR observations in our sample of LIRGs (see Sect.~1), four are within 1~kpc from the host nucleus. Based on the near-IR light curve fitting to templates we find \\object{SN 2010cu} to be consistent with either a Type IIP or a near-IR bright Type IIn/L SN with a host galaxy extinction of $A_{V}\\lesssim1$~mag. Taking into account also the extinction map \\object{SN 2011hi} is found to be most consistent with a Type IIn/L SN during the near-IR excess phase with host galaxy extinction of $A_{V}\\sim0$~mag. A Type Ib/c classification was found unlikely for both the SNe. Future near-IR and radio follow-up observations of \\object{SN 2011hi} will be able to better constrain the nature and line-of-sight extinction of this potentially very interesting event The surprisingly low derived line-of-sight extinctions appear to be in conflict with the expectation of the occurrence of highly obscured CCSNe in the nuclear region of LIRGs. For example, in \\citet{kankare08} $A_{V}=16$~mag of host galaxy extinction was derived with near-IR template light curve fitting for \\object{SN 2008cs}, located at a projected distance of 1.5~kpc from the host LIRG nucleus. Based on this it seems more likely that \\object{SNe 2010cu} and \\object{2011hi} are located in the outer parts of the LIRG system which we see projected close to the LIRG nucleus. Additionally the majority of LIRGs at the high IR luminosity end are complex interacting systems and their dust extinction distributions can be very complex compared to normal spiral galaxies with symmetric exponential disks, see e.g., \\citet{alonso-herrero06} and \\citet{vaisanen08}. Despite the low extinctions such SNe occurring close to the LIRG nuclei remain undiscovered by the current SN searches and will therefore contribute to the 'SN rate problem' recently suggested by \\citet{horiuchi11}." }, "1112/1112.4859_arXiv.txt": { "abstract": "{} {We develop a statistical analytical model that predicts the occurrence frequency distributions and parameter correlations of avalanches in nonlinear dissipative systems in the state of a slowly-driven self-organized criticality (SOC) system.} {This model, called the fractal-diffusive SOC model, is based on the following four assumptions: (i) The avalanche size $L$ grows as a diffusive random walk with time $T$, following $L \\propto T^{1/2}$; (ii) The energy dissipation rate $f(t)$ occupies a fractal volume with dimension $D_S$, (iii) The mean fractal dimension of avalanches in Euclidean space $S=1,2,3$ is $D_S \\approx (1+S)/2$; and (iv) The occurrence frequency distributions $N(x) \\propto x^{-\\alpha_x}$ based on spatially uniform probabilities in a SOC system are given by $N(L) \\propto L^{-S}$, with $S$ being the Eudlidean dimension. We perform cellular automaton simulations in three dimensions ($S=1,2,3$) to test the theoretical model.} {The analytical model predicts the following statistical correlations: $F \\propto L^{D_S} \\propto T^{D_S/2}$ for the flux, $P \\propto L^{S} \\propto T^{S/2}$ for the peak energy dissipation rate, and $E \\propto F T \\propto T^{1+D_S/2}$ for the total dissipated energy; The model predicts powerlaw distributions for all parameters, with the slopes $\\alpha_T=(1+S)/2$, $\\alpha_F=1+(S-1)/D_S$, $\\alpha_P=2-1/S$, and $\\alpha_E=1+(S-1)/(D_S+2)$.\tThe cellular automaton simulations reproduce the predicted fractal dimensions, occurrence frequency distributions, and correlations within a satisfactory agreement within $\\approx 10\\%$ in all three dimensions.} {One profound prediction of this universal SOC model is that the energy distribution has a powerlaw slope in the range of $\\alpha_E=1.40-1.67$, and the peak energy distribution has a slope of $\\alpha_P=1.67$ (for any fractal dimension $D_S=1,...,3$ in Euclidean space $S=3$), and thus predicts that the bulk energy is always contained in the largest events, which rules out significant nanoflare heating in the case of solar flares.} ", "introduction": "The statistics of nonlinear processes in the universe often shows powerlaw-like distributions, most conspicously in energetic dynamic phenomena in astrophysics (e.g., solar and stellar flares, pulsar glitches, auroral substorms) and in catastrophic events in geophysics (e.g., earthquakes, landslides, or forest fires). The most widely known example is the distribution of earthquake magnitudes, which has a powerlaw slope of $\\alpha \\approx 2.0$ for the differential frequency distribution (Turcotte 1999), the so-called Gutenberg-Richter (1954) law. Bak, Tang, and Wiesenfeld (1987, 1988) introduced the theoretical concept of self-organized criticality (SOC), which has been initially applied to sandpile avalanches at a critical angle of repose, and has been generalized to nonlinear dissipative systems that are driven in a critical state. Comprehensive reviews on this subject can be found for applications in geophysics (Turcotte 1999), solar physics (Charbonneau et al.~2001), and astrophysics (Aschwanden 2011). Hallmarks of SOC systems are the scale-free powerlaw distributions of various event parameters, such as the peak energy dissipation rate $P$, the total energy $E$, or the time duration $T$ of events. While the powerlaw shape of the distribution function can be explained by the statistics of nonlinear processes that have an exponential growth phase and saturate after a random time interval (e.g., Willis and Yule 1922; Fermi 1949; Rosner and Vaiana 1978; Aschwanden et al.~1998; Aschwanden 2004, 2011), no general theoretical model has been developed that predicts the numerical value of the powerlaw slope of SOC parameter distributions. Simple analytical models that characterize the nonlinear growth phase with an exponential growth time $\\tau_G$ and the random distribution of risetimes with an average value of $t_S$, predict a powerlaw slope of $\\alpha_P = 1 + t_S/\\tau_G$ for the energy dissipation rate (e.g., Rosner and Vaiana 1978; Aschwanden et al.~1998), but cellular automaton simulations suggest a much more intermittent energy release process than the idealized case of an avalanche with a single growth and decay phase. An alternative theoretical explanation for a slope $\\alpha_E=3/2$ was put forward by a dimensional argument (Litvinenko 1998), which can be derived from the definition of the kinetic energy of convective flows, but this model entails a specific physical mechanism that has not universal validity for SOC systems. In this Paper we propose a more general concept where the powerlaw slope of the occurrence frequency distribution of SOC parameters depends on the fractal geometry of the energy dissipation domain. We aim for a universal statistical model of nonlinear energy dissipation processes that is independent of any particular physical mechanism. The fractal structure of self-organized processes has been stressed prominently from beginning (Bak, Tang, \\& Wiesenfeld 1987, 1988; Bak and Chen 1989), but no quantitative theory has been put forward that links the fractal geometry to the size distribution of SOC events. Fractals have been studied independently (e.g., Mandelbrot 1977, 1983, 1985), while the fractal geometry of SOC avalanches was postulated (e.g., see textbooks of Bak 1996; Sornette 2004; Aschwanden 2011), but no general self-consistent model has been attempted. This paper presents an analytical theory that derives a theoretical framework to quantitatively link the concept of fractal dimensions to the occurrence frequency distributions of SOC avalanche events (Section 2), tests of the analytical theory with numerical simulations of cellular automaton models in three Euclidean dimensions (Section 3), a comparison and application to solar flares (Section 4), and a summary of the model assumptions and conclusions (Section 5). ", "conclusions": "We developed an analytical theory for the statistical distributions and correlations of observable parameters of SOC events, which includes the avalanche length scale $L$, the time duration $T$, the peak $P$ and energy dissipation rate $F$. The basic assumptions of our analytical model, which we call the {\\sl fractal-diffusive SOC model}, are the following: \\begin{enumerate} \\item{{\\bf Diffusive expansion of SOC avalanches:} The radius $r(t)$ or spatial length scale $L$ of an avalanche grows with time like the average of a diffusive random walk, which predicts a statistical correlation $L \\propto T^{1/2}$ between the length scale $L$ and time duration $T$ of the avalanche.} \\item{{\\bf Fractal Energy Dissipation Rate:} The complexity of random next-neighbor interactions in a critical SOC state can be characterized approximately with a fractal geometry. The volume (or area) of the instantaneous energy dissipation rate is assumed to have a fractal dimension $D_S$. The predicted statistical correlations are: $F \\propto T^{D_S/2}$, $P \\propto T^{S/2}$, and $E \\propto T^{1+D_S/2}$.} \\item{{\\bf Mean Fractal Dimension:} The mean fractal dimension $D_S$ for different Euclidean space dimensions $S=1,2,3$ can be estimated from the arithmetic mean of the minimum dimension for a propagating avalanche, $D_{S,min} \\approx 1$, and the maximum (Euclidean) dimension $D_{S,max} = S$, which yields $D_S \\approx (1+S)/2$.} \\item{{\\bf Occurrence Frequency Distributions:} Equal probability of avalanches with size $L$ at various spatial locations in a uniform, slowly-driven SOC system predicts a probability distribution of $N(L) \\propto L^{-S}$. A direct consequence of this assumption, together with the other assumptions made above, yields a powerlaw function $N(x) \\propto x^{-\\alpha_x}$ for the occurrence frequency distributions of all parameters, which is the hallmark of a SOC system. The predicted powerlaw indices are: $\\alpha_T=(1+S)/2$, $\\alpha_F=1+(S-1)/D_S$, $\\alpha_P=2-1/S$, and $\\alpha_E=1+(S-1)/(D_S+2)$. Specifically, for applications to 3-D phenomena, absolute values are predicted for the powerlaw slopes $\\alpha_L=3$ and $\\alpha_T=2$, and $\\alpha_P=1.67$, while a range of $\\alpha_E=1.4,...,1.67$ is expected for any fractal dimension in the range of $1 \\le D_3 \\le 3$.} \\end{enumerate} We have validated our theory by a detailed comparison with a set of SOC simulations using a specific form of a cellular automaton avalanche model (connectivity, stability threshold, redistribution rule, etc), and found a good agreement between theory and numerical simulations, in the order of $\\approx 10\\%$ for the powerlaw slopes ($\\alpha_L, \\alpha_T, \\alpha_E, \\alpha_P$), the power indices of correlated parameters ($\\beta_{TL}, \\beta_{TP}, \\beta_{TE}$), and the fractal dimensions $D_S$, for all three Euclidean space dimensions $S=1,2,3$. Yet, at its most general level our theory is saying that the self-similarity of energy release statistics in such models is a direct reflection of the fractal nature of avalanches. The SOC flare model recently proposed by Morales \\& Charbonneau (2008, 2009) offers an interesting test of this conclusion. Their model, defined on a set of initially parallel magnetic flux strands, contained in a plane and subjected to random sideways deformation, with instability and readjustment occurring when the crossing angle of two flux strands exceeds some threshold angle. This model is thus strongly anisotropic, with pseudo-local stability and redistribution, in the sense that these operators now act on nearest-neighbors nodes located along each flux strand, rather than in the immediate spatial vicinity of the unstable sites. This is very different from the isotropic Lu et al.~(1993)-type SOC model used here for validation. Yet, the results compiled in Table 2 of Morales \\& Charbonneau (2008) for their highest resolution simulations reveal that the theoretical occurrence frequency distributions obtained herein, do hold within the stated uncertainties on the power-law fits. Likewise, the power indices of the correlated parameters also agree with theory within the inferred uncertainties. This provides additional empirical support to our conjecture that the fractal-diffusive SOC model does represent a robust characterization of avalanche energy release in SOC systems in general. However, it should be remembered that the inferred scaling laws are only valid for a slowly-driven SOC system, while alternative SOC systems with time-variable drivers or non-stationary input rates exhibit modified occurrence frequency and waiting time distributions (Charbonneau et al.~2001; Norman et al.~2001), which was also found in solar observations extending over multiple solar cycles (Crosby et al.~1993; Biesecker 1994; Bai 1993; Aschwanden 2010a). What other predictions can be made from our analytical model? For SOC processes in 3-D space, which is probably the most common application in the real world, the mean fractal dimension is predicted to be $D_3\\approx 2.0$, which can be tested by measurements of fractal dimensions in observations. The 20 largest solar flares observed with TRACE have been analyzed in this respect and an area fractal dimension of $D_2=1.89\\pm0.05$ was found at the flare peaks, which translates into a value of $D_3=2.10\\pm0.14$ if we use an anisotropic flare arcade model (Aschwanden and Aschwanden 2008a). The distribution of flare energies is predicted to have a powerlaw slope of $\\alpha_E=1.50$, which closely matches the observed statistics of solar flare hard X-ray emission ($\\alpha_E \\approx 1.49-1.56$). Since this value is undisputably below the critical limit $\\alpha=2$ of the energy integral, the total released energy is contained in the largest flares and thus rules out any significant nanoflare heating of the solar corona. Another prediction, that we did not test here with solar flare data, is the diffusive flare size scaling. Straightforward tests could be carried out by gathering statistics of the flare size evolution during individual flares (which are predicted to scale as $x(t) \\propto t^{1/2}$), as well as from the statistics of a large sample of flares, which is predicted to show a correlation $L \\propto T^{1/2}$. The application of our fractal-diffusive SOC model to solar flares implies that the subsequent triggering of local magnetic reconnection events during a flare occurs as a diffusive random walk. A similar finding of diffusive random walk was also found in the turbulent flows of magnetic bright points in the lanes between photospheric granular convection cells (Lawrence et al.~2001). The spatio-temporal scaling of the diffusive random walk predicts also the size, duration, and energy of the largest flare, which is likely to be constrained by the size $L_{AR} \\propto T_{max}^{1/2}$ of the largest active region." }, "1112/1112.3965_arXiv.txt": { "abstract": "{ We present \\swift\\ follow-up observations of a sample of 35 unclassified faint X-ray sources drawn from the \\asca\\ Galactic centre and Galactic plane surveys. Our short, pointed XRT observations allow detections down to a limiting 0.3--10 keV flux of $F_X \\sim10^{-13}~\\flux$, which translates into a luminosity of $L_X\\sim10^{33}~\\lum$ for an assumed distance of $D=8$~kpc. The brightest source in our sample reaches a maximum 0.3--10 keV luminosity of $L_X\\sim2\\times10^{36}~(D/8~\\mathrm{kpc})^2~\\lum$ during our \\swift\\ observations. We detect 16 ($46\\%$) of the \\asca\\ sources with the XRT, while 19 were not detected during our program. Since we are probing the faint end of the \\asca\\ source populations, we expect a large fraction of the non-detections to be due to the Eddington bias. This is strengthened by the fact that we find the observed XRT count rates to be predominantly lower than expected based on the reported \\asca\\ intensities. Nevertheless, investigation of the \\asca\\ properties and any possible long-term X-ray variability leads us to conclude that the non-detections likely include two spurious \\asca\\ detections and three objects that are highly variable or transient X-ray sources. For the 16 XRT-detected sources we obtain positional accuracies of $\\sim2-4''$, which significantly improves upon their \\asca\\ uncertainties of $\\sim1'$. We use the X-ray spectra and variability to characterise these objects. Most appear to be faint, persistent X-ray emitters that have highly absorbed spectra. Based on their X-ray properties we identify three accreting compact objects: one confirmed (\\ucxb) and one candidate (\\porb) X-ray binary, and one possible magnetically accreting white dwarf (\\polar). Furthermore, we use the improved positions of XRT-detected sources to search for counterparts in simultaneously obtained \\swift/UVOT images and possible associations with catalogued sources at various wavelengths. This reveals three possible main sequence stars amongst our sample. The other sources remain unclassified, but our improved XRT positions provide good prospects for dedicated follow-up observations that have the potential to elucidate the nature of these faint, unclassified \\asca\\ sources. } ", "introduction": "The Advanced Satellite for Cosmology and Astrophysics (\\asca) was launched in 1993 and operated until 2000. It was the first mission that combined imaging capabilities with a relatively broad energy passband (0.7--10 keV). During its lifetime, the \\asca\\ satellite pointed towards the Galactic plane and the inner bulge of the Galaxy many times. In particular, two dedicated surveys were performed that detected about 200 X-ray sources with 0.7--10 keV fluxes of $F_X\\gtrsim 3\\times 10^{-13}~\\flux$ \\citep[][]{sugizaki01,sakano02}. Both surveys used the GIS (gas imaging spectrometer) detector, which had a spatial resolution of $\\sim30\\arcsec$ and a field of view (FOV) of $\\sim50'$. The \\asca\\ Galactic plane survey entailed a systematic study of X-ray sources located in the inner Galactic disk \\citep{sugizaki01}. The region of $| l | \\lesssim45^{\\circ}$, $| b |\\lesssim 0.4^{\\circ}$ was covered between 1996 and 1999 using successive 10-ks observations. Investigation of three different energy bands (0.7--2, 2--10 and 0.7--7 keV) resulted in the identification of 163 individual X-ray sources detected with a significance of $\\gtrsim 4 \\sigma$. \\asca\\ targeted a region of $5^{\\circ} \\times 5^{\\circ}$ around the centre of our Galaxy between 1993 and 1999 \\citep{sakano02}. Searching the 0.7--3 and 3--10 keV images yielded a total of 52 point sources that were detected with a significance of $\\gtrsim5\\sigma$. In both \\asca\\ surveys, spectral analysis was performed for sources that were detected with a significance of $\\gtrsim5\\sigma$. The GIS data were fitted to a simple absorbed powerlaw model to characterise the X-ray spectra \\citep[][]{sugizaki01,sakano02}. Source positions could be determined with a $90\\%$ confidence positional accuracy of $\\sim1'$ \\citep{ueda1999,sugizaki01,sakano02}. After cross-correlation with the \\einstein\\ and \\rosat\\ bright source catalogues, about 2/3 of the \\asca-detected X-ray sources remained unclassified \\citep{sugizaki01,sakano02}. These objects have 0.7--10 keV X-ray fluxes in the range of $F_X \\sim 10^{-13} - 10^{-11}~\\flux$, which translates into a luminosity of $L_X \\sim 10^{33}-10^{35}~\\lum$ for a distance of $D=8$~kpc. This is right in between the well-known bright X-ray point sources ($L_X>10^{36}~\\lum$) that are mostly active X-ray binaries \\citep[e.g.,][]{chen97}, and dim X-ray sources ($L_X<10^{33}~\\lum$) that consist of a collection of source types such as accreting white dwarfs, dormant X-ray binaries, active stars or young pulsars \\citep[e.g.,][]{muno2009}. The luminosity range traced by the unclassified \\asca\\ sources suggests that a variety of source types might be amongst them. For example, it covers the intensities typically seen for strongly magnetised neutron stars (magnetars), bright magnetically accreting white dwarfs (polars and intermediate polars), sub-luminous X-ray binaries and X-ray emitting massive stars \\citep[e.g.,][]{maeda1996,ezuka1999,wijnands06,chelovekov07,gelfland2007,heras08,kaur2010,degenaar09_gc,anderson2011}. In addition, the sample may contain foreground stars or background active galactic nuclei (AGN). The main difficulty in classifying the \\asca\\ sources is their relatively large positional uncertainties of $\\sim$1$'$. This inhibits follow-up observations aiming to search for counterparts at other wavelengths. We launched a program to observe unclassified \\asca\\ sources with the \\swift\\ satellite \\citep[][]{gehrels2004}. The primary instrument for our observations is the X-ray telescope \\citep[XRT;][]{burrows05}. The XRT is sensitive in the 0.3--10 keV energy range and has a FOV of $\\sim23' \\times 23'$. Furthermore, we used data obtained with the Ultra-Violet/Optical Telescope \\citep[UVOT;][]{roming05} that has a FOV of $\\sim17' \\times 17'$ and can be operated using a variety of filters in a wavelength range of $\\sim1500-8000$ \\AA{} \\citep[][]{poole2008}. In addition to these narrow-field instruments, \\swift\\ is equipped with the Burst Alert Telescope \\citep[BAT;][]{barthelmy05}: a hard X-ray (15--150 keV) monitoring instrument that has a wide FOV ($2$~steradian). The combination of its pointing flexibility, good sensitivity in the soft X-ray band (below 10 keV) and multi-wavelength coverage makes \\swift\\ an ideal tool to perform follow-up observations of X-ray sources discovered by other missions \\citep[e.g.,][]{rodriguez09,starling2011,kennea2011}. By observing unclassified \\asca\\ sources with \\swift, our primary aim is to improve their positional uncertainties to a few arcseconds, so that dedicated searches for counterparts at other wavelengths become feasible. In addition, we strive to obtain X-ray spectral information and to study possible variability in their X-ray emission. This approach provides a basis for further investigation of the nature of individual sources, as well as the population of faint \\asca\\ sources as a whole.\\footnote{See also the {\\it ChIcAGO} project, which targets a (different) sample of unclassified \\asca\\ sources using \\chan\\ and multi-wavelength follow-up observations \\citep[][]{anderson2011}.} ", "conclusions": "\\label{sec:discuss} We carried out \\swift\\ follow-up observations of a sample of 35 unclassified faint X-ray sources detected during \\asca\\ surveys of the Galactic plane and centre between 1993 and 1999 \\citep[][]{sugizaki01,sakano02}. A total of 16 sources from our sample were detected with the XRT, which allows us to improve their positional accuracy from $\\sim$1$'$ to $\\sim$2--4$''$. The mean offset of the improved XRT position from the \\asca\\ coordinates is $47.9''$ ($0.8'$), which is relatively close to the reported 90\\% confidence \\asca\\ positional uncertainties. We investigated the X-ray spectra of the detected sources and studied any possible long-term X-ray variability of all our 35 targets. Our program was conducted between 2006 and 2010, hence $\\sim7-17$ years since the \\asca\\ detections. In general, it is not possible to classify the sources based on their X-ray spectral properties, particularly given the fact that the total number of counts detected during our observations was often low (see Table~\\ref{tab:spec}). Using the improved \\swift/XRT position information, we therefore searched optical and infrared source catalogues to assess any possible associations. Below we discuss the possible nature of the sample of unclassified \\asca\\ sources covered by our program. Proposed classifications and possible associations are summarised in Table~\\ref{tab:class}. \\subsection{The nature of the \\asca\\ sources detected with \\swift}\\label{subsec:nat_detect} \\subsubsection{Accreting compact objects}\\label{subsubsec:compactobjects} {\\bf \\ucxb} was observed with \\swift/XRT on two different epochs in 2007 July and 2008 August--July, during which it is detected at an average 2--10 keV luminosity of $L_X \\sim 6\\times10^{34}~(D/9.2~\\mathrm{kpc})^2~\\lum$. This is considerably lower than typically observed for neutron star LMXBs and classifies the source as a sub-luminous X-ray binary \\cite[][]{muno05_apj622,wijnands06,degenaar09_gc,campana09}. The \\swift/XRT data show that the quiescent phase signalled by \\chan\\ in 2008 May must have been short \\citep[$\\lesssim11$~months, see][]{bassa08,jonker08}. Its ability to spend long times accreting at a low X-ray luminosity, combined with its thermonuclear X-ray burst properties and low absolute optical magnitude ($M_i \\gtrsim 2$), led to the suggestion that \\ucxb\\ could be an ultra-compact X-ray binary \\citep[][]{bassa08}. Such systems harbour hydrogen-poor donor stars in small orbits of $\\lesssim 90$~min \\citep[][]{nelson86}. The XRT spectrum of \\ucxb\\ is very soft with a photon index of $\\Gamma=2.7$. Such a high spectral index is quite unusual, although other LMXBs accreting at similarly low luminosities have also been found to display very soft X-ray spectra \\citep[e.g.,][]{zand05,delsanto07,degenaar09_gc,degenaar2010_burst,armas2011}. An in-depth investigation of high-quality \\xmm\\ data of the sub-luminous (candidate) LMXB \\xte\\ revealed that the X-ray spectrum contained a soft (thermal) component in addition to the hard powerlaw emission \\citep[][]{armas2011}. In contrast, the XRT spectrum of \\ucxb\\ can be adequately fit by an absorbed powerlaw alone and does not require the addition of a soft spectral component. \\\\ \\noindent {\\bf \\porb} is the brightest source in our sample and was detected with the XRT at an average 2--10 keV luminosity of $L_X \\sim 2\\times10^{35}~(D/8~\\mathrm{kpc})^2~\\lum$. It has a highly absorbed ($N_H \\sim 8 \\times 10^{22}~\\nh$) X-ray spectrum, that can be described by a simple powerlaw model with an index of $\\Gamma=2.3$. The XRT light curve shows that the source intensity varies considerably by a factor of $\\sim10$ on a timescale of days. Given its X-ray spectral characteristics, its average intensity and the timescale of the observed variability we tentatively identify \\porb\\ as an X-ray binary. During our program \\porb\\ spanned a luminosity range of $L_X\\sim1\\times10^{34}-5\\times10^{35}~(D/8~\\mathrm{kpc})^2~\\lum$. Such large variations in X-ray intensity (a factor $\\sim50$) are not uncommon for X-ray binaries and are usually ascribed to considerable variations in the mass-accretion rate onto the compact primary. For example, strong X-ray variability is a characteristic feature of Supergiant Fast X-ray Transients (SFXTs), in which the compact primary accretes matter from the varying wind of a massive ($\\gtrsim10~\\Msun$) supergiant companion star \\citep[e.g.,][]{sidoli08,romano2011}. Similar to what we observe for \\porb, these kind of systems are typically highly absorbed. The comparison breaks down, however, when considering the X-ray spectral shape: SFXTs typically have much harder X-ray spectra ($\\Gamma \\sim 1$) in the luminosity range that we observe for \\porb\\ \\citep[][]{romano2011}. Variations of similar magnitude and on a similar timescale as seen for \\porb\\ have also been observed from LMXBs in which the companion star overflows its Roche lobe and accretion is governed via an accretion disk \\citep[e.g.,][]{wijnands2001_1808,linares2008,degenaar2011_1701,fridriksson2011}. Although the average 2--10 keV X-ray luminosity of this source is lower than is typically seen for active LMXBs \\citep[e.g.,][]{chen97}, there is a growing group of sub-luminous sources that accrete at similar X-ray intensities to \\porb\\ \\citep[e.g., \\ucxb\\ discussed above, but see also][]{cornelisse02,muno05_apj622,wijnands06,zand05,zand07,delsanto07,degenaar09_gc,campana09,degenaar2010_burst}. Based on its X-ray spectral characteristics we consider it more likely that \\porb\\ belongs to the class of LMXBs and thus harbors a low-mass ($\\lesssim1~\\Msun$) companion star as opposed to a more massive donor. Dedicated follow-up observations at longer wavelengths have the potential to rule out a high-mass X-ray binary nature and thus provide a test for our proposed classification.\\\\ \\noindent {\\bf \\polar} stands out by displaying a much less absorbed ($N_H\\lesssim3\\times10^{21}~\\nh$) and much flatter ($\\Gamma \\sim 0$) X-ray spectrum than observed for the majority of X-ray sources in our sample (see Table~\\ref{tab:spec}). The hard X-ray spectrum of \\polar\\ (see Fig.~\\ref{fig:wd}) resembles that of magnetically accreting white dwarfs \\citep[polars and intermediate polars;][]{ezuka1999,muno2003,hong2009}. Although wind-accreting neutron stars in high-mass X-ray binaries (HMXB) may show similarly hard X-ray spectra, the low hydrogen absorption column density inferred from fitting the X-ray spectrum and the lack of any UVOT or catalogued optical/infrared counterpart within the $2.8''$ XRT positional uncertainty, renders a HMXB nature less likely. Based on its X-ray spectral properties and lack of a counterpart at longer wavelengths we tentatively classify \\polar\\ as a magnetically accreting white dwarf. The limited statistics of the XRT data does not allow us to investigate the timing properties of the source. Deeper X-ray observations may test our proposed source classification by searching for common features of magnetically accreting white dwarfs, such as prominent iron emission lines in the X-ray spectrum or coherent X-ray pulsations with a periodicity on the order of minutes to hours \\citep[e.g,][]{ritter2003,muno2003,hong2009,kaur2010}. Follow-up observations at other wavelengths (optical/infrared) can further aid in the classification of this source. \\subsubsection{Main sequence stars}\\label{subsubsec:ms_stars} AX J1738.4--2902 is positionally coincident with an optical/infrared catalogued source. We also detect a likely counterpart at this position in our UVOT images. Based on the fact that we a find very soft X-ray spectrum with little absorption, we tentatively associate AX J1738.4--2902 with the catalogued optical/infrared object. As such, this \\asca\\ source is identified as a K0V pre-main sequence star (see Sect.~\\ref{subsec:oir}). For another 7 of our XRT detections we find possible associations with X-ray/UV/optical/infrared/radio sources (see Tables~\\ref{tab:oir} and~\\ref{tab:class}). In Fig.~\\ref{fig:ir} we plot the approximate {\\it 2MASS} colours for the 6 possible infrared counterparts to detected \\asca\\ sources in our sample (magnitudes are not corrected for extinction).\\footnote{Taking into account extinction would push the data points down and towards the left in the colour-colour diagram (Fig.~\\ref{fig:ir}).} Circles represent sources which also have optical information, whereas squares denote those which are only detected in the infrared bands (see Table~\\ref{tab:oir}). The dashed line in Fig.~\\ref{fig:ir} indicates the expectation for A- through M-type main-sequence stars \\citep[][]{bessell1988,tokunaga2000}. The labels to the data points correspond to the source numbers used in Table~\\ref{tab:oir}. We can see in Fig.~\\ref{fig:ir} that three sources lie along the main-sequence line: AX J1504.6--5824, AX J1720.8--3710 and AX J1738.4--2902. Our proposed counterpart for the latter is indeed classified as a K0V star (see above). If AX J1504.6--5824 and AX J1720.8--3710 are related to the infrared objects located within their XRT error circles, both are likely main sequence stars as well. The low absorption column density ($N_H\\lesssim3\\times10^{21}~\\nh$) inferred from fitting the X-ray spectrum of AX J1720.8--3710 makes this association likely. The statistics of the spectral data are limited, however, and the offset from the XRT position to the {\\it 2MASS} source is relatively large (see Tables~\\ref{tab:spec} and~\\ref{tab:oir}). In the case of AX J1504.6--5824, there is also a considerable offset between the XRT position and the UV/optical/infrared object, whereas the X-ray spectrum appears relatively hard and highly absorbed (Tables~\\ref{tab:spec} and~\\ref{tab:oir}). This casts doubt on the possible classification as a foreground star. We emphasise that the probability of a chance detection of an optical/infrared source within the XRT error circles is high (typically $\\sim50\\%$) as the source fields are crowded (see Sect.~\\ref{subsec:oir}). It may not be a coincidence that the sources for which we find possible associations are often the ones that were detected at low significance in the XRT data and consequently have relatively large positional errors compared to the brighter XRT detections (see Table~\\ref{tab:pos}). Our proposed possible associations may be confirmed or rejected by obtaining even more accurate X-ray positions (which is currently only possible with \\chan), or by conducting dedicated follow-up observations at optical/infrared wavelengths. \\begin{figure} \\begin{center} \\includegraphics[width=8cm]{2MASS_ASCA_custom.eps} \\end{center} \\caption[]{Approximate {\\it 2MASS} colours for possible near-infrared counterparts of detected sources in our sample (magnitudes are not corrected for extinction). Circles represent sources that also have optical information, whereas squares indicate objects that are only detected in the infrared bands. The data points are labelled to indicate the corresponding source numbers (see Table~\\ref{tab:oir}). The dashed line indicates the expectation for A- through M-type main-sequence stars. Correcting for extinction would move the data points down and towards the left in this diagram. } \\label{fig:ir} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[width=8cm]{rates_sigma2.eps} \\end{center} \\caption[]{Observed versus expected XRT count rates for both detected and undetected sources in our sample (represented by squares and upper limit symbols, respectively). The solid line corresponds to a ratio of observed over expected count rates that equals unity, whereas the dashed lines represents $1\\sigma$ deviations from this relation. } \\label{fig:rates} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[width=8cm]{asca_par_histo_custom.eps} \\end{center} \\caption[]{The source detection significances of the \\asca\\ Galactic plane survey are shown for the 3 energy bands investigated in that work \\citep[][]{sugizaki01}. \\swift/XRT non-detected sources are represented by the solid black line, the firmly detected sources are coloured red and tentative detections are indicated by hatched lines. These three different groups are overlaid in the graph (i.e., it is not cumulative). } \\label{fig:ascapar} \\end{figure} \\subsection{The nature of non-detected sources: a comparison with the \\asca\\ results}\\label{subsec:nat_nondetect} In Fig.~\\ref{fig:rates}, we compare the observed and expected XRT count rates for our sample sources. This figure shows that to within 1$\\sigma$ (dashed lines) only one source was brighter than predicted, while for 46\\% of our sample the count rates lay below expectations. Among these fainter than expected sources are many of the XRT non-detections, implying either transient or highly variable sources, significant Eddington bias effects, extended rather than point-like sources or spurious detections in the \\asca\\ surveys. The Eddington bias, which boosts the true count rate of sources at or near the detection limit \\citep[][]{eddington1913}, is likely to be at work in this sample as we are probing the faint end of the \\asca\\ catalogues. This idea is strengthened by the fact that we find the observed XRT count rates to be predominantly lower than expected. Without knowing the intrinsic source population we cannot estimate the magnitude of this effect, but we note that work on the \\xmm\\ hard band survey (2--10 keV) has shown that for the \\xmm\\ slew survey, the true count rate distribution peaks a factor of 2 below the observed count rate for detections with low numbers of counts \\citep[][Warwick, R. et al., submitted to A\\&A]{starling2011}. Three of our XRT non-detected sources have count rate upper limits that are a factor 5--10 lower than predicted from the \\asca\\ observations (see Table~\\ref{tab:nondet}), which strongly suggests these are not faint, steady sources with count rates boosted by the Eddington bias, but are either variable (possibly transient) sources or spurious detections. Of these three, AX J1833.9--0822 was a firm 6$\\sigma$ detection above 2 keV with \\asca\\ and AX J1836.3--0647 was a 5$\\sigma$ detection in the 0.7--7 keV band dominated by its soft flux, while no similar information is available for AX J1751.1--2748. These detection significances disfavour a non-astrophysical, spurious nature. Various classes of X-ray sources (e.g., X-ray binaries, accreting white dwarfs, magnetars, stars, AGN) vary in intensity by a factor of a few. X-ray sources are generally denoted as transient when the variations are as large as $>2$ orders of magnitude. We thus tentatively identify AX J1751.1--2748, AX J1833.9--0822 and AX J1836.3--0647 as strongly variable or transient X-ray sources, although we cannot classify them further. While it is not possible to classify the sources \\swift\\ has not detected, we may learn something of their nature from further examining the details of their \\asca\\ measurements. A total of 26 objects in our sample were drawn from the \\asca\\ Galactic plane survey, which were evaluated in 3 energy bands (0.7--2 keV, 2--10 keV and 0.7--7 keV) and assigned a detection significance \\citep[][]{sugizaki01}. We plot these values in Fig.~\\ref{fig:ascapar}, distinguishing between the XRT non-detections (solid black lines), detections (coloured red) and tentative detections (hatched lines). It can be seen that just 3 sources were significantly detected in the soft energy band (0.7--2 keV), and these were also detected both in other \\asca\\ bands and later with the XRT. This is unsurprising given the high absorbing columns in these directions (see Tables~\\ref{tab:spec} and \\ref{tab:nondet}). This causes the broad band (0.7--7 keV) to largely follow the results of the hard band (2--10 keV), where XRT-detected sources generally represent the higher-significance \\asca\\ detections while XRT non-detections are found below 5--6$\\sigma$. Our tentative XRT detections span a wider range of \\asca\\ detection significances, and are concentrated in the 0.7--7 keV band over the 4--5$\\sigma$ range. It is of note that Galactic plane survey sources with $>$4$\\sigma$ detections in all 3 \\asca\\ energy bands number just two, AX J1538.3--5541 and AX J1651.0--4403, both of which are well detected with \\swift/XRT (see Sect.~\\ref{subsubsec:porb} for further details of AX J1538.3--5541). There are 3 hard-band only sources: 1 was tentatively detected with the XRT and the other 2 were not. Broad-band only detections generally have the fewest total counts and are therefore most likely to be spurious. We have two, AX J1835.1--0806 and AX J1857.3+0247. Both went undetected with the XRT adding further uncertainty to assignment of an astrophysical nature to these sources. We thus tentatively mark the two as spurious \\asca\\ detections. However, we note that their XRT count rate limits are not more than a factor of 2 below the expected rates. This is certainly possible within the Eddington bias effect and a large number of sources are variable at this level. Furthermore, our expected XRT count rates are overestimated if the sources are extended rather than point-like. As mentioned in Sect.~\\ref{subsec:hard}, AX J1857.3+0247 may possibly be related to an extended X-ray source. Spectral analysis was attempted for those sources with \\asca\\ detections above 5$\\sigma$ in both \\citet{sugizaki01} and \\citet{sakano02}, over the full energy range of 0.7--10 keV and adopting an absorbed power law model, although only \\citet{sugizaki01} provide error estimates on the parameters. For the five firm XRT detections with well-measured \\asca\\ spectral parameters, $N_H$ and $\\Gamma$ are consistent within the 90\\% uncertainties with one exception: AX J1538.3--5541. For this source $N_H$ is consistent between the \\asca\\ and \\swift/XRT averaged spectra, while the photon index $\\Gamma$ is just consistent at the 3$\\sigma$ level possibly having been harder during the \\asca\\ observation. We found this source to be highly time variable from multiple XRT observations, and possible spectral evolution is investigated in Sect.~\\ref{subsubsec:porb}. The overall consistency between spectral measurements validates our use of the \\asca\\ spectra to make count rate predictions for XRT as used in Tables~\\ref{tab:spec} and~\\ref{tab:nondet}. In conclusion, the 31\\% of sample sources firmly detected with {\\it Swift}/XRT were generally found with greater confidence in the \\asca\\ Galactic plane and Galactic centre surveys than the XRT non-detected sources. Very few sources were detected below 2 keV with \\asca, which is likely due to high foreground absorbing columns (see Tables~\\ref{tab:spec} and~\\ref{tab:nondet}). However, the greatest deviation from the expected count rate among the XRT non-detections corresponds to three sources that were largely well-detected with \\asca, strengthening the probability that they are astrophysical sources of variable or transient nature. We identify two sources that most likely have been spurious \\asca-detections given that these were weakly detected only in the full \\asca\\ energy band and not detected with XRT. For sources well-detected with both \\asca\\ and XRT we find their spectra to be unchanged within the given errors between the two epochs. \\subsection{Summary}\\label{subsec:summary} Our sample of 35 unclassified \\asca\\ sources includes one confirmed and one candidate X-ray binary, and likely one magnetically accreting white dwarf. We assign these classifications based on their X-ray spectral and temporal properties. Furthermore, we used the improved XRT positions of the 16 sources detected in our sample to search for possible associations with catalogued sources at X-ray, UV, optical, infrared and radio wavelengths. This results in the identifications of possible counterparts for 8 of our targets, amongst which are 3 likely main sequence stars (see Table~\\ref{tab:class}). The probability of an optical/infrared chance detection is, however, considerable in these crowded fields ($\\sim50\\%$). By assessing any possible long-term variability and the reported \\asca\\ properties of the 19 sources that were not detected by XRT, we identify two potentially spurious \\asca-detections amongst them and three X-ray sources that appear to be variable or transient (see Table~\\ref{tab:class}). A substantial fraction of the XRT non-detections are expected to be due to the Eddington bias, and might thus involve weak, persistent X-ray sources with intensities that lie near the detection limit of our \\swift/XRT observations. With our improved $\\sim2-4''$ X-ray positions obtained for the 16 XRT-detected sources dedicated follow-up observations at different wavelengths become feasible. These have the potential to further classify these faint X-ray sources. As can be seen in Table~\\ref{tab:class}, our study shows that the unclassified \\asca\\ sources harbour a variety of astrophysical objects. \\begin{table*} \\begin{threeparttable}[t] \\begin{center} \\caption[]{{Proposed source associations/classifications.}} \\begin{tabular}{l l l} \\toprule $\\#$ & Name & Comments \\\\ \\midrule \\multicolumn{3}{l}{{\\bf Firm \\swift/XRT detections}} \\\\ 2 & AX J1504.6--5824 & Possible association with an optical/infrared object: main sequence star \\\\ 5 & AX J1538.3--5541 & Candidate X-ray binary, most likely an LMXB \\\\ 7 & \\polar & Candidate accreting magnetised white dwarf\\\\ 16 & AX J1738.4--2902 & Likely associated with 1RXS J173826.7--290140/2MASS J17382620--2901494: pre-main sequence star (K0V) \\\\ 19 & AX J1742.6--2901 & Possible association with 2RXP J174241.8--290215, optical/infrared/radio object \\\\ 22 & AX J1754.2--2754 & Confirmed neutron star LMXB, proposed ultra-compact X-ray binary \\\\ 29 & AX J1846.0--0231 & Possible association with an infrared object\\\\ \\midrule \\multicolumn{3}{l}{{\\bf Tentative \\swift/XRT detections}} \\\\ 11 & AX J1717.2--3718 & Possible association with an infrared object \\\\ 13 & AX J1720.8--3710 & Possibly related to 1RXS 172051.9--371033, UV/optical/infrared object: main sequence star \\\\ 17 & AX J1739.5--2730 & Possible association with an optical object \\\\ 30 & AX J1846.1--0239 & Possible association with an optical/infrared object \\\\ \\midrule \\multicolumn{3}{l}{{\\bf \\swift/XRT non-detections}} \\\\ 20 & AX J1751.1--2748 & Strongly variable/transient X-ray source \\\\ 25 & AX J1833.9--0822 & Strongly variable/transient X-ray source \\\\ 27 & AX J1835.1--0806 & Potential spurious \\asca\\ detection \\\\ 28 & AX J1836.3--0647 & Strongly variable/transient X-ray source \\\\ 34 & AX J1857.3+0247 & Potential spurious \\asca\\ detection \\\\ \\bottomrule \\end{tabular} \\label{tab:class} \\begin{tablenotes} \\item[]% \\end{tablenotes} \\end{center} \\end{threeparttable} \\end{table*}" }, "1112/1112.3648_arXiv.txt": { "abstract": "{\\it Chandra} spectroscopy of transient stellar-mass black holes in outburst has clearly revealed accretion disk winds in soft, disk--dominated states, in apparent anti-correlation with relativistic jets in low/hard states. These disk winds are observed to be highly ionized, dense, and to have typical velocities of $\\sim$1000~km/s or less projected along our line of sight. Here, we present an analysis of two {\\it Chandra} High Energy Transmission Grating spectra of the Galactic black hole candidate IGR J17091$-$3624 and contemporaneous EVLA radio observations, obtained in 2011. The second {\\it Chandra} observation reveals an absorption line at 6.91$\\pm$0.01~keV; associating this line with He-like Fe XXV requires a blue-shift of $9300^{+500}_{-400}$~km/s (0.03$c$, or the escape velocity at 1000~R$_{Schw}$). This projected outflow velocity is an order of magnitude higher than has previously been observed in stellar-mass black holes, and is broadly consistent with some of the fastest winds detected in active galactic nuclei. A potential feature at 7.32 keV, if due to Fe XXVI, would imply a velocity of $\\sim 14600$~km/s (0.05$c$), but this putative feature is marginal. Photoionization modeling suggests that the accretion disk wind in IGR J17091$-$3624 may originate within 43,300 Schwarzschild radii of the black hole, and may be expelling more gas than accretes. The contemporaneous EVLA observations strongly indicate that jet activity was indeed quenched at the time of our {\\it Chandra} observations. We discuss the results in the context of disk winds, jets, and basic accretion disk physics in accreting black hole systems. ", "introduction": "A detailed observational account of how black hole accretion disks drive winds and jets remains elusive, but the combination of high resolution X-ray spectroscopy, improved radio sensitivity, and comparisons across the black hole mass scale hold great potential. The broad range in X-ray luminosity covered by transient stellar-mass black holes makes it possible to trace major changes in the accretion flow as a function of the inferred mass accretion rate; this is largely impossible in supermassive black holes. Disk winds and jets, for instance, appear to be state-dependent and mutually exclusive in sources such as H 1743$-$322 \\citep{Miller06, Blum10}, GRO J1655$-$40 \\citep{Miller08,Luketic10,Kallman09}, and GRS 1915$+$105 \\citep{Miller08, Neilsen09}. This may offer insights into why many Seyfert AGN, which are well known for their disk winds, are typically radio--quiet \\citep[though not necessarily devoid of jets; see][]{King11, Jones11,Giroletti09}. The proximity of Galactic black hole binaries (BHB) ensures a high flux level and spectra with excellent sensitivity in the Fe K band. This is of prime importance because He-like Fe XXV and H-like Fe XXVI lines can endure in extremely hot, ionized gas \\citep[see, e.g.][]{Bautista01}, and therefore trace the wind region closest to where it is launched near the black hole. Studies of some stellar-mass black hole disk winds find that the gas is too ionized, too dense, and originates too close to the black hole to be expelled by radiative pressure or by thermal pressure from Compton heating of the disk, requiring magnetic pressure \\citep{Miller06, Miller06b, Kubota07}. Winds that may originate close to the black hole and carry high mass fluxes are also observed in AGN \\citep[e.g.,][]{Kaspi02, Chartas02,King11b,Tombesi10}. In this Letter, we present evidence of a particularly fast disk wind in the black hole candidate IGR J17091$-$3624. The current outburst of IGR J17091$-$3624 was first reported on 2011 January 28 \\citep{Krimm11}. Our observations caught IGR J17091$-$3624 in the high/soft state, but it is important to note that the source has also showed low/hard state episodes with flaring and apparent jet activity in radio bands \\citep{Rodriguez11}. X-ray flux variations in IGR J17091$-$3264 bear similarities to the microquasar GRS 1915$+$105 \\citep[e.g.,][]{Altamirano11}. ", "conclusions": "At ionizations above $10^{3}$, radiation pressure is inefficient, and it is not able to drive these winds \\citep[e.g.,][]{Proga00b}. Thus, although the UV components of disk winds in AGN are driven at least partially by radiation pressure, the wind in IGR J17091$-$3624 likely cannot be driven in this way. A thermal wind can arise at radii greater than $0.2~R_{C}$ \\citep{Woods96}, where $R_{C} = (1.0\\times 10^{10})\\times (M_{BH}/M_{\\odot})/T_{C8}$, where $T_{C8}$ is the Compton temperature of the gas in units of $10^{8}$~K. The spectrum observed in the second observation gives $R_{C} \\simeq 5\\times 10^{12}$~cm. Therefore, if we assume our conservative estimate of the launching radius, it is possible for IGRJ17091$-$3624 to have a thermally driven wind. However, if the wind originates closer to the black hole, then it is likely that magnetic processes -- either pressure from magnetic viscosity within the disk \\citep[e.g.,][]{Proga03} or magneto-centrifugal acceleration \\citep[e.g.,][]{Blandford82} -- must play a role in launching the wind observed in IGR J17091$-$3624. Fast X-ray disk winds are not only seen in BHB like IGR J17091$-$3624, but also in AGN and quasars \\citep[e.g.,][]{King11b, Chartas02}. The fastest UV winds observed in AGN are pushed to high velocities by radiation pressure. It remains to be seen whether a common driving mechanism works across the black hole mass scale to drive fast, highly ionized X-ray disk winds. \\cite{Chartas02} show that in the quasar APM 08279+5255 there are broad absorption features, which are likely highly relativistic Fe XXV and/or Fe XXVI lines. In these regards, it bears some similarities to the most extreme winds in BHB's. Observations of BHB point to an anti-correlation of wind and jet outflows from accretion disks \\citep{Miller06,Miller08, Blum10, Neilsen10}. Winds appear to only be detected, or at least are considerably stronger, in soft, disk--dominated states, and absent in hard states where compact, steady jets are ubiquitous \\citep{Fender06}. In H 1743$-$322, in particular, there is evidence that the absence of winds in hard states is {\\it not} an artifact of high ionization hindering the detection of absorption lines, but instead represents a real change in the column density (and thus the mass outflow rate) in any wind \\citep{Blum10}. It appears that our coordinated {\\it Chandra} and EVLA observations of IGR J17091$-$3624 support this anti-correlation. The EVLA observations place very tight limits on the radio flux when the disk wind is detected, orders of magnitude below the level at which IGR J17091$-$3624 was detected in radio during its low/hard state only a few months prior \\citep{Rodriguez11}. \\cite{Neilsen09} suggested that the production of winds may be responsible for quenching jets in GRS 1915$+$105. It might then be the case that jets should be observed whenever winds are absent. In our first observation of IGR J17091$-$3624, however, neither a wind nor a jet is detected, with tight limits. Instead, the apparent dichotomy between winds and jets may signal the magnetic field topology in and above the disk is state-dependent. This is broadly consistent with multi-wavelength studies suggesting synchrotron flares above the disk, but only in the hard state \\citep[e.g. GX 339$-$4, XTE J1118$+$480,][]{DiMatteo99, Gandhi10}. It is interesting to speculate that the magnetic field might be primarily toroidal in the soft state, where a Shakura-Sunyaev disk is dominant, but primarily poloidal in the hard state, when the mass accretion rate is lower \\citep[e.g.,][]{Beckwith08}. The type of outflow that is observed may also depend greatly on how much mass is loaded onto magnetic field lines; that could depend on variables including the mass accretion rate through the disk. \\vspace{0.2in} We would like to thank the anonymous referee. We thank Michael Nowak for his instrumental help as well. ALK gratefully acknowledges support through the NASA Earth and Space Sciences Fellowship. JMM gratefully acknowledges support through the {\\it Chandra} Guest Observer program. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. \\begin{deluxetable}{l l l l l} \\tablecolumns{5} \\tablewidth{0pc} \\tabletypesize{\\scriptsize} \\tablecaption{Spectral Modeling Parameters of the 2$^{nd}$ HEG observation} \\tablehead{Parameter & Model 1 & Model 2 & Model 3 & Model 4 \\\\ \\hline & diskbb + po & (diskbb + po) & comptt & (comptt) \\\\ & + Gauss + Gauss & $\\times$ Xstar $\\times$ Xstar & + Gauss + Gauss & $\\times$ Xstar $\\times$ Xstar } \\startdata N$_H$ (10$^{22}$ cm$^{-2}$) & 1.14$\\pm0.06$ & 1.13$\\pm0.06$ & 0.475 $^{+0.017}_{-0.018}$ & 0.558$^{+0.025}_{-0.028}$ \\\\ -\\\\ T$_{in}$ (keV) & 1.53 $\\pm0.09$ & 1.51 $^{+0.11}_{-0.09}$& - & -\\\\ Norm & 13.1$^{+3.6}_{-2.5}$ & 13.8$^{+4.0}_{-1.7}$ & - &- \\\\ $\\Gamma$ & 1.93 $^{+0.15}_{-0.16}$ & 1.91 $\\pm0.17$& - & -\\\\ Norm &0.35 $^{+0.07}_{-0.08}$ & 0.34 $\\pm0.08$ & - & -\\\\ -\\\\ T$_0$ (keV) & - &- & 0.58$\\pm0.01$ & 0.59 $\\pm0.01$ \\\\ kT (keV) & -& -& 10.5$^{+30}_{-1.7}$ & 9.8 $\\pm0.02$ \\\\ $\\tau_{plasma}$ & - & - & 2.24$\\pm0.01$ &2.28 $\\pm0.01$ \\\\ Norm & -&-& 0.0584$\\pm0.0001$ & 0.062$^{+0.002}_{-0.05}$ \\\\ -\\\\ E$_{Fe XXV}$ (keV) & 6.91$\\pm0.01$ & -& 6.91$^{+0.02}_{-0.01}$&-\\\\ FWHM (keV) & 0.091$^{+0.022}_{-0.049}$ & - & 0.13$^{+0.19}_{-0.04}$ &-\\\\ EW (keV)& 0.021$^{+0.005}_{-0.002}$ & -&$0.040^{+0.007}_{-0.009}$&-\\\\ Norm ($10^{-4}$) & 3.5 $^{+0.8}_{-0.6} $& -&6.0$^{+1.1}_{-1.3}$ &-\\\\ v (km/s) & 9300$^{+500}_{-400}$ & -& 9300$^{+400}_{-800}$&-\\\\ -\\\\ E$_{Fe XXVI}$ (keV) &7.32$^{+0.02}_{-0.06}$& -& 7.30$\\pm$0.02 &-\\\\ FWHM (keV) & 0.081$^{+0.079}_{-0.027}$ & - & 0.25 $^{+0.13}_{-0.01}$&-\\\\ EW (keV) & 0.032 $^{+0.018}_{-0.004}$& -&$0.089^{+0.013}_{-0.014}$&-\\\\ Norm ($10^{-4}$) & 3.4$^{+1.9}_{-0.4}$ & -&11.8$^{+1.7}_{-1.6}$ &-\\\\ v (km/s) &14600$^{+2500}_{-800}$& -& 13800$\\pm800$&-\\\\ -\\\\ N (10$^{22}$ cm$^{-2}$) & - & 0.47$^{+0.17}_{-0.19}$ & - & 0.45$^{+0.33}_{-0.17}$ \\\\ $\\log \\xi$(ergs cm s$^{-1}$) & - & 3.3 $^{+0.2}_{-0.1}$ &-&3.4 $^{+0.2}_{-0.1}$ \\\\ v (km/s) & - & 9600$^{+400}_{-500}$ & - & 9600$\\pm300$ \\\\ -\\\\ N (10$^{22}$ cm$^{-2}$) & - & 1.66 $^{+1.18}_{-0.83}$ & - & 1.97$^{+1.26}_{-0.51}$ \\\\ $\\log \\xi$(ergs cm s$^{-1}$) & - & 3.9$^{+0.5}_{-0.3}$ &-&3.7 $^{+0.3}_{-0.1}$ \\\\ v (km/s) & - & 15400$\\pm400$& - & 15400 $^{+400}_{-300}$\\\\ -\\\\ $\\chi^2/\\nu$ & 2725/3408 = 0.80 & 2731/3408 = 0.80 & 2793/3408 = 0.82 & 2761/3408 = 0.81 \\\\ \\enddata \\label{lines} \\tablecomments{This Table lists the line detections using Gaussian functions as well as more self-consistent, photoionization components created with XSTAR, assuming two different continuum models. {\\it TBabs} is applied to all the models and the errors are 1$\\sigma$ confidence level.} \\end{deluxetable} \\clearpage" }, "1112/1112.3681_arXiv.txt": { "abstract": "{Dark matter particles may be captured by a star and then thermalized in the star's core. At the end of its life a massive star collapses suddenly and a compact object is formed. The dark matter particles redistribute accordingly. In the inelastic dark matter model, an extended dense dark matter mini-halo surrounding the neutron star may be formed. Such mini-halos may be common in the Galaxy. The electron/positron flux resulting in the annihilation of dark matter particles, however, is unable to give rise to observable signal unless a nascent mini-halo is within a distance $\\sim ~{\\rm a~few}~0.1$ pc from the Earth.} ", "introduction": "Dark matter (DM) is a form of matter necessary to account for gravitational effects observed in very large scale structures such as anomalies in the rotation of galaxies and the gravitational lensing of light by galaxy clusters that cannot be accounted for by the quantity of observed matter \\cite[e.g.,][]{lsp,lkp}. The existence of dark matter is a piece of solid evidence for new physics that is beyond the standard model and the relevant researches are the focus of modern physics and astrophysics. Among the hypothetical particles proposed so far, the weakly interacting massive particles (WIMPs) are the leading candidate \\cite{lsp,lkp}. The direct-detection searches look for the signal of WIMP-nuclei scattering in underground detectors. The indirect-detection experiments, instead, aim to catch the annihilation/decay products of DM particles. Currently there are some anomalous signals in favor of WIMPs but the evidence is not conclusive \\cite{atic,pamela,fermi,fan}. The DM particles may be captured by the stars. The capture of WIMPs by the sun, earth and other astrophysical objects has been extensively investigated \\cite{gould1,gould2,press85}. The captured particles will eventually be thermalized and centered in the core of the star, namely, a DM core forms in the center of a star \\cite{peter47}. After the death of a (massive) star, a compact object is formed. The collapse of the star lasts for a very short time and the density distribution of the DM core will be modified. If the compact object is small enough, a significant part of the DM core may be left outside the compact object, i.e., a DM mini-halo is formed. As the outer shell of the stellar remnant eventually disappears, the annihilation products of the DM matter, such as electrons and photons, are visible for the observers. In this work we investigate the formation and evolution of such a mini-halo and calculate the flux of the annihilated electrons/positrons. This work is structured as follows. In section 2 we describe the distribution of WIMPs inside the star. In section 3 we investigate the redistribution of WIMPs in the collapse process. In section 4 we calculate the evolution of the DM mini-halo and in section 5 we estimate the electron/positron flux detectable on the earth. We summarize our results in section 6. ", "conclusions": "In this work we have investigated the formation, evolution and the annihilation of the particles in dark matter mini-halo surrounding the compact object. In our model, the dark matter particles are captured by the stellar matter and then distributed in the core of the star. At the end of its life the star uses up the fuel and a compact object, either a white dwarf, a neutron star or a black hole, is formed. The dark matter particles will redistribute around the newly formed compact object. In the elastic dark matter model and if the elastic cross section is large enough in the inelastic dark matter model, the dark matter particles will be thermalized and no prominent dark matter mini-halo can be formed. In the inelastic dark matter model, however, if the elastic cross section is very small (for example $\\lesssim 10^{-47}~{\\rm cm^{2}}$) an extended dense dark matter mini-halo surrounding neutron star may be formed. Though such mini-halos may be very common in the Galaxy, the electron/positron flux resulting in the annihilation of these dark matter particles is negligible compared with the observational data unless a nascent mini-halo is within $\\sim$ a few $0.1~{\\rm pc}$ to the Earth." }, "1112/1112.4774_arXiv.txt": { "abstract": "The Hubble radius is a particular manifestation of the Universe's gravitational horizon, $R_{\\rm h}(t_0) \\equiv c/H_0$, the distance beyond which physical processes remain unobservable to us at the present epoch. Based on recent observations of the cosmic microwave background (CMB) with WMAP, and ground-based and HST searches for Type Ia supernovae, we now know that $R_{\\rm h}(t_0)\\sim$$13.5$ Glyr. This coincides with the maximum distance ($ct_0\\approx 13.7$ Glyr) light {\\it could} have traveled since the big bang. However, the physical meaning of $R_{\\rm h}$ is still not universally understood or accepted, though the minimalist view holds that it is merely the proper distance at which the rate of cosmic recession reaches the speed of light $c$. Even so, it is sometimes argued that we can see light from sources beyond $R_{\\rm h}$, the claim being that $R_{\\rm h}$ lies at a redshift of only $\\sim$$2$, whereas the CMB was produced at a much greater redshift ($\\sim$$1100$). In this paper, we build on recent developments with the gravitational radius by actually calculating null geodesics for a broad range of FRW cosmologies, to show---at least in the specific cases we consider here, including $\\Lambda$CDM---that no photon trajectories reaching us today could have ever crossed $R_{\\rm h}(t_0)$. We therefore confirm that the current Hubble radius, contrary to a commonly held misconception, is indeed the limit to our observability. We find that the size of the visible universe in $\\Lambda$CDM, measured as a proper distance, is approximately $0.45ct_0$. ", "introduction": "The standard model of cosmology is confronted with several unpalatable coincidences, suggesting that we do not yet have a fully consistent picture of the Universe's dynamical expansion (see, e.g., Melia \\& Shevchuk 2011). Part of the problem is that cosmological observations can only be interpreted from within the context of a pre-assumed model, and the data can be quite compliant. The $\\Lambda$CDM (Cold Dark Matter with a cosmological constant $\\Lambda$) model has been without peer in cosmology (see, e.g., Spergel et al. 2003, and Tegmark et al. 2004). For example, this model has been used with complementary measurements of the cosmic microwave background (CMB) radiation to determine that the Universe is flat, so its energy density $\\rho$ is at (or very near) the ``critical\" density $\\rho_{\\rm c}\\equiv 3c^2H^2/ 8\\pi G$. But among the many peculiarities of this description of the universe is the inference, based on current observations, that the density $\\rho_{\\rm d}$ of dark energy must itself be of order $\\rho_{\\rm c}$. (In the context of $\\Lambda$CDM, the best fit to the WMAP data indicates that dark energy represents approximately $73\\%$ of the total $\\rho\\approx \\rho_{\\rm c}$; see Spergel et al. 2003.) Dark energy is often thought to be the manifestation of the aforementioned cosmological constant, $\\Lambda$, though no reasonable explanation has yet been offered as to why such a fixed, universal density ought to exist at this scale. It is well known that if $\\Lambda$ is associated with the energy of the vacuum in quantum theory, it should have a scale representative of phase transitions in the early Universe---120 orders of magnitude greater than $\\rho_{\\rm c}$. Many workers have attempted to circumvent these difficulties by proposing alternative forms of dark energy, including Quintessence (Ratra \\& Peebles 1988; Wetterich 1988), which represents an evolving canonical scalar field with an inflation-inducing potential, a Chameleon field (see, e.g., Mota \\& Barrow 2004; Khoury \\& Weltman 2004; Brax et al. 2004) in which the scalar field couples to the baryon energy density and varies from solar system to cosmological scales, and modified gravity, arising out of both string motivated, or General Relativity modified actions (Capozziello et al. 2003; Nojiri \\& Odintsov 2003; Carroll et al. 2004), which introduce large length scale corrections modifying the late time evolution of the Universe. The actual number of suggested remedies is far greater than this small, illustrative sample. An equally perplexing puzzle with $\\Lambda$CDM has been dubbed the ``coincidence problem,\" arising from the peculiar near-simultaneous convergence of the matter energy density $\\rho_{\\rm m}$ and $\\rho_{\\rm d}$ towards $\\rho_{\\rm c}$ in the present epoch. Though $\\rho_{\\rm m}$ and $\\rho_{\\rm d}$ are expected to change at different rates as the Universe expands (particularly if dark energy is a cosmological constant) they are nearly equal in the present epoch, implying that we live at a special time in cosmic history. In a recent paper (Melia \\& Shevchuk 2011), we proposed an explanation for yet another disturbing coincidence, having to do with the apparent equality of our gravitational (or Hubble) radius $R_{\\rm h}$ with the distance $ct_0$ light could have traveled since the big bang (in terms of the presumed current age $t_0$ of the Universe). This equality has received some scrutiny in recent years (Melia 2003, 2007, 2009, Melia \\& Abdelqader 2009, van Oirschot et al. 2010; see also Lima 2007 for a related, though unpublished, work). Unfortunately, there is still some confusion regarding the properties of $R_{\\rm h}$ due to a misunderstanding of the role it plays in our observations. For example, it is sometimes suggested (see, e.g., Davis \\& Lineweaver 2004; van Oirschot et al. 2010) that sources beyond $R_{\\rm h}(t_0)$ are observable today, which is certainly not the case. We will therefore begin by elaborating upon what the gravitational radius $R_{\\rm h}$ is---and what it is not. Though first defined in Melia (2007), an unrecognized form of $R_{\\rm h}$ actually appeared in de Sitter's (1917) own account of his spacetime metric. And we will advance the discussion further by actually calculating photon trajectories for various well-studied Friedmann-Robertson-Walker (FRW) Cosmologies, demonstrating that the null geodesics reaching us at $t_0$ have never crossed $R_{\\rm h}(t_0)$. Some come close, and in one case---the de Sitter model---they approach $R_{\\rm h}$ asymptotically as $t$ recedes to our infinite past. Our conclusion in this paper will be that $R_{\\rm h}(t_0)$ is a real limit to our observability at the present time $t_0$. ", "conclusions": "Throughout this paper, we have made a conscious effort to discuss the properties of null geodesics in FRW cosmologies without resorting to conformal diagrams. This approach, also used to great effect by Ellis and Rothman (1993), makes it easier to think in terms of familiar quantities (proper distances and proper time) that are not always straightforward to interpret otherwise. Misconceptions often arise from the misinterpretation of coordinate-dependent effects. In their paper, Ellis and Rothman clearly delineated true horizons from apparent horizons, and extended the definitions, first introduced by Rindler (1956), in a clear and pedagogical manner. In this paper, we have paid particular attention to the gravitational horizon, also manifested as the Hubble radius, which is time-dependent and may or may not turn into an event horizon in the asymptotic future, depending on the equation of state. It is useful at this point to be absolutely clear about how far light has traveled in reaching us. It is quite evident from our results that sources whose light we see today were at a proper distance $R(t_e)\\rightarrow 0$ when $t_e\\rightarrow 0$. For the FRW cosmologies we have considered here (which include $\\Lambda$CDM) the light reaching us today---including that from the recombination region associated with the CMB---has traveled a {\\it net} proper distance of at most $\\sim$0.3--$0.4ct_0$. It is therefore not correct to claim that the size of the visible universe in these cosmologies is $ct_0$ (or even greater in some interpretations). Because all causally connected sources in an expanding universe began in a vanishingly small volume as $t\\rightarrow 0$, the maximum proper distance from which we receive light today must necessarily be less than $ct_0$, since presumably there were no pre-existing sources at a non-zero proper distance prior to $t=0$ with which we were in causal contact. It is true, however, that more and more sources become visible to us as time advances, since for $t>t_0$, the geodesic curves terminating in our future all rise above their current counterparts shown in figures~1,2,3,4, and 5d. In our future, we will see light from sources that radiated at proper distances greater than those shown here. Of course, $R_{\\rm h}(t)$ will also continue to increase, and it is not difficult to convince oneself from Equations~(5) and (8) that the limits of observability will always be $R_{\\rm h}(t)$, since $R_\\gamma(t_e)/R_{\\rm h}(t)$ is smaller than $1$ for all $t_e-1$, Equation~(5) shows that $\\dot{R_{\\rm h}}>0$, and therefore $R_{\\rm h}(t)>R_{\\rm h}(t_{e,\\;{\\rm max}})$ for $t>t_{e,\\;{\\rm max}}$, which also means that $R_{\\rm h}(t)> R_\\gamma(t_{e,\\;{\\rm max}})$ for all $t>t_{e,\\;{\\rm max}}$." }, "1112/1112.2708_arXiv.txt": { "abstract": "% Kuiper Belt object 2007 TY430 is the first wide, equal-sized, binary known in the 3:2 mean motion resonance with Neptune. The two components have a maximum separation of about 1 arcsecond and are on average less than 0.1 magnitudes different in apparent magnitude with identical ultra-red colors ($g-i=1.49\\pm0.01$ mags). Using nearly monthly observations of 2007 TY430 from 2007-2011, the orbit of the mutual components was found to have a period of $961.2\\pm4.6$ days with a semi-major axis of $21000\\pm160$ km and eccentricity of $0.1529\\pm0.0028$. The inclination with respect to the ecliptic is $15.68\\pm0.22$ degrees and extensive observations have allowed the mirror orbit to be eliminated as a possibility. The total mass for the binary system was found to be $7.90\\pm0.21 \\times10^{17}$ kg. Equal-sized, wide binaries and ultra-red colors are common in the low inclination ``cold'' classical part of the Kuiper Belt and likely formed through some sort of three body interactions within a much denser Kuiper Belt. To date 2007 TY430 is the only ultra-red, equal-sized binary known outside of the classical Kuiper belt population. Numerical simulations suggest 2007 TY430 is moderately unstable in the outer part of the 3:2 resonance and thus 2007 TY430 is likely an escaped ``cold'' classical object that later got trapped in the 3:2 resonance. Similar to the known equal-sized, wide binaries in the cold classical population, the binary 2007 TY430 requires a high albedo and very low density structure to obtain the total mass found for the pair. For a realistic minimum density of 0.5 g/cm$^{3}$ the albedo of 2007 TY430 would be greater than 0.17. For reasonable densities, the radii of either component should be less than 60 km, and thus the relatively low eccentricity of the binary is interesting since no tides should be operating on the bodies at their large distances from each other. The low prograde inclination of the binary also makes it unlikely the Kozai mechanism could have altered the orbit, making the 2007 TY430 binary orbit likely one of the few relatively unaltered primordial binary orbits known. Under some binary formation models, the low inclination prograde orbit of the 2007 TY430 binary indicates formation within a relatively high velocity regime in the Kuiper Belt. ", "introduction": "Small solar system bodies, such as asteroids and Kuiper Belt Objects (KBOs), may be leftover remnants of the planetesimals that went into the formation of the terrestrial and giant planets. These small bodies' orbital and physical properties give insight into the Solar System's origins and evolution. KBO binaries are particularly informative about the conditions at the time when they formed, which must have been at an earlier epoch of Solar System history. The Kuiper Belt is dynamically structured with three main dynamical classes (Figures~\\ref{fig:kboeabinary} to~\\ref{fig:kboiabinaryblowup}). 1) Classical KBOs have semi-major axes between about 40 and 50 AU with moderate eccentricities and inclinations. These objects may be regarded as the population originally predicted for the Kuiper Belt (Fernandez and Ip 1981), but they have higher eccentricities and inclinations than expected (Jewitt et al. 1998). The dynamics of the classical KBOs have shown that the outer Solar System has been highly modified through the evolution of the planets (Hahn and Malhotra 2005). There appears to be two subsets of classical KBOs. The low inclination ``cold'' classical population has generally smaller and redder objects compared to the high inclination ``hot'' classical population (Tegler and Romanishin 2000; Brown 2001; Levison and Stern 2001; Trujillo and Brown 2002; Peixinho et al. 2008). These two classical populations likely represent different formation regions from the outer solar system. Through migration and scattering of the planets very early in the Solar System the two classical populations came to reside where we see them now (Gomes 2003; Levison and Morbidelli 2003; Batygin et al. 2011). 2) Scattered disk objects have large eccentricities with perihelia near the orbit of Neptune ($q \\sim 30-45$ AU). The scattered disk objects are likely to have been moved to their current orbits through interactions with Neptune (Gomes et al. 2008). 3) Resonant KBOs are in mean motion resonances with Neptune and generally have higher eccentricities and inclinations than classical KBOs. Most of the resonance objects were likely captured into their resonances from the outward migration and circularization of Neptune's orbit (Malhotra 1995; Levison et al. 2008). Where the resonance objects came from and how and when this capture occurred is still highly debated. The 3:2 resonance, called Plutinos since Pluto is in this resonance, appears to be the most populated resonance (Jewitt et al. 1998; Chiang and Jordan 2002; Sheppard et al. 2011). The other resonances with a significant number of known objects are the outer mean motion resonances 5:3, 7:4, 2:1 and 5:2 (Elliot et al. 2005; Hahn and Malhotra 2005; Gladman et al. 2008). Binary small bodies are a particularly good way to obtain clues about the Solar System's past. Remarkably, the small bodies in the various Solar System reservoirs appear to have significantly different binary characteristics (Richardson and Walsh 2006; Noll et al. 2008a). In the main asteroid belt, most secondaries are close to and significantly smaller than the primaries (Merline et al. 2002; Margot 2002; Marchis et al. 2006b). This is indicative of formation through direct collisions. The few binaries known in the Jupiter Trojan population hint at these objects being of very low density, implying they may be similar to comets and Trans-Neptunian Objects (TNOs) (Marchis et al. 2006a). Starting with the discovery of the first KBO binary other than Pluto, it has become apparent that most of the low inclination classical Kuiper Belt binaries have large separations and similar sized components (Veillet et al. 2002; Noll et al. 2008a). Because of the large angular momentum in these equal-sized, wide binary systems, they cannot be formed by direct collisions. Direct formation by gravitational collapse, the dynamical friction of a sea of KBOs, three body interactions, some other collisionless interactions, or a combination of these are the best viable scenarios for binary formation in the classical Kuiper Belt (Nesvorn{\\'y} et al. 2010; Goldreich et al. 2002; Weidenschilling 2002; Funato et al. 2004; Astakhov et al. 2005; Lee 2007; Schlichting and Sari 2008a; Gamboa Suarez et al. 2010). The cold classical disk also has a high fraction of binaries, irrespective of separation and difference in magnitude of the individual components, compared to the hot classical, resonant and scattered populations, about 30 percent versus 5 percent binary fraction, respectively (Stephens and Noll 2006; Noll et al. 2008b). Unlike the low inclination classical KBOs, many of the other known KBO binaries do not have such equal-sized and wide components. The different types of binary formation mechanisms proposed will each tend to create a particular subset of satellite semi-major axes, eccentricities, inclinations and secondary diameters (Kern and Elliot 2006a; Noll et al. 2008a). Thus, understanding the orbital elements and physical characteristics of the components of a binary is essential to constraining how the binary may have formed. Understanding the formation mechanism of binaries gives insight into the original solar nebula and the collisional environment in the distant past. The two distinct binary populations seen in the cold and hot dynamical classes are strong evidence of a different history. For example, Parker and Kavelaars (2010) use the fragile nature of the wide binaries in the cold classical region as evidence to show that cold classical KBOs were not likely to have been implanted dynamically, as suggested by Levison et al. (2008). Binaries also likely played an important role in planet formation by acting as a heat source in the planetesimal disk, tranforming gravitational potential energy into kinetic energy of planetesimals in the disk (Perets 2011). 2007 TY430 is a unique binary KBO because it is the first known equal-sized, wide binary in the 3:2 mean motion resonance with Neptune, which is the innermost well populated Neptune resonance. The other known 3:2 resonance binary objects (including Pluto, Orcus, (47171) 1999 TC36 and (208996) 2003 AZ84) have significantly smaller and relatively closer secondaries (magnitude differences greater than 2 mags) that are more indicative of formation through direct collisions (Stern 2002; Brown et al. 2006; Canup 2011). Because 2007 TY430 indicates a different binary formation mechanism operating within the 3:2 resonance population, 2007 TY430 was extensively observed over the past several years in order to determine the orbital and physical properties of its two components. ", "conclusions": "The ultra-red color and wide equal-sized binary nature of 2007 TY430 makes it very similar to the unique characteristics of the low inclination classical Kuiper Belt. This strongly favors the argument that 2007 TY430 originated as a typical low inclination ``cold'' classical Kuiper Belt object (assuming that the cold classical population forms in situ). Weak chaos over long timescales (or stronger chaos on faster timescales) slowly removed 2007 TY430 from the cold classical region into a dynamically excited orbit in the scattered disk (see Volk and Malhotra 2011). Eventually 2007 TY430 became trapped in the unstable outskirts of the 3:2 resonance where it is seen today. During this process, the binary remained stable. This would suggest that 2007 TY430's formation environment is likely the cause of its ultra-red color and wide binary nature and not its current orbital characteristics. This bolsters the argument of Benecchi et al. (2009) who suggest that similar colors for binary components indicate common formation conditions, regardless of present location. \\subsection{Stability of 2007 TY430} The Hill sphere of an object orbiting the Sun is the region where satellites are generally stable around the object, given by: \\begin{equation} r_{H} = a_{p} \\left[\\frac{(m_{p1}+m_{p2})}{3M_{\\odot}}\\right]^{1/3} \\label{eq:hill} \\end{equation} where $a_{p}$ is heliocentric distance of the object in AU, $M_{\\odot}$ is the mass of the Sun and $m_{p1}$ and $m_{p2}$ are the mass of the first and second components of the binary, respectively. The Hill radius for 2007 TY430 is about $3.015\\times 10^{5}$ km. Using the heliocentric semi-major axis of 2007 TY430 from Table 1 and a total mass of $7.90\\pm0.21 \\times10^{17}$ kg for the binary, gives $a_{bin}/r_{H}$ of about 0.07 for 2007 TY430. While this is still well within the stable region of orbital phase space, it is among the widest known binaries (Nesvorn{\\'y} et al. 2003, Nicholson et al. 2008, Noll et al 2008a). Widely separated binaries were probably much more common in the past as they are unstable over the age of the solar system from dynamical perturbations (Petit and Mousis 2004; Noll et al. 2006, 2008a; Petit et al. 2008; Parker and Kavelaars 2010). Its possible the cold classical KBOs have had lower numbers of collisions or fewer planet interactions allowing the largest population of surviving wide, equal-sized binaries in the Kuiper belt. 2007 TY430 has similar separation and component sizes as the cold classical KBOs 1998 WW31 and 2000 CF105 (Table 5). Petit and Mousis (2004) determined these two binary KBOs were only stable for about 1-2 Gyrs. The Petit and Mousis (2004) result is likely a lower limit on the binary lifetimes since they used a steeper size frequency distribution for the smaller Kuiper belt objects than currently believed to exist, allowing for more small impactors to disrupt the binaries than is currently likely (Fraser et al. 2008; Fuentes and Holman 2008; Fuentes et al. 2009; Fraser 2009). The 2007 TY430 binary probably has a similar timescale of stability, though 2007 TY430 is in the Plutino population and not the low inclination classical belt. The timescales of destruction for wide binary TNOs by Parker and Kavelaars (2010) also suggest 2007 TY430 would not have a high probability for surviving for the age of the Solar System, and thus 2007 TY430 could just be one of a few binaries left of a once much larger population of binaries. \\subsubsection{2007 TY430: The Case for a Primordial Binary Orbit} The inclinations of many TNO binary components are generally high (Naoz et al. 2010). The Kuiper Belt binary 2001 QW322 was found to have a fairly low eccentricity ($e<0.4$) yet a very large inclination ($i\\sim125$ degrees) and semi-major axis ($\\sim 120000$ km) (Petit et al. 2008). Because of 2001 QW322's large inclination, its relatively low eccentricity and large semi-major axis can be explained by the Kozai mechanism (Perets and Naoz 2009). Most binary minor planets do not appear to have primordial orbits as they could have been affected by either tidal forces or have high eccentricities or high inclinations at which the Kozai mechanism may operate (Ragozzine 2009, Naoz et al. 2010). The binary orbit of 2007 TY430 is compared to other known equal-sized binary orbits in Figures ~\\ref{fig:binarymutualrhi} and ~\\ref{fig:binarymutualrhe}. 2007 TY430 does not have orbital elements similar to other TNO binaries shown in a similar Figure 3 of Naoz et al. (2010), which compares the normalized separations of the binary components to their eccentricities. The $log(a_{bin}/r)$ for 2007 TY430 is slightly greater than 2.5, but still in a stable region of the Naoz et al. Figure 3. Interestingly, when including recent results from Parker et al. (2011), it appears most of the equal-sized binaries with very large semi-major axes appear to have moderate to low eccentricities and low inclinations, making them unsusceptible to the Kozai resonance and tides. This is in contrast to many of the lower semi-major axis, equal-sized binaries that appear susceptible to the Kozai mechanism because of their large mutual inclinations and higher eccentricities. This result is likely because the large semi-major axis objects would become unstable to perturbations if their eccentricities or inclinations were too large. 2007 TY430 is the lowest eccentricity, wide component, equal-sized binary known. Because the mutual orbital elements of the 2007 TY430 binary has low inclination, low eccentricity and large semi-major axis (Figures~\\ref{fig:binarymutualrhi} and ~\\ref{fig:binarymutualrhe}), the Kozai mechanism and tidal interactions have not likely modified the primordial orbit of 2007 TY430 (Chyba et al. 1989; Murray and Dermott 1999; Perets and Naoz 2009). Modification of the 2007 TY430 binary could have occurred over the age of the solar system from direct collisions on either of the two components, from relatively massive bodies passing within the Hill sphere of the two components or interactions with the giant planets (Petit and Mousis 2004; Nesvorn{\\'y} et al. 2011). The high eccentricities and inclinations of other TNO binary components suggest formation in a dense collisional environment. In this environment, gravitational encounters can create these high eccentricities and inclinations along with secular Kozai effects and tidal evolution (Naoz et al. 2010). Other binaries observed to date could have formed in a similar manner as 2007 TY430, but later three body encounters or direct collisions after binary formation changed the binary components orbits to be highly inclined and eccentric. It is thus possible that 2007 TY430 simply has escaped any significant collisions or gravitational encounters while other known binaries have not or formed in a different collisional environment. This could be because 2007 TY430 diffused out of the low inclination classical belt early on and the Plutino orbit made it less susceptible to collisions. Thus 2007 TY430's binary orbit could be more primordial than many of the other known equal-sized wide binary objects observed to date. \\subsection{Formation of the 2007 TY430 binary system} The proposed binary formation by Funato et al. (2004) from exchanged binaries is unlikely since TNOs appear to lack the many high-eccentricity binaries suggested by this mechanism (Naoz et al. 2010), of which 2007 TY430 is another example of a low eccentricity binary. Weidenschilling (2002) proposed a hybrid mechanism of a direct collision between two bodies while in the Hill Sphere of a third body, which can directly form large separation binaries. This mechanism would only be likely to occur during the formation of the Kuiper Belt, as many more large bodies than currently observed are required as well as low velocities are needed. A similar mechanism, the L$^{3}$ mechanism proposed by Goldreich et al. (2002), has a third body strongly interacting with two other bodies while they each are within the other body's Hill sphere. The L$^{3}$ mechanism should dominate over the Weidenschilling (2002) mechanism because strong gravitational interactions should occur much more frequently than the actual collisions of objects. Chaos assisted capture (Astakhov et al. 2005, Lee et al. 2007), is effectively the same as the L$^{3}$ mechanism; in chaos assisted capture, two bodies are temporarily trapped in their mutual Hill Spheres and can become permanently trapped when a third 'intruder' body is scattered by the pair. This mechanism would create wide separation binaries with equal-size and moderate eccentricity. A second binary formation mechanism proposed by Goldreich et al. (2002), the L$^{2}$s mechanism, involves two objects forming an unbound transient binary that becomes bound with the aid of the dynamical friction from a sea of small objects. Any binary formation that involves dissipation of energy in a smooth and gradual manner, like the L$^{2}$s mechanism, will likely form retrograde binaries (Schlichting and Sari 2008b). Because of 2007 TY430's prograde orbit, it is unlikely to have formed by the L$^{2}$s mechanism. Unlike the L$^{2}$s mechanism, the L$^{3}$ mechanism of Goldreich et al. (2002) should form equal populations of prograde and retrograde binaries (Schlichting and Sari 2008). So 2007 TY430 could have formed through the L$^{3}$ mechanism which would only be likely if the Kuiper Belt objects had relatively large velocities, called super-Hill velocities (Schlichting and Sari 2008a, 2008b). The Hill velocity (Rafikov 2003; Goldreich et al 2004; Murray-Clay and Chiang 2006; Lee et al. 2007) is defined as: \\begin{equation} v_H = \\left[\\frac{G(m_{p1}+m_{p2})}{r_{H}}\\right]^{1/2} \\label{eq:hill} \\end{equation} where $G$ is the gravitational constant. For 2007 TY430, the Hill velocity is about 0.4 m s$^{-1}$ which is much less than the Keplerian velocity of its Plutino type orbit that is about 4 km s$^{-1}$. Formation through the L$^{3}$ mechanism would require large velocities between KBOs that several authors do not think were prevalent in the early Kuiper Belt (Goldreich et al. (2002), Schlichting and Sari (2008b), Murray-Clay and Schlichting (2011)). Goldreich et al. (2002) suggest the velocities of KBOs were about a third of the Hill velocity, which would make retrograde orbits through the L$^{2}$s mechanism the dominant binary formation mechanism. This is because binary formation efficiency from gravitational encounters decreases significantly with higher relative velocities since the sphere of influence or time an object affects another is decreased (Noll et al. 2008a; Schlichting and Sari 2008b). Murray-Clay and Schlichting (2011) argue that sub-Hill velocities and the binary formation from dynamical friction (L$^{2}$s mechanism) is the best mechanism for the equal-sized binary formations. 2007 TY430 does not fall in along these formation lines because it is not retrograde, meaning large velocities during formation were likely and thus dynamical friction unlikely to be the capture mechanism. A weakness to the high velocity scenario is that Schlicting and Sari (2008b) suggest that the velocity would need to be finely tuned and stay in such a state for a considerable amount of time in order for a high velocity regime to form these prograde, equal-sized binaries. To date, not enough equal-sized binary orbital inclinations are known to draw a strong conclusion. Only 17 equal-sized KBO binaries have known inclinations (Figure~\\ref{fig:binarymutualrhi}). Of these five are retrograde and twelve are prograde, but three of the prograde objects have $\\Delta Mag>1$ mag and thus are not quite equal-sized binaries. Several authors have suggested that the observed inclinations of TNO binaries are consistent with them being randomly distributed (Noll 2003, Chiang et al. 2006, Grundy et al. 2011). This suggests the Kuiper Belt disk was in the dispersion-dominated (dynamically hot) regime during binary formation (see Stewart and Ida 2000; Collins and Sari 2006; Schlichting and Sari 2008b). Lee et al. (2007) point out there are more prograde than retrograde binaries and determine this as a sign of how the binaries formed. Schlichting and Sari (2008b) suggest this sign is showing the velocity regime in which the binaries formed. Schlichting and Sari (2008b) predicted that over 97 percent of binaries with comparable masses will have retrograde orbits if the relative velocities of the Kuiper Belt objects were low. Since it appears to be well below 50 percent, the relative velocities must have been significantly higher, as discussed above for the formation of 2007 TY430. Binary formation from direct gravitational collapse of an over density of concentrated cm to meter sized solids was suggested by Nesvorn{\\'y} et al. (2010). This mechanism appears to be able to produce a wide range of distant equal-sized binaries with a large range of eccentricities with most having prograde inclinations. This mechanism could explain 2007 TY430, but may have trouble producing the retrograde binaries found to date. \\subsection{Plutino Resonance and the Cold Classical Belt} The Neptune mean motion resonance populations could have been emplaced by one of two favored mechanisms. 1) The slow migration outwards of Neptune by over 10 AU from its formation location from planetesimal scattering would have allowed Neptune to sweep many objects into the resonances we see today (Malholtra 1995; Hahn and Malhotra 2005). This scenario would likely mean that the inner 3:2 resonance would have a significantly different population of objects than the more outer resonances that would have swept through the classical Kuiper Belt population. This scenario would also suggest the resonance populations should have a cold component representative of the cold classicals along with a hot component representing the objects from closer in. 2) The chaotic population of the resonances could have occurred if Neptune was scattered with a relatively large eccentricity out from near the current Saturn region to near its present semi-major axis (Levison and Morbidelli 2003; Tsiganis et al. 2005; Levison et al. 2008). As Neptune circularized its orbit through interactions with planetesimals, objects in the resonance areas would become trapped. In this scenario, the resonance populations would be more uniform and not have a cold component, unlike the slow migration scenario. The resonance populations should have a cold component if Neptune experienced extensive slow, smooth migration. This cold component should be similar in characteristics to the low inclination classical Kuiper Belt, that is, equal-sized binaries should be common as well as ultra-red material (Murray-Clay and Schlichting 2011). This cold component should also have lower eccentricities compared to other objects, meaning there could be a correlation between binaries, color and eccentricity as well as with inclination in the resonance populations. In the Levison et al. (2008) scenario, Chaotic scattering of Neptune and then circularization, the resonance populations should match the scattered disk in characteristics. No obvious cold component is observed in the 3:2 resonance population and the statistics are too low for the other resonance populations to determine if there is a cold component (Murray-Clay and Schlichting 2011). The 3:2 having no cold component is still consistent with slow, smooth migration since the 3:2 does not sweep through the cold classicals (Murray-Clay and Schlichting 2011). Thus the 3:2 resonance is not as strong as a marker as the other resonances, which should have cold components as they swept through the classical region of the Kuiper Belt. It is possible that sweeping of the $\\nu_{8}$ resonance through the 3:2 resonance and inner classical Kuiper belt cleared out much of these regions (Petit et al. 2011). If this is true, the $\\nu_{8}$ resonance likely did not sweep through until after the formation of any equal-sized binaries since the binaries are unlikely to have formed as the density of such objects would have been too low. 2007 TY430 has a moderately inclined and eccentric heliocentric orbit, and thus would be considered part of the hot component of the 3:2 resonance on dynamical grounds. Although 2007 TY430 could easily have been emplaced recently by dynamical scattering and resonance sticking, it is interesting to speculate on the meaning of this cold-classical like binary if it was placed in the resonance primordially. As discussed above, if 2007 TY430 was captured from an originally low eccentricity, low inclination orbit by slow, smooth migration, it would have originated around 31 AU by conservation of Brouwer's integral. This would suggest that the primoridal cold classical region extended much further inward than is seen today, but the observational evidence for a cold region of the 3:2 resonance is insufficient (Murray-Clay and Schlichting 2011). If its dynamical history was well-known, the relatively high eccentricity and inclination of 2007 TY430 could have been seen as an indication that the slow, smooth migration model is insufficient. More binary objects need to be discovered in the 3:2, but the existence of 2007 TY430 in the hot region of the 3:2 and no other known equal-sized 3:2 resonance objects, would favor chaotic scattering by Neptune over slow smooth migration. \\subsection{Other Known Equal-Sized Wide Binaries} Table 5 shows the other known equal-sized TNO binaries. Here, equal-sized is defined as the components having less than 1 magnitude difference between components. This magnitude difference corresponds to mass ratios less than 4, assuming similar albedos and densities for the two components and the lack of significant rotational brightness variations. As shown in Table 5 and Figures~\\ref{fig:kboeabinary} to~\\ref{fig:kboiabinaryblowup}, the equal-sized binary population is mostly in the low inclination classical belt. Interestingly, most of the equal-sized binary TNOs not in the low inclination classical belt appear to be in mean motion resonances. We see no obvious increase in the binary population between the semi-major axes of 43.5 and 44.5 AU, which has been called the ``Kernel'' area by Petit et al. (2011) as this region appears to be somewhat more dense than the rest of the classical Kuiper belt (Figures~\\ref{fig:kboeabinaryblowup} and~\\ref{fig:kboiabinaryblowup}). It does appear that there is an absence of equal-sized binaries around 43.5 AU, which is near the 7:4 resonance, as well as very few with eccentricities greater than 0.10. The colors of the equal-sized binaries are not all ultra-red, as many are near neutral in color (Tegler and Romanishin 2000; Trujillo and Brown 2002; Gulbis et al. 2006; Peixinho et al. 2008; Petit et al. 2008; Benecchi et al. 2009; Sheppard 2010). As Figure~\\ref{fig:binarycolors} shows, the only equal-sized binary that is ultra-red in color ($S>25$ as defined in Sheppard (2010)) and not in the classical population is 2007 TY430 (colors are from the database described in Hainaut and Delsanti 2002). In all, only 1 of 6 equal-sized binaries with known colors outside of the classical population has an ultra-red color, which is 2007 TY430 (Table 5). In contrast 13 of 17 equal-sized binaries in the classical population have spectral gradients near or within the ultra-red color region ($S>20$). Numerically integrating the non classical equal-sized binaries from Table 5 (2000 QL251, 2000 FE8, 2006 SF369, 1998 WV24, 2000 CM114, 2001 QC298) with ten clones each for 500 MYr found all but 2000 CM114 were fairly stable. Some of the clones of 2000 QL151 ($6/10$, 1 a little chaotic) and 2006 SF369 ($10/10$, 3 a little chaotic) were experiencing Kozai oscillations (not unusual for resonant objects). Thus only 2000 CM114 is likely to have come from the classical region, but its inclination is currently very large and it is not ultra-red in color. Since all of the equal-sized binaries outside the classical belt, except for 2007 TY430, do not have ultra-red colors, it suggests that equal-sized binaries formed in multiple locations and not just in the cold classical belt. As shown in Figure~\\ref{fig:binarycolors}, the inner classical Kuiper belt is likely an extension of the main classical Kuiper belt as both have low inclination, ultra-red, equal-sized binaries. If the ultra-red, equal-sized binary of 1999 OJ4 formed in the inner classical belt, where we see it today, it must have formed before any clearing through the $\\nu_{8}$ resonance sweeping mechanism since the desnity of KBOs would have been to low for likely equal-sized binary formation after any significant clearing of the region." }, "1112/1112.0011_arXiv.txt": { "abstract": "In this proceeding, we show that when we combined WMAP and the most recent results of XENON100, the invisible width of the Higgs to scalar dark matter is negligible($\\lesssim 10 \\%$), except in a small region with very light dark matter ($\\lesssim 10$ GeV) not yet excluded by XENON100 or around 60 GeV where the ratio can reach 50\\% to 60\\%. The new results released by the Higgs searches of ATLAS and CMS set very strong limits on the elastic scattering cross section. ", "introduction": "Two of the most important issues in particle physics phenomenology are the nature of the dark matter and the mechanism to realize spontaneously the electroweak symmetry breaking of the Standard Model (SM). The observations made by the WMAP collaboration show that the matter content of the universe is dark, making up about 85 \\% of the total amount of matter whereas the XENON collaboration recently released its constraints on direct detection of Dark Matter . These constraints are the most stringent in the field nowadays, and begin to exclude a significant part of the parameter space of the Weakly Interacting Massive Particle (WIMP) paradigm. On the other front, the accelerator collaborations ATLAS, CMS and D0/CDF \\cite{ATLAS} have obtained important results concerning the Higgs searches. . It is obvious that the Higgs hunting at LHC is intimately linked with measurement of elastic scattering on nucleon, especially in Higgs-portal like models where the Higgs boson is the key particle exchanged through annihilation/scattering processes. It has already been showed recently that a combined LEP/TEVATRON/XENON/WMAP analysis can restrict severely the parameter space allowed in generic constructions \\cite{Mambrini:2011pw}. In this work, we apply such analysis in the specific context of a scalar singlet dark matter extension of the Standard Model and show that most of the region allowed by WMAP will be excluded/probed by LHC and XENON100 by the end of next year. ", "conclusions": "" }, "1112/1112.2478_arXiv.txt": { "abstract": "% We consider the evolution of an initially FLRW universe after the formation of a long, straight, cosmic string with arbitrary tension and mass per unit length. The birth of the string sources scalar and tensor-type perturbations in the background metric and both density and velocity perturbations in the background fluid, which compensate for the string mass and maintain energy conservation. The former generate the deficit angle within the light cone of the string and a gravitational shock front at the cosmological horizon, whereas the latter are confined within the sound cone. We study the properties of the metric within each region of the resulting spacetime and give the explicit coordinate transformations which demonstrate non-violation of causality. This paper generalizes the work of previous studies for the Nambu-Goto string. ", "introduction": "Cosmic strings are linear concentrations of energy which may have formed during symmetry-breaking phase transitions in the early Universe \\cite{Nielsen_Olesen, hep-ph/9411342, VS1994,hep-th/0508135v2, Preskill}. Though the string width is determined by the inverse of the symmetry-breaking energy scale, this is small compared to cosmological distances, and they may be approximated as one-dimensional objects for many purposes \\cite{Goto1,Anderson2003}. Specifically, strings may have been produced at the epoch of electro-weak symmetry-breaking \\cite{Nambu1}, or the GUT scale, but their formation is also a generic feature of the phase transitions inherent in many extensions of the standard model, (c.f. \\cite{hep-ph/0308134} and references therein). In field theory they are a type of topological defect, analogous to the magnetic flux tubes and other vortex-type defects created in certain condensed matter systems \\cite{Abrikosov1, Zurek1, Zurek2, Zurek3, Bowick1, Williams1, Hendry1, Chuang1, Annett1}, and are produced via the Kibble mechanism \\cite{ICTP/75/5}, if the vacuum manifold ($\\mathcal{M}$) possesses a nontrivial first homotopy group (e.g. $\\pi_1(\\mathcal{M})=\\mathbb{Z}$, in which each element, a nonzero integer, corresponds to an allowed winding number for the string vortex cross-section). In recent years, the formation of ``cosmic\", i.e. horizon-sized, fundamental strings ($F$-strings) and one-dimensional $D$-branes ($D$-strings) has also been extensively studied in string theory \\cite{Witten:1985, Polchinski_Intro, CMP_FD1,CMP_FD2,0911.1345v3, 0811.1277v1, hep-th/0505050v1, astro-ph/0410073v2, Sakellariadou:2009, Rajantie:2007, Copeland:2005}, particularly in the context of brane inflation (see \\cite{Cline1, Gasperini1, Tye1, Carroll:TASI, Quevedo:Lectures, Danielsson1} for reviews), in which such defects can be copiously produced \\cite{Sarangi1, Jones_etal1, Pogosian_Obs1}. However, regardless of the precise details of individual models, the main phenomenological, and observationally relevant, parameters that characterize the string are the energy of the per unit length, which we denote $U$, and the tension, $\\mu = qU$. \\footnote{This is not strictly true in the case of models containing compact extra-dimensions in which the intercommuting probability for strings which ``cross\" in the noncompact space can be much less than unity ($P<<1$). In this case the dynamics of string networks and their implications for cosmological observations can be substantially altered \\cite{Sarangi1,Jones_etal1,Pogosian_Obs1,Avgoustidis:2004,Copeland:2005}. Extra-dimensional scenarios, especially those motivated by, or directly embedded into, string theory also allow for the formation of cosmic ``necklaces\" of various kinds, which introduce significant alterations to the network dynamics along with additional phenomenological parameters, such as the bead mass \\cite{Leblond:2007, Dasgupta1, Martins1, Matsuda2, Lake1, Lake3}. The analogues of a superstring necklaces in field theory are series of monopoles connected by string-segments, which may also be formed in some symmetry-breaking models \\cite{Siemens1, Blanco-Pillado1}. The authors hope to address the important question of the inborn metric for a cosmic necklace in a future article.} The Nambu-Goto (NG) action represents the simplest model and enforcing Lorentz symmetry along the string corresponds to setting $q=1$. If the NG string is wiggly at microscopic scales, it may be approximated macroscopically by a simpler embedding with an effective mass per unit length which differs from the intrinsic tension. Thus in the coarse-grained limit, small-scale structure, even on pure NG strings, can lead to a value of $q$ different from unity \\cite{vilenkin,carter}. For current carrying strings, the mass-density and tension may differ, as originally shown by Witten \\cite{witten,FPRINT-92-39}, so that $q \\neq 1$ in general. \\\\ \\indent The gravitational field around a static, straight string exhibits an interesting feature not seen in fields surrounding spherically-symmetric distributions of matter. In a perturbative analysis of a string in the wire approximation (to first order in $G\\mu$), Vilenkin \\cite{167218} showed that test particles within the string light-cone experience a gravitational acceleration along the radial direction equal to $2GU(q-1)/r$, where $r$ is a distance from the string, and that the surrounding space has a conical structure with deficit angle $4\\pi (1+q)GU$. For the NG string therefore, the gravitational force vanishes and the spacetime becomes locally Minkowski. These results were found to hold in a nonperturbative analysis (i.e. to all orders in $G\\mu$), independently, by Gott \\cite{Gott1} and Hiscock \\cite{Hiscock1}, even for a string of finite width. One observational consequence is the formation of double images from bright sources lying behind the string in relation to an observer. The images have exactly the same shape and brightness and for NG strings their angular separation on the sky is of the order of the deficit angle, though for $q>>1$ it may be substantially smaller \\cite{Pe94,Uz01}. This feature has been used to search for long strings \\cite{Gott1, Vilenkin_lensing1,Hogan_lensing1,Shlaer_lensing1,Huterer_lensing1,Hindmarsh_lensing1, Dyda_lensing1,Gasperini_lensing1, Morganson_lensing1, deLaix_lensing1} and loops \\cite{Cowie_lensing1, deLaix_lensing2, Mack_lensing1} via astronomical observations. Bounds on phenomenological parameters such as the string tension and number density have also been obtained using the predicted effects of strings on CMB anisotropy \\cite{Zeldovich_CMB1, Kaiser_CMB1,Stebbins_CMB1,Gangui_CMB1,Allen_CMB1,Perivolaropoulos_CMB1,Benabed_CMB1,Landriau_CMB1, Landriau_CMB2, Pogosian_CMB1, Wyman_CMB1, Battye_CMB1, Battye_CMB2, Contaldi_CMB1, Bevis_CMB1, Bevis_CMB2, Bevis_CMB3, Bevis_CMB4,Suyama_CMB1,Suyama_CMB2}, the expected emission of high-energy rays from cusps and kinks \\cite{Bhattacharjee_CosRay1, Sigl_CosRay1, Berezinsky_CosRay1, Wichoski_CosRay1} and possible signatures in the 21cm line \\cite{Khatri_21cm}. Interestingly, it has recently been proposed that cusp-emission from superconducting strings may be responsible for anomalous gamma-ray burst observations \\cite{Gr09,Ta09,Sa09} and there remains ongoing controversy on that point (c.f. \\cite{Ba87,Pa88,Br93,Pl94,Be01,Be04,Ch10,Wa11,Ch11} and references therein). With the advent of the Planck experiment \\cite{Planck}, it is also hoped that the polarization of CMB $B$-modes caused by string networks in the early universe could confirm, or rule out, the existence of GUT-scale strings in the near future \\cite{Seljak_Pol1,Pogosian_Pol1,Pogosian_Pol2,Bevis_Pol1,Garcia_Pol1}. Finally, gravitational radiation from strings has been extensively studied \\cite{Vachaspati_GravRad1,Hindmarsh_GravRad1,Darmour_GravRad1,Darmour_GravRad2,Darmour_GravRad3} and future gravitational wave detections from experiments such as LISA/NGO \\cite{LISA/NGO} or LIGO \\cite{LIGO} may place even stronger bounds on string parameters. \\\\ \\indent It is interesting to consider how the conical structure of the spacetime surrounding the string evolves from the initial Friedmann-Lema\\^\\i tre-Robertson-Walker (FLRW) background, which exists before the symmetry-breaking epoch, and a detailed analysis of this problem was originally given by Magueijo \\cite{Magueijo:1992tt,mag-unpublish} for the case of an NG string. He introduced a phenomenological string energy-momentum tensor which is localized at the origin and appears only after the symmetry-breaking time. The generation of string mass-energy is compensated by a decrease in the mass-energy of the perfect fluid that dominates the Universe, so that conservation of energy and momentum are not violated. \\footnote{In previous analyses this was not always the case, though particle production by cosmic strings was studied in detail in the following sources, \\cite{Parker:1987qx,Sa88,Da88}.} Inclusion of the background fluid is therefore compulsory and the problem of string generation must be considered on an (initially) FLRW background. In his analysis, Magueijo obtained solutions of the linearized Einstein equations on the FLRW background metric by assuming the dimensionless parameter $GU$ to be small and constant. This is consistent with known observational bounds on the string tension in various cosmological models, which give an upper limit of $GU \\leq \\mathcal{O}(10^{-6})$ for field-theoretic strings \\cite{VS1994, Bevis_CMB1,Bevis_CMB2,Bevis_CMB3,Bevis_CMB4} and $10^{-11} \\leq GU \\leq 10^{-6}$ for the warped tension of cosmic $F$/$D$-strings \\cite{Jones_etal1,Pogosian_Obs1} (see also \\cite{Hindmarsh_rev} for a review of observational constraints up to 1990). He showed that, at late times and in the region close to the string, the spacetime takes the form of that surrounding an eternal string having deficit angle $8\\pi GU$ \\cite{167218,Gott1,Hiscock1} and that a gravitational shock propagates along the lightcone, with the spacetime outside the horizon remaining unperturbed so that the FLRW metric remains valid, though detailed calculations confirming this picture remain unpublished \\cite{mag-unpublish}. \\\\ \\indent In this paper, we reconsider this issue and also extend the original analysis by considering the more general case of $q \\ne 1$. This allows us to include the physically interesting class of ``conical\" spacetimes which are not locally flat and to consider the gravitational acceleration of test particles in the vicinity of a noneternal string. We follow the same approach taken in \\cite{Magueijo:1992tt} to evaluate the metric perturbations induced by string formation and find that the general picture remains qualitatively similar. However, as shown in detail in Sec.~IV, our solutions do not exactly match those obtained previously, even for $q=1$. We believe that this is because of the choice of boundary condition for the Poisson equation governing the gravitational potential given in \\cite{Magueijo:1992tt} is not consistent with trace part of the Einstein equations. We also provide concrete coordinate transformations both for inside and outside the sound horizon of the perfect fluid, as well as in the region between the sound horizon and the edge of the light cone, which result in an FLRW metric beyond the causal horizon and the metric for an eternal string deep inside the sound horizon, respectively, in accordance with previous results \\cite{167218,Gott1,Hiscock1,Magueijo:1992tt,mag-unpublish}. \\\\ \\indent The structure of the paper is as follows: In Sec.~II we define the perturbation variables and decompose them into scalar and tensor components. In Sec.~III we derive the master equations for each type, while analytic solutions are given in Sec.~IV together with a comparison of the corresponding solutions obtained in \\cite{Magueijo:1992tt}. Section V is devoted to studying the properties of the resulting class of spacetimes and a brief summary of our results and discussion of the prospects for future work is given in the conclusion, Sec.~VI. ", "conclusions": "We derived the analytic form of the metric perturbations on an FLRW background, excited by the formation of a straight cosmic string under the wire approximation. Our results are valid for a general string with differing tension and energy per unit length. For the Nambu-Goto string, our metric does not coincide with that given in previous work due to a differing choice of boundary conditions for the Poisson equation. We found that, as a result of this choice, the solutions given in \\cite{Magueijo:1992tt} for the gravitational potential and curvature perturbation variables do not satisfy the trace part of the Einstein equations. \\\\ \\indent By performing the appropriate coordinate transformation, we explicitly verified that the spacetime outside the light cone remains unperturbed and governed by the FLRW metric. Whereas it is nontrivial to demonstrate nonviolation of causality in the original metric, it is manifestly respected in the new coordinate system. At the cosmological horizon, we found that there exists a gravitational shock giving rise to a divergent tidal force along both the angular direction and the direction of the string axis, $z$. The shock distorts a small ring of test particles placed in the $\\theta-z$ plane by stretching it along $\\theta$-direction and shrinking it along the $z$-direction. The degree of divergence is weak, so that the resultant distortion remains finite. Inside the cosmological horizon, the metric perturbations exhibit self-similarity at late time, i.e. they depend only on a ratio of the radial coordinate to the conformal time. This is reasonable since the Hubble distance is the only characteristic length scale in the system. Deep inside the horizon, we found that the metric tends to that for an eternal string at late times, which has been used to search for the signatures of cosmic strings in astronomical observations." }, "1112/1112.5599_arXiv.txt": { "abstract": "{We constrain the energy at which the transition from Galactic to extragalactic cosmic rays occurs by computing the anisotropy at Earth of cosmic rays emitted by Galactic sources. Since the diffusion approximation starts to loose its validity for $E/Z \\gtrsim 10^{16-17}$\\,eV, we propagate individual cosmic rays using Galactic magnetic field models and taking into account both their regular and turbulent components. The turbulent field is generated on a nested grid which allows spatial resolution down to fractions of a parsec. Assuming sufficiently frequent Galactic CR sources, the dipole amplitude computed for a mostly light or intermediate primary composition exceeds the dipole bounds measured by the Auger collaboration around $E \\approx 10^{18}$\\,eV. Therefore, a transition at the ankle or above would require a heavy composition or a rather extreme Galactic magnetic field with strength $\\gtrsim 10\\,\\mu$G. Moreover, the fast rising proton contribution suggested by KASCADE-Grande data between $10^{17}$\\,eV and $10^{18}$\\,eV should be of extragalactic origin. In case heavy nuclei dominate the flux at $E \\gtrsim 10^{18}$\\,eV, the transition energy can be close to the ankle, if Galactic CRs are produced by sufficiently frequent transients as e.g.\\ magnetars. } ", "introduction": "\\label{Introduction} The question at which energy the transition from Galactic to extragalactic cosmic rays (CRs) takes place is one of the major unresolved issues of cosmic ray physics. Two promising possibilities are to associate the transition with one of the two evident features of the cosmic ray spectrum: The second knee around $E\\simeq 5\\times10^{17}$\\,eV or the ankle at $E\\simeq3\\times 10^{18}$\\,eV. Since the chemical composition of galactic and extragalactic CRs should differ in general, both because of propagation effects and of the different nature of their sources, the transition may be detected experimentally studying the chemical composition of CRs as function of energy. In the case of a transition around the second knee, Galactic CR sources such as e.g.\\ supernova remnants would accelerate CRs up to the rigidity-dependent knee, which is close to $10^{17}$\\,eV for iron. If the extragalactic CR flux dominating at higher energies would consist mainly of protons, the ankle could be explained as a dip in the extragalactic CR spectrum due to the pair-production losses of protons on cosmic microwave background (CMB) photons $p + \\gamma_{\\rm CMB} \\rightarrow p + e^+ + e^-$~\\cite{dip}. Below $\\sim10^{17-18}$\\,eV, the extragalactic CR flux may be suppressed because of CR propagation in extragalactic magnetic fields~\\cite{Lemoine:2004uw,Kotera:2007ca}. On the other hand, the scenario of Ref.~\\cite{Allard:2005cx} would favour a transition at the ankle. The composition of the CR flux at high energies is the subject of current debate due to the facts that hadronic physics must be extrapolated from lower energies and that the complex experimental analyses for different experiments are not yet completely reconciled. The scenario of Ref.~\\cite{dip} is supported by the composition measurements of HiRes~\\cite{HiRes} and the first results of the Telescope Array~\\cite{TA}, which are consistent with a light composition around the ankle and above. On the other hand, recent results from the Pierre Auger Observatory~\\cite{Abraham:2010yv,auger:2011pe} indicate a composition becoming heavier with increasing energy above the ankle, and the Yakutsk EAS array muon data suggests a non negligible fraction of heavy nuclei above $\\simeq10^{19}$\\,eV~\\cite{Yakutsk}. Moreover, the measurements of the KASCADE-Grande~\\cite{KASCADE-Grande} collaboration are consistent with a dominantly heavy composition up to $10^{18}$\\,eV. However, the KASCADE-Grande data indicate a fast rising proton contribution above $10^{17}$\\,eV. Thus at present the experimental data on the CR composition do not allow us yet to determine the transition energy between Galactic and extragalactic CRs. In this paper we suggest to use instead experimental limits on the anisotropy of the arrival directions of UHECRs to constrain the maximal contribution of Galactic CRs at $E\\gtrsim 10^{18}$\\,eV. At energies below $10^{17}$\\,eV, the diffusive propagation of Galactic cosmic rays and their resulting anisotropy at Earth was studied in details in Refs.~\\cite{Blasi:2011fi,Blasi:2011fm}. Since the propagation of CRs in the Galactic magnetic field (GMF) is not longer diffusive at $E\\gtrsim 10^{17}$\\,eV, we directly propagate UHECRs in the GMF using the numerical code developed in Refs.~\\cite{Giacinti:2010dk,Giacinti:2011uj}. We present also a way to generate the turbulent field on a nested grid without limitation on its spatial resolution. This method allows us to include magnetic field fluctuations spanning the required large dynamical range of scales, from negligible compared to the CR Larmor radii up to 300\\,pc. As main result of this work we show that the existing limits on CR anisotropies strongly restrict the contribution of the CNO element group to the Galactic CR component above $E\\gtrsim1$\\,EeV, while the contribution of iron is restricted above $E\\gtrsim3$\\,EeV. Details of the method to generate turbulent magnetic fields are discussed in the Section~\\ref{Method_TF}. In Section~\\ref{Method_Anisotropy}, we review the GMF models used and discuss how the CR anisotropy is calculated. Results of numerical simulations are presented in the Sections~\\ref{Anisotropy} and \\ref{Spectrum} for anisotropies and the spectrum of UHECR. ", "conclusions": "\\label{Conclusions} In this work we studied the consistency of a transition from Galactic to extragalactic CRs with existing anisotropy limits as a function of energy above $E=10^{18}$\\,eV. The diffusion approximation predicts a dipole anisotropy $\\delta=-3D_{ij}\\nabla_j\\ln(n)$ increasing with energy, since both the diffusion tensor $D_{ij}$ and the relative CR gradient $\\nabla_j\\ln(n)$ increase with energy. However, this approximation becomes unreliable at ${\\cal O}(E/Z)\\sim 10^{16}$\\,eV, and therefore we studied the propagation of CRs in the Galactic magnetic field directly by backtracking trajectories. We simulated the turbulent magnetic field on nested grids which allows one to include turbulent field modes $\\Bk$ with arbitrary small wave-lengths. For the regular Galactic magnetic field we used up-to-date models from Ref.~\\cite{Pshirkov:2011um}. Because the global structure of the GMF is still rather uncertain, we studied the dependence of the resulting anisotropy on the magnetic field parameters such as its strength $B_0$, scale height $z_0$, correlation length $L_{\\rm c}$ and exponent $\\alpha$ of its power-spectrum. We also examined the dependence of our results on the width and height of the disk in which sources are located. The main results of this study are presented in the Figs. 3--5. They show that the anisotropy mostly depends on the amplitude $B_0$ of the magnetic field in the disk. As our main conclusion from this study, we found that existing anisotropy limits are not compatible with light (proton) and intermediate (CNO) nuclei of Galactic origin as dominant contribution to the CR flux above 1\\,EeV. By contrast, Galactic iron nuclei as CR primaries are consistent with the existing limits even up to 10--20\\,EeV, if the strength of the turbulent field is as large as $B_{\\rm rms}\\sim 8\\,\\mu$G. This finding implies that determining the chemical composition of the CR flux around $10^{18}$\\,eV settles also the question of the transition energy between Galactic and extragalactic component: As light nuclei at this energy are not sufficiently isotropized, they have to be extragalactic. Therefore the fast increasing proton contribution indicated by the KASCADE-Grande collaboration between $10^{17}$\\,eV and $10^{18}$\\,eV suggests the beginning of an extragalactic component. We also studied qualitatively the dependence of the anisotropy on the effective density of sources, see Figs.~6--8. The average escape time of iron nuclei with 10\\,EeV energy from the Galaxy is $\\sim 10^{5}$\\,yr. Assuming for magnetars a rate of $10^{-3}/$yr, the effective density of magnetars as sources of CR at 10\\,EeV is $\\sim 100$/Galaxy. Thus magnetars satisfy the anisotropy constraint and can be natural candidates for the sources of the high-energy end of the Galactic CR flux in the scenario where the transition from Galactic to extragalactic cosmic rays occurs at the ankle, provided they are able to accelerate iron up to few$\\;\\times 10^{18}$\\,eV. In summary, we conclude that models with a transition from Galactic to extragalactic cosmic rays around the ankle are consistent with the existing anisotropy limits if the composition of Galactic cosmic rays at $E \\gtrsim 10^{18}$\\,eV is dominated by heavy nuclei. In contrast, if the chemical composition at these energies turns out to be light or intermediate, a transition at the ankle would be very strongly disfavoured." }, "1112/1112.4255.txt": { "abstract": "In this paper we revisit the dynamical dark energy model building based on single scalar field involving higher derivative terms. By imposing a degenerate condition on the higher derivatives in curved spacetime, one can select the models which are free from the ghost mode and the equation of state is able to cross the cosmological constant boundary smoothly, dynamically violate the null energy condition. Generally the Lagrangian of this type of dark energy models depends on the second derivatives linearly. It behaves like an imperfect fluid, thus its cosmological perturbation theory needs to be generalized. We also study such a model with explicit form of degenerate Lagrangian and show that its equation of state may cross $-1$ without any instability. ", "introduction": "The recent data from type Ia Supernovae (SNIa) and cosmic microwave background (CMB) radiation and so on have provided strong evidences for a spatially flat and accelerated expanding universe. In the context of Friedmann-Robertson-Walker (FRW) cosmology with Einstein gravity, this acceleration is attributed to the domination of a component with negative pressure, called dark energy. So far, the nature of dark energy remains a mystery. Theoretically, the simplest candidate for such a component is a small positive cosmological constant, but it suffers the difficulties associated with the fine tuning and the coincidence problems. Therefore, many physicists are attracted by the idea of dynamical dark energy models, such as quintessence \\cite{Ratra:1987rm,Wetterich:1987fm,Caldwell:1997ii}, phantom \\cite{Caldwell:1999ew}, k-essence \\cite{ArmendarizPicon:2000dh}, quintom \\cite{Feng:2004ad} and so on (see Refs. \\cite{Copeland:2006wr, Frieman:2008sn, Caldwell:2009ix, Cai:2009zp, Li:2011sd} for recent reviews). Although the recent fits to the data \\cite{Li:2011dr}, in combination of the 7-year WMAP \\cite{Komatsu:2010fb}, the Sloan Digital Sky Survey \\cite{Reid:2009xm}, and the recently released \u00a1\u00b0Union2\u00a1\u00b1 SNIa data \\cite{Riess:2011yx}, show remarkably the consistence of the cosmological constant, it is worth noting that a class of dynamical models with the equation-of-state (EoS) across $-1$, dubbed as {\\it quintom} which dynamically violates the null energy condition (NEC), is mildly favored \\cite{Feng:2004ad}. However, it was noticed that consistent single field models realizing the quintom scenario are difficult to be constructed. For example, the EoS of quintessence is limited to be in the region $-1\\leq w \\leq 1$, while for phantom $w$ is always smaller than $-1$. It was proved that in models described by a single perfect fluid or a single scalar field with a Lagrangian of k-essence form \\cite{ArmendarizPicon:1999rj}, the cosmological perturbations encounter a divergence when the background EoS crosses $-1$ \\cite{Feng:2004ad, Vikman:2004dc, Hu:2004kh, Caldwell:2005ai, Zhao:2005vj}. This statement was explicitly proven in Ref. \\cite{Xia:2007km} as a ``No-Go\" theorem for dynamical dark energy models. To realize a viable quintom scenario, one usually needs to add more degrees of freedom into the dark energy budget. The simplest and also the first quintom model was constructed by a combination of a canonical scalar and a phantom scalar field \\cite{Feng:2004ad}. However theorists are still interested in pursuing single field quintom models. The first single scalar field quintom model was realized by introducing higher order derivative terms \\cite{Li:2005fm}, also see \\cite{Zhang:2006ck} for generalization. It was also obtained in the frame of nonlocal string theory \\cite{Aref'eva:2005fu} and decaying tachyonic branes \\cite{Cai:2007gs}. However, there is a quantum instability due to an unbounded vacuum state in such type of models \\cite{Carroll:2003st, Cline:2003gs}. A possible approach to stable violations of NEC is the ghost condensation of Ref. \\cite{ArkaniHamed:2003uy}, in which the negative kinetic modes are bounded via a spontaneous Lorentz symmetry breaking, although it might allow for superluminal propagation of information in some cases \\cite{Dubovsky:2005xd}. Various theoretical realizations of quintom scenarios and their implications for early universe physics were reviewed in Ref. \\cite{Cai:2009zp} (see also \\cite{Qiu:2010ux}). Recently, a scalar field model which stably violates the NEC has been studied extensively, which is the so-called Galileon \\cite{Nicolis:2008in}. It was viewed as a local infrared modification of General Relativity, generalizing an effective field description of the DGP model \\cite{Dvali:2000hr}. The key feature of these models is that they contain higher order derivative terms in the action while the equation of motion remains second-order in order to avoid the appearance of ghost modes, realizing the idea pioneered by Horndeski thirty years ago \\cite{Horndeski:1974}. Later on, various phenomenological studies of this type of models were performed, namely, see Refs. \\cite{deRham:2010eu, Deffayet:2010qz, Pujolas:2011he, Deffayet:2009wt, Deffayet:2011gz, Gao:2011qe,Chow:2009fm, Silva:2009km, DeFelice:2010pv, Kobayashi:2010cm, Hinterbichler:2010xn, Creminelli:2010ba}. Motivated by the feature of the Galileon model, in this paper we revisit quintom dark energy models containing higher derivative terms. We start from a general covariant Lagrangian of single scalar field involving higher derivative terms and pursue how to keep the model free from extra degree of freedom. We show that it is able to eliminate the ghost mode by imposing a degenerate condition that the Lagrangian only depends on the second derivative terms linearly. Based on the degenerate Lagrangian, we build an explicit dark energy model and study its dynamics of its homogeneous background and its perturbations. Our numerical calculations show that the EoS is able to cross the cosmological constant boundary smoothly and the perturbation modes are well controlled when the crossing takes place. We understand the reason of realizing the single field quintom without ghosts is that, such a single scalar field model is no longer be able to correspond to a perfect fluid. Thus, this model does not conflict with the ``No-Go\" theorem for quintom dark energy model building as proposed in \\cite{Xia:2007km}. This paper is organized as follows. In Section II, we simply review the difficulty of constructing a single field dark energy model which gives rise to quintom scenario. In Section III, we start with a single field action involving higher derivative terms, and present the general analysis. We also discuss under what condition the higher derivative terms do not bring a ghost mode to the effective field description in Section IV. Section V is devoted to the study of the background and perturbation dynamics of single field dark energy model with degenerate higher derivatives in the frame of flat FRW universe. Specifically, we present an explicit model of quintom dark energy with degenerate higher derivatives. We perform numerical computation to illustrate such a model can realize the EoS across $-1$ smoothly. Section VI is the summary. ", "conclusions": "The model building of dark energy which is possible to cross the cosmological constant boundary, i.e., the quintom dark energy has attracted a lot of attention in the literature (for example see Refs. \\cite{Quintom_sum}). A very interesting question is how to build a single scalar model which gives rise to the crossing without any instabilities. The past studies have shown that a single scalar field satisfying a generic k-essence Lagrangian cannot give rise to the stable violation of NEC. However, it becomes possible when higher derivative operators are introduced. The non-degenerate higher derivative model can eliminate the classical instabilities but is still plagued by the ghost mode. In this paper we considered the degenerate higher derivative model inspired by the Galileon theory and its generalizations. We started with a single scalar of which the Lagrangian contains generic higher derivative operator, and then suggested a degenerate condition to eliminate the extra degrees of freedom brought by these higher derivative operators. Using the degenerate condition in the curved spacetime, the Lagrangian can be reduced to the form (\\ref{kg}). We studied the cosmological application of this model, including the background and perturbation analysis. Particularly, we chose an explicit example and performed a detailed numerical computation. Our numerical results verified that this model is able to realize the EoS to cross the cosmological constant boundary and the perturbation evolves smoothly without any pathologies. A quintom model which successfully violates the NEC without leading to quantum instability also has important implications to the early universe physics. In Refs. \\cite{Cai:2007}, the authors found that a quintom model can avoid the big bang singularity widely existing in standard and inflationary cosmologies. In this picture, the moment of initial singularity can be replaced by a big bounce. Based on this scenario, the corresponding perturbation theory has been extensively developed, namely, on adiabatic perturbations \\cite{Cai:2008qb, Cai:2008qw, Cai:2009hc}, non-Gaussianities\\cite{Cai:2009fn}, entropy fluctuations\\cite{Cai:2008qw, Cai:2011zx}, and the related preheating phase\\cite{Cai:2011ci}. Recently, it was observed that the Galileon model with the EoS across $-1$ exactly leads to a bouncing solution in the frame of the flat FRW universe\\cite{Qiu:2011cy}. Consequently, we expect that bouncing cosmologies can be realized in a generic quintom model with degenerate higher derivative operators." }, "1112/1112.5882_arXiv.txt": { "abstract": "{DBV stars are pulsating white dwarfs with atmospheres rich in He. Asteroseismology of DBV stars can provide valuable clues about the origin, structure and evolution of hydrogen-deficient white dwarfs, and may allow to study neutrino and axion physics. Recently, a new DBV star, KIC 8626021, has been discovered in the field of the \\emph{Kepler} spacecraft. It is expected that further monitoring of this star in the next years will enable astronomers to determine its detailed asteroseismic profile.}{We perform an asteroseismological analysis of KIC 8626021 on the basis of fully evolutionary DB white-dwarf models.} {We employ a complete set of evolutionary DB white-dwarf structures covering a wide range of effective temperatures and stellar masses. They have been obtained on the basis of a complete treatment of the evolutionary history of progenitors stars. We compute $g$-mode adiabatic pulsation periods for this set of models and compare them with the pulsation properties exhibited by KIC 8626021.}{On the basis of the mean period spacing of the star, we found that the stellar mass should be substantially larger than spectroscopy indicates. From period-to-period fits we found an asteroseismological model characterized by an effective temperature much higher than the spectroscopic estimate.}{In agreement with a recent asteroseismological analysis of this star by other authors, we conclude that KIC 8626021 is located near the blue edge of the DBV instability strip, contrarily to spectroscopic predictions. We also conclude that the mass of KIC 8626021 should be substantially larger than thought.} ", "introduction": "\\label{intro} V777 Her (or DBV) stars are $g$-mode variable white dwarfs with He-rich atmospheres (DB) and effective temperatures in the range $21\\,500 \\lesssim T_{\\rm eff} \\lesssim 29\\,000$ K that pulsate with periods between 100 and $1100$ s (Winget \\& Kepler 2008; Althaus et al. 2010). They are the hotter cousins of the ZZ Ceti (or DAV) stars, which are pulsating H-rich atmosphere white dwarfs that define an instability strip centered at $T_{\\rm eff} \\approx 12\\,000$ K. As an excellent demonstration of the validity of the stellar pulsation theory, the existence of the DBV class of compact pulsators was predicted on theoretical grounds (Winget et al. 1982a) before it were confirmed observationally (Winget et al. 1982b). Pulsations in V777 Her are thought to be driven by a combination of the $\\kappa-\\gamma$ mechanism acting in the He partial ionization zone ---and thus setting the blue edge of the DBV instability strip (Winget et al. 1983; Bradley \\& Winget 1994; C\\'orsico et al. 2009), and the ``convective driving'' mechanism (Brickhill 1991; Goldreich \\& Wu 1999) which is thought to be dominant once the outer convection zone has deepened enough. White-dwarf asteroseismology ---the comparison between the pulsation periods of white dwarfs and the periods computed for appropriate theoretical models--- allows us to infer details of the origin, internal structure and evolution of white dwarfs. In particular, estimates of the stellar mass, He and H layer mass, core composition, magnetic field, rotation rate, seismological distance, and properties of the outer convection zone can be inferred from the observed pulsations of DAV and DBV stars (Winget \\& Kepler 2008; Althaus et al. 2010). Finally, an eventual measurement of the temporal changes in the observed stable periods of DBVs could allow to study neutrino emission (Winget et al. 2004) and also \\emph{axion} emission (see Isern et al. 2010 and references therein). Unlike the ZZ Ceti variables, that constitute the most numerous class of pulsating white dwarfs (about 150 objects are known today; Castanheira et al. 2010), V777 Her are quite rare and difficult to find, and until recently, only 20 stars of this class were known (Beauchamp et al. 1999; Nitta et al. 2009; Kilkenny et al. 2009). The list is now a bit enlarged with KIC 8626021 ($=$ GALEX J192904.6$+$444708, $T_{\\rm eff}= 24\\,950 \\pm 750$ K and $\\log g= 7.91\\pm 0.07$ dex), a very recently discovered DBV star located in the field of view of the \\emph{Kepler Mission} by {\\O}stensen et al. (2011) (hereinafter {\\O}EA11)\\footnote{Even more recently, Hermes et al. (2011) have reported the discovery of the first DAV star in the \\emph{Kepler Mission} field.}. This star exhibits at least five periodicities in the range $197-376$ s and amplitudes between 1 and 5.2 mma, with the three strongest modes showing a triplet structure due possibly to rotation. It is expected that additional modes with higher radial orders (longer periods) will be detected in future runs of \\emph{Kepler}, thus increasing the potential of asteroseismology to infer its interior structure. Given the values of $T_{\\rm eff}$ and $\\log g$ quoted above and the DB models of Althaus et al. (2009a), KIC 8626021 should have a mass of $M_*= 0.56 \\pm 0.03 M_{\\odot}$. Motivated by the exciting discovery of the first pulsating white dwarf in the \\emph{Kepler} field of view, we present an asteroseismological analysis of KIC 8626021 on the basis of the DB white-dwarf models presented in Althaus et al. (2009a). This grid of evolutionary models was computed for a wide range of stellar masses on the basis of a complete treatment of the evolutionary history of progenitors stars, including those stages relevant for the chemical profiles of the white dwarf, such as the core H and He burning phases, the thermally pulsing asymptotic giant branch phase, and the born-again episode that is responsible for the H deficiency. By considering the mean period spacing exhibited by the star, we found that KIC 8626021 should be substantially more massive than spectroscopy indicates. We also found that period to period fits favour an asteroseismological solution characterized by an effective temperature much higher than the spectroscopic estimate. While writing this paper, Bischoff-Kim \\& {\\O}stensen (2011) (hereinafter BK{\\O}11) announced the results of their own independent asteroseismological analysis on KIC 8626021. They found that this star is actually a hot DBV. Using entirely different models and methods, we arrive at the same conclusion. The fundamental difference between our approach and that of BK{\\O}11 is that we run evolutionary models starting on the Zero Age Main Sequence, accounting for nuclear burning, time dependent diffusion of the elements and other physical processes that take place in the course of the evolution of a star. BK{\\O}11 perform ``fast white dwarf asteroseismology'', where they guess and parameterize the internal chemical composition profiles and build static models. That method allows them a fuller exploration of parameter space. For KIC 8626021, they find very good fits (hotter than what the preliminary spectroscopy suggests) but also conclude that their internal Oxygen composition profiles are in poor agreement with stellar evolution calculations. They suggest further studies to see what the physical parameters of models that agree with stellar evolution calculations would be. In essence, this is the kind of study we present here. While our best fit models are evidently different, we concur with the conclusion that KIC 8626021 is hotter (and more massive) than suggested by the initial spectroscopic study. The conclusion that this star is residing at the blue edge of the DBV instability strip appears to be robust, and calls for the necessity of a new and improved spectroscopic determination of its effective temperature. In Section \\ref{modeling} we present some details about our models and methods. In Section \\ref{avgspacing}, we use the average period spacing of KIC 8626021's period spectrum to draw some conclusions about its mass and effective temperature. We contrast our models to the grid of models computed by BK{\\O}11. In Section \\ref{best-fit}, we present our best fit models. Again, we contrast our results with those of BK{\\O}11. In Section \\ref{discussion} we discuss the discrepancies between the results coming from spectroscopy and from asteroseismology, and the differences between our asteroseismological results and those of BK{\\O}11. We conclude in Section \\ref{conclusions}. ", "conclusions": "\\label{conclusions} In this paper we have presented a detailed asteroseismic analysis of KIC 8626021, the first pulsating DB white dwarf star discovered by the \\emph{Kepler Mission}, on the basis of the full evolutionary DB white-dwarf models presented in Althaus et al. (2009a) which were computed for a wide range of stellar masses and He envelopes. These DB white dwarf models are characterized by consistent chemical profiles for both the core and the envelope. These chemical profiles are the result of the computation of the full and complete evolution of the progenitor stars from the zero age main sequence, including the core H and He burning phases, the thermally pulsing asymptotic giant branch phase, the born-again episode that is responsible for the H deficiency, and from time-dependent element diffusion predictions during the white-dwarf stage. By considering the mean period spacing exhibited by the star, we found that KIC 8626021 should have a stellar mass in the range $0.60 \\lesssim M_*/M_{\\odot} \\lesssim 0.87$, substantially larger than those derived by previous spectroscopic ($M_* \\sim 0.56 M_{\\odot}$; {\\O}EA11) and asteroseismic ($M_* \\sim 0.57 M_{\\odot}$; BK{\\O}11) studies. We also found that period-to-period fits point to an asteroseismological model with an effective temperature of $\\sim 27\\,300$ K, in strong conflict with the spectroscopic estimate ($T_{\\rm eff} \\sim 24\\, 900$ K). Our results are in agreement with the recent asteroseismic analysis of BK{\\O}11 on KIC 8626021, in particular regarding its effective temperature. In fact, these authors conclude that KIC 8626021 \\emph{must be} a hot DBV with $T_{\\rm eff} \\sim 29\\,200$ K. It would be interesting to see what a spectroscopic analysis based on higher signal-to-noise spectra will tell about the surface gravity and effective temperature of the star. If KIC 8626021 is a hot DBV, as first found by BK{\\O}11 and confirmed now by our results, then further monitoring of KIC 8626021 with \\emph{Kepler} in the next years probably will allow a measurement of $\\dot{\\Pi}$, which in turn could open the possibility to constrain the plasmon neutrino emission rate (Winget et al. 2004; BK{\\O}11). This endeavour is evidently dependent on the models, which can always be improved. Uncertainties in the models can also be assessed (Bischoff-Kim et. al. 2008)." }, "1112/1112.0005_arXiv.txt": { "abstract": "{ Gravitationally lensed quasars can be used as powerful cosmological and astrophysical probes. We can (i) infer the Hubble constant $H_{0}$ based on the so-called time-delay technique, (ii) unveil substructures along the line-of-sight toward distant galaxies, and (iii) compare the shape and the slope of baryons and dark matter distributions in the inner regions of galaxies. To reach these goals, we need high-accuracy astrometry of the quasar images relative to the lensing galaxy and morphology measurements of the lens. In this work, we first present new astrometry for 11 lenses with measured time delays, namely, JVAS~B0218+357, SBS~0909+532, RX~J0911.4+0551, FBQS~J0951+2635, HE~1104-1805, PG~1115+080, JVAS~B1422+231, SBS~1520+530, CLASS~B1600+434, CLASS~B1608+656, and HE~2149-2745. These measurements proceed from the use of the Magain-Courbin-Sohy (MCS) deconvolution algorithm applied in an iterative way (ISMCS) to near-IR HST images. We obtain a typical astrometric accuracy of about 1-2.5 mas and an accurate shape measurement of the lens galaxy. Second, we combined these measurements with those of 14 other lensing systems, mostly from the COSMOGRAIL set of targets, to present new mass models of these lenses. The modeling of these 25 gravitational lenses led to the following results: 1) In four double-image quasars (HE0047-1746, J1226-006, SBS~1520+530, and HE~2149-2745), we show that the influence of the lens environment on the time delay can easily be quantified and modeled, hence putting these lenses with high priority for time-delay determination. 2) For quadruple-image quasars, the difficulty often encountered in reproducing the image positions to milli-arcsec accuracy (astrometric anomaly problem) is overcome by explicitly including the nearest visible galaxy/satellite in the lens model. However, one anomalous system (RXS~J1131-1231) does not show any luminous perturber in its vicinity, and three others (WFI~2026-4536, WFI~2033-4723, and B2045+265) have problematic modeling. These four systems are the best candidates for a pertubation by a dark matter substructure along the line-of-sight. 3) We revisit the correlation between the position angle (PA) and ellipticity of the light and of the mass distribution in lensing galaxies. As in previous studies, we find a significant correlation between the PA of the light and of the mass distributions. However, in contrast with these same studies, we find that the ellipticity of the light and of the mass also correlate well, suggesting that the overall spatial distribution of matter is not very different from the baryon distribution in the inner $\\sim$5\\,kpc of lensing galaxies. This offers a new test for high-resolution hydrodynamical simulations. } ", "introduction": "The measurement of the time delay between the lensed images of a gravitationally lensed quasar offers one of the most elegant ways to measure the \\textit{Hubble constant}, $H_{0}$ \\citep{Refsdal1964}. This cosmological application of strongly lensed quasars motivated many of the early gravitational lensing studies and time-delay measurement campaigns. Unfortunately, the precision on $H_{0}$ obtained with the time delay method has been, until now, challenged by systematic errors. Recently, \\citet{Suyu2009, Suyu2010} showed that a detailed study of the lensed quasar CLASS~B1608+656, combining different high-resolution data and advanced lens modeling techniques, could lead to an estimate of $H_{0}$ with very small random and systematic uncertainties: $H_{0}= 70.6 \\pm 3.1 \\ \\rm km \\ \\rm s^{-1} \\rm Mpc^{-1}$. The determination of $H_{0}$ based on a large sample of lensed quasars can also reach a precision competitive with the one of standard methods \\citep{Oguri2007,Coles2008}. In order to use the time-delay method as a precision-cosmology tool, it is necessary to derive accurate time-delay measurements and to properly understand the uncertainties associated to the lens environment and to the lens mass distribution. The COSMOGRAIL project, i.e. the COSmological MOnitoring of GRAvItational Lenses, aims to provide the community with exquisite lightcurves and accurate time-delay measurements for a sample of more than 20 lensed quasars \\citep{Cosmograil1, Cosmograil5, Cosmograil7, Courbin2011}. To understand degeneracies affecting lens models, we started to model all systems monitored by COSMOGRAIL as well as those monitored by earlier campaigns in a uniform way. Because the relative astrometry of the lensed images is the main constraint on the lens models, we first worked on deriving an updated astrometry of these systems by applying the \\textit{Iterative Strategy coupled with the MCS{\\footnote{MCS is an abbreviation build based on the name of the three designers of the method, Magain, Courbin, and Sohy.}} deconvolution algorithm} \\citep{MCS98}, i.e. ISMCS \\citep{Chantry2007} to HST images. This allows us to obtain accurate relative astrometry and morphological information of the lensing galaxy (i.e. ellipticity and position angle PA of its major axis). We presented the first part of this work in \\citet[][ hereafter Paper I]{Chantry2010}, where seven systems currently monitored by COSMOGRAIL were analyzed. We present here the deconvolution of a sample of an additional eleven lensed quasars with published time delays: JVAS~B0218+357, SBS~0909+532, RX~J0911.4+0551, FBQS~J0951+2635, HE~1104-1805, PG~1115+080, JVAS~B1422+231, SBS~1520+530, CLASS~B1600+434, CLASS~B1608+656, and HE~2149-2745. High-accuracy relative astrometry of lensed quasars allows one to test the ability of a smooth mass model of the galaxy to reproduce the observed image configuration. The failure of a smooth lens model to do this, which means that we identify an astrometric anomaly, may be caused by a perturbation of the potential by dark matter clump(s) along the line-of-sight, and likely located in the halo of the main lens. During the past five years, the occurrence and amplitude of astrometric anomalies caused by dark matter clumps has been questioned and their use to study the amount of substructures in galaxies has been investigated \\citep[][ Paper I and references therein]{Zackrisson2009}. We use the seven quads studied here and in Paper I plus seven other quads with astrometry derived with ISMCS or VLA+HST imaging, to revisit the evidence for astrometric anomalies. We also quantitatively address the role of the nearby lens environment in producing these anomalies. The deconvolution method we used also provides accurate morphological information on the lensing galaxies. With these data, we compared the mass and the light distribution in these systems. This problem has first been approached by \\citet{Keeton1998} who compared the mass and the light distribution for 14 multiply-imaged quasars (seven doubles and seven quads). More recently, \\citet{Ferreras2008} and \\citet{Treu2009} studied a sample of nine (resp. 25) lensing galaxies from the SLACS sample \\citep{Bolton2006}. Although these works probe different populations of lenses, the ``SLACS lenses'' lying in typical elliptical galaxy environments while lensed quasars often lie in richer environments \\citep{Huterer2005, Oguri2005b, Suyu2009}, they find that the PA of the light and of the mass do agree within 10 degrees. Conversely, no correlation has been found between the ellipticity of the light and of the mass distributions in \\citet{Keeton1998} and in \\citet{Ferreras2008}. The outline of the paper is the following. Section \\ref{material11} presents the eleven lensed systems deconvolved with the ISMCS technique. The deconvolution method and the derived astrometry are described in Sect. \\ref{dec11}. Section~\\ref{sec:model} explains the mass modeling strategy. We discuss the lens models for doubly imaged quasars in Sect.~\\ref{sec:doubles}. In Sect.~\\ref{sec:quads} we present an enlarged sample of quadruply imaged systems for which we provide additional mass models and a comparison of the shape of the mass and of the light distribution. Finally, our conclusions are presented in in Sect. \\ref{Conc11}. ", "conclusions": "\\label{Conc11} We applied the iterative strategy coupled with the MCS deconvolution algorithm, i.e. ISMCS, to high-resolution near-IR images of a sample of eleven lensed quasars. The ISMCS allowed us to obtain accurate astrometry, in most cases with 1 to 2.5 mas error bars, and shape parameters for the light distribution of the lensing galaxy. In three cases, the deconvolution process was not entirely satisfactory, i.e. HE~1104-1805, PG~1115+080 and CLASS~B1608+656, but still provided better results than simple PSF subtraction. For one special case, i.e. JVAS~B0218+357, the small angular separation and the poor resolution of NICMOS prevents us from reaching the desired accuracy. For one system, SBS~0909+532, we were able to detect the lensing galaxy for the first time, which is angularly small and relatively bright. Two objects reveal partial Einstein rings (PG~1115+080, CLASS~B1608+656). We provided, for these systems, simple mass models (isothermal and de Vaucouleurs mass distributions) that give an overview of the ability of smooth mass models to reproduce the known lensed quasars. These models give some hints about the effect of the radial profile on the flux ratios and on the time delays. For the sample of ten doubly imaged quasars presented here and in Paper I \\citep[i.e.][]{Chantry2010}, we have also studied the role of intrinsic ellipticity and external shear in the models. For half of these systems{\\footnote{HS~2209+1914 is more uncertain because of the lack of detailed environmental studies}} (HE0047-1746, J1226-006, SBS~1520+530, HE~2149-2745, and HS~2209+1914) one should be able to unfold intrinsic and extrinsic shear and therefore minimize systematic biases on $H_0$ based on the time-delay method. This identification of the most promising systems is partly subjective because detailed environmental studies of the whole sample are lacking. Exhaustive studies of lens environment are fortunately ongoing \\citep{Williams2006, Momcheva2006, Wong2011} and should allow one to address this question more quantitatively. In addition to the ten doubly imaged quasars, we have performed modeling with singular isothermal ellipsoid (SIE) of a sample of 14 quadruply imaged systems for which we have relative astrometry with accuracy typically better than 2 mas. This sample allowed us to investigate two questions: \\begin{enumerate} \\item {\\it {How efficient are SIE models in reproducing the image and galaxy position of the observed quads and how does the local environment of the lens influence this result ? }} We found that the environment of most of the quads is rich and that at least the nearest galaxy or a nearby group has to be included explicitly in the model to recover the image and galaxy positions. Following this procedure leads to a formally perfect fit with an SIE model of the main lens for 11 of the 14 systems. When this perturber is not included, models generally tend to predict the position of the main lensing galaxy significantly offset with respect to the observed position, the lensed image positions being less affected. For the two systems RXS~J1131-1231 and WFI~2033-4723, we were not able to obtain an acceptable fit. The problems encountered with WFI~2033-4723 might be caused by other nearby galaxies not included explicitly in our model. RXS~J1131-1231 remains the best candidate of our sample for which astrometric anomalies may be caused by a dark matter substructure. Two other systems (WFI~2026-4536 and B2045+265) show a significant mismatch between the $\\theta_e$ of the light and of the mass. The origin of this discrepancy is unclear and might also be the signature of dark matter substructures. We mention that astrometric anomalies may become proeminent when higher accuracy images are available. This is the case for B0128+437, which is perfectly reproduced by our model but shows astrometric anomalies once sub-mas astrometry of the lensed images is considered \\citep{Biggs2004}. \\\\ \\item {\\it {How do the morphology of the mass and of the light profile compare ?}} Our analysis of 12 out of 14 quads with measured structural parameters allowed us to find that the PA of the mass and of the light distribution are aligned within typically 10 degrees, in agreement with previous studies \\citep{Keeton1998, Ferreras2008, Treu2009}. We also discovered that the ellipticity of the light generally agrees well with that of the total mass distribution (8 out of 12 lenses show a good correlation). Overall, this indicates that the light provides a good proxy to the total mass distribution in these lensing galaxies{\\footnote{The radial distribution of the total mass disagrees however with that of the light, as illustrated by the inability of deVaucouleurs mass profiles to model quads.}}, a result particularly interesting in the light of the rich environment of the lenses. Understanding why these correlations take place is important for both our comprehension of dark matter and galaxy formation and for time-delay studies. Therefore, an increase in the sample of lenses with shape measurements and a comparison with numerical simulations, including problems caused by selection biases \\citep{Vandeven2009, Mandelbaum2009}, are necessary. This will provide important clues on the role of dark matter in shaping galaxies. \\\\ \\end{enumerate}" }, "1112/1112.3006_arXiv.txt": { "abstract": "Cosmological data have provided new constraints on the number of neutrino species and the neutrino mass. However these constraints depend on assumptions related to the underlying cosmology. Since a correlation is expected between the number of effective neutrinos $N_{eff}$, the neutrino mass $\\sum m_\\nu$, and the curvature of the universe $\\Omega_k$, it is useful to investigate the current constraints in the framework of a non-flat universe. In this paper we update the constraints on neutrino parameters by making use of the latest cosmic microwave background (CMB) data from the ACT and SPT experiments and consider the possibility of a universe with non-zero curvature. We first place new constraints on $N_{eff}$ and $\\Omega_k$, with $N_{eff} = 4.03 \\pm 0.45$ and $10^3 \\, \\Omega_k = -4.46 \\pm 5.24$. Thus, even when $\\Omega_k$ is allowed to vary, $N_{eff} = 3$ is still disfavored with 95\\% confidence. We then investigate the correlation between neutrino mass and curvature that shifts the $95 \\%$ upper limit of $\\sum m_\\nu < 0.45$ eV to $\\sum m_\\nu < 0.95$ eV. Thus, the impact of assuming flatness in neutrino cosmology is significant and an essential consideration with future experiments. ", "introduction": "Throughout the previous decades experimental cosmology has benefited from accurate measurements of the cosmic microwave background (CMB). The data have determined constraints on several cosmological parameters to remarkable accuracy and the ability to constrain new physics with the CMB continues to improve. Future CMB experiments might even be able to measure B-mode polarization and distinguish between neutrino hierarchy models. However, when constraining new parameters one must be careful when constraints depend on assumptions about the underlying cosmology. For example, a correlation between the neutrino properties and the curvature of the universe is clearly expected since, a higher number of neutrino species or large mass would introduce pre-recombination effects, shifting the positions of the peaks in the angular CMB spectrum (cf.~\\cite{cormelk,Mantz:2009rj}). Here we present an update on the constraints of the number of neutrino species $N_{eff}$ and the sum of neutrino masses $\\Sigma m_{\\nu}$ in the framework of non-flat universes with $\\Omega_k \\neq 0$ combining the Wilkinson Microwave Anisotropy Probe (WMAP) 7-year~\\cite{Komatsu:2010fb}, South Pole Telescope (SPT)~\\cite{Keisler:2011aw} and Atacama Cosmology Telescope (ACT)~\\cite{Dunkley:2010ge} datasets. % The paper is organized as follows. In Section~\\ref{sec:theory} we % give theoretical arguments for why the $N_{eff}$ and $\\Omega_k$ % parameters should be correlated. In Section~\\ref{sec:method} we discuss our method of constraining the parameters $N_{eff}$, $\\Sigma m_{\\nu}$, and $\\Omega_k$. We present the results of the analysis in Section~\\ref{sec:imp}. Finally, in Section~\\ref{sec:conclusion} we conclude and discuss the implications of assuming flatness in neutrino cosmology. ", "conclusions": "\\label{sec:conclusion} The resolution of the high effective neutrino number in cosmology remains an open question. However, additional neutrinos % may be due to parameter degeneracy or other issues in statistical analysis rather than new physics. % The focus of this paper has been an argument for correlation between the number of effective neutrinos $N_{eff}$ and the curvature of the Universe $\\Omega_k$, which arises from the effect of these parameters on distance measurements. The qualitative argument is confirmed by a statistical analysis of CMB anisotropy measurements using CosmoMC. In this paper we have shown that there is a correlation between $N_{eff}$ and $\\Omega_k$ that gets stronger when SPT and ACT datasets are added to WMAP alone. However, even when $\\Omega_k$ is % allowed to vary, $N_{eff} = 3$ is still disfavored by the data with 95\\% confidence. Although the correlation favors a closed universe with $\\Omega_k < 0$, % if CMB data were to favor open models then the neutrino number would decrease as predicted. Perhaps % the same element of the data that favors a closed universe may also be responsible for the trend toward a higher $N_{eff}$. More importantly, % we find a strong correlation between curvature and % the sum of the neutrino masses. % Future experiments will provide further insight into both $N_{eff}$ and $\\sum m_\\nu$~\\cite{Carbone:2010ik}. Our results are consistent with the current understanding of the data available. The strongest constraints on these parameters from the statistical analysis assuming a flat universe are $N_{eff} = 3.89 \\pm 0.41$ and $\\sum m_\\nu < 0.45$ eV with 95\\% confidence level using WMAP7+ACT+SPT+BAO+$H_0$. The constraints are weakened by % degeneracy with the curvature parameter $\\Omega_k$. However, this still represents the continued effort toward significant improvements on parameter constraints in cosmology. Although the sum of the neutrino masses is significantly improved from the WMAP 7-year result of $\\sum m_\\nu < 0.57$ eV, the constraint is far from being sensitive enough to rule out one of the mass hierarchies. Furthermore, we have shown that the mass uncertainty more than doubles when $\\Omega_k \\neq 0$. Based on our results and the estimated quality of data for Planck and other experiments, it should be possible to determine the existence or nonexistence of sterile radiation to much greater confidence in the near future." }, "1112/1112.2882_arXiv.txt": { "abstract": "{ \\object{Comet P/2010~A2 LINEAR} is an object on an asteroidal orbit within the inner Main Belt, therefore a good candidate for membership with the Main Belt Comet family. It was observed with several telescopes (ESO New Technology Telescope, La Silla, Chile; Gemini North, Mauna Kea, Hawai`i; University of Hawai`i 2.2~m, Mauna Kea, Hawai`i) from 14~Jan. until 19~Feb. 2010 in order to characterize and monitor it and its very unusual dust tail, which appears almost fully detached from the nucleus; the head of the tail includes two narrow arcs forming a cross. No evolution was observed during the span of the observations. Observations obtained during the Earth orbital plane crossing allowed an examination of the out-of-plane 3D structure of the tail. The immediate surroundings of the nucleus were found dust-free, which allowed an estimate of the nucleus radius of 80--90~m, assuming an albedo $p=0.11$ and a phase correction with $G=0.15$ (values typical for S-type asteroids). A model of the thermal evolution indicates that such a small nucleus could not maintain any ice content for more than a few million years on its current orbit, ruling out ice sublimation dust ejection mechanism. Rotational spin-up and electrostatic dust levitations were also rejected, leaving an impact with a smaller body as the favoured hypothesis. This is further supported by the analysis of the tail structure. Finston-Probstein dynamical dust modelling indicates the tail was produced by a single burst of dust emission. More advanced models (described in detail in a companion paper), independently indicate that this burst populated a hollow cone with a half-opening angle $\\alpha\\sim 40\\degr$ and with an ejection velocity $v_{\\rm max}\\sim 0.2$~m~s$^{-1}$, where the small dust grains fill the observed tail, while the arcs are foreshortened sections of the burst cone. The dust grains in the tail are measured to have radii between $a=1$--20~mm, with a differential size distribution proportional to $a^{-3.44\\pm0.08}$. The dust contained in the tail is estimated to at least $8\\times 10^8$~kg, which would form a sphere of 40~m radius (with a density $\\rho=3\\,000$~kg~m$^{-3}$ and an albedo $p=0.11$ typical of S-type asteroids). Analysing these results in the framework of crater physics, we conclude that a gravity-controlled crater would have grown up to $\\sim 100$~m radius, i.e. comparable to the size of the body. The non-disruption of the body suggest this was an oblique impact. ", "introduction": "Habitability within our solar system is determined by the distribution of water and volatiles, yet the origin of this distribution is currently a fundamental unresolved planetary science issue \\citep{KC03_habitable}. There are three leading scenarios for the origin of terrestrial planetary water, including: ($i$) nebular gas adsorption on micron-sized dust grains inside the snow line \\citep[][the distance from the Sun outside of which water the temperature is low enough for water ice to condense]{Mur+08_water}, ($ii$) chemical reactions between an early hydrogen envelope and oxides in a magma ocean \\cite{GI07_DHratio, GI08_ocean}, or ($iii$) delivery by volatile-rich planetesimals (asteroids or comets) formed beyond the snow line \\citep{Mor+00_water}. The first two processes probably contributed to, but may be unable to account for all of Earth's water \\citep{Mot+07_water}. Within the broader context of icy bodies, main belt comets (MBCs), a newly discovered class of ``comets having stable orbits completely confined to the main asteroid belt'' \\citep{HsiehJ06_MBC}, present a subclass of particular significance to the history of terrestrial water and other important volatiles. In addition to possibly P/2010~A2 (LINEAR), there are currently five known MBCs: 133P/Elst-Pizarro, 176P/LINEAR, 238P/Read, C/2008 R1 (Garradd) and P/2010 R2 (La Sagra). The main belt comets are defined \\citep{HsiehJ06_MBC} by ($i$) a semi-major axis that is less than Jupiter's, ($ii$) Tisserand parameters significantly greater than 3 ---meaning they are dynamically decoupled from Jupiter, like ordinary asteroids \\citep{Vaghi73, Kresak80}--- and ($iii$) mass loss with a cometary appearance. As comets in near-circular orbits within the asteroid belt, these objects likely still harbor nebular water frozen out from beyond the primordial ``snow line'' \\citep{Encrenaz08, GaraudL07_snow} of the young solar system. Dynamical simulations suggest they probably formed in-situ \\citep{Haghighipour09_MBC} at a different temperature from either comets or the asteroidal reservoir that has been sampled through our meteorite collections. Thus, understanding their chemistry can provide unique insight into the distribution of volatiles in the early stages of planet formation. P/2010 A2 (LINEAR) was discovered by the LINEAR project on 7~Jan., 2010 \\citep{CBET2114} and was described as ``a headless comet with a straight tail, and no central condensation'' \\citep{IAUC9105}. On 11--12 Jan., observers at the WYNN 3.5-m and at the 2.5-m Nordic Optical Telescope reported that the object had a asteroidal-like body $\\sim$ 150--200m in diameter which was connected to the tail by an unresolved light bridge \\citep{IAUC9109}. Based on the orbital semi-major axis and on the Tisserand parameter, $T_J=3.6$, Jewitt \\citep{IAUC9109} concluded that P/2010~A2 was a new MBC. With the smallest perihelion distance of any of the MBCs ($q=2.29$~AU), these elements suggested a membership in the Flora family. Flora is a large asteroid family of $\\sim$500 members which can be broken up into many sub-families or clans and which are broadly compositionally consistent with space-weathered S-type asteroidal spectra \\citep{Flo+98}. Jewitt \\citep{IAUC9109} and Licandro \\citep{CBET2134} suggested that the location of the nucleus outside the coma might be the consequence of an impact. Is P/2010~A2 a genuine, sublimating comet (whose activity was possibly triggered by an impact), or is the tail the signature of an impact (or another alternative process), with a dispersion of the dust ejecta but no sublimation? If it can be demonstrated that the object is a comet, it would indicate that at least one object in the inner asteroid belt still contains volatile material. The water snow line in the proto-solar nebula is estimated to have been around 2.5~AU from the Sun \\citep{Jon+90} or possibly even as close as 1.5~AU \\citep{Lec+06,Mac+10,Min+11}. Initially parent bodies beyond the snow line are believed to be mixtures of water ice and silicates, which are then heated due to the radioactive decay of $^{26}$Al within 1--2 million years after nebular collapse \\citep{Krot+06}. \\citet{Grimm+89} find that once the water is liquid, it is consumed by hydration reactions, preferentially in the interior \\citep{Cohen+00,Wilson+99}, possibly leaving ice in the outer layers. In the inner asteroid belt, it is likely that most of that water has been converted in hydrated material observed on S-type asteroids \\citep{Riv+02}. MBCs may be the frozen components of the outer edges of asteroid parent bodies that have survived the age of the solar system. Four of the MBCs are located in the outer main belt, where asteroids with hydrated material are less common, suggesting water ice could still exist. The orbital elements of the MBCs are listed in Table~\\ref{tab:ele} and displayed in Fig.~\\ref{fig:ele}. 133P/Elst-Pizarro, the first MBC discovered, is a member of the Themis family; 176P/LINEAR was found in a survey targeting objects with orbits similar to that of 133P/Elst-Pizarro \\citep{HHH09}. 238P/Read, however, was discovered serendipitously (i.e. not within a MBC-dedicated program) on a similar orbit. A fourth MBC, P/2008~R1 Garradd, was discovered serendipitously in the central region of the main belt. The fifth MBC, P/2010~R2 La Sagra, was also discovered serendipitously with a semi-major axis similar to that of 133P/Elst-Pizarro \\citep{Mar+10}. More recently, asteroid (596)~Scheila presented a comet-like appearance \\citep{Lar10}. \\citet{Bod+11} and \\citet{Jew+11} conclude however that this dust cloud was likely caused by the impact of a small asteroid on Scheila. The survival of an icy body in the inner asteroid belt would therefore give exciting constraints on the evolution of water in the belt. It would also raise fundamental questions on how to shield water ice in that area of the solar system in a rather small body. \\begin{figure} \\caption{(not available on astroPh) The orbital elements of P/2010~A2 (green square to the left), of the 5 known MBCs (circles, 133P and 238P are indistinguishable) and of the numbered main belt asteroids (small dots). When available, the proper elements were used (source: the Asteroid Dynamics Site, http://hamilton.dm.unipi.it/astdys/). The inclination of P/2010~R2 La Sagra is off scale; its $a$ is marked by a triangle on the top of the plot; $i=21.39$ . } \\label{fig:ele} \\end{figure} \\begin{figure}[h] {\\bf a.} \\includegraphics[width=8.8cm]{fig_2010-01-14_NTT_OH.eps}\\\\ {\\bf b.} \\includegraphics[width=8.8cm]{fig_2010-01-15_NTT_HH.eps}\\\\ {\\bf c.} \\includegraphics[width=8.8cm]{fig_2010-01-16_NTT_HH.eps}\\\\ {\\bf d.} \\includegraphics[width=8.8cm]{fig_2010-01-17_NTT_NN.eps}\\\\ {\\bf e.} \\includegraphics[width=8.8cm]{fig_2010-01-19_GN_OH.eps} \\caption{P/2010~A2, images from UT~14.2 (a), 16.2 (b), 17.2 (c), 18.2 (d) Jan. 2010 using the NTT, UT~19.5 Jan. 2010 using GN (e). The linear gray scale covers the range of (0--3) $\\times 10^{-8} Af$, a proxy for the amount of dust present (see Section~\\ref{sec:dust}). Each panel includes a 40$\\times 30''$ zoomed portion to show details of the inner structure. The positions of N and E are indicated, as are the anti-solar direction and the heliocentric velocity vector. } \\label{fig:image}% \\end{figure} \\addtocounter{figure}{-1} \\begin{figure}[h] {\\bf f.} \\includegraphics[width=8.8cm]{fig_2010-01-22_UH_BY.eps}\\\\ {\\bf g.} \\includegraphics[width=8.8cm]{fig_2010-01-23_UH_BY.eps}\\\\ {\\bf h.} \\includegraphics[width=8.8cm]{fig_2010-01-25_UH_BY.eps}\\\\ {\\bf i.} \\includegraphics[width=8.8cm]{fig_2010-02-02_GN_OH.eps}\\\\ {\\bf j.} \\includegraphics[width=8.8cm]{fig_2010-02-19_UH_JP.eps} \\caption{continued. P/2010~A2, images from UT~22.3 (f), 23.4 (g), 25.4 (h) Jan. 2010 using UH~2.2-m, UT~2.3 Feb. 2010 using GN (i), UT~19.5 Feb. 2010 using UH~2.2-m (j).} \\end{figure} \\citet{Mor+10} presented observations obtained with the GTC, WHT and NOT on La Palma, which they modeled with an extended period of water-driven cometary activity. \\citet{Jew+10} acquired a series of HST images; from the orientation and geometry of the tail, they favor the disruption of an asteroid (either by collision or spin up) in Feb.--Mar. 2009. Finally, \\citet{Sno+10} secured observation from the Rosetta spacecraft. Thanks to the position of the space probe, they could observe the object with a very different geometry (but with a much more modest resolution), from which they concluded that the observed dust tail was caused by an impact. In order to investigate the process that generated the observed dust tail, we acquired deep images of P/2010~A2 over various epochs. The observations are presented in Section~\\ref{sec:obs}. The analysis and modelling of the nucleus and dust are described in Section~\\ref{sec:ana}. In Section~\\ref{sec:con}, we discuss the conclusions and summarize the results. A companion paper \\citep{PAPER2} is devoted to the details of the dust models developed for this study, and a second follow-up article \\citep{PAPER3} focuses on the characteristics of the impact process and of the nucleus based on the interpretation of the dust as an impact plume. ", "conclusions": "\\label{sec:con} The dust tail of P/2010~A2 was observed and analyzed so to cast some light on the possible cometary nature of that object, as its appearance and orbit could make it a member of the small family of Main Belt Comets. The main results and conclusions of this study are listed hereafter. \\begin{itemize} \\item Observations were acquired with GN, NTT, UH~2.2-m from 14 Jan. until 19 Feb. 2010. % The tail did not show any measurable evolution over that time span, which included the crossing of the orbital plane, which indicates the tail is an out-of-plane 3D structure, suggesting the dust was released with a significant orthogonal velocity. \\item The nucleus, detached from the tail, appears not to be surrounded by dust (at most 3\\% of the light corresponds to near-nucleus dust); dynamical dust models also indicate that no recent dust is present, suggesting no cometary activity at the time of the observations. The absolute magnitude of the comet was estimated to SDSS $r'(1,1,0)=21.74\\pm0.04$ on 19~Jan. 2010 and $21.55\\pm0.05$ on 2~Feb. Assuming an albedo $p=0.11$, a value typical for S-type asteroids, this converts into a nucleus of radius $r=80$--90~m. Using a density of $\\rho=3$\\,000~kg~m$^{-3}$ (also typical of a S-type asteroid), the escape velocity of the body is $v_e=0.10$--0.12~m~s$^{-1}$. \\item A thermal model of the nucleus indicates that water ice (as well as any more volatile ice) would not survive more than a few million years in the object on its current orbit. As we have no reason to suspect a recent insertion on the current orbit, this rules out ice sublimation as the source of the observed dust. Rotational spin up and electrostatic lifting were also rejected as possible source for the dust tail. \\item A Finston-Probstein-type dynamical dust modelling of the morphology, light distribution and detachment of the tail indicates that a short burst of dust emission (a day long or less) represents best the observed tail, with an ejection velocity in the range $v_e=0.20$--0.30~m~s$^{-1}$. The ejection date is in agreement with other studies of this object \\citep{Jew+10, Sno+10}; we continue this study using Snodgrass' date, 10 Feb. 2009. The model, together with direct measurements of the tail within the camera's field of view, indicate it is constituted of dust grains with radii $a$ ranging between a few tenths of millimetres to 20~mm, with a number of particles proportional to $a^{-3.44\\pm0.08}$ {(differential size distribution index).} The total dust content of the tail is estimated to at least $8\\times 10^8$~kg, which could be packed in a sphere with a radius of 40~m (assuming the same density $\\rho=3$\\,000~kg~m$^{-3}$ as above). \\item The position and shape of the tail's envelope was modeled as originating from dust emitted on a cone with a half-opening angle $\\alpha \\sim 40\\degr$ pointing below the orbital plane, toward the forward direction, with a ejection velocity $v_e \\le 0.20$~m~s$^{-1}$. The X-shaped arcs were modeled as a collection of large particles, with $v_e \\ge 0.20$~m~s$^{-1}$; they were independently found to be emitted on a section of the same hollow cone. The are seen as an arc as a consequence of the geometric foreshortening of the cone section (after one year of evolution); other sections of the cone, which are not foreshortened in the same way, contribute to the overall tail. The Rosetta observations \\citep{Sno+10} did not see other sections of the cone as arcs because of the coarser resolution of its camera. \\item Because ice sublimation, electrostatic lifting and rotational spin up were rejected as possible causes for the dust emission, and because of the very short release time and the geometry of the dust emission on a hollow cone with an half-opening angle $\\alpha = 40\\degr$, we conclude that the dust release was caused by an impact by a small object. \\item Considering the volume of ejected dust in the framework of a gravity-dominated crater formation process, a crater of $\\sim 100$~m radius was produced in a total time of the order of seconds. While that crater is large compared to the body, other bodies show craters as large as themselves. An oblique impact ---which is statistically the most likely--- would explain how P/2010~A2 escaped complete, catastrophic disruption. \\end{itemize} In summary, the tail of P/2010~A2 can be explained by an impact on a $\\sim 80$--90~m asteroid. The impact released dust in a hollow cone, which evolved for about one year to form the observed dust distribution. The fine-grained dust formed a long tail that was extended by solar radiation pressure, and the large grains (un-affected by the solar radiation pressure) retained the conical shape. The viewing geometry revealed sections of that cone as the arcs in the head of the dust cloud. Therefore, what we witnessed was likely an event of collisional grinding within the Flora collisional family, and not a genuine, water-ice driven Main Belt Comet. With increasingly sensitive sky surveys covering larger fractions of the sky during each dark period, we can expect to detect more and more of these collisions, as illustrated by the recent discovery of the dust cloud around (596)~Scheila. While the original goal of this project was to probe the water ice content in the asteroid main belt, its end result was the observation of a natural impact on a small asteroid." }, "1112/1112.0555_arXiv.txt": { "abstract": "The ultrahigh energy neutrino cross section is a crucial ingredient in the calculation of the event rate in high energy neutrino telescopes. Currently there are several approaches which predict different behaviors for its magnitude for ultrahigh energies. In this contribution is presented a summary of current predictions based on the non-linear QCD evolution equations, the so-called perturbative saturation physics. In particular, predictions are shown based on the parton saturation approaches and the consequences of geometric scaling property at high energies are discussed. The scaling property allows an analytical computation of the neutrino scattering on nucleon/nucleus at high energies, providing a theoretical parameterization. ", "introduction": "The investigation of ultrahigh energy (UHE) cosmic neutrinos provides an opportunity for study particle physics beyond the reach of the LHC \\cite{neu_review}. As an example, nowadays the Pierre Auger Observatory is sensitive to neutrinos of energy $\\ge 10^8$ GeV \\cite{pao}. A crucial ingredient in the calculation of attenuation of neutrinos traversing the Earth and the event rate in high energy neutrino telescopes is the high energy neutrino-nucleon cross section, which provides a probe of Quantum Chromodynamics (QCD) in the kinematic region of very small values of Bjorken-$x$. The typical $x$ value probed is $x \\approx m_W^2/2m_NE_{\\nu}$, which implies that for $E_{\\nu} \\approx 10^8 - 10^{10}$ GeV one have $x \\approx 10^{-4} - 10^{-6}$ at $Q^2 \\approx 10^4$ GeV$^2$. This kinematical range was not explored by the HERA measurements of the structure functions \\cite{hera}. The description of QCD dynamics in such very high energy limit is a subject of intense debate \\cite{hdqcd}. Theoretically, at high energies (small Bjorken-$x$) one expects the transition of the regime described by the linear dynamics, where only the parton emissions are considered, to a new regime where the physical process of recombination of partons becomes important in the parton cascade and the evolution is given by a non-linear evolution equation. This regime is characterized by the limitation on the maximum phase-space parton density that can be reached in the hadron wavefunction (parton saturation), with the transition being specified by a typical scale, which is energy dependent and is called saturation scale $Q_{\\mathrm{s}}$. Moreover, the growth of the parton distribution is expected to saturate, forming a Color Glass Condensate (CGC), whose evolution with energy is described by an infinite hierarchy of coupled equations for the correlators of Wilson lines \\cite{hdqcd}. In the mean field approximation, the first equation of this hierarchy decouples and boils down to a single non-linear integro-differential equation: the Balitsky-Kovchegov (BK) equation \\cite{BAL,KOVCHEGOV}. Experimentally, possible signals of parton saturation have already been observed both in $ep$ deep inelastic scattering at HERA and in deuteron-gold collisions at RHIC \\cite{hdqcd}. Currently, there are predictions of the neutrino nucleon cross sections with structure functions constrained by HERA data are based on linear dynamics \\cite{ccs}, using DGLAP or an unified DGLAP/BFKL evolution, or phenomenological models that resembles the expected behavior predicted by the non-linear QCD dynamics \\cite{bhm} (i.e., the proton structure function saturating the Froissart bound at asymptotic energies, $F_2^p \\propto \\ln^2 (1/x)$). As a general feature, the nonlinear QCD dynamics predicts sizable suppression of UHE neutrino cross section in comparison with standard approaches. Here, we summarize the main results of works presented in Refs. \\cite{magno1,magno2}. In Ref. \\cite{magno1} the contribution of non-linear effects was estimated considering distinct phenomenological models based on saturation physics. An update on those calculations have been done recently \\cite{victor}. In Ref. \\cite{magno2}, the geometric scaling property (which is a natural consequence of the asymptotic solutions of the nonlinear QCD evolution equations) is considered to obtain an analytical parameterization for the UHE neutrino cross sections. In what follows we introduce the theoretical and phenomenological tools and present the main results and predictions ", "conclusions": "" }, "1112/1112.5236_arXiv.txt": { "abstract": "We analyze \\textit{in situ} measurements of solar wind velocity obtained by the \\textit{Advanced Composition Explorer} (ACE) spacecraft during the solar activity cycle $23$. We calculated a robust complexity measure, the permutation entropy ($S$) of solar wind time series at different phases of a solar activity cycle. The permutation entropy measure is first tested on the known dynamical data before its application to solar wind time series. It is observed that complexity of solar wind velocity fluctuations at $1$ AU shows hysteresis phenomenon while following the ascending and descending phases of the activity cycle. This indicates the presence of multistability in the dynamics governing the solar wind velocity over a solar activity cycle. ", "introduction": "\\label{S-intro} The corona changes its shape enormously during a solar activity cycle resulting in temporal and structural properties of the solar wind velocity variation in a solar cycle \\cite{schw07}. \\textit{In situ} solar wind plasma observations show that its local properties at $1$~AU are modulated by the solar activity cycle \\cite{hapg91,rich08}. Around solar activity minimum, the structures of the corona and the solar wind are rather simple and remain so for several months \\cite{schw07}. At solar activity maximum, slow solar wind dominates at all helio latitudes \\cite{mcco00}. Recently it was reported that slow solar wind velocity just before the maximum of the solar activity cycle is least correlated to data obtained from the rest of the solar activity cycle \\cite{suya11}. Observations of solar wind velocity made by different spacecraft have been analyzed in considerable detail and reported by several authors. Values of complexity measures such as entropy (\\opencite{mace97}, \\citeyear{maco98}; \\opencite{mace00}; \\opencite{reda01}), correlation dimension \\cite{mace97,gupt08} and Lyapunov exponents \\cite{mace98,reda01,gupt08} show that solar wind velocity fluctuations are a consequence of complex nonlinear dynamical processes. Inherent changes in the dynamics governing the solar wind velocity at $0.3$ AU have been observed \\cite{gupt08}. \\inlinecite{mila04} analyzed magnetic and bulk velocities measured by ACE and found an anisotropy in the velocity, magnetic, and cross helicity correlation functions and power spectra. \\inlinecite{mcco00} analyzed Ulysses observations to demonstrate that the mid-latitude solar wind structure becomes increasingly complex as solar activity increases. \\inlinecite{cons09} investigated the emergence of spatio-temporal complexity in the $11$-year solar cycle monitored by sunspot activity. They showed that spatio-temporal or dynamical complexity is an intrinsic property of the solar cycle. Using information entropy approach to the sunspot number time series, they showed how the dynamical complexity increases during the maximum phase of the solar cycle. Hysteresis occurs in several phenomena in physics, chemistry, biology, and engineering. It is a nonlinear phenomenon observed in systems from diverse areas of science, \\textit{e.g}., electromagnetism, electro-plasticity, superconductivity, and granular motion \\cite{bert98,guye99,katz02,zhar02}. This phenomenon occurs when a nonlinear system has at least two coexisting stable states in the hysteresis region where the system is found to depend on the history of the dynamics. Here, in dynamical systems, history corresponds to the system's initial conditions. For example, hysteresis is observed in van der Pol system, Duffing system \\cite{thom86}, and Lorenz system \\cite{alfs85}. The hysteresis has been observed in coupled nonlinear systems as well \\cite{pras05}. Hysteresis phenomenon has been noticed in various solar indices. \\inlinecite{bach94} observed the presence of hysteresis patterns among many pairs of activity indices during solar cycle $21$ and $22$. They found that this hysteresis can be expressed approximately as a hierarchy of delay times behind the leading index, the sunspot number. \\inlinecite{reye98} analyzed the low-degree $p$-mode frequency shifts and solar activity indices (radio flux at $10.7$ cm and magnetic index) over solar cycle $22$ and observed a hysteresis phenomenon. \\inlinecite{more00} suggested that high latitude fields are necessary to produce a significant difference in hysteresis between odd and even-degree $p$-modes frequencies. \\inlinecite{trip00} reported that the intermediate degree $p$-mode frequencies of solar cycle $22$ show a hysteresis phenomenon with the magnetic indices whereas no such effect exists for the radiative indices. \\inlinecite{ozgu01} showed the presence of hysteresis between the solar flare index and some solar activity indicators such as total sunspot area, mean magnetic field, and coronal index during solar cycles $21$ and $22$. They found that these indices follow different paths for ascending and descending phases of the solar cycles while saturation effect exists at the extreme phases. In the present work, we attempt to understand the dynamics of ascending and descending phases of a solar cycle. We use permutation entropy $(S)$ of hourly averaged solar wind velocity time series to capture the complexity trend over a cycle. We use the data obtained from ACE during $1998-2010$. This period belongs to solar activity cycle $23$. We use permutation entropy $(S)$ to detect the hysteresis in a dynamical system and calculate it for time series obtained from simulated as well as solar wind velocity data. In the next section, we review the algorithm to calculate the permutation entropy $(S)$ of a time series. In Section~\\ref{S-dynamical_hysteresis} we describe how $S$ detects the multistability present in the modeled dynamical system. In Section~\\ref{S-solar_wind} we analyze solar wind data using permutation entropy. This is followed by conclusions in Section~\\ref{S-conc}. ", "conclusions": "The long term sunspot time series shows an average cycle length of $11$ years. Although nearly periodic, the period as well as amplitude of the cycle varies irregularly. Apart from the sunspot data, the intrinsic irregularity of the solar cycle is seen in other observable variables like surface flows, solar irradiance (solar constant), the solar wind, and so on. The Maunder Minimum and several ancient periods from solar proxy-data suggest that the Sun exhibits quasi-periodic or intermittent behavior \\cite{femi97}. Owing to changes in magnetic activity, many aspects of the solar wind change over a solar cycle, including the speed, the density, the dynamic pressure, the composition, and the temperature \\cite{rich08}. In this paper, we present the analysis of solar wind velocity data during the solar activity cycle $23$. We obtained $18$ different time series of hourly averaged solar wind velocity, measured by ACE spacecraft at $1$ A.U. To quantify the randomness of these time series, we use a robust, conceptually simple and computationally efficient measure called permutation entropy. A smaller value of permutation entropy indicates a more regular time series. We observe that as the solar cycle $23$ progresses towards maximum, the permutation entropy increases, saturates around the peak of activity and then decreases as the activity of cycle $23$ subsides. We also note (Figure. \\ref{fig5}) that the value of permutation entropy follows different paths in the ascending and descending phases of the solar activity cycle. In addition, while the ascent is fluctuating, the descent is smooth. This hysteresis phenomenon shows the multistability in the dynamics of solar wind, over the solar activity cycle. The behavior is similar to the one observed for hysteresis phenomenon of other solar indices \\cite{bach94,reye98,ozgu01} and confirms (\\textit{cf}. \\opencite{cons09}) that spatio-temporal or dynamical complexity is an intrinsic property of the solar cycle. \\label{S-conc} \\begin{acks} The authors thank the ACE Science Center and instrument teams for making available the ACE data used here. VS and AP thank CSIR for SRF and DST Govt. of India for financial supports respectively. \\end{acks}" }, "1112/1112.4837_arXiv.txt": { "abstract": "{ We propose a fully nonlinear framework to construct consistency relations for testing generic cosmological scenarios using the evolution of large scale structure. It is based on the covariant approach in combination with a frame that is purely given by the metric, the normal frame. As an example, we apply this framework to the $\\Lambda$CDM model, by extending the usual first order conditions on the metric potentials to second order, where the two potentials start to differ from each other. We argue that working in the normal frame is not only a practical choice but also helps with the physical interpretation of nonlinear dynamics. In this frame, effective pressures and anisotropic stresses appear at second order in perturbation theory, even for ``pressureless'' dust. We quantify their effect and compare them, for illustration, to the pressure of a generic clustering dark energy fluid and the anisotropic stress in the DGP model. Besides, we also discuss the effect of a mismatch of the potentials on the determination of galaxy bias. } ", "introduction": "\\label{sec:intro} Current cosmological observations provide evidence for a recent onset of a phase of accelerated expansion of the Universe \\cite{Riess:1998cb, Perlmutter:1998np, Komatsu:2010fb,Suzuki:2011hu,Sherwin:2011gv}. Under the assumption of large-scale homogeneity and isotropy, the observations can only be described in the framework of General Relativity (GR) by invoking some form of Dark Energy (DE) with negative pressure. Apart from the Standard Model (SM) of particle physics, the preferred model of the content of the Universe also assumes the presence of Cold Dark Matter (CDM) and DE in the form of a cosmological constant ($\\Lambda$) and is therefore named $\\Lambda$CDM. The dynamics of the Universe is based on the Einstein equations and on the assumption that the average geometry is described by a Friedmann-Lema\\^itre-Robertson-Walker (FLRW) metric. The formation of structures is then modelled using cosmological perturbation theory \\cite{Bardeen:1980kt, Kodama:1985bj, Mukhanov:1990me, Durrer:1993db, Ma:1995ey, Bernardeau:2001qr} which describes the evolution of small fluctuations in the energy-momentum content and the metric. The $\\Lambda$CDM model comes with a minimal set of parameters that are increasingly well constrained by the data and, to date, no significant departures from its predictions on cosmological scales have been found \\cite{Komatsu:2010fb, Bean:2010zq, Zhao:2010dz}. However, the difficulty in explaining the required tiny value of $\\Lambda$ has motivated many alternative scenarios with dynamical DE fields or modifications to GR aka Modified Gravity (MG)~\\cite{Wetterich:1987fm, Ratra:1987rm, ArmendarizPicon:1999rj, Dvali:2000hr, Deffayet:2001pu, Copeland:2006wr, Durrer:2008in, Tsujikawa:2010zza, Clifton:2011jh, Capozziello:2011et,Kunz:2012aw}. Given the existing tight constraints, it is essential to understand very accurately what the theoretical predictions are and how to interpret the observables, in order to maximise the benefits of the future data. A possible falsification of $\\Lambda$CDM would then be the first step towards a complete understanding of the dark energy. In practice, we can choose between comparing $\\Lambda$CDM to a large representative number of alternative scenarios, parameterise departures from it in a sufficiently flexible (but still economical) way \\cite{Caldwell:2007cw, Amendola:2007rr, Hu:2007pj, Song:2008vm, Bertschinger:2008zb, Song:2010rm, Pogosian:2010tj, Daniel:2010yt, Song:2010fg, Hojjati:2011ix, Zhao:2011te, Dossett:2011zp, Dossett:2011tn, Baker:2011jy, Zuntz:2011aq}, or identify key relations whose breakdown would hint towards a falsification of the standard paradigm \\cite{Huterer:2006mva, Song:2008vm, Song:2008xd}. All three strategies are useful in their own right: the third option is initially most powerful for falsifying or confirming $\\Lambda$CDM, while the second can be helpful as a guidance for the first, more traditional option. In any case, exquisite precision in the model predictions and understanding of the observables are needed to make decisive statements. In this work we propose and start to explore a program for constructing fully nonlinear consistency relations between geometry and matter content, not assuming GR but allowing for generic modifications of the field equations. This can be seen as a (nonlinear) step beyond the popular modified growth parameterisations \\cite{Caldwell:2007cw, Amendola:2007rr, Hu:2007pj, Song:2008vm, Bertschinger:2008zb, Song:2010rm, Pogosian:2010tj, Daniel:2010yt, Song:2010fg, Hojjati:2011ix, Zhao:2011te, Dossett:2011zp, Dossett:2011tn, Baker:2011jy, Zuntz:2011aq}. Such relations can either be implemented as consistency checks to falsify $\\Lambda$CDM (or some other paradigm), or can be parameterised to explore a possible departure from $\\Lambda$CDM. To construct the consistency relations we make use of the fact that at late times matter (baryonic and dark) is well described by a pressureless perfect fluid (dust) on cosmological scales. The absence of isotropic and anisotropic pressures in the matter rest frame directly implies two consistency relations between the geometry and the physics that gives rise to the late time accelerated expansion. However, we point out some important subtleties that arise due to the choice of frame and that can become relevant on nonlinear scales. We first formulate the consistency relations in the exact nonlinear theory employing the covariant approach \\cite{Ellis:1971pg, Ellis:1989jt, Bruni:1992dg, Ehlers:1993gf, Maartens:1998xg,Clarkson:2010uz}. Then, we discuss how the conditions come about after matter-radiation equality in first and second order cosmological perturbation theory, using the longitudinal (or conformal Newtonian) gauge. At first order, the resulting conditions are well known. They simply tell us that the two metric potentials are equal, $\\phi_1=\\psi_1$, and their evolution is governed by a simple second order differential equation that can be solved exactly: the Bardeen equation \\cite{Bardeen:1980kt}. At second order, the total measured pressure and anisotropic stress depend on the observer frame that is used to define these quantities, and this is where the subtleties come into play \\cite{Hwang:2005hd}. The pressure measured in the frame comoving with the matter is zero by definition (in analogy to the Lagrangian picture in the Newtonian approximation, see for instance \\cite{Hwang:2006iw,Villa:2011vt}). However, in a non-comoving frame, effective pressures and anisotropic stresses are induced by the relative velocity field, even if the source is a pressureless perfect fluid \\cite{Hwang:2005hd}. As we shall discuss in detail, there are good reasons to work in a specific non-comoving frame, called the \\emph{normal frame}. This frame is orthogonal to the surfaces of constant time and is given purely by the geometry, i.e.\\ the (perturbed) metric. An added advantage of this framework is that the density and pressure of matter (and of any other component) in the normal frame are then immediately related to the \\emph{geometrical pressures} that can be read off from the Einstein tensor. This means that projecting on the normal frame lets us cleanly separate the geometrical fluctuations from those of the stress-energy content. Consequently, the consistency relations in the normal frame take a particularly simple form. The change of frame from comoving to non-comoving induces effective matter pressures that we quantify. We find them to be small in comparison to typical non-standard sources of pressure and anisotropic stress in models beyond $\\Lambda$CDM. The outline of the paper is as follows. First, in section \\ref{sec:covariant}, we review the effects of changing the observer frame from comoving to non-comoving, and discuss why the normal frame is particularly useful. In section \\ref{sec:metric_linear} we state the consistency relations in the exact nonlinear case, discuss the connection between the covariant and the metric perturbations approach and review the first order consistency relations. In section \\ref{sec:nonlinear} we derive the second order consistency relations and compute the effective matter pressure and anisotropic stress. Section \\ref{sec:discussion} is devoted to the discussion of a number of typical cases where the effective matter stresses could potentially lead to wrong physical interpretations of the data. We consider the cases of a general clustering DE fluid and the DGP (Dvali-Gabadadze-Porrati \\cite{Dvali:2000hr}) model as a prototypical example of MG, because for both their respective intrinsic pressure and/or anisotropic stress perturbations could, in principle, be confused with the contributions from matter. Further, we discuss the impact on the determination of the galaxy bias from weak lensing and galaxy clustering. We summarise our work and conclude in section \\ref{sec:conclusions}. We include three appendices. In appendix \\ref{app:conventions} we state our conventions on notation and units. Appendix \\ref{app:covariant} briefly reviews the covariant approach and appendix \\ref{app:metric} gives more details on the metric perturbation calculations. ", "conclusions": "\\label{sec:conclusions} With current cosmological observations tightening constraints on the $\\Lambda$CDM scenario and not (yet) hinting at strong deviations from it, or from GR, it becomes increasingly important to devise ways of falsifying or confirming the standard paradigm. To be able to do so, a very precise understanding of the observations and the corresponding predictions is needed not only on linear but also on nonlinear scales. In recent years, much progress has been made in nonlinear Newtonian perturbation theory to understand and complement numerical analyses. Simulation efforts in non-standard scenarios are also underway. On the other hand, relativistic perturbation theory predictions and consistency relations are mostly restricted to linear theory. But to be able to confidently falsify or confirm the standard paradigm it is imperative to understand the implications of dynamics in curved space on weakly nonlinear scales where Newtonian theory may not be sufficient, even more so if modifications to GR are allowed. A framework based on casting beyond-$\\Lambda$CDM structure growth in terms of two functions, that can either be parameterised in generic ways or predicted from a given theory, has recently received considerable attention. This tool is often referred to as Parameterised-Post-Friedmann approach or modified growth parameterisations. In this paper, we propose an alternative framework deriving fully nonlinear consistency relations between the geometry, the matter content and any possible dark component. Since our framework requires neither the assumption of GR nor minimal interactions, but only relies on the symmetries of the Einstein tensor, it can be applied to different scenarios and models. The framework derives from the covariant approach, employing a very practical choice of 4-vector field (the normal frame) to decompose the Einstein tensor, the energy-momentum tensor and their difference, $X^\\mn\\equiv M_P^2 G^\\mn - T^\\mn$, that describes the physics beyond GR and/or beyond the standard constituents of the cosmological model. The normal frame, $n^\\mu$, is defined to be orthogonal to the surfaces of constant time. Because its definition is purely geometric it is only given by the metric itself and so are the irreducible components of $M_P^2G^\\mn$: the \\emph{geometrical} dynamic quantities. This is the reason why the normal frame is such a convenient choice, as it naturally allows us to write consistency relations that separate into a geometrical part, a matter part and a dark energy part: \\be \\nn p_X = p_G - \\frac{1}{3}v_m^2\\rho_m \\,, \\qquad \\pi_X^\\mn = \\pi_G^\\mn - \\rho_m v_m^{\\la\\mu}v_m^{\\nu\\ra} \\ee where the $G$-variables are only given by the metric, see section \\ref{sec:fullconsistency}. These fully nonlinear consistency relations can either be taken as conditions for $\\Lambda$CDM+GR to hold by requiring $p_X=-M_P^2 \\Lambda$, $\\pi_X^\\mn=0$, or the $X$-quantities can be parameterised and reconstructed from data. The advantage of this very general framework is that it is a top-down approach that a priori takes all non-linearities and all scalar, vector and tensor degrees of freedom into account, and can be simplified in perturbation theory at will. As a consequence of selecting the normal frame, the dynamical quantities on the matter side (the energy density, the pressure, the energy flux and the anisotropic stress) do not take the values that one would expect, or find when measured in the matter rest frame. The transformations from the rest frame to the normal frame can be written fully nonlinearly. These show that a non-vanishing effective pressure and anisotropic stress are measured in the normal frame, even if the matter is assumed to be a perfect pressureless fluid in its rest frame, see section \\ref{sec:framechange}. This is the origin of the terms $\\sim v_m^2$ in the consistency relations. However, these effects are only relevant at second order in perturbation theory and they can be accounted for, knowing that they are present. Furthermore, we point out that cosmological observations do not actually measure the density field in its comoving frame, as we do observe the effect of the velocity field of the sources through the projection onto the photon wave vector. This means that there are non-comoving effects that need to be taken into account. More importantly, apart from gravitational nonlinearities, we identify the pressure and anisotropic stress measured in the normal frame as the physical quantities that enter as sources in the Bardeen equation for the second order Weyl potential and the constraint for the second order gravitational slip. This is a consequence of the fact that the normal frame is purely defined in terms of the metric. Finally, let us recall that the normal frame helps to recover the Eulerian picture of Newtonian perturbation theory from GR. Although GR is a theory of curved spacetime, the normal frame is related to the Eulerian picture in Newtonian nonlinear dynamics. We compute the consistency relations for $\\Lambda$CDM in first and second order scalar perturbation theory in the generalised longitudinal gauge (that is the gauge where $n^\\mu$ has zero spatial components and does not see shear or vorticity). At first order we recover the usual conditions: the gravitational slip vanishes, $\\Pi_1\\equiv\\phi_1-\\psi_1=0$, and the Weyl potential, $\\vf_1\\equiv\\frac{1}{2}(\\phi_1+\\psi_1)$, follows the standard Bardeen equation, see section \\ref{sec:linear}. We solve the Bardeen equation analytically to provide the basis for a careful study of the matter pressure and anisotropic stress. We compute $\\frac{1}{3}v_m^2\\rho_m$ and $\\rho_m v_m^{\\la\\mu}v_m^{\\nu\\ra}$ at second order and find these terms to be small on large scales. They become relevant only on scales $k\\gtrsim 0.1\\, h/\\Mpc$. The second order anisotropic stress consistency relation reveals a non-vanishing gravitational slip due to the effective matter anisotropic stress and non trivial combinations of products of spatial derivatives of $\\vf_1$, see equation \\eqref{eq:k4pi} and also refs. \\cite{Acquaviva:2002ud,Vernizzi:2004nc}. Here we clarify the origin of the different contributions to $\\Pi_2$. The pressure consistency, on the other hand, yields the second order analogue of the Bardeen equation for $\\vf_2$, see equation \\eqref{eq:bardeen2}. It acquires different source terms which we analyse, compute and compare and we find subtle cancellations between the contributions from the effective matter stresses and $\\partial^2\\vf_1^2$--type terms. The fact that the source term of the second order Bardeen equation depends non trivially and sensitively on the pressure and anisotropic stress perturbations of a possible dark energy makes the weakly non-linear evolution a promising tool to test the $\\Lambda$CDM paradigm. Combining different cosmological probes enables us to reconstruct the consistency relations, as all observations probe the metric potentials, either directly or indirectly via the Poisson equation, see also e.g.\\ \\cite{Song:2008xd}. Let us emphasise that the effective matter pressure and anisotropic stress are seen by the second order metric perturbations in any case, irrespectively of the yet to be solved issue of which precise frame is the one that corresponds to us as observers. Finally, we discuss a set of applications: we compare the effective matter pressure with the pressure perturbations in clustering dark energy, concluding that the matter pressure is negligible on large scales and becomes relevant on weakly nonlinear scales. Thus, the matter pressure only needs to be taken into account when deriving constraints on the dark energy sound speed from small scale observations. Furthermore, we compare the matter anisotropic stress to typical anisotropic stresses arising in MG scenarios, specifically the DGP model. Anisotropic stress is normally considered a smoking gun for the detection of MG. We find that, typically, the matter anisotropic stress is by far subdominant and not an issue. Finally, we assess the impact of the anisotropic stress on the determination of galaxy bias from combining WL and galaxy clustering data. We find that the gravitational slip affects the Lagrangian galaxy bias and the velocity bias in opposite ways, but only marginally. Using multiple tracers at several redshifts and different combinations of observables, we expect that one could be able to filter out or strongly constrain the gravitational slip." }, "1112/1112.1415_arXiv.txt": { "abstract": "We examine to what extent the inferred surface temperature of magnetars in quiescence can constrain the presence of a superfluid in the neutron star core and the role of magnetic field decay in the core. By performing detailed simulations of neutron star cooling, we show that extremely strong heating from field decay in the {\\it core} cannot produce the high observed surface temperatures nor delay the onset of neutron superfluidity in the core. We verify the results of Kaminker et al., namely that the high magnetar surface temperatures require heating in the neutron star {\\it crust}, and crust heating is decoupled from cooling/heating in the core. Therefore, because crust heating masks core heating, it is not possible to conclude that magnetar cores are in a non-superfluid state purely from high surface temperatures. From our interior temperature evolutions and after accounting for proton superconductivity in the core, we find that neutron superfluidity in the core occurs less than a few hundred years after neutron star formation. This onset time is unaffected by heating due to core field decay at fields $\\lesssim 10^{16}\\mbox{ G}$. Thus all known neutron stars, including magnetars, without a core containing exotic particles, should have a core of superfluid neutrons and superconducting protons. ", "introduction": "\\label{sec:intro} Neutron stars (NSs) begin their lives very hot (with temperatures $T>10^{11}\\mbox{ K}$) but cool rapidly through the emission of neutrinos. Neutrino emission processes depend on uncertain physics at the supra-nuclear densities ($\\rho>\\densnuc\\approx 2.8\\times 10^{14}\\mbox{ g cm$^{-3}$}$) of the NS core (see \\citealt{tsuruta98,yakovlevpethick04,pageetal06}, for review). Current theories indicate that the core may contain a neutron superfluid and proton superconductor or even exotic particles, such as hyperons and deconfined quarks (see, e.g., \\citealt{lattimerprakash04,haenseletal07}, for review). The recent observation of rapid cooling \\citep{heinkeho10,shterninetal11} of the NS in the Cassiopeia~A supernova remnant provides the first constraints on the critical temperatures for the onset of superfluidity of core neutrons $\\Tcnt$ (in the triplet state) and protons $\\Tcp$ (in the singlet state), i.e., $\\Tcnt\\approx(5-9)\\times 10^8$~K and $\\Tcp\\sim (2-3)\\times 10^9$~K \\citep{pageetal11,shterninetal11}. Anomalous X-ray pulsars and soft gamma-ray repeaters form the magnetar class of NSs, i.e., systems which possess superstrong magnetic fields ($B\\gtrsim 10^{14}\\mbox{ G}$)\\footnote{However, there exists an apparently low surface magnetic field magnetar \\citep{reaetal10}.}, and their strong fields likely power the activity seen in these objects (see \\citealt{woodsthompson06,mereghetti08}, for review). One notable property of magnetars is that their observed surface temperatures $\\Ts$ in quiescence are significantly higher than those of other NSs of a similar age. In fact, they are too high for NSs that cool passively, i.e., without an additional source of internal heat (accretion heating can be excluded by, e.g., non-detections of binary companion or disk emission). An interesting problem concerns the heat generated from magnetic field decay, which has been proposed to be the source for the high temperatures of magnetars \\citep{thompsonduncan96,heylkulkarni98,colpietal00,aguileraetal08b}. This heat can strongly influence the time/age at which the core becomes superfluid if heating/field decay occurs in the core \\citep{thompsonduncan96,arrasetal04,dallossoetal09}. The problem is important since the presence of superfluid components has a strong impact on magnetar interior dynamics, such as the mechanism for producing glitches (see, e.g., \\citealt{sauls89}), fluid oscillations (see, e.g., \\citealt{passamontiandersson11}), magnetic field and rotational evolution (see, e.g., \\citealt{glampedakisandersson11}), and magnetohydrodynamical equilibrium \\citep{glampedakisetal12,landeretal12}. We address this problem by conducting detailed calculations of the thermal evolution of a NS with various prescriptions for an internal heat source that can be associated with magnetic field decay. We show that, regardless of the magnetic field strength and detailed mechanism for field decay in the core, the heat generated by the decay is insufficient to power the surface emission of magnetars. Furthermore, by accounting for the effects of proton superconductivity on field decay, we find that this core heating is not strong enough to balance neutrino cooling and cannot delay the onset of core neutron superfluidity for fields $\\lesssim 10^{16}\\mbox{ G}$ (onset at age $\\lesssim\\mbox{a few}\\times 100\\mbox{ yr}$). Thus the cores of all currently known magnetars should be in a superfluid state. We briefly describe past works and note key findings and assumptions made in these works that we improve upon here. \\citet{arrasetal04} consider coupled magnetic field decay and thermal evolution of magnetars. The internal field and superfluid temperatures are assumed to be $>10^{15}\\mbox{ G}$ and $\\Tcp=5\\times 10^9\\mbox{ K}$ and $\\Tcnt=(5-9)\\times 10^8\\mbox{ K}$, respectively. \\citet{arrasetal04} find that magnetic field decay can delay the transition to core neutron superfluidity and maintain a relatively high surface temperature to ages $\\approx 10^3-10^5\\mbox{ yr}$, depending on $\\Tcnt$. However, their calculation only considers volume-averaged quantities and thus assumes that the NS interior is isothermal (see Section~\\ref{sec:nscooleq}), which cannot be the case if there is a localized heating source such as field decay in the crust or core due to the large but finite thermal conductivity of NS matter. \\citet{kaminkeretal06} calculate the evolution of the temperature profile [i.e., $T(\\rho)$] by solving the energy balance and flux equations [see eqs.~(\\ref{eq:energybalancegr}) and (\\ref{eq:heatfluxgr})]. They demonstrate that, in order to explain the observed high surface temperatures, magnetars require a heat source and, most importantly, this heat source (e.g., from field decay; \\citealt{ponsetal09}) must be located in the outer crust; if the heat source is located too deep in the NS interior (e.g., in the core), then neutrino emission efficiently removes the heat locally, and the surface temperature cannot be increased sufficiently to match the observed values. Furthermore, the outer crust is thermally decoupled from the core so that heating of the crust does not affect (neutrino) cooling of the core. Because of this thermal decoupling, the results and conclusions of \\citet{kaminkeretal06} are not particularly sensitive to the state of matter in the core, e.g., superfluidity of the core nucleons. Though no quantitative results are shown, \\citet{kaminkeretal06} state that cooling calculations with the effects of inner crust superfluidity and neutrino emission by Cooper pair formation produce different temperature profiles, but these effects do not change the surface temperature. In follow-up work, \\citet{kaminkeretal09} improve their calculations and examine the effects of light-element accreted envelopes and anisotropic heat conduction due to the magnetic field in the envelope and outer crust. While accreted envelopes can give higher surface temperatures for the same core temperatures, core heating is still unable to produce surface temperatures that are high enough to explain magnetar observations. \\citet{dallossoetal09} criticize the work of \\citet{kaminkeretal06}, arguing that the phenomenological heating model considered is independent of magnetic field and decays on the wrong timescale [see eq.~(\\ref{eq:crustheat})] and that the correct heating should depend inversely on temperature. This last point implies that more heat is generated at lower temperatures and can lead to an equilibrium between neutrino cooling and heating from field decay (see also \\citealt{thompsonduncan96}). The equilibrium temperature (above $10^9\\mbox{ K}$) can be maintained for $\\approx 10^4\\mbox{yr}$. The results of \\citet{dallossoetal09} suggest that the core of magnetars are not superfluid until after this time, since the critical temperature for the onset of neutron superfluidity is $(5-9)\\times 10^8\\mbox{ K}$ \\citep{pageetal11,shterninetal11}. However no cooling calculation is performed by \\citet{dallossoetal09}. Here we perform NS cooling simulations to determine the role of core heating by magnetic field decay. We do not examine in detail field evolution and heating in the crust (see, e.g., \\citealt{ponsetal09}). Rather we use the phenomenological model of \\citet{kaminkeretal06,kaminkeretal09} to demonstrate the effect of {\\it crust heating} on the NS cooling behavior. This will be adequate for our purposes since the goal here is to assess the importance of {\\it core heating} in magnetars. As we will show, as long as the magnetic field decay time is longer than the cooling time, the core heating we use is the maximum one can use. While this maximal heating can delay the onset of neutron superfluidity [on the timescale estimated by \\citet{thompsonduncan96,dallossoetal09}], it cannot reproduce the high surface temperatures seen for magnetars. We arrive at this same conclusion using the prescription for heating and field decay given in \\citet{dallossoetal09}. Furthermore, the above works neglect the effects of proton superconductivity; when this is taken into account, there is no delay in superfluidity onset for any reasonable core magnetic field strength. In Section~\\ref{sec:model}, we describe the thermal evolution equations and input physics, including superfluid properties and prescriptions for internal heat sources. Section~\\ref{sec:results} presents the results of our calculations. We summarize our results and discuss their implications in Section~\\ref{sec:discuss}. ", "conclusions": "\\label{sec:discuss} We performed detailed calculations of neutron star cooling, including the effects of superfluidity and additional heating (due to magnetic field decay) in the crust and core. We find that magnetic field decay in the neutron star core cannot be the sole source powering the high observed surface temperature of magnetars in quiescence; the high temperatures require outer crust heating \\citep{kaminkeretal06,kaminkeretal09}. Because of crust heating and effective thermal decoupling between the outer crust and core, the state of matter in the core cannot be deduced from these surface temperature measurements. By computing the evolution of the temperature profile $T(\\rho)$, we determine the time when core neutrons first become superfluid, i.e., when $T<\\Tcnt(\\rho)$. We find that heating by field decay in the core (with fields $\\lesssim 10^{16}\\mbox{ G}$) cannot balance neutrino cooling and thus cannot maintain relatively high core temperatures (c.f., \\citealt{thompsonduncan96,dallossoetal09}). As a result, onset of superfluidity for neutrons in the core cannot be delayed, and neutron stars possess superfluid and superconducting cores after a few hundred years; this does not strongly depend on the nuclear EOS. Since core heating is not significant, the temperature profiles $T(\\rho)$ and surface temperature evolution $\\Ts^\\infty(t)$ for magnetars are just those with crust heating (see Figs.~\\ref{fig:tempprofcrust} and \\ref{fig:timets}, respectively), and crust heating does not affect the onset of core neutron superfluidity because of thermal decoupling between the outer crust and core. Magnetar activity may be driven by field decay in the core \\citep{thompsonduncan95,thompsonduncan96} or by processes in the crust \\citep{thompsonduncan93,thompsonduncan95,glampedakisetal11,priceetal12}. Although the former may still be true, it must be accompanied by heating in the (outer) crust. On the other hand, field evolution in the crust easily couples to surface emission. Thus in order to understand the high surface temperature of magnetars, detailed studies should focus on magnetic field evolution and heating in the crust (see, e.g., \\citealt{ponsetal09,cooperkaplan10,priceetal12}). We note that several magnetars may have very similar X-ray luminosities \\citep{durantvankerkwijk06}; this could suggest that their crustal field strengths are similar and the field decay timescale is longer than the age of the oldest of these sources. Finally, the (normal versus superfluid) state of the core has important consequences for magnetic and rotational evolution of magnetars \\citep{glampedakisandersson11}, as well as their glitching behavior and possible stellar oscillations; studies of these effects may be effective probes of the neutron star core. We showed that the core can be treated as being in a superfluid and superconducting state after the neutron star is a few hundred years old." }, "1112/1112.1765_arXiv.txt": { "abstract": "The current direct observations of brown dwarfs and exoplanets have been obtained using instruments not specifically designed for overcoming the large contrast ratio between the host star and any wide-separation faint companions. However, we are about to witness the birth of several new dedicated observing platforms specifically geared towards high contrast imaging of these objects. The Gemini Planet Imager, VLT-SPHERE, Subaru HiCIAO, and Project 1640 at the Palomar 5m telescope will return images of numerous exoplanets and brown dwarfs over hundreds of observing nights in the next five years. Along with diffraction-limited coronagraphs and high-order adaptive optics, these instruments also will return spectral and polarimetric information on any discovered targets, giving clues to their atmospheric compositions and characteristics. Such spectral characterization will be key to forming a detailed theory of comparative exoplanetary science which will be widely applicable to both exoplanets and brown dwarfs. Further, the prevalence of aperture masking interferometry in the field of high contrast imaging is also allowing observers to sense massive, young planets at solar system scales ($\\sim$3-30 AU)--- separations out of reach to conventional direct imaging techniques. Such observations can provide snapshots at the earliest phases of planet formation---information essential for constraining formation mechanisms as well as evolutionary models of planetary mass companions. As a demonstration of the power of this technique, I briefly review recent aperture masking observations of the HR 8799 system. Moreover, all of the aforementioned techniques are already extremely adept at detecting low-mass stellar companions to their target stars, and I present some recent highlights. ", "introduction": "In recent years, astronomers have identified more than 400 planets outside our solar system, launching the new and thriving field of exoplanetary science (Marcy \\etal~2005). The vast majority of these objects have been discovered indirectly by observing the variations induced in their host star's light. The radial velocity surveys can provide orbital eccentricity, semi-major axes, and lower limits on the masses of companion planets while observations of transiting planets can provide fundamental data on planet radii and limited spectroscopy of the planets themselves. However, studying those objects out of reach to the radial velocity and doppler methods will reveal completely new aspects of exoplanetary science in great detail. {\\it Direct imaging} of exoplanets provides a complementary set of parameters such as photometry (and hence luminosity), as well as detailed spectroscopic information. \\begin{figure} \\centering \\resizebox{13.5cm}{!}{\\includegraphics[angle=-90]{f1.ps}} \\caption[]{{\\it Left:} A schematic demonstration taken from \\cite{oh09} demonstrating the challenges associated with high contrast imaging. Hypothetical orbits of a Jupiter and an Earth have been overlaid on the Point Spread Function of a nearby star. Planetary mass companions typically have angular separations of a fraction of an arcsecond, and contrast ratios spanning several orders of magnitude ($\\sim$10$^4$-10$^7$). {\\it Right:} The Mass vs. Orbital separation for planetary mass companions detected through transits (filled circles), radial velocity measurements (squares), solar system planets (triangles), as well as those objects detected through direct imaging (open circles). } \\label{highcontrast} \\end{figure} Moreover, the direct imaging of planetary mass companions will allow researchers to more fully probe the mass-separation parameter space (Figure~\\ref{highcontrast}, right panel) occupied by these objects (Oppenheimer \\& Hinkley 2009). At young ages (\\S 3), the orbital placement of planetary mass companions serves as a birth snapshot, lending support to models (Pollack \\etal~1996) that may be more efficient at building a massive core at only a few AU, or models that allow for the formation of massive objects at tens of AU though fragmentation of the gaseous disk (Boss 1997). Finally, imaging of multi-planet systems at any age will serve as dynamical laboratories for studying the planetary architectures. \\begin{figure} \\centering \\resizebox{14.5cm}{!}{\\includegraphics[angle=90]{f2.ps}} \\caption[]{{\\it Top Panel:} Example Spectra for M, L, and T-dwarfs, as well as Jupiter with prominent absorption bands marked, taken from Marley \\etal~(2009). {\\it Lower Panels:} The points with error bars are simulated measurements gathered with a low resolution ($\\lambda/\\Delta\\lambda$$\\sim$45) spectrograph, shown along with higher resolution spectra from NASA IRTF (black curves) for the labelled spectral types. Such low spectral resolution measurements will be obtained by GPI, SPHERE and Project 1640, and can easily sample broad absorption features present in late type stars and exoplanets. Lower panels are taken from Rice \\etal~(2011) in prep. Images used courtesy of Michael Cushing, Mark Marley, and Emily Rice. } \\label{spectra} \\end{figure} Perhaps just as essential, spectroscopy provides clues to the atmospheric chemistry, internal physics, and perhaps may even shed light on non-equilibrium chemistry associated with these objects. More robust classification schemes for planets in general will arise from observing as many planets as possible at different ages, in different environments, and with a broad range of parent stars. Figure~\\ref{spectra} provides an illustration of the diversity of the spectra of late-M, L and T dwarfs as well as a spectrum of Jupiter showing broad absorption bands due to e.g. H$_2$O, CH$_4$, and NH$_3$. Spectral characterization of such objects can be accomplished even with the somewhat low spectral resolution ($\\lambda/\\Delta\\lambda\\sim$30-50) of the instruments described in this review. These pieces of information not only reveal the detailed properties of the objects themselves, they serve as key benchmarks for competing evolutionary models describing the thermal and atmospheric properties of these objects. It is often overlooked, but should be emphasized, that the recent spectacular images (Marois \\etal~2008, Kalas \\etal~2008, and Lagrange \\etal~2010) of Jovian planets were obtained with Adaptive Optics, hereafter ``AO'', systems and infrared cameras not specifically designed for the task of overcoming the high contrast ratio between the planets and their host stars. These were obtained using existing instrumentation, but with observing and data reduction strategies customized for high contrast imaging. Further, the recent L-band images of HR 8799 and $\\beta$ Pic b (Marois \\etal~2010, Lagrange \\etal~2010) were obtained without the use of a coronagraph! These studies have demonstrated that direct imaging of planetary mass companions as well as disks (e.g. Oppenheimer \\etal~2008, Hinkley \\etal~2009) is now a mature technique and may become routine using ground-based observatories. More so, the handful of coming instruments dedicated to detailed spectroscopic characterization of planetary mass companions may make these kinds of discoveries routine, initiating an era of comparative exoplanetary science. \\subsection{The Challenge of High Contrast Imaging} The major obstacle to the direct detection of planetary companions to nearby stars is the overwhelming brightness of the host star. If our solar system were viewed from 20 pc, Jupiter would appear $10^8-10^{10}$ times fainter than our Sun in the near-IR (Barraffe \\etal~2003) at a separation of 0.25$^{\\prime\\prime}$, completely lost in its glare (See Figure~\\ref{highcontrast}, left panel). The key requirement is the suppression of the star's overwhelming brightness through precise starlight control (Oppenheimer \\& Hinkley 2009). A promising method for direct imaging of stellar companions involves two techniques working in conjunction. The first, high-order AO, provides control and manipulation of the image by correcting the aberrations in the incoming stellar wave front caused by the Earth's atmosphere. AO has the effect of creating a nearly diffraction-limited point spread function, with the majority of the stellar flux concentrated in this core. Second, a Lyot coronagraph (Sivaramakrishnan \\etal~2001) suppresses this corrected light. Together, these two techniques can obtain contrast levels of $10^4$-$10^5$ at 1$^{\\prime\\prime}$ (Leconte \\etal~2010). Improvements in coronagraphy, specifically the apodization of the telescope pupil (Soummer 2005), as well as post-processing to suppress speckle noise (Hinkley \\etal~2007, Crepp et al 2011), can significantly improve the achieved contrast, especially at high Strehl ratios. Below we briefly describe some of the instrumentation being built at the time of this writing, or currently in place on large telescopes. ", "conclusions": "" }, "1112/1112.6373_arXiv.txt": { "abstract": "We have performed extensive two-dimensional magnetohydrodynamic simulations to study the amplification of magnetic fields when a supernova blast wave propagates into a turbulent interstellar plasma. The blast wave is driven by injecting high pressure in the simulation domain. The interstellar magnetic field can be amplified by two different processes, occurring in different regions. One is facilitated by the fluid vorticity generated by the ``rippled\" shock front interacting with the background turbulence. The resulting turbulent flow keeps amplifying the magnetic field, consistent with earlier work \\citep{Giacalone2007}. The other process is facilitated by the growth of the Rayleigh-Taylor instability at the contact discontinuity between the ejecta and the shocked medium. This can efficiently amplify the magnetic field and tends to produce the highest magnetic field. We investigate the dependence of the amplification on numerical parameters such as grid-cell size and on various physical parameters. We show the magnetic field has a characteristic radial profile that the downstream magnetic field gets progressively stronger away from the shock. This is because the downstream magnetic field needs a finite time to reach the efficient amplification, and will get further amplified in the Rayleigh-Taylor region. In our simulation we do not observe a systematic strong magnetic field within a small distance to the shock. This indicates that if the magnetic-field amplification in supernova remnants indeed occurs near the shock front, other processes such as three-dimensional instabilities, plasma kinetics and/or cosmic ray effect may need to be considered to explain the strong magnetic field in supernova remnants. ", "introduction": "Powerful shocks associated with supernova remnants (hereafter SNRs) sweeping through the interstellar medium (ISM) are remarkable high-energy phenomena in astrophysics. It is widely believed that these high-Mach number shocks are the sources of galactic cosmic rays with energies up to at least $10^{15}$ eV. SNRs are also sources of strong radio and/or X-ray emissions. In these high-energy processes, the magnetic field is of great importance. Moreover, it provides information on energetic charged particles, which are presumably accelerated by the supernova shocks. The ISM is known to be turbulent. Measurements of the ISM radio-wave scintillation have established the existence of large-scale density turbulence which has a Kolmogorov-like power spectrum spanning more than ten decades of spatial scale with an outer scale of several parsecs \\citep[e.g.,][]{Lee1976ApJ,Armstrong1981,Rickett1990,Armstrong1995,Minter1996}. This has been called ``the big power law in the sky.\" \\citep[see,][]{Spangler2007} The galactic magnetic field is observed to be a few micro-Gauss and has uniform and fluctuating components that are roughly in equipartition \\citep[e.g.,][]{Beck1996,Minter1996,Han2004}. The turbulent magnetic field can interact with the shock waves, distorting their surfaces leading to shock ripples \\citep{Neugebauer2005} and the enhanced downstream magnetic fluctuations \\citep{Lu2009}. It is also important for efficient particle acceleration \\citep{Giacalone2005,Jokipii2007,Guo2010a}. Turbulence in the upstream medium has also been considered \\citep[][]{Balsara2001} to explain the irregular and patchy emission morphology observed in SNRs \\citep[e.g.,][]{Anderson1996}. Recently, it has been inferred from observations that the magnetic field in young SNRs is strongly enhanced to a magnitude much greater than the compression given by the shock jump condition. For example, by assuming that the so-called X-ray ``thin rims\" seen in several young SNRs \\citep[e.g.,][]{Bamba2005} are caused by shock accelerated electrons rapidly losing energy in strong magnetic field through synchrotron radiation, the associated magnetic field may be more than $50-100\\mu G$ in order to explain the thickness of the ``thin rims\" \\citep{Berezhko2003,Volk2005,Ballet2006,Parizot2006}. For SNR shocks with higher shock speeds propagating in more inhomogenous media such as Cas A and Tycho, the downstream magnetic fields are inferred to be more enhanced. ``Thin rims\" are also seen in radio emissions \\citep{Reynoso1997}, which cannot be explained by the electrons losing energy in strong magnetic field. This indicates some other mechanism, e.g., decay of magnetic fluctuations may need to be considered \\citep{Pohl2005}. Further downstream of the shock, the magnetic field may possibly be even higher than the region right behind the shock \\citep{Vink2003}. The morphology of X-ray emission in SNRs shows filamentary structure and rapid time variation, which indicates that the magnetic field could be as high as $1 mG$ \\citep{Uchiyama2007}, over two orders of magnitude higher than the background fluid. It should be pointed out this rapid variation of synchrotron emission could be well reproduced in the case of strong magnetic fluctuations \\citep{Bykov2008}. Since the effect of turbulence is not fully understood at this point, the actually magnetic-field amplification factors in young SNRs remain uncertain. \\citet{Bell2001} and \\citet{Bell2004} proposed that cosmic rays accelerated by the SNR forward shock waves provide a current that leads to an instability that can amplify the magnetic field close to the shock front. Numerical simulations show the evidence of this instability, however, it is found to easily saturate and the amplification factor may be limited \\citep[e.g.,][]{Riquelme2009}. Recently, \\citet{Giacalone2007} proposed an alternative mechanism, in which the interaction between the warped shock front and large-scale density fluctuations produces fluid vorticity downstream of strong shocks. That fluid vorticity can stretch, distort and amplify the magnetic field. The magnetic-field amplification in this mechanism relies on the dynamics of a magnetized fluid rather than the cosmic-ray kinetic physics. It is interesting to note, however, \\citet{Balsara2001} performed three-dimensional MHD blast wave simulation with moderate resolution which does not show magnetic-field enhancement larger than $50 \\mu G$, whereas two-dimensional Cartesian geometry simulations with high numerical resolution give strong amplification \\citep{Giacalone2007,Inoue2009} with maximum values larger than $100\\mu G$. This discrepancy warrants further investigation. In addition, a critical constraint from several observed ``thin rims\" is that the required magnetic-field amplification should occur within a narrow distance $\\lesssim 0.01$ pc of the supernova shocks \\citep[for a review, see][]{Reynolds2011}. This is supported by some coincidences of X-ray ``thin rims\" and shock locations inferred from H$\\alpha$ observations \\citep{Winkler2003} and radio polarimetry \\citep{Gotthelf2001}. This places an important constraint on various field amplification models, although it should be noted that there is no one-to-one correspondence between X-ray ``thin rims\" and inferred shock locations. To our knowledge, this constraint has not been taken into account previously in comparing models and simulations with observations. In this study, we perform a series of two-dimensional ideal MHD simulations with high spatial resolution to study strong supernova blast shock waves propagating into the ISM containing pre-specified large-scale density and magnetic fluctuations. We investigate how the amplification depends on a variety of parameters, including the explosion energy, the level of background turbulence, and the numerical resolution, etc. The paper is organized as follows. In Section 2, we describe our numerical model and simulation setup. We present the simulation results in Section 3, and in Section 4 we summarize and discuss our results. ", "conclusions": "The inferred strong magnetic field in young SNRs is a significant result and can be important in the high-energy processes including particle acceleration and thermal/nonthermal emissions. The origin of this process, however, is still under debate. In this work we study the interaction between a supernova blast wave with a turbulent upstream medium which contains density and magnetic-field fluctuations. The vorticity produced at the rippled shock front can stretch and distort the magnetic field lines, and this leads to a strong magnetic-field amplification downstream \\citep{Giacalone2007,Inoue2009}. Using two-dimensional MHD simulations of a blast wave, we confirm the key features of this process. Based on our simulations, we conclude that the increase of magnetic field is dependent on shock speed and background density turbulence amplitudes. Furthermore, the numerical resolution used in the simulations can play an important role as well. Previous work \\citep{Balsara2001} using three-dimensional MHD simulation with moderate resolution shows no magnetic-field amplification beyond $50\\mu G$. Here we show the magnetic evolution downstream is sensitive to the resolutions used in the simulation. For high resolutions, the simulations allow rapid growth at small scales, this leads to efficient field amplification. Furthermore, we find that there are two different processes and spatial regions where magnetic field is amplified. One is associated with the shock amplification immediately downstream and the other is associated with the RTI at the interface between the ejecta and the shocked medium. However, in our simulations, we did not observe a systematic strong magnetic field within a thin region immediate downstream of the supernova shock. For example, using the results of the highest resolution case, within $0.15$ pc downstream of supernova shock, we observe only about $0.8\\%$ region which has magnetic field larger than $30 \\mu G$. This lack of strong magnetic field can be understood as the downstream dynamo process requires an efficient stretching to produce strong magnetic field. The time scale for the growth of magnetic field depends on the eddy turnover time. Only after a certain time can the field get sufficient amplification. If the thin rims ($0.01 - 0.1$ pc) observed in young SNRs are indeed caused by the electrons losing energy in strong fields ($\\sim$ several hundred $\\mu G$), some other processes such as three-dimensional instabilities, plasma kinetics or the effect of cosmic rays might be needed to explain the magnetic-field amplification in young SNRs. We note that in observation there is no one-to-one correspondence between X-ray ``thin rims\" and inferred shock locations. In fact many ``thin rims\" and filamentary structures seen in X-ray observation can hardly be related to shock front due to observation limitation. Further understanding about the relationship between the small scale X-ray structure and shock locations is needed to further constraint and distinguish the different mechanisms for amplification of magnetic field. We also note that the two-dimensional simulation in this study could be significantly different from three-dimensional simulation. It is known that in two-dimensional simulation the dynamics of MHD flow is very different from that for three-dimensional simulation. For example the inverse cascade of enstrophy can causes strong intermittency due to two-dimensional effect \\citep{Biskamp2003}. There are several recent studies show that magnetic field amplification behind high-mach number shock \\citep{Inoue2011a,Inoue2011b} and in Rayleigh-Taylor region \\citep{Stone2007} can still operate in three-dimensions, which confirm the results found in two-dimensional simulations. Also, in three-dimensional simulation the MHD flow could develop other types of instabilities with larger growth rates. Further three-dimensional MHD simulation with high resolution will be useful in confirming the conclusions of this paper." }, "1112/1112.1881_arXiv.txt": { "abstract": "{In most of particle dark matter (DM) models, the DM candidate injects sizable fluxes of high-energy electrons and positrons through its annihilations or decays. Emitted in regions with magnetic field, they in turn give raise to a synchrotron radiation, which typically covers radio and infrared bands. We discuss the possibility of detecting signatures of Galactic and extra-galactic DM in the total intensity and small-scale anisotropies of the radio background.} \\FullConference{XXIst International Europhysics Conference on High Energy Physics\\\\ 21-27 July 2011\\\\ Grenoble, Rhones Alpes France} \\begin{document} ", "introduction": "From the title of my talk, one may wonder how it can be possible to infer something completely unknown, as non-gravitational signals of dark matter (DM), by means of something quite well-known, as radio-astronomy. Discovery of new physics normally goes through designing new experiments. Here the idea is very simple, namely, that the technique of radio observations is potentially very promising in the quest for particle DM, but requires improved flux and angular sensitivities with respect to current capabilities. The development of ASKAP, EVLA, MeerKAT, and, in particular, SKA makes the near future promisingly bright in this respect. Those new radio telescopes will certainly discover signals of previously unknown physical mechanisms, and we discuss the possibility that an emission induced by particle DM will be among them. Weakly interacting massive particles (WIMPs) are the most investigated class of DM candidates in the literature (for a review, see, e.g.,~\\cite{Bertone:2010zza}). One of the routes to test the hypothesis of WIMP DM stems from the the bases of the framework themselves. Indeed, given the weak interaction, there is a (\"weak\" but finite) probability that WIMPs in DM halos annihilate in pairs or decay into detectable species. In particular, and with the exception of WIMP models annihilating/decaying into neutrinos only, a sizable branching ratio of annihilation/decay into electrons and positrons is a general feature of WIMP models (see, e.g., Fig.~4 in \\cite{Regis:GC}). Interactions of high-energy $e^+/e^-$ with the interstellar magnetic field in astrophysical structures give rise to magnetic bremsstrahlung called synchrotron radiation. In the monochromatic approximation of synchrotron power, the energy of an electron emitting at frequency $\\nu$ is given by $E\\simeq 15\\sqrt{\\nu_{GHz}/B_{\\mu G}}$ GeV (where $\\nu_{GHz}$ is the frequency in GHz and $B_{\\mu G}$ is the magnetic field in $\\mu G$). It follows that, assuming a magnetic field of few $\\mu G$ (as typical for galaxies), emissions at radio frequencies are mostly generated by electrons with energy around 1-10 GeV. Therefore, assuming a DM mass $\\gtrsim 10$ GeV, a sizable DM-induced radio synchrotron emission is a general prediction of WIMP models (with, of course, spectrum and absolute flux depending on the specific model). ", "conclusions": "" }, "1112/1112.6328_arXiv.txt": { "abstract": "We discuss the possible impact of strange quark matter on the evolution of core-collapse supernovae with emphasis on low critical densities for the quark-hadron phase transition. For such cases the hot proto-neutron star can collapse to a more compact hybrid star configuration hundreds of milliseconds after core-bounce. The collapse triggers the formation of a second shock wave. The latter leads to a successful supernova explosion and leaves an imprint on the neutrino signal. These dynamical features are discussed with respect to their compatibility with recent neutron star mass measurements which indicate a stiff high density nuclear matter equation of state. ", "introduction": "A core-collapse supernova (SN) explosion marks the disruption of a massive star by an energetic shock wave followed by the formation of a neutron star or a black hole. In the first hundreds of milliseconds (ms) of a supernova, temperatures in the range of tens of MeV, densities beyond nuclear matter saturation density $n_0 \\sim 0.145\\:$fm$^{-3}$ and proton fractions $Y_p \\leq 0.3$ are reached. With such properties, supernovae (SNe) are astrophysical laboratories for dense nuclear matter which, in the phase diagram of strongly interacting matter, possess overlap regions with heavy-ion experiments, such as in the future FAIR facility at GSI, Darmstadt (Germany) and the NICA facility at the JINR in Dubna (Russia). The modeling of core-collapse SNe represents a great computational challenge. It requires as input amongst others a nuclear matter equation of state (EoS), which provides information about the thermodynamic properties and compositions of matter for a large range of baryon number densities $n_b$, temperatures $T$, and isospin states, characterized by $Y_p$. Being discussed to populate neutron star interiors, hyperons and quark matter can also be included in SN equations of state (EoSs). At present, both components are tested on their impacts in SN simulations. Hereby, the applied quark and hyperon EoSs should be compatible with observed neutron star properties, e.g. pulsar masses. As we will argue in the next section, the latter indicate a stiff nuclear matter EoS at high density. \\newline In the following we will give a short summary on recent pulsar mass measurements and their implications for quark matter in neutron star interiors. We will proceed with an overview of quark matter studies in core-collapse SNe. In the last section, we will focus on the impact of low density quark-hadron phase transitions on the gravitational collapse of light and intermediate mass progenitor stars. ", "conclusions": "The onset of strangeness in core-collapse supernovae (SNe) can be studied by implementing hyperons and strange quark matte in SN equations of state (EoSs). Hereby, the chosen parameter sets must fulfill restrictions from the recent finding of a two solar mass pulsar PSR J1614-2230, which is only compatible with a stiff high density nuclear matter EoS. Within currently applied quark EoS models, high critical densities were shown to shorten the time to black hole formation in the gravitational collapse of massive progenitor stars. For the onset of quark matter with a soft EoS at critical densities $n_{crit}$ around nuclear matter saturation density $n_0\\sim 0.145$fm$^{-3}$, the proto neutron star collapses to a more compact hybrid star configuration within hundreds of ms after core-bounce. This launches a shock wave which leads to the SN explosion and releases a neutrino burst, dominated by $\\bar{\\nu}_e$. The properties of the neutrino burst are dependent on $n_{crit}$ as well as the model of the progenitor star. A first study with a stiff quark EoS shows no impacts on the SN dynamics as the critical densities are too high to be reached in the early post-bounce phase. However, studies are on the way, in which we apply a stiff quark EoS parameter set where SN matter enters the quark-hadron mixed phase at $n_{crit} \\sim n_0$.\\\\ \\newline \\textit{Acknowledgements}\\\\ The project was funded by the Swiss National Science Foundation (SNF) under project numbers PP00P2 - 124879/1, 200020 - 122287. T.F is support by HIC for FAIR and by the SNF under project~no.~PBBSP2-133378. The work of G.P. is supported by the DFG under Grant No. PA 1780/2-1 and J.S.-B. is supported by the DFG through the Heidelberg Graduate School of Fundamental Physics. M. H. is supported by the SNF under project number no. 200020-132816/1. M. H. is also grateful for participating in the EuroGENESIS collaborative research program of the European Science Foundation (ESF) and the ENSAR/THEXO project. I.S. is supported by the AvH foundation via a Feodor Lynen fellowship and wishes to acknowledge the HPCC of MSU and the iCER. The authors are additionally supported by CompStar, a research networking program of the ESF, and the scopes project funded by the SNF grant. no. IB7320-110996/1." }, "1112/1112.0574_arXiv.txt": { "abstract": "We report \\emph{Warm Spitzer} full-orbit phase observations of WASP-12b at 3.6 and 4.5 $\\mu$m. This extremely inflated hot Jupiter is thought to be overflowing its Roche lobe, undergoing mass loss, accretion onto its host star, and has been claimed to have a C/O ratio in excess of unity. We are able to measure the transit depths, eclipse depths, thermal and ellipsoidal phase variations at both wavelengths. The large amplitude phase variations, combined with the planet's previously-measured day-side spectral energy distribution, is indicative of non-zero Bond albedo and very poor day--night heat redistribution. The transit depths in the mid-infrared ---$(R_{p}/R_{*})^{2} = 0.0123(3)$ and $0.0111(3)$ at 3.6 and 4.5 $\\mu$m, respectively--- indicate that the atmospheric opacity is greater at 3.6 than at 4.5~$\\mu$m, in disagreement with model predictions, irrespective of C/O ratio. The secondary eclipse depths are consistent with previous studies: $F_{\\rm day}/F_{*} = 0.0038(4)$ and 0.0039(3) at 3.6 and 4.5 $\\mu$m, respectively. We do not detect ellipsoidal variations at 3.6 $\\mu$m, but our parameter uncertainties ---estimated via prayer-bead Monte Carlo--- keep this non-detection consistent with model predictions. At 4.5 $\\mu$m, on the other hand, we detect ellipsoidal variations that are much stronger than predicted. If interpreted as a geometric effect due to the planet's elongated shape, these variations imply a 3:2 ratio for the planet's longest:shortest axes and a relatively bright day--night terminator. If we instead presume that the 4.5 $\\mu$m ellipsoidal variations are due to uncorrected systematic noise and we fix the amplitude of the variations to zero, the best fit 4.5 $\\mu$m transit depth becomes commensurate with the 3.6 $\\mu$m depth, within the uncertainties. The relative transit depths are then consistent with a Solar composition and short scale height at the terminator. Assuming zero ellipsoidal variations also yields a much deeper 4.5 $\\mu$m eclipse depth, consistent with a Solar composition and modest temperature inversion. We suggest future observations that could distinguish between these two scenarios. ", "introduction": "Thermal phase variations are a powerful way to constrain the climate on exoplanets. Such observations have been made for non-transiting short-period planets \\citep{Cowan_2007, Crossfield_2010}, but are most potent when combined with transit and eclipse observations for edge-on systems, because of the additional knowledge of the planet's inclination, mass and radius \\citep{Knutson_2007a, Knutson_2009a, Knutson_2009b}. Secondary eclipse depths provide a constraint on the planet's day-side temperature. Thermal phase variations probe the day--night temperature contrast and hence the planet's heat redistribution efficiency. If the observational cadence and signal-to-noise ratio are sufficiently high, phase variations are also sensitive to the offset between the noon meridian and the planet's hottest local stellar time, hence constraining wind speed and direction. By considering eclipse depths at a variety of wavelengths for a sample of 24 transiting planets, \\cite{Cowan_2011b} estimated their day-side effective temperatures, hence placing a joint constraint on the Bond albedo and heat recirculation efficiency of these planets. That study found that typical hot Jupiters exhibit a variety of albedo/recirculation efficiencies, but planets with substellar equilibrium temperatures greater than $T_{0}\\approx2700$~K all seem to have lower albedo and/or recirculation efficiency. In other words, the hottest transiting giant planets have a qualitatively different climate than the merely hot Jupiters, but it is not known whether this is due to a difference in albedo, circulation, or both. Direct measurements of hot Jupiter geometric albedos from optical secondary eclipse observations span more than an order of magnitude and do not resolve this degeneracy. In this paper, we break the albedo-recirculation degeneracy for WASP-12b \\citep{Hebb_2009}, one of the very hottest known exoplanets, with a day-side temperature of $\\sim3000$~K: the amplitude of thermal phase variations is a direct measure of the planet's day-night temperature contrast and hence heat transport efficiency. If the night-side temperature is high, then the planet's albedo must be exceedingly low to be consistent with its high day-side temperature. If, on the other hand, the night-side temperature is low, then the planet has an albedo in the tens of percent. WASP-12b has been a fascinating planet since its discovery. The discrepant timing of its secondary eclipse indicated that the planet had a slight eccentricity \\citep{Lopez-Morales_2010}, but subsequent eclipse \\citep{Campo_2011, Croll_2011} and radial velocity \\citep{Husnoo_2011} observations have all but ruled this out. Nevertheless, the planet's short-period orbit (1.1 days; just outside its star's Roche limit) and inflated radius (1.8 $R_{J}$) led to the prediction that it is tidally distorted \\citep{Ragozzine_2009, Leconte_2011, Budaj_2011}, and undergoing Roche-lobe overflow followed by accretion onto its host star \\citep{Li_2010, Lai_2010}. The putative early ingress of an ultraviolet transit observed by \\cite{Fossati_2010} seems to support this prediction, but may also be explained in terms of a leading bow shock from material streaming off the planet \\citep{Vidotto_2010, Llama_2011}. More recently, \\cite{Madhusudhan_2011} used the wavelength-dependance of mid-infrared eclipse depths of \\cite{Croll_2011} and \\cite{Campo_2011} to constrain the atmospheric composition of WASP-12b, and found it has a carbon to oxygen ratio (C/O) greater than unity, unlike Solar System planets, or the assumed composition of extrasolar planets. Those findings rested heavily on the relative eclipse depths at 3.6 and 4.5 $\\mu$m. Our observations of eclipses and transits at these wavebands should be able to reinforce or rule out the high C/O scenario. ", "conclusions": "We obtained \\emph{Warm Spitzer} full-orbit phase observations of WASP-12b at 3.6 and 4.5 $\\mu$m, allowing us to measure the transit depths, eclipse depths, thermal and ellipsoidal phase variations at both wavelengths. We are able to push \\emph{Warm Spitzer} photometry to within 10--20\\% of the Poisson limit, but there are two important caveats: A) Removing intra-pixel sensitivity variations (IPSVs) from the data is inherently a model-dependent endeavor. This means that we must specify not only an IPSV model, but also an astrophysical model before getting close to the quoted precision. The simultaneous fit to astrophysical and systematic effects makes it difficult to produce a ``clean'' lightcurve independent of astrophysical assumptions. For example, we obtain very different thermal phase variation parameters depending on how we correct for systematics, and it is difficult to distinguish between these scenarios based solely on goodness-of-fit. Instead, we must resort to a number indirect clues as to which IPSV-removal scheme is more trustworthy. B) There is still red noise in our residuals, no matter how we remove IPSVs. This remaining red noise is the dominant source of uncertainty for all of our astrophysical parameters. We find that WASP-12b exhibits large-amplitude thermal phases ---indicative of poor day--night heat transport and a moderate Bond albedo--- but also an unexpectedly large phase offset at 3.6 $\\mu$m. We do not detect ellipsoidal variations at 3.6 $\\mu$m, while we detect an unexpectedly strong signal at 4.5 $\\mu$m. This leads us to two possible hypotheses: 1) If we take the 4.5~$\\mu$m ellipsoidal variations at face value, we find: deeper transits at 3.6 $\\mu$m as compared to 4.5 $\\mu$m, inconsistent with either Solar or enhanced CO models; eclipse depths consistent with previous studies. If the 4.5 $\\mu$m ellipsoidal variations are astrophysical in nature, it indicates that the planet is far more distorted than predicted, and exhibits a bright terminator. In this scenario, the 3.6 $\\mu$m ellipsoidal variations are attenuated due to detector systematics, possibly throwing off the 3.6 $\\mu$m transit depth as well. 2) If instead we presume that the 4.5 $\\mu$m ellipsoidal variations are caused by detector systematics and set them to zero ---the null hypothesis--- we find: transit depths consistent with a Solar composition and short atmospheric scale height at the planet's terminator; eclipse depths consistent with a Solar composition and a modest temperature inversion; ellipsoidal variations in line with predictions. The null hypothesis is attractive in its simplicity, but requires that we were very unlucky; follow-up \\emph{Warm Spitzer} observations would have different systematics (the PSF would fall on different regions of the pixels) and could settle the question of ellipsoidal variations. It is likely that near infrared transit spectroscopy could break the composition degeneracy, or at least determine the atmospheric structure of WASP-12b; if the planet has a short scale-height at the terminator it will lend credence to the null hypothesis. Further optical transit photometry will be useful in pinning down the transmission spectrum and refining geometrical parameters; if $a/R_{*}<3$, then the planet could very well be more distorted than predicted, making the large ellipsoidal variations more plausible. Optical eclipse measurements from the ground or from space might confirm the moderate albedo of the planet. The planet is hypothesized to be losing mass to its host star. If this is indeed the case, the presence of an accretion disk, accretion stream and impact hot spot may necessitate a more holistic model to properly interpret observations." }, "1112/1112.6214_arXiv.txt": { "abstract": "The number of methods used to study the properties of galaxies is increased, and testing these methods is very important. Galactic globular clusters (GCs) provide an excellent medium for such test, because they can be considered as simple stellar populations. We present ages and metallicities for 40 Galactic GCs as determined from three publicly available techniques, including colour, Lick-index and spectrum-fitting methods, based on Bruzual \\& Charlot evolutionary population synthesis (EPS) models. By comparing with the ages obtained from colour-magnitude diagrams (CMDs) and metallicities obtained from spectra of stars, we are able to estimate the ability of these methods on determination of GCs$^{'}$ parameters, which is absolutely necessary. As a result, we find that: (i) for the metallicity, our derived metallicities agree with those derived from the spectra of stars, Lick-index method is suitable to study metallicity for the stellar population systems in the range of $-1.5\\la$[Fe/H]$\\la-0.7$ and spectrum-fitting method is suitable to study metallicity for the stellar population systems in the range of $-2.3\\la$[Fe/H]$\\la-1.5$; (ii) for the age, these three methods have difficulties in age determination, our derived ages are smaller (about 2.0\\,Gyr, on average) than the results of CMDs for all these three methods. We use Vazdekis and Maraston models to analyze whether our results are dependent on EPS models, and find that the tendency of these two models is the same as that of Bruzual \\& Charlot models. Our results are independent of the EPS models. In addition, our test is based on the old GCs and our conclusions may hold for old stellar population systems. Besides the age-metallicity degeneracy, we examine the possible effects of other factors (horizontal branch morphology, blue straggler stars, binary interactions and $\\alpha-$enhancement) and give quantitative analysis of the influences of these factors on age determinations (except for $\\alpha-$enhancement). For colour and spectrum-fitting methods, the age can be underestimated about $0.0-3.0$\\,Gyr, $0.0-2.0$\\,Gyr, and $0.0-3.0$\\,Gyr due to influences of horizontal branch, blue straggler and binary stars, respectively. And for Lick-index method, the lower limit of maximal change of age is 6.0\\,Gyr, 5.0\\,Gyr and 3.0\\,Gyr due to influences of horizontal branch, blue straggler and binary stars, respectively. ", "introduction": "\\end{flushleft} \\end{figure} \\begin{figure} \\begin{flushleft} \\includegraphics[bb=80 35 470 785,height=8.5cm,width=4cm,clip,angle=270]{zhy2.ps} \\caption{H$\\beta$ versus three indices (Mg\\,$b$, $<$Fe$>$ and [MgFe]$^{'}$) for the models of Thomas, Maraston \\& Johansson (2011a), the ages are from $1$ to $15$ Gyr and [Fe/H] are from $-2.3$ to $0.67$ that are labeled in the figure. The gray solid and red dashed lines stand for [$\\alpha$/Fe]$=0.0$ and 0.5, respectively.}\\label{entropy-rlof} \\end{flushleft} \\end{figure} Fig.\\,1 shows an example of the method of using contour map of $\\chi^2$ to obtain age and metallicity for NGC\\,6235. The star stands for the $\\chi^2_{min}$ and corresponds to the obtaining $\\tau$ and [Fe/H]. The shaded area maps 1$\\sigma$ confidence region, and the $\\Delta$$\\tau$ and $\\Delta$[Fe/H] are obtained based on this shaded area. \\subsection{Lick-index method} Lick/IDS indices have been used to break the age-metallicity degeneracy (Trager et al. 2000; Terlevich \\& Forbes 2002; Gallazzi et al. 2005) of SP systems. In this paper, we select the Balmer (H$\\beta$, H$\\delta_{A}$, H$\\delta_{F}$, H$\\gamma_{A}$ and H$\\gamma_{F}$), Mg\\,\\emph{b}, Fe5270 and Fe5335 indices to study the parameters of GCs (Balmer indices are age-sensitive indices, Mg\\,\\emph{b}, Fe5270 and Fe5335 are metal-sensitive indices). The H$\\delta$ and H$\\gamma$ are high order Balmer indices which measured with a narrower ($\\sim$\\,20\\,${\\rm\\AA}$; marked by subscript F) and a wider ($\\sim$\\,40\\,${\\rm\\AA}$; marked by subscript A) central bandpass \\citep{wort97}. Among these five Balmer indices, H$\\beta$ and H$\\gamma_{F}$ are little sensitive to $\\alpha/$Fe and the other three indices are sensitive to $\\alpha/$Fe. And in Section 5.4, we will discuss the influence of $\\alpha$-enhancement on age determination for Lick-index method. For the three metal-sensitive indices, we adopt the following definition: [MgFe]$^{'}$\\,=\\,[Mg\\,\\emph{b}(0.72\\,$\\times\\,$Fe5270\\,+\\,0.28\\,$\\times$\\,Fe5335)]$^{1/2}$, which defined by Thomas, Maraston \\& Bender (2003). The [MgFe]$^{'}$ has been found to be good tracer of metallicity. Adopting the models of Thomas, Maraston \\& Johansson (2011a) we present this character in Fig.\\,2, the grids of H$\\beta$ versus Mg\\,$b$, $<$Fe$>$ ($<$Fe$>=$1/2(Fe5237$+$Fe5335)) and [MgFe]$^{'}$ with two types of $\\alpha$/Fe ([$\\alpha$/Fe]$=0.0$ and $0.5$), the gray solid and red dashed lines stand for [$\\alpha$/Fe]$=0.0$ and $0.5$, respectively. From them we can see that the influence of $\\alpha$/Fe ratio is obvious for H$\\beta$ versus Mg\\,$b$ and $<$Fe$>$ grids, and is small for H$\\beta$ versus [MgFe]$^{'}$ grid. The Mg\\,$b$ increases and $<$Fe$>$ decreases with increasing $\\alpha$/Fe, and the combination index [MgFe]$'$ counteract the effect of $\\alpha$/Fe ratio on Mg$b$, Fe5237 and Fe5335. This certify that the [MgFe]$'$ is independent of $\\alpha$/Fe and can be considered as good tracer of metallicity \\citep{thom03}. \\begin{figure*} \\includegraphics[bb=20 20 600 780,height=16cm,width=9.5cm,clip,angle=270]{zhy3.ps} \\caption{The distribution of GCs on the Balmer indices and [MgFe]$^{'}$ planes. Model grides from BC03 are shown: age is constant for solid line (top to bottom, 1, 2, 4, 5, 8, 10, 12 and 15 Gyr), and metallicity is constant for dashed line (left to right: [Fe/H]$=-2.3, -1.7, -0.7, -0.4$, $0.0$ and $+0.4$). Line strengths of Galactic GCs are plotted on SSP grides. The red open squares, blue open circles and purple open pentacles stand for GCs with red, blue and unknown HBR, respectively. In the top left plane (H$\\beta$ versus [MgFe]$^{'}$ plane), we give these values of age and metallicity.}\\label{entropy-rlof} \\end{figure*} In Fig.\\,3, we show the distribution of GCs on five Balmer (H$\\beta$, H$\\delta_{A}$, H$\\delta_{F}$, H$\\gamma_{A}$ and H$\\gamma_{F}$) indices versus [MgFe]$^{'}$ planes. We can see that the majority of GCs lie beyond the grid covered by H$\\beta$ (this problem does not involve all Balmer indices, Thomas et al. 2011a), which would be inconsistent with the typical universe age 13.6\\,Gyr. This is a general problem (see Fig.\\,2 of Mendel et al. 2007) for all studies of using H$\\beta$ index to study the GCs based on existing models. This problem is also clearly demonstrated by the $'$zero-point$'$ offset (some GCs lie beyond the grid of H$\\beta$ and most GC ages are greater than 14\\,Gyr). Although these problems have been pointed by some groups (Vazdekis et al. 2001; Schiavon et al. 2002; Cenarro et al. 2008), but few works have been done on studying the specific sources of this problem or quantified their effects on the derived ages and metallicities. Similar to the technique described in Puzia et al. (2005), the ages and metallicities for individual GCs are computed as the weighted mean of the parameters derived from five Balmer indices versus [MgFe]$^{'}$ grids. Due to the influence of hot blue HB stars, the isochrones for old ($\\tau$\\,$>$\\,8.0\\,Gyr) stellar populations with metallicity below [Fe/H]$\\,=-0.7$ tend to overlap (Maraston \\& Thomas 2000; Mendel et al. 2007), this nature is reflected by H$\\beta$ in panel (a) of Fig.\\,3. This brings an ambiguity in determining ages and metallicities of GCs and unnaturally extends the age distributions for old ages for Lick-index method. Note that, for the GCs lying beyond the Lick indices grids, we assume the values of GCs are 15.0\\,Gyr. \\subsection{spectrum-fitting method} With the increase of spectroscopic data, people use spectra to study the properties of SP systems. So far there are many kinds of full spectrum-fitting techniques used in such study (e.g. Cid Fernandes et al. 2005; Mathis, Charlot \\& Brinchmann 2006; Tojeiro et al. 2007; K08). We employ a modified version of the \\tiny{STARLIGHT} \\normalsize{code} by Cid Fernandes et al. (2005) to analyze the observed spectra of Galactic GCs. The \\tiny{STARLIGHT} \\normalsize{}code is originally used to study the properties of galaxy, and is achieved by fitting the observed spectrum \\emph{F$\\rm_{O}$} with a model spectrum \\emph{F$\\rm_{M}$} that mixed by \\emph{N$_{\\star}$} SSPs with different ages and metallicities from the BC03 models. The code is carried out with a simulated annealing plus Metropolis scheme (Cid Fernandes et al. 2001), which searches for the minimum \\begin{equation} \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\chi^2= \\sum\\limits_{\\lambda}\\, [(F\\rm_{O}-\\emph{F}\\rm_{M})\\omega_{\\lambda}]^2, \\end{equation} where $\\omega$$^{-1}_{\\lambda}$ is the error in \\emph{F$\\rm_{O}$}. The line-of-sight stellar motions are modeled by a Gaussian distribution centered at velocity \\emph{v}$_{\\star}$ and with dispersion \\emph{$\\sigma$}$_{\\star}$. The output includes the distributions of stellar age, metallicity, extinction, velocity dispersion and stellar mass. ", "conclusions": "We use colour, Lick-index and spectrum-fitting methods based on BC03 model to obtain the ages and metallicities for Galactic GCs, and compare them with those determined from spatially resolved observations (such as CMDs, spectra of stars). We also compare the ages and metallicities derived from other two models (Vazdekis and Maraston models) to test whether our results are dependent on EPS models. Before displaying the results of this work, we should introduce some pretreatments of model, sample data and methods. The details are given as follows. \\subsection{Pretreatments} \\begin{itemize} \\item For the EPS model, as said above, the BC03 model provides six metallicities, we linearly interpolate colours, Lick indices and spectra of SSPs that span [Fe/H] from $-$2.3 to $+$0.4 in increments of 0.1 dex. \\item In order to discuss the influence of HB morphology on age determinations in Section 5.1, we divide the GCs into three groups according to the value of HB ratio (HBR \\footnote{ HBR\\,$=(B-R)/(B+R+V)$, B is the number of blue HB stars, R is the number of red HB stars and V is the number of RR Lyrae variable.}) from the Harris$'$ catalogue. One is the blue HBR type (HBR\\,$>$0), the second is red HBR type ( HBR\\,$<$0) and the third is unknown HBR (HBR is not given). \\item For the colour method, we do extinction corrections for observed magnitudes of GCs by adopting the extinction curves of \\citet{schl98} and the $E(B-V)$ of Harris$'$ catalogue (the 8th column of Table A1). \\item For the Lick-index method, in this work we calculate the Lick indices from the spectra of Galactic GCs by degrading to wave-dependent resolution of the Lick system (see Section 4.4 of BC03), without adopting the results by using fitting functions. \\begin{figure*} \\includegraphics[bb=20 20 580 730,height=16.0cm,width=9.5cm,clip,angle=270]{zhy4.ps} \\caption{Top panels: the comparisons between the metallicities obtained by three methods and the spectra of stars for Galactic GCs. Symbols have the same meanings in Fig.3. Bottom panels are the comparisons between the ages obtained by three methods and the CMDs. Left, middle and right panels are comparison results for methods of colour, Lick-index and spectrum-fitting, respectively. }\\label{entropy-rlof} \\end{figure*} \\item For the spectrum-fitting method, we adopt a modified version of the \\tiny{STARLIGHT}\\normalsize{} code. Because it is generally accepted that the GC can be represented by a SSP, only one component in the base. We fit the observed GC spectrum \\emph{F$\\rm_{O}$} with each of the SSP spectrum \\emph{F$_{i}$}, do not fit the observed spectrum with a linear combination of \\emph{N$_{\\star}$} SSPs. In this study, following \\citet{cid05}, we find the best fitting SSP, which matches a given observed spectrum of GC, through a standard $\\chi^2$ minimization procedure: \\begin{equation} \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\chi^2_{i}= \\sum\\limits_{\\lambda}\\, [(F\\rm_{O}-\\emph{F$_{i}$})\\omega_{\\lambda}]^2, \\end{equation} instead of equation (2), where \\emph{F$_{i}$} corresponds to the spectrum of i-th SSP model. And in this work we construct a base of $18$\\,ages\\,$\\times\\,24$\\,metallicities\\,$=432$\\,SSP model spectra. The use of \\tiny{STARLIGHT}\\normalsize{} to study the integrated spectra of clusters has also been extensively discussed by Cid Fernandes \\& Gonz$\\acute{a}$lez Delgado (2010). In the spectrum-fitting process, we use the spectra in the 3700$-$5700\\,${\\rm\\AA}$ range because the red and blue parts of the observations have poor quality. Errors on the age and metallicity estimation are also determined by the $\\chi^2$ contours. \\end{itemize} \\subsection{Comparison of parameters with other studies } Using the methods and pretreatments outlined in Sections\\,3 and 4.1, we obtain ages, metallicities and corresponding errors for GCs. In order to distinguish the ages and metallicities derived from different methods, we define some parameters in this paper. $\\tau\\rm_{C}$ and [Fe/H]$\\rm_{C}$ stand for the age and metallicity determined from colour method, $\\tau\\rm_{I}$ and [Fe/H]$\\rm_{I}$ represent those obtained from Lick-index method, and $\\tau\\rm_{S}$ and [Fe/H]$\\rm_{S}$ stand for those determined from spectrum-fitting method. The parameters of literatures are represented by $\\tau\\rm_{L}$ and [Fe/H]$\\rm_{L}$. In Table A2 we list the GC IDs in the first column. In the 2nd to 7th columns, $\\tau\\rm_{C}$, [Fe/H]$\\rm_{C}$, $\\tau\\rm_{I}$, [Fe/H]$\\rm_{I}$, $\\tau\\rm_{S}$ and [Fe/H]$\\rm_{S}$ are given with errors, respectively. And in 8th and 9th columns we list the $\\tau\\rm_{L}$ and [Fe/H]$\\rm_{L}$. In the top panels of Fig.\\,4, we give the comparison between our GC metallicities and those given in the literatures that obtain metallicities from the spectra of stars. And (a), (b) and (c) panels are the comparison results for colour, Lick-index and spectrum-fitting methods, respectively. The vertical bars are the errors in metallicities derived from three methods. From them we see that the metallicities obtained by three methods are in agreement with literature value for the whole sample. Meanwhile we also find that all GCs with blue HBR have lower metallicity than those with red HBR for these three methods, which agrees with the study of Lee, Yoon \\& Lee (2000). Our specific analysis are given as follows.\\\\ (a) For colour method, [Fe/H]$\\rm_{C}$ agrees with [Fe/H]$\\rm_{L}$ for the whole sample, but there exists discreteness. We can see that [Fe/H]$\\rm_{C}$ is a bit smaller than [Fe/H]$\\rm_{L}$ in the range of [Fe/H]$\\rm_{L}$$\\la-$1.0 and [Fe/H]$\\rm_{C}$ agrees with [Fe/H]$\\rm_{L}$ for the range of [Fe/H]$\\rm_{L}$$>-$1.0. But for the value of [Fe/H]$\\rm_{L}$ is about $-$1.1, there exist three GCs (NGC 6171, 6342 and 6652) having lager [Fe/H]$\\rm_{C}$ than [Fe/H]$\\rm_{L}$. From the [Fe/H]$\\rm_{C}$ and $\\tau\\rm_{C}$ of these GCs listed in Table A2, we see that $\\tau\\rm_{C}$ is smaller than $\\tau\\rm_{L}$, so these GCs can be affected by age-metallicity degeneracy.\\\\ (b) For the Lick-index method, the [Fe/H]$\\rm_{I}$ is perfectly in agreement with [Fe/H]$\\rm_{L}$ in the range of $-1.5\\la$\\,[Fe/H]\\,$\\la-0.7$. But there have difficulties in metallicity determination for the range of [Fe/H]\\,$\\la-1.5$, because it tends to overlap for Balmer and [MgFe]$^{'}$ indices at low-metallicity range for BC03 model. And we can see this phenomenon clearly in the panel (a) of Fig.\\,3. This indicates that Lick-index method is suitable to study metallicity for SP systems in the range of $-1.5\\la$\\,[Fe/H]\\,$\\la-0.7$ and has some difficulties in studying SP systems for the range of [Fe/H]\\,$\\la-1.5$.\\\\ (c) For the spectrum-fitting method, we find that [Fe/H]$\\rm_{S}$ matches [Fe/H]$\\rm_{L}$ in the range of $-2.3\\la$\\,[Fe/H]\\,$\\la-1.5$, and [Fe/H]$\\rm_{S}$ is smaller than [Fe/H]$\\rm_{L}$ for the range of [Fe/H]\\,$\\ga-1.5$. All these results imply that the spectrum-fitting method may be suitable to study metallicity for SP systems in the range of $-2.3\\la$\\,[Fe/H]\\,$\\la-1.5$ and [Fe/H]$\\rm_{S}$ may be smaller than [Fe/H]$\\rm_{L}$ for the range of [Fe/H]\\,$\\ga-1.5$. On the whole, our metallicities obtained from three methods match those determined from spectra of stars in the entire metallicity range spanned by the GCs. The Lick-index method is suitable to study the metallicity in the range of $-1.5\\la$\\,[Fe/H]\\,$\\la-0.7$ and spectrum-fitting method is suitable to study the metallicity in the range of $-2.3\\la$\\,[Fe/H]\\,$\\la-1.5$. In the bottom panels of Fig.\\,4, we compare our derived GC ages of three methods with the values of literature that determine ages from CMDs. And (d), (e) and (f) panels are corresponding to colour, Lick-index and spectrum-fitting methods, respectively. From them we can see there exists large discrepancy in age determinations, and the errors are also large for these three methods. Our specific analysis are given as follows.\\\\ (d) For the colour method, most of the GCs with blue and unknown HBR have relative lower $\\tau\\rm_{C}$ than $\\tau\\rm_{L}$, but there has no obvious tendency for those with red HBR. The age of GCs with blue HBR can be influenced by HB stars for the existence of HB stars making GCs look younger. Five GCs are found to possess extreme low $\\tau\\rm_{C}$, including two with unknown HBR (NGC 6388 and 6441) and three with red HBR (NGC 6171, 6342, and 6652). Among these GCs, three have relative lager [Fe/H]$\\rm_{C}$ than [Fe/H]$\\rm_{L}$, which can be affected by the age-metallicity degeneracy (see Fig.\\,4a). Meanwhile from the 4th and 5th columns of Table A1, we know that the other two GCs are close to the Galactic bulge, which are badly contaminated by the field stars or affected by differential reddening, so we can not get reliable parameters for them.\\\\ (e) For the Lick-index method, we can see there exist some GCs with extreme large $\\tau\\rm_{I}$ (about 15.0\\,Gyr), and except these GCs most of GCs have smaller $\\tau\\rm_{I}$ which are influenced by HB stars, especially for those with blue HBR. About eight GCs have large $\\tau\\rm_{I}$ of 15.0\\,Gyr which lie beyond the grids, including two with red HBR (NGC 1851 and 6637), one with unknown HBR (NGC 5946) and five with blue HBR (NGC 1904, 2298, 3201, 6254 and 7078). Similar to above, Lick-index method has some difficulties in parameter determinations for the range of [Fe/H]\\,$\\la-1.5$, and most of these eight GCs are metal-poor ([Fe/H]$\\rm_{I}$$\\la-$1.5), so we obtain extreme large $\\tau\\rm_{I}$ for them.\\\\ (f) For the spectrum-fitting method, most GCs with blue HBR have relative smaller $\\tau\\rm_{S}$ than $\\tau\\rm_{L}$, but there has no obvious tendency for those with red and unknown HBR. From panel (c), we know that [Fe/H]$\\rm_{S}$ is consistent with [Fe/H]$\\rm_{L}$ in the range of [Fe/H]\\,$\\la-1.5$ and most of GCs with blue HBR are metal-poor, so the $\\tau\\rm_{S}$ of GCs with blue HBR are affected by HB stars rather than age-metallicity degeneracy. For the GCs with red and unknown HBR, one part has large $\\tau\\rm_{S}$ which can be influenced by age-metallicity degeneracy, and the other part has small $\\tau\\rm_{S}$ which may be influenced by HB stars and BSs. \\begin{table*} \\begin{center} \\caption[]{Comparison of main model ingredients in each EPS model. The first, second and third rows list the stellar evolution track, stellar spectral library and IMF of the EPS models. The 4th and 5th rows show the metallicity ($Z$) and age range that these models cover, and the number of metallicities is also given in the parenthesis of the 4th row. } \\tabcolsep=0.15in \\label{Tab:frac} \\renewcommand{\\arraystretch}{1.5} \\begin{tabular}{|l|l|l|l|r|r|r|r|r|r|c|c|c|c|c|c|c|c|c|c|} \\hline\\hline\\noalign{\\smallskip} Models &BC03 &Vazdekis &Maraston \\\\ \\hline\\noalign{\\smallskip} Stellar evolution track &Padova 1994 &Padova 2000 &Cassisi+Geneva \\\\ Stellar spectral library &STELIB &MILES &MILES\\\\ IMF &Chabrier &Salpeter &Chabrier\\\\ $Z$(number) &$0.0001-0.0500(6)$&$0.0001-0.0300(7)$&$0.0001-0.0400(5)$\\\\ Age(Gyr) &$0.0001-20.0000$ &$0.0630-17.1800$ &$0.0060-15.0000$\\\\ \\hline \\end{tabular} \\end{center} \\end{table*} Note that, relevant to this test is that CMD-derived ages also can carry their own problems for this test, because the CMD-derived ages depend on the adopted tracks and on whether element ratios be taken into account, which has also been discussed in \\citet{mara11}. On the whole, all of these three methods have difficulties in age determination and we can not directly say which method is better on age determination. Except the influences of CMD-derived ages and age-metallicity degeneracy in the entire age range, there exist some uncertainties in age determinations, some factors (e.g. HB stars, BSs, binary stars and $\\alpha$-enhancement) can affect the age determinations and we will discuss them in the next Section. \\begin{figure*} \\includegraphics[bb=20 15 580 770,height=16.0cm,width=9.5cm,clip,angle=270]{zhy5.ps} \\caption{Comparisons between the results from three EPS models and the literatures. Top and bottom panels are the comparisons for metallicity and age, respectively. The open squares, pentacles and circles stand for the results of BC03, Vazdekis and Maraston EPS models, respectively. Left, middle and right panels are comparison results for methods of colour, Lick-index and spectrum-fitting, respectively. The dispersion (rms) of each model is also listed in the lower right corner of each panel. }\\label{entropy-rlof} \\end{figure*} \\subsection{Consistency checks} We use Vazdekis\\footnote{http://miles.iac.es/} \\citep[the updated v9.1 version,][]{vazd10,falc11} and Maraston \\citep{mara11} models to investigate whether our results are dependent on EPS models. The Vazdekis models use the Padova 2000 theoretical isochrones \\citep{gira00}, the MILES library \\citep{sanc06} and four types of IMF with stellar mass limits (Vazdekis et al. 2003) of 0.1 and 100 M$_{\\odot}$. This version covers the wavelength range of 3540$-$7410\\,${\\rm\\AA}$ with a spectral resolution (FWHM) of $\\sim$2.5\\,${\\rm\\AA}$ \\citep{falc11}, seven metallicities $0.0001\\leqslant Z \\leqslant0.03$ and 50 ages across the range of 0.0630$-$17.1800\\,Gyr. In this work, we select the unimodal Salpeter IMF \\citep{salp55} to study the parameters of GCs for spectrum-fitting and Lick-index methods. For the colour method, because the MILES library has relatively limited wavelength coverage and does not extend into I-band, we select the colours from previous models of Vazdekis \\citep{vazd99} to study the parameters of GCs. This does not influence our results (private discussion). The Maraston models assume a stellar evolution prescription, which consists of the isochrones and stellar tracks by Cassisi, Castellani \\& Castellani (1997) for ages larger than $\\sim$30.0Myr and by Ganeva \\citep{scha92} for younger populations. Sets of models have been also computed with Padova stellar evolutionary models \\citep{gira00}. Four different libraries of flux-calibrated empirical stellar spectra and three types of IMF have been considered. We select the MILES library and the \\citet{chab03} IMF to study the parameters of GCs. These updated models cover the wavelength range of 3500$-$7430\\,${\\rm\\AA}$ with a spectral resolution (FWHM) of $\\sim$2.54\\,${\\rm\\AA}$ \\citep{beif11} and five metallicities $0.0001\\leqslant Z \\leqslant0.04$ with different ages across the range of 0.0060$-$15.0000\\,Gyr (Maraston \\& Str\\\"{o}mb\\\"{a}ck 2011, and references therein). Just as said above, we use the updated models of \\citet{mara11} based on MILES library for spectrum-fitting and Lick-index methods. And for the colour method we select the models of \\citet{mara05}, which based on the $BaSeL$ library, to study the parameters of GCs. In Table 1 we list the comparison of main model ingredients in each EPS model. The first, second and third rows list the stellar evolution track, stellar spectral library and IMF of the EPS models that we used in this study. The 4th and 5th rows show the metallicity ($Z$) and age range that these models cover, and the number of metallicities is also given in the parenthesis of the 4th row. We use these two EPS models to obtain the parameters of GCs based on three methods described above and compare the results of these two models with that of the literatures. In Fig.\\,5, we show the comparisons between the parameters of GCs obtained from tree methods based on three EPS models (BC03, Vazdekis and Maraston) and the literatures, in which the open squares, pentacles and circles stand for the results of BC03, Vazdekis and Maraston EPS models, respectively. The dispersion (rms) is also listed in the lower right corner of each panel, rms$\\rm_{bc03}$, rms$\\rm_{vazd}$ and rms$\\rm_{mara}$ stand for the rms of BC03, Vazdekis and Maraston EPS models, respectively. We find that the metallicities obtained from three methods of these two EPS models (Vazdekis and Maraston) have an agreement with those of the literatures in the entire metallicity range. From the top panels (a, b and c panels), we find that the comparisons between metallicities obtained by three methods and the spectra of stars for these two models are similar to that of BC03 models. For Lick-index method in panel (b), we see that the rms$\\rm_{vazd}$ and rms$\\rm_{mara}$ are smaller than rms$\\rm_{bc03}$, this may be due to the adoption of MILES for these two models and STELIB for BC03 models \\citep{mara11}. For the spectrum-fitting method in panel (c), the [Fe/H]$\\rm_{S}$ of Vazdekis model is larger than that of the other two models, so the rms$\\rm_{vazd}$ is smaller than rms$\\rm_{bc03}$ and rms$\\rm_{mara}$. From these rms, we can know that the dispersion of colour method is larger than that of other two methods. All these can not influence our conclusions, Lick-index method is suitable to study metallicity in the range of $-1.5\\la$\\,[Fe/H]\\,$\\la-0.7$ and spectrum-fitting method is suitable to study metallicity in the range of $-2.3\\la$\\,[Fe/H]\\,$\\la-1.5$. From bottom panels, we can see that all these three methods have difficulties in age determinations for these two EPS models and the rms is larger than that of metallicity. There also exist some GCs with extreme large $\\tau\\rm_{I}$ (about 15.0\\,Gyr) for these two models, and most of GCs have relative smaller $\\tau\\rm_{C}$, $\\tau\\rm_{I}$ and $\\tau\\rm_{S}$ than $\\tau\\rm_{L}$, which is similar to that of BC03 models. The comparisons of derived metallicities and ages of individual GCs show that these three models make different parameter predictions. However, the whole tendency for these three methods of these three EPS models is nearly the same. This indicates that our conclusions are independent of the EPS models. \\begin{figure} \\includegraphics[bb=95 45 590 690,height=8.0cm,width=6.0cm,clip,angle=270]{zhy6.ps} \\caption{Taking NGC 7078 as an example to illustrate the method of selecting HB stars and BSs from the CMD. The HB stars are marked with red asterisks and BSs are marked with blue circles.}\\label{entropy-rlof} \\end{figure} \\begin{figure*} \\includegraphics[bb=35 15 555 780,height=16.cm,width=9.5cm,clip,angle=270]{zhy7.ps} \\caption{Theoretical Balmer index$-$\\emph{T}$\\rm_{eff}$ relations obtained from the B\\tiny{LUERED} \\normalsize{} library for different log \\emph{g} (solid, dashed, dotted, and dash-dot-dot-dot lines stand for log \\emph{g}$=3.0, 4.0, 4.5$ and $5.0$, respectively) at [Fe/H]$=-1.0$.}\\label{entropy-rlof} \\end{figure*} We have investigated the utility of three methods (colour, Lick-index and full spectrum-fitting) with the same EPS models (BC03 models) for estimating ages and metallicities of Galactic GCs. We also compared our results with those estimated in the literatures from other methods, and the main results of this study are as follows. (1) Our results show that Galactic GCs are almost old metal-poor SP systems. (2) Metallicities determined from these three methods are in agreement with those of literatures in the entire range spanned by GCs. The Lick-index method is suitable to obtain metallicity for SP systems in the range of $-1.5\\la$[Fe/H]$\\la-0.7$ and spectrum-fitting method is suitable to study metallicity for SP systems in the range of $-2.3\\la$[Fe/H]$\\la-1.5$. (3) There exist some discrepancies between our ages and the literatures, we can not directly say which method is more suitable for age determination. Through comparing with the literatures we find our results are younger than the literatures on average, especially for GCs with blue HBR. Many factors can affect the age determinations and make GCs become younger (such as the HB morphology, BSs, binary stars, $\\alpha-$enhancement). (4) We use Vazdekis and Maraston models to investigate whether our results are dependent on EPS models. The comparisons of derived metallicities and ages for individual GCs show that three models make different parameter predictions. However, the whole tendency for these three methods of these three EPS models is nearly the same. All these indicate that our above conclusions are independent of the EPS models. (5) In this work we use the old Galactic GCs to test these three methods, and our results may hold for old SP system. (6) We also study the quantitative influences of HB, BSs and binary stars on age determinations for three methods. The existence of all these stars can make the GCs look younger. For the colour and spectrum-fitting methods, the age can be underestimated about 0.0$-$3.0\\,Gyr, 0.0$-$2.0\\,Gyr and 0.0$-$3.0\\,Gyr due to influences of HB, BSs and binary stars, respectively. And for Lick-index method, the lower limit of maximal change of age is 6.0\\,Gyr, 5.0\\,Gyr and 3.0\\,Gyr due to influences of horizontal branch, blue straggler and binary stars, respectively. (7) For the Lick-index method, we also investigate the influence of $\\alpha-$enhancement on age determination, and find that the $\\alpha-$enhancement has little effect on age determination for Galactic GCs." }, "1112/1112.4211_arXiv.txt": { "abstract": "We are extending our search for faint PNe in the LMC to include the outer 56 deg$^{2}$ area not covered in the original UKST survey of the central 25deg$^{2}$ region. Candidate PNe have been selected using the Magellanic Cloud Emission Line Survey (MCELS) and the first round of observations has yielded 93 new LMC PNe while confirming the 102 previously known PNe in the outer LMC. We plan to continue our spectroscopic object identification program until we cover all our remaining candidates in the survey area. These observations, providing medium and high resolution spectra from $\\lambda$3650\\AA~to $\\lambda$6900\\AA~will additionally be used to measure fluxes for a series of research projects including luminosity functions, abundances and LMC kinematics. ", "introduction": "In 2006 Reid and Parker added 460 planetary nebulae (PNe) to the 169 previously known across the central 25deg$^{2}$ of the LMC. These candidates were assigned a probability rating which relied on spectroscopic and optical confirmation. With the assistance of increased high resolution spectroscopy and NIR imaging from Spitzer Space Telescope SAGE data (see \\cite[Meixner et al. 2006]{Meixner2006}) we have been able to either confirm or re-classifying most of the lower classed ``possible PNe''. The resulting large number of PNe now known in the central LMC (\\cite[Reid \\& Parker, 2006a,b]{Reid2006b}) have yielded significant advances in our knowledge of the central LMC's kinematical sub-structure (rotation, inclinations, transverse velocity) and raised interesting questions regarding the kinematical structure of the outer regions. Our access to the MCELS survey (\\cite[Smith et al. 1998]{Smith1998}) provides the opportunity to achieve equivalent results across a much larger 84deg$^{2}$ of the LMC, critical for a complete determination of kinematics, abundance gradients and central star properties (eg. \\cite[Reid \\& Parker, 2010b]{Reid2010b}) . We have already used AAOmega to spectroscopically verify 93 new PNe to add to the 102 previously known in the outer LMC (eg. Fig.\\,\\ref{fig1}). New PN radial velocities are being compared to other tracers and the HI gas disk. These are being added to existing kinematic data to create gradients and verify models. With a near complete LMC PNe census to V=22, an unbiased LMC PNe luminosity function (PNLF) is being built in order to identify any population sub-trends or `dips' while providing an accurate bright-end cut off which is used as a standard candle. ", "conclusions": "" }, "1112/1112.4482_arXiv.txt": { "abstract": "The usage of the high-level scripting language Python has enabled new mechanisms for data interrogation, discovery and visualization of scientific data. We present \\yt{}\\footnote{Available at \\texttt{http://yt-project.org/}}, an open source, community-developed astrophysical analysis and visualization toolkit for data generated by high-performance computing (HPC) simulations of astrophysical phenomena. Through a separation of responsibilities in the underlying Python code, \\yt{} allows data generated by incompatible, and sometimes even directly competing, astrophysical simulation platforms to be analyzed in a consistent manner, focusing on physically relevant quantities rather than quantities native to astrophysical simulation codes. We present on its mechanisms for data access, capabilities for MPI-parallel analysis, and its implementation as an \\textit{in situ} analysis and visualization tool. ", "introduction": "\\label{sec:introduction} In the last decade, multiphysics astrophysical simulations have increased exponentially in both sophistication and size \\cite{ 2005Natur.435..629S, % 2008JPhCS.125a2008K, % 2009JCoPh.228.6833R, % 2008MNRAS.390.1326O, % 2007arXiv0705.1556N, % 2010arXiv1008.4368K, % 2010arXiv1008.2801A, % 2010arXiv1002.3660K, % 2000ApJS..131..273F, % 2007ApJ...659L..87A}; % however, the software tools to mine those simulations have not kept pace. Typically, methods for examining data suffer from a lack of agility, discouraging exploratory investigation. To accommodate this, massively parallel visualization tools such as VisIT and ParaView \\cite{visit_paper, paraview_paper} have been repurposed as domain-specific astrophysical tools. This repurposing, while effective, does not benefit from domain-specific analysis or data structures. The lack of domain-specific quantitative analysis tools designed for astrophysical data leads to the development of specialized tools by individual researchers or research groups, most of which are never shared outside the research group. This can substantially inhibit collaboration between different groups--even those using the same simulation code. This fractionation of the astrophysical community demonstrates a clear need for a flexible and cross-code software package for quantitative data analysis and visualization. In this paper we present \\yt{} \\cite{yt_full_paper}, a data analysis and visualization package that works with several astrophysical simulation codes. \\yt{} is developed openly and is freely available at \\texttt{http://yt-project.org/}. It has been designed to be a common platform for simulation analysis, so that scripts can be shared across groups and analysis can be repeated by independent scientists\\footnote{A platform for sharing scripts is provided with yt, with command-line helpers, at \\texttt{http://hub.yt-project.org}.}. By making this tool available, we hope not only to encourage cross-group collaboration and validation of results, but to remove or at least greatly lower the barrier to entry for exploratory simulation analysis. \\yt{} provides mechanisms for conducting complete analysis pipelines resulting in publication quality figures and data tables, as well as the necessary components for constructing new methods for examining data. The concepts for data handling and representation in \\yt{} are certainly not new, but their application to astrophysical data enables complex, detailed analysis pipelines to be shared between individuals studying disparate phenomena using disparate methods. This enables and even encourages reproducibility and independent verification of results. We have built this analysis and visualization code in Python, using NumPy \\cite{numpy_paper} for fast mathematical operations, \\texttt{mpi4py} for MPI-parallelism \\cite{MPI4PY:PAPER1, MPI4PY:PAPER2}, and optionally Matplotlib for 2D visualization \\cite{matplotlib_paper}. Additionally, several core library routines in \\yt{} such as the AMR volume rendering, multi-dimensional binning, and file access routines, are written in Cython. In addition to utilizing community-developed Python modules, \\yt{} is itself a Python module suitable for direct scripting or access as a library. A community of users and developers has grown around the project, with over 20 committers in the history of the project, and it has been used in numerous published papers and posters. (See, for example, \\cite{ 2010ApJ...715.1575S, 2010ApJ...721.1105B, 2009Sci...325..601T, 2011ApJ...738...54K, 2011arXiv1108.4427Z, 2011ApJ...737...63A, 2011MNRAS.414.2297I, 2011ApJ...735...49M, 2011ApJ...731...59C}.) In order to accomodate the diverse computing environments on which astrophysical simulations are run, \\yt{} was designed to use primarily off-screen rendering and scripting interfaces, although several smaller tools are provided for specific, interactive visualization tasks. The former method is well-suited to remote visualization and can be run via a job execution queue on a batch-compute cluster, such as those on which the underlying simulation are run. \\yt{} is subdivided into several sub-packages for data handling, data analysis, and plotting. This modularity encourages the creation of reusable components for multi-step analysis operations. While work continues on the exploratory, post-processing methods of data analysis and visualization that \\yt{} was originally designed for, current development has focused on analysis during the course of a simulation, or so-called \\textit{in situ} analysis. This allows for high-cadence analysis to be conducted without writing data to disk. Future simulations, such as those to be conducted on petascale machines, will require analysis and visualization during the simulation rather than exclusively as a post-processing technique. In this paper, we will describe the mechanisms that \\yt{} provides for accessing data (\\S \\ref{sec:mechanisms}), methods of interacting with \\yt{} (\\S \\ref{sec:interacting}), the visualization techniques offered by \\yt{} (\\S \\ref{sec:visualization}), the parallelism strategy for data analysis and generation of visualizations (\\S \\ref{sec:parallelism}), and end with a discussion of the process of embedding \\yt{} in running simulation codes and how this will be inverted in the future (\\S \\ref{sec:embedding}). ", "conclusions": "\\label{sec:conclusions} The \\yt{} project is fully free and open source software, released under the GNU General Public License, with no dependencies on external code that is not also free and open source software. The development process occurs completely in the open at \\texttt{http://yt-project.org/}, with publicly-accessible source control systems, bug tracking, mailing lists, and regression tests. Building a community of users has been a priority of the \\yt{} development team, both to encourage collaboration and to solicit contributions from new developers; both the user and developer communities are highly distributed around the world. \\yt{} is developed using Mercurial\\footnote{\\texttt{http://mercurial.selenic.com/}}, a distributed version control system that enables local versioned development and encourages users to make and contribute changes upstream. Many of the operations conducted in \\yt{}: fluid analysis, phase diagrams, volume rendering, parallelism, and in situ analysis could feasibly be applied to domains other than astrophysics. We intend to generalize the underlying code base such that it can be applied to many other data formats in astrophysics, and ultimately we hope to provide these tools and techniques to domains other than astrophysics. Our first steps toward this, providing a generic and arbitrary data loader, have shown that it is feasible. Future versions of \\yt{} will generalize fields and particle handling, and should make this process much easier. The creation of a freely available, publicly inspectable platform for simulation analysis allows the community to disentangle the coding process from the scientific process. Simultaneously, by making this platform public, inspectable and freely available, it can be openly improved and verified. The availability and relatively approachable nature of \\yt{}, in addition to the inclusion of many simple analysis tasks, reduces the barrier to entry for young scientists. Rather than worrying about the differences between Enzo and FLASH hierarchy formats, or row versus column ordering, or HDF4 versus HDF5 versus unformatted fortran data formats, researchers can focus on understanding and exploring their data. More generally, however, by orienting the development of an analysis framework as a community project, the fragmentation of methods and mechanisms for astrophysical data analysis is greatly inhibited. Future generations of simulations and simulation codes will not only benefit from this collaboration, but they will require it." }, "1112/1112.3094_arXiv.txt": { "abstract": "There is great interest in numerical relativity simulations involving matter due to the likelihood that binary compact objects involving neutron stars will be detected by gravitational wave observatories in the coming years, as well as to the possibility that binary compact object mergers could explain short-duration gamma-ray bursts. We present a code designed for simulations of hydrodynamics coupled to the Einstein field equations targeted toward such applications. This code has recently been used to study eccentric mergers of black hole-neutron star binaries. We evolve the fluid conservatively using high-resolution shock-capturing methods, while the field equations are solved in the generalized-harmonic formulation with finite differences. In order to resolve the various scales that may arise, we use adaptive mesh refinement (AMR) with grid hierarchies based on truncation error estimates. A noteworthy feature of this code is the implementation of the flux correction algorithm of Berger and Colella to ensure that the conservative nature of fluid advection is respected across AMR boundaries. We present various tests to compare the performance of different limiters and flux calculation methods, as well as to demonstrate the utility of AMR flux corrections. ", "introduction": "\\label{intro} The interface between strong field gravity and matter dynamics promises to be one of the important frontiers in the coming years. A new generation of gravitational wave detectors (LIGO~\\cite{LIGO}, GEO~\\cite{GEO},TAMA~\\cite{TAMA}, and VIRGO~\\cite{VIRGO}) are now operational, and within the next few years are expected to reach sensitivities that will allow observations of the Universe in gravitational radiation for the first time. The prime targets of these observations are compact object (CO) binaries composed of combinations of black holes (BHs) and neutron stars (NSs). Modeling of such sources is a crucial ingredient to realize the promise of gravitational wave astronomy. Even if an event is detected with a high signal-to-noise ratio, reconstructing the dynamics of the system that produced the signal cannot be done directly but instead will require template banks of theoretical waveforms informed by numerical simulations. Compact object mergers involving NSs are expected to be significant sources of not only gravitational radiation, but also possible progenitors for short-gamma-ray bursts (SGRBs)~\\cite{npp92,2005Natur.437..851G,2005ApJ...630L.165L} and other electromagnetic and neutrino counterparts~\\cite{Metzger:2011bv}. Efforts are already underway to use potential gravitational wave sources as triggers for searches for electromagnetic transients~\\cite{LIGO_EM,LIGO_EM2}. Observations would help constrain evolutionary models for the progenitor stars and their environments. Perhaps most intriguingly, the observations would give clues to the equation of state (EOS) of matter at nuclear densities (as in NS interiors), which cannot be probed in laboratories on Earth and is not fully understood at the theoretical level (for a broad discussion see for example~\\cite{glendenning}). The reason that the gravitational wave signature could contain information about the matter EOS (and other details about the internal structure of neutron stars) is that the EOS in general has a significant effect on the bulk motion of matter, and it is this bulk motion that is the mechanism by which gravitational waves are produced. Several studies to date have looked into this issue, suggesting the imprint of the EOS on the gravitational waves will be strong enough to detect~\\cite{Kiuchi:2011re,Pannarale:2011pk,jocelyn,Oechslin:2007gn,Tsui:2006tr,Kokkotas:2005vr,Shibata:2005xz,Shibata:2005ss,Faber:2004fs,Bejger:2004zx,Lai:1996sv,Xing:1994ak} (though, in some cases, the expected frequencies are higher than the range to which the current generation of ground-based detectors are most sensitive, thus limiting the information which can be extracted). While CO binaries containing NSs are a particularly interesting class of sources involving general relativistic (GR) hydrodynamics, they are by no means the only such systems. Examples of additional systems that have already been considered include BH accretion tori~\\cite{montero,korobkin,2011PhRvD..84b4024F,2012ApJ...744...45B} and NS-white dwarf mergers~\\cite{vasileos}. Thoroughly modeling systems like those described above would require evolution of the spacetime, the photon and neutrino radiation fields, and the magnetized, relativistic fluid. Even a minimalistic treatment, with the Einstein equations coupled to the equations of relativistic hydrodynamics, represents a complex, nonlinear system of partial differential equations. Numerical simulations are thus essential for exploring such strong field, dynamical systems. There is a long history of adapting successful techniques for simulating Newtonian hydrodynamics to relativistic and general relativistic fluids which we will not attempt to summarize (see~\\cite{grhydro_lr} for a review of general relativistic hydrodynamics). Instead, we will briefly attempt to place the code described in the present paper in the context of other recent codes developed for fluids on evolving spacetimes. \\footnote{Note that our focus is restricted to codes which handle dynamically evolving gravitational fields. Such codes, however, frequently owe much to earlier, fixed-background evolution codes (see~\\cite{grhydro_lr}). In addition, advancements such as GR-hydro with multipatch grids~\\cite{thor} and with GPUs~\\cite{horizon} have recently been made with fixed-background codes.} Several of these codes~\\cite{mhdcode1,SACRA,Giacomazzo:2007ti,Jena_nsns,Bode:2009mt} solve the field equations in the BSSN formulation~\\cite{sn,bs}. The remainder~\\cite{matt,Anderson:2006ay} use the generalized-harmonic formulation~\\cite{garfinkle,gh3d} which we also employ; unlike our code, however, these groups convert to a fully first-order formulation~\\cite{new_lindblom_et_al}. Most groups use finite-difference methods for the metric evolution and a conservative, high-resolution shock-capturing (HRSC) scheme for the hydro evolution; these unigrid algorithms are then interfaced with some sort of adaptive mesh refinement (AMR). A notable exception for the metric evolution is~\\cite{matt}, which employs pseudospectral methods for the metric and then interpolates to a finite-volume grid for the fluid. Some groups have implemented the MHD equations in full GR; since these codes all make use of conservative HRSC methods, they may be principally differentiated by how they meet the challenge of preserving the $\\nabla \\cdot {\\mathbf B} = 0$ constraint. (A straightforward finite-difference evolution of the magnetic field would generically lead to magnetic monopoles and, hence, unphysical behavior.) {\\tt WhiskyMHD} employs constrained transport~\\cite{Giacomazzo:2007ti} for this purpose, which preserves the constraint to machine accuracy, whereas the code of~\\cite{chawla} uses hyperbolic divergence cleaning. Constrained transport, however, requires special interpolation at refinement level boundaries in order to preserve the constraint. The Illinois group found that a vector-potential formulation of the MHD equations works well when coupled to AMR~\\cite{illinoisNewMHD}. This is because the constraint is preserved by construction with the vector-potential, even with the restriction and prolongation operations of AMR (see also~\\cite{illinoisMHD2} for a thorough examination of the electromagnetic gauge condition). Studies indicate that magnetic fields do not significantly affect the gravitational dynamics of CO mergers (see {\\em e.g.} \\cite{chawla}), but they could be critical for understanding EM counterparts including the possible formation of a SGRB engine. A new method to treat the MHD equations was recently presented in~\\cite{lehner_new}, where ideal MHD is used in high matter density regions ({\\em e.g.} inside a NS), while the force-free approximation is used elsewhere ({\\em e.g.} the magnetosphere of a NS). The authors applied the method to study the collapse of magnetized hypermassive NSs (which could be formed via binary NS mergers) and suggested that intense EM outbursts could accompany such events. Besides MHD, the other major advances in the physical model for numerical relativity codes have been in the arena of microphysics. While the $\\Gamma=2$ EOS was the community standard for quite some time, most codes now allow for a nuclear theory-based EOS~\\cite{matt2,shibataBNS} and/or use various parametrized, piecewise polytropic EOSs inspired by the range of plausible nuclear EOSs~\\cite{shibataBHNS4,illinoisWDNS}. These advances in EOS description primarily affect the cold NS structure, but the group developing the {\\tt SACRA} code has also begun to account for neutrino transport via a simplified leakage scheme~\\cite{Kiuchi:2011re,2011PhRvL.107e1102S}. The same group has also made available a formulation for a more accurate truncated moment scheme with a variable Eddington factor closure~\\cite{truncated_moment}, which shows much promise for numerical relativity simulations with neutrino physics beyond the leakage approximation. Another category of GR hydrodynamics codes employs the conformal-flatness approximation, which is particularly useful when supernova simulations are the target application. An example is CoCoNuT/VERTEX, which incorporates relativistic hydrodynamics, conformally flat gravity, and ray-by-ray neutrino transport~\\cite{coconut}. The code of~\\cite{cerda_duran} employs a similar scheme for hydrodynamics and gravity but adds a test magnetic field; this code has been used to study the magnetorotational instability in supernovae. Newtonian (and semi-Newtonian)~\\cite{lee_kluzniak,magma}, conformally flat~\\cite{faber,oechslin}, and fixed-background~\\cite{laguna} SPH codes represent an important, orthogonal approach to studying CO interactions. SPH has an advantage over Eulerian schemes when a large range of spatial scales is involved. Such a situation may arise in CO mergers when material is stripped from a star in a tidal interaction and forms an extended tail. On the other hand, Eulerian codes are the standard approach when strong shocks are present, as would arise in binary NS mergers or disk circularization. (Recent progress has been made, however, in applying SPH to situations with relativistic shocks~\\cite{Rosswog_SRHD}.) In addition, SPH has not (to our knowledge) yet been coupled to a code which evolves the full Einstein equations. Nonetheless, comparisons between Eulerian and SPH results could prove very useful on a problem-by-problem basis to characterize the errors in both methods. Though current efforts in GR simulations involving matter tend to focus on increasingly complex physical models, there remain many unanswered questions in the astrophysics of compact objects that can be addressed with a code which solves the Einstein equations coupled to perfect fluid hydrodynamics. We have thus focused our code development on hydrodynamics in full GR, while maintaining a flexible infrastructure to accommodate additional physics modules in the future. We evolve the field equations in the generalized-harmonic formulation using finite differences. The fluid is evolved conservatively using one of several different shock-capturing techniques we test here. We have also implemented the hydrodynamical equations in a manner that is independent of EOS. We make use of AMR by dynamically adapting the mesh refinement hierarchy based on truncation error estimates of a select number of the evolved variables. We also utilize Berger and Colella~\\cite{bc89} style flux corrections (also known as ``refluxing'') in order to make the use of AMR compatible with the conservative nature of the hydrodynamic equations. Though AMR flux corrections have been implemented in other astrophysical hydrodynamics codes (such as Athena~\\cite{ATHENA}, CASTRO~\\cite{CASTRO}, {\\tt Enzo}~\\cite{EnzoMHD}, and FLASH~\\cite{FLASH}), to our knowledge this algorithm has not been used previously for hydrodynamics simulations in full general relativity.\\footnote{Note that ``flux correction'' here refers to the enforcement of conservation at AMR boundaries, not the recalculation of fluxes with a more dissipative scheme to preserve stability as in Athena~\\cite{athena_rmhd}.} A further noteworthy feature of our implementation is that we store corrections to the corresponding fluid quantity integrated in the volume of a given cell instead of the flux, allowing for easy implementation within a computational infrastructure that supports cell-centered but not face-centered distributed data structures. The code described here has recently been applied to studying BH-NS mergers with eccentricity as may arise in dense stellar systems such as galactic nuclear clusters and globular clusters~\\cite{ebhns_letter,ebhns_paper}. In the remainder of this paper we outline our computational methodology for simulating hydrodynamics coupled to the Einstein field equations and describe tests of this methodology. In Sec.~\\ref{methods} we review the generalized-harmonic approach to solving the field equations and present our methods for conservatively evolving a perfect fluid coupled to gravity, including our method for inverting the conserved quantities to obtain the primitive fluid variables and the implementation of flux corrections to enforce the conservation of fluid quantities across AMR boundaries. In Sec.~\\ref{tests} we present simulation results which test these methods, highlight the strengths and weaknesses of various shock -capturing techniques, and demonstrate the utility of the flux correction algorithm. ", "conclusions": "Numerous scenarios that fall within the purview of general relativistic hydrodynamics are still mostly unexplored---especially CO mergers involving neutron stars. There is a rich parameter space, of which large areas remain uncharted due to uncertainty or potential variability in BH and NS masses, BH spin and alignment, the NS EOS, and other aspects. Beyond the pure hydrodynamics problem, the roles of magnetic fields and neutrino physics are just beginning to be explored by various groups, and we expect to add support for such physics to our code in the future. The potential applications of robust and flexible numerical algorithms for evolving hydrodynamics together with the Einstein field equations are manifold. With this in mind, we have implemented methods for conservatively evolving arbitrary EOSs, in particular for converting from conserved to primitive variables without knowledge of derivatives; and we have implemented numerous reconstruction and flux calculation methods that can be used interchangeably to meet problem specific requirements. Though accurate treatment of shocks may not be crucial for BH-NS mergers (where shocks are not expected to be dynamically important), the same is not true of NS-NS binaries, especially eccentric ones where the stars may come into contact during nonmerger close encounters~\\cite{Roman}. We have also taken care to implement a flux correction algorithm that preserves the conservative nature of hydrodynamical advection across AMR boundaries. Though strict conservation is not, strictly speaking, essential (since any nonconservation would be at the level of truncation error), it is an especially appealing property when studying, for example, CO mergers as potential SGRB progenitors. After merger, material that did not fall into the black hole --- typically on the order of a few percent of the original NS mass --- will fill a large volume making up an accretion disk and potentially unbound material. Though accurately tracking this material is not important for the gravitational dynamics, it is critical for characterizing potential EM counterparts to the merger." }, "1112/1112.1058_arXiv.txt": { "abstract": "{} {We present the analysis of a large sample of gamma-ray burst (GRB) X-ray light curves in the rest frame to characterise their intrinsic properties in the context of different theoretical scenarios.} {We determine the morphology, time scales, and energetics of $64$ long GRBs observed by \\emph{Swift}/XRT \\emph{without} flaring activity. We furthermore provide a one-to-one comparison to the properties of GRBs \\emph{with} X-ray flares.} {We find that the steep decay morphology and its connection with X-ray flares favour a scenario in which a central engine origin. We show that this scenario can also account for the shallow decay phase, provided that the GRB progenitor star has a self-similar structure with a constant envelope-to-core mass ratio $\\sim 0.02-0.03$. However, difficulties arise for very long duration ($t_p\\gtrsim10^4$ s) shallow phases. Alternatively, a spinning-down magnetar whose emitted power refreshes the forward shock can quantitatively account for the shallow decay properties. In particular we demonstrate that this model can account for the plateau luminosity vs. end time anticorrelation.} {} ", "introduction": "Since the launch of the \\emph{Swift} satellite in 2004 \\citep{2004ApJ...611.1005G}, the evolution of the gamma-ray burst (GRB) X-ray light curves has appeared to be more complex than previously thought. The different shapes of the light curves \\citep{2006ApJ...642..389N} and the existence of X-ray flares superimposed on the smooth continuum \\citep{2007ApJ...671.1903C,chinca10} up to very late times \\citep{2008A&A...487..533C,2011A&A...526A..27B} have prompted a large number of studies conducted on different samples of X-ray light curves in order to single out their main properties \\citep{2006ApJ...642..354Z,2006ApJ...647.1213O,2007ApJ...662.1093W,2007ApJ...666.1002Z,2007ApJ...670..565L,2008ApJ...675..528L,2009ApJ...707..328L,2008ApJ...689.1161G,2009ApJ...698...43R,2009MNRAS.397.1177E}. Analysis of the observational properties of X-ray flares \\citep{2007ApJ...671.1903C,chinca10,giantflares10,2011A&A...526A..27B,2011MNRAS.410.1064M} allowed us to establish a connection with the prompt emission. In particular, the extension of the prompt lag-luminosity relation to X-ray flares strongly suggests that X-ray flares and prompt emission pulses are produced by the same mechanism \\citep{giantflares10}. The nature of the underlying X-ray \\emph{continuum} is still debated. It was immediately clear soon after the first set of observations by \\emph{Swift}/XRT \\citep{2005SSRv..120..165B} and redshift measurements that the light curves differed from a simple power-law behaviour and that some of the characteristics repeat in a systematic way \\citep[see][]{2006ApJ...642..389N,2006ApJ...642..354Z,2006ApJ...647.1213O}. After six years of data collection, we now have a much larger sample for investigating the morphology of the X-ray light curves, and a rather large subsample of redshift measurements for this analysis in the rest frame. The rest frame allows us to understand the intrinsic properties, the timescales involved, and the energetics. In fact, the parameters derived are directly related to the physics of the phenomenon, and the distributions are not affected by the distribution of redshift. This is the approach we adopt in this work. We select light curves observed by XRT without flares, in order to avoid any kind of contamination from other emission components. When necessary we compare our results with the sample analysed in \\citet{2011MNRAS.410.1064M} of GRBs with flares, providing for the first time a one-to-one comparison of light curves \\emph{with} and \\emph{without} flares. The initial steep decay of the afterglow is commonly due to the tail of the prompt emission \\citep{2000ApJ...541L..51K}, and the plateau phase is often connected with external shock emission with an injection of energy. However, the standard model for GRBs suffers from several problems \\citep[for a review see][]{2009arXiv0911.0349L}. Here we aim at exploring alternative mechanisms to shape the light curve, such as the accretion of the stellar material left behind by the collapse of the massive star \\citep{2008MNRAS.388.1729K} or the rotational energy of a newly-born magnetar \\citep[see e.g.][and references therein]{2010arXiv1012.0001M}. We show indeed that an accretion model would explain the X-ray light curves well enough, as well as the properties of flares as discussed by \\citet{2011MNRAS.410.1064M}. The plateau phase agrees with the assumption that new energy is injected from a spinning down neutron star into the forward shock. In particular within this framework we demonstrate that this model can account for the anticorrelation between the plateau luminosity and its end time found by \\citet{2008MNRAS.391L..79D,2010ApJ...722L.215D}. In Sect.~\\ref{data_analysis} we describe the light curve extraction and the luminosity calibration procedures. In Sect.~\\ref{sample} we describe the sample selection criteria, the fitting procedure and the main results about the morphology and spectral properties of the sample. In Sect.~\\ref{energy} we present the analysis of the energy of the sample compared with the prompt emission energy. In Sect.~\\ref{flares} we compare the present sample with the light curves with flares analysed in \\citet{2011MNRAS.410.1064M}. In Sect.~\\ref{discussion} we discuss our findings and in Sect.~\\ref{conclusions} we draw our main conclusions. We adopt standard values of the cosmological parameters: $H_\\circ=70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_M=0.27$, and $\\Omega_{\\Lambda}=0.73$. Errors are given at $1\\, \\sigma$ confidence level unless otherwise stated. Times are in rest frame unless otherwise stated. ", "conclusions": "We analysed $64$ long GRB X-ray light curves observed by XRT with redshift measurement that do not exhibit flaring activity. This allowed us to characterise the morphology and energetics of the sample in order to constrain the mechanism that produces the X-ray continuum. The light curves of the sample can be divided into four morphological Types, with similarities between different parts. When necessary, we compared our sample with the one considered in \\citet{2011MNRAS.410.1064M} of $44$ GRBs that exhibit flaring activity. We found that \\begin{itemize} \\item a large fraction of the light curves are on average consistent with the high latitude emission from a structured jet, but all of them show a strong or moderate spectral evolution. The accretion model accounts for the observed range of temporal decays and the spectral evolution of the steep decay phases of Types Ib and II; \\item the injection phase observed in Type Ia and II can be interpreted as power emitted from spinning-down ultramagnetised neutron star that refreshes the forward shock. We applied such a model to our sample and found that it fits the data well enough; \\item this scenario accounts for the observed $L_p-t_p$ anticorrelation \\citet{2008MNRAS.391L..79D,2010ApJ...722L.215D}: we reproduced the normalisation, slope, and scatter of the anticorrelation starting from plausible values of the parameters. The consistency of Type Ia shallow decay with the anticorrelation within $99\\%$ confidence level suggests that they can be reproduced by this model with the same distribution of $B$ and $P$ as for Type II; \\item a second phase of accretion of a progenitor with a structure core envelope can originate the shallow phase in Types Ia and II. The correlation between the prompt emission total isotropic energy with the shallow decay X-ray energy implies a self-similar structure for the progenitor star, with $\\sim 0.02-0.03$. However in this scenario, difficulties arise for the very long duration ($t\\gtrsim 10^4$ s) of Type Ib shallow decay; \\item $70\\%$ of the light curves with flares are Type II, while only $23\\%$ of the non-flare sample are, confirming our previous finding \\citep{giantflares10}. This suggests that Type II light curves are related to the suitable conditions for the occurrence of flares. The accretion model has the merit of proposing a unique mechanism to explain the steep decay and the flaring activity; \\end{itemize} The magnetar model is able to explain the different morphologies of the X-ray continuum and the properties of the shallow decay. The main limit in the accretion model is the difficulty of achieving very long timescales as observed in Type Ib shallow decay. The behaviour of the steep decay and the suggested connection with the flaring activity \\citep[see also][]{2009MNRAS.395..955B,giantflares10,2011MNRAS.410.1064M} indicates that the steep decay may not be viewed as simply the tail of the prompt emission." }, "1112/1112.6164_arXiv.txt": { "abstract": "We study the role of viscosity and the effects of a magnetic field on a rotating, self-gravitating fluid, using Newtonian theory and adopting the ideal magnetohydrodynamic approximation. Our results confirm that viscosity can generate vorticity in inhomogeneous environments, while the magnetic tension can produce vorticity even in the absence of fluid pressure and density gradients. Linearizing our equations around an Einstein-de Sitter cosmology, we find that viscosity adds to the diluting effect of the universal expansion. Typically, however, the dissipative viscous effects are confined to relatively small scales. We also identify the characteristic length bellow which the viscous dissipation is strong and beyond which viscosity is essentially negligible. In contrast, magnetism seems to favor cosmic rotation. The magnetic presence is found to slow down the standard decay-rate of linear vortices, thus leading to universes with more residual rotation than generally anticipated. ", "introduction": "Most, if not all, astrophysical systems rotate and it is conceivable that vorticity can play an important role in a number of situations. So far, however, the majority of the related studies has been confined to relatively small-scale systems. Although the question of universal rotation and how large this may be has been posed more than once (see, e.g.,~\\cite{BJS,KO,KHB,GSFB,JBEGH,SC} and references therein), in cosmology the role of vorticity is usually bypassed. One of the reasons for doing so is that typical inflationary scenarios are irrotational, since the vacuum fluctuations do not generate vorticity~\\cite{LL}. Another, is that linear rotational perturbations are known to decay as the inverse square of the cosmological scale factor, which implies strong dilution for any vortex-like distortions that could happen to exist. Dissipative effects during the subsequent radiation era, like Silk damping and neutrino free-streaming, may also have a negative impact on large-scale rotation, though (to the best of our knowledge) studies of this nature are not yet available in the literature. In cosmology, where the terms vorticity and rotation are used interchangeably, vorticity is usually generated from nonlinear effects. This is one additional reason for not taking vorticity into account in many cosmological perturbation studies. Inhomogeneous non-barotropic (e.g., non-adiabatic) media, for example, can generate rotation at the second perturbative level~\\cite{CMM,CM}, and so can fluid viscosity~\\cite{MB,D-SB,Fetal}. Nevertheless, resorting to nonlinear effects is not always necessary, since there are agents that can induce rotation still at the linear level. For instance, a coherent magnetic field (`$B$-field') can act as a source of cosmic vorticity~\\cite{W,TM}. The opposite is also theoretically possible. Thus, following~\\cite{H}, a number of authors have explored the likelihood that rotational distortions can lead to cosmic magnetogenesis (see, e.g.,~\\cite{Metal,MMNR,KMST,GR} and references therein). In addition, both the fluid viscosity and the $B$-field can affect the evolution of rotational perturbations and thus change the standard picture, which usually assumes non-magnetized, perfect media. Therefore, in principle at least, the effects of the aforementioned sources could lead to more residual vorticity than previously anticipated. The present paper uses Newtonian gravity to study the effects of viscosity and those of a magnetic field on rotation. We adopt the covariant, Lagrangian approach to fluid dynamics, which has been proved helpful in several (mainly relativistic) cosmological studies (see~\\cite{E1} for an introduction to the Newtonian version of the formalism). First, we look into the effects of fluid viscosity and then consider those of a coherent magnetic field. Our results, confirm those of earlier studies claiming that both of the aforementioned agents can act as sources of cosmic rotation. We show, in particular, that viscosity can only generate vorticity in media with an inhomogeneous density distribution. This is not the case for the $B$-field, which does not (in principle at least) require the presence of density perturbations. Assuming the Newtonian analogue of the Einstein-de Sitter universe as our background model, we find that viscosity can only generate vorticity at second order. Magnetism, on the other hand, can do the same at the linear perturbative level. Moreover, the irreducible decomposition of variables and equations that the covariant formalism introduces, allows us to identify the agents leading to vorticity generation, as well as the way they do so. Viscosity, for example, sources rotational distortions via the coupling of the shear with certain types of density perturbations. Alternatively, the magnetic field acts primarily through the tension component of the Lorentz force. Our next step is to investigate how viscous and/or magnetized media can affect the linear evolution of rotational perturbations. In the first case, we find that viscosity always leads to vorticity dissipation, though the effects are generally scale dependent. We identify, in particular, a characteristic length beyond which viscosity is essentially negligible and below which its role becomes progressively more important. In particular, well inside this ``viscous scale'' vorticity dissipates very quickly, while far beyond that scale the fluid kinematics are dominated by the background Hubble expansion. There, one recovers the standard inverse-square decay law for rotational perturbations. The size of the aforementioned viscous scale depends on the specifics of the cosmic fluid, such as the average velocity and the mean free-path of the species involved. When dealing with non-relativistic matter, however, the typical domain affected by viscosity is very small (in cosmological terms). In contrast to viscosity, the presence of magnetic fields seems to have a favorable effect on linear rotational perturbations. In addition to generating vorticity at the linear order, the $B$-field slows down the standard decay rate of rotation in perturbed Friedmann-Robertson-Walker (FRW) universes. As mentioned earlier, vorticity typically depletes as $a^{-2}$, with $a$ representing the cosmological scale factor. When the linear magnetic effects are accounted for, however, the vorticity decay-rate changes to $a^{-3/2}$. The implication of this effect, which comes entirely from the magnetic tension, is that magnetized universes, can have more residual vorticity than their non-magnetic counterparts. An analogous, conclusion was also reached indirectly, through the relativistic analysis of linear density inhomogeneities in magnetized cosmologies~\\cite{TM}. To the best of our knowledge, this is the first direct indication that magnetism ``favors'' both the generation and the survival of cosmic rotation. ", "conclusions": "Realistic media do not behave like perfect fluids and magnetic fields, of various scale and strength, appear to be everywhere in the universe. At the same time, cosmic rotation still poses a number of fundamental questions that remain as yet unanswered. In this respect, it would be interesting to see whether viscosity and magnetism have something new to say about vorticity in general. So far, both agents have been known to act as sources of rotation, though their effects on pre-existing vorticity are not clear yet. The present article adopts the Newtonian approach to investigate the rotational behavior of a viscous fluid and subsequently that of a magnetized medium. At first, our analysis is fully nonlinear and confirms that both viscosity and magnetism can generate vorticity. We also show that the former of these two agents does so through the coupling of the shear with certain types of density inhomogeneities. Magnetism, on the other hand, does not necessarily need an inhomogeneous matter distribution to source rotational distortions. The $B$-field generates vorticity by means of its Lorentz force, especially via the tension component of the latter. Assuming the Newtonian analogue of the Einstein-de Sitter universe and then linearizing our equations around this background, we looked into the effects of viscosity and magnetism on rotational cosmological perturbations. We found that the former effects are negative. Viscosity generally leads to vorticity dissipation, though (for non-relativistic species) its role is typically confined to relatively small lengths. Nevertheless, further study is required to establish the actual range of the dissipative effects. Here, we have taken the first step in this direction by identifying the characteristic ``viscous scale'', below which viscosity dominates and beyond which the universal expansion takes over. In contrast, the magnetic presence seems to favor rotation. More specifically, after linearizing our equations around the aforementioned Einstein-de Sitter background, we found that the $B$-field slows down the standard decay rate of rotational cosmological perturbations. In particular, vorticity no longer obeys the standard $a^{-2}$ decay-law, but drops as $a^{-3/2}$ ($a$ is the cosmological scale factor). Also, the magnetic effects are scale-independent, which means that their domain of influence is decided by the coherence length of the $B$-field. Analogous results were also obtained when studying vortex-like perturbations in the density distribution of a magnetized medium. Overall, it appears as though magnetism favors both the emergence and the survival of cosmic rotation. Put another way, magnetized universes should contain more residual vorticity than their magnetic-free counterparts. In closing, we should also note that it would be worth-studying the above described effects using a relativistic approach. This will enable one to involve highly energetic particles, look deep into the pre-decoupling universe and also allow for spacetimes more complex than the Einstein-de Sitter universe." }, "1112/1112.4343_arXiv.txt": { "abstract": "{The composition of Galactic Cosmic Rays (GCR) presents strong similarities to the standard (cosmic) composition, but also noticeable differences, the most important being the high isotopic ratio of \\neo, which is $\\sim$5 times higher in GCR than in the Sun. This ratio provides key information on the GCR origin.} {We investigate the idea that GCR are accelerated by the forward shocks of supernova explosions, as they run through the presupernova winds of the massive stars and through the interstellar medium.} { We use detailed wind and core yields of rotating and non-rotating models of massive stars with mass loss, as well as simple models for the properties of the forward shock and of the circumstellar medium.} { We find that the observed GCR \\neo \\ ratio can be explained if GCR are accelerated only during the early Sedov phase, for shock velocities $>$1500-1900 km/s. The acceleration efficiency is found to be of the order of 10$^{-6}$-10$^{-5}$, i.e. a few particles out of a million encountered by the shock escape the SN at GCR energies. We also show quantitatively that the widely publicized idea that GCR are accelerated in superbubbles fails to account for the high \\neo \\ ratio in GCR.} {} ", "introduction": "Supernova (SN) shocks are generally thought to be the main accelerator of the bulk of Galactic Cosmic Rays (GCR). Indeed, the power of GCR in the Milky Way is estimated to 10$^{41}$ ergs/s, corresponding to 10-20 \\% of the kinetic power of Galactic supernovae (assuming canonical values of 2 SN per century, each one releasing 10$^{51}$ ergs of kinetic energy). The site of the acceleration of GCR remains debatable today, despite more than five decades of theoretical and observational studies (e.g. Strong et al. 2007 and references therein). Over the years, it has been suggested that GCR are accelerated in 1) SN remnants (either by the forward or the reverse shock or both), 2) the interstellar medium (ISM), 3) the winds of massive stars, 4) the interiors of superbubbles, excavated by the massive star winds and the subsequent SN explosions of an OB association. Each one of the proposed sites has its own advantages and shortcomings, regarding the energetics and/or the composition of accelerated matter. For instance, it has been argued that the hot, low density environment of a superbubble minmizes radiative losses of SN shocks and energy losses of accelerated particles, thus allowing the latter to reach substantial energies, up to the \"knee\"of the GCR spectrum (e.g. Parizot et al. 2004). On the other hand, reverse SN shocks running into the SN interior carry insufficient energy to explain the bulk GCR energetics (Ramaty et al. 1997). Moreover, they should accelerate $^{59}$Ni, a product of explosive nucleosynthesis which is unstable to e$^-$-capture (with a lifetime of 10$^4$ yr) and which has not been detected in GCR (Wiedenbeck et al. 1999), while a rapid acceleration would render it practically stable and thus detectable. It was realized early on that the {\\it elemental} composition of GCR differs significantly from the one of the ISM. Those differences may provide valuable information on the origin of GCR particles and, perhaps, on the site - and even the mechanism - of acceleration. Volatiles behave differently from refractories: the former display a mass-dependent enrichment with respect to H, which reaches a factor of 10 for the heaviest of them; the latter are all overabundant (w.r.t. H) by a factor of 20, while C and O display intermediate overabundances, by factors of 9 and 5, respectively (e.g. Wiedenbeck 2007 and references therein). This complex pattern is now thought to result not from ionization effects (as suggested in Cass\\'e and Goret 1978, and further developped by Meyer 1985) but rather from effects related to elemental condensation temperature (Meyer et al. 1997): refractories are locked in dust grains, which are sputtered by repeated SN shocks and the released ions are easily picked-up and accelerated (Ellison et al. 1997). This, quite elaborate, scheme, which builds on earlier ideas by e.g. Cesarsky and Bibring (1981), accounts quantitatively for most of the observed features of GCR source composition; still, it leaves unanswered the key issue about the acceleration site of GCR (and how it affects the composition of accelerated matter). The most conspicuous feature of GCR source composition is undoubtely the high isotopic \\neo \\ ratio. Its value was measured since the late 1970ies (Garcia-Munoz, Simpson and Wefel 1979, Wiedenbeck and Greiner 1981). The most accurate measurement today, obtained from analysis of the CRIS instrument, leads to a best estimate (Binns et al. 2008) of 0.387 $\\pm$ 0.007 (statistical) $\\pm$ 0.022 (systematic). This is 5.3$\\pm$0.3 times the value of the (\\neo)$_{\\odot}$ \\ ratio in the solar wind. Contrary to the case of the elemental source GCR abundances, which may be affected by various physico-chemical factors (first ionization potential, condensation temperature, etc.) isotopic ratios can only be affected by nucleosynthetic processes and thus provide crucial information on the origin of cosmic ray particles. It should be noticed that up to now there is now clear evidence for any other GCR isotopic ratio to differ from solar, with the potential exception of $^{58}$Fe/$^{56}$Fe, which is estimated to 1.5$\\pm$0.3 times solar (Binns et al. 2008). Soon after the discovery of the anomalous GCR \\neo \\ ratio, Cass\\'e and Paul (1982) suggested that it could be explained by a mixture of $\\sim$2\\% of material from the wind of a WC star to 98\\% of material with standard composition. In early He-burning, $^{14}$N (produced through the CNO cycle in the previous H-burning phase) is transformed almost totally in \\neb \\ through $^{14}$N($\\alpha,\\gamma$)$^{18}$F($\\beta^+$)$^{18}$O($\\alpha,\\gamma$)$^{22}$Ne. He-burning products (like $^{12}$C and \\neb) are expelled by the stellar winds of massive stars during their WC phase. The observed GCR \\neo \\ ratio is obtained by assuming dilution of WC material with matter of standard composition. Subsequent studies put the aforementionned idea on a quantitative basis, with the use of detailed models of the evolution and nucleosynthesis of massive, mass losing stars (Maeder 1983, Prantzos 1984, Meyer 1985, Prantzos et al. 1987). In those studies, the acceleration site of GCR was considered as decoupled from the nucleosynthesis site, and unrelated to the fraction of admixtured WC material. Higdon and Lingenfelter (2003) evaluated quantitatively the \\neo \\ ratio within a superbubble, created by the collective action of stellar winds and SN shockwaves. They adopted stellar wind yields for \\nea \\ and \\neb \\ from the models of Schaller et al. (1992) and SN yields from the models of Woosley and Weaver (1995). They found that the \\neo \\ ratio in the superbubble decreases with time (since \\neb \\ from the winds dominates the evolution of \\neo \\ at early times) and that its time average value is compatible with the GCR source \\neo \\ inferred from observations. In a subsequent paper, Lingenfelter and Higdon (2007) recognised that the Schaller et al. (1992) yields of \\neb \\ were highly overestimated\\footnote{The reason was the excessively high mass loss rates adopted in that work.} and, consequently, \"... new detailed calculations of the expected GCR isotopic ratio are called for...\", but they did not attempt such a re-evaluation. In the meantime, Binns et al. (2005), using updated wind yields of massive stars with rotation (from the Geneva group, see Sec. 2.2), found good agreement between the observed \\neo \\ ratio and an admixture of $\\sim$20\\% material from WR stars with 80\\% material of standard composition. According to Binns et al. (2008), since WR stars are evolutionary products of OB stars, such an agreement \"...suggests that OB associations within superbubbles are the likely source of at least a substantial fraction of GCR\". Howewer, theoretical studies in the past 10 years are based mostly on the paradigm of GCR being accelerated in SN remnants, not in superbubbles, e.g. \\cite{Ptuskin05, Berezhko06,Berezhko09,Ptuskin10,Caprioli10,Schure10,Ellison11} \\ and references therein. The kinetic energy of the bulk motion of the forward shock of the SN explosion is converted to GCR energy through diffusive shock acceleration. The process is highly non-linear and involves the dynamical reaction of both the accelerated particles and of the magnetic field on the system. Those studies usually take into account the fact that the SN explosion often occurs within the cavity excavated in the interstellar medium (ISM) by the wind of the massive star prior to the explosion (\\cite{Biermann01}); however, the structure of the circumstellar environment in that case is quite complex and simplified models are used for its description. Although \\cite{Caprioli11b} considered the composition of GCR (H,He, CNO, MgSiAl, Fe) resulting from such an acceleration site, none of those studies considered the \\neo \\ ratio. In this work we study the \\neo \\ ratio of GCR accelerated by the forward shocks of SN explosions, as they run through the presupernova winds of massive stars and through the interstellar medium. We consider the whole mass spectrum of massive stars (from $\\sim$10 to 120 \\ms), including stars with either small or large mass losses prior to their explosions. We consider stellar properties (masses of winds, ejecta, yields etc.) from recent models with mass loss and or without rotation (from Hirschi et al. 2005 and Limongi and Chieffi 2006, respectively), the former having larger \\neb \\ enhancements in their winds. We adopt a simplified prescription (suggested in Ptuskin and Zirakashvili 2005 and reformulated in Caprioli 2011) to describe the structure of the circumstellar medium at the time of the explosion and we consider that GCR start being accelerated in the Sedov-Taylor (ST) phase of the SN remnant (see e.g. Ptuskin et al. 2010). By requiring the resulting IMF averaged \\neo \\ ratio to equal the observed one $R_{Obs}$=(\\neo)$_{GCR}$/(\\neo)$_{\\odot}$=5.3$\\pm$0.3 we are able to constrain the forward shock velocity to values $>$1900 km/s for rotating stars (and to $>$2400 km/s for non rotating ones), i.e. we find that GCR are accelerated during the early ST phase, lasting for a few 10$^ 3$ yr. Assuming that 10\\% of the SN kinetic energy is converted to GCR, we find that during the acceleration period a few particles out of a million encountered by the forward shock are accelerated. Finally, we reassess the superbubble paradigm for the origin of GCR, by evaluating consistently the \\neo \\ ratio with the aforementioned stellar yields. We find that it can not be as high as observed, unless some extremely favorable assumptions are made (only the early period of the superbubble lifetime considered, no gas left over from the formation of the OB association). We conclude that superbubbles cannot be at the origin of the bulk of GCR. The plan of the paper is as follows. In Sec. 2 we present the general \"set-up\" of our model: the adopted stellar models (Sec. 2.2) and their wind yields (Sec. 2.3), the description of the circumstellar environment (Sec. 2.4) and the evolution of the forward shock in the ST phase (Sec. 2.5). In Sec. 3 we present our results for the (time-dependent) composition of the accelerated particles, the limits imposed on the shock velocity by the observed \\neo \\ ratio and the efficiency of the particle acceleration. Finally, in Sec. 4 we explore the \\neo \\ ratio of GCR, assumed to be accelerated inside a superbubble, and we show that it cannot match the oberved one (unless extreme assumptions are made). The results are summarized in Sec. 5. ", "conclusions": "In this work we explore some implications of the idea that cosmic rays are accelerated by the forward shock in supernova remnants, during their ST phase. We focus on the chemical composition resulting from such an acceleration and, in particular, on the \\neo \\ ratio, which is the most characteristic feature of the observed GCR source composition and is unaffected by atomic effects. For that purpose, we adopt recent models of the nucleosynthesis and evolution of massive stars with mass loss: those of LC06 with no rotation and those of HMM05 with rotation. In Sec. 2 we present a detailed summary of the properties of those models and, in particular, of the chemical composition of their winds, insisting on the fact that rotating models release more \\neb \\ in their winds than non-rotating ones. We also present the adopted model for the evolution of a SN remnant within a stellar wind, based on the ideas of Ptuskin and Zirakashvili (2005) and the equations summarized in Caprioli (2011). In the framework of the simple model adopted here, we follow the time-dependent composition of GCR, accelerated by the forward shock as it runs through either the ISM (in the case of stars with M$<$25-35 \\ms, depending on rotation) or through the stellar wind (in the case of more massive stars). In fact, during the largest part of the ST phase, the shock runs through ISM and encounters a $\\sim$solar composition. In order to reproduce the observed high value of (\\neo)$_{CR}$ ($R_{Obs}$=5.3$\\pm$0.3 in solar units) after accounting for the stellar IMF, we have then to assume that acceleration is efficient only during a short early period in the ST phase. We chose to use the shock velocity as a criterion for efficient acceleration and, based on the aforementioned CR composition argument, we find that shock velocities larger than $\\sim$1900 km/s (for the rotating stellar models) or 2400 km/s ( for the non-rotating ones) are required. This result is obtained by assuming a step function for the efficiency $f$ of particle acceleration ($f$=0 before the ST phase and after $\\upsilon_{min}$, and $f$=1 between the two). For the - perhaps, more realistic - assumption of a velocity-dependent efficiency $f\\propto \\upsilon^2$, we find slightly lower values for $\\upsilon_{min}$ (1600 km/s for HMM05 yields and 2150 km/s for CL06 yields, respectively). In the framework of the adopted models, this corresponds to a circumstellar mass of several tens of \\ms \\ encountered by the forward shock. Assuming, furthermore, that 10\\% of the SN kinetic energy is used in acceleration of escaping cosmic rays with standard energy spectra, allows us to evaluate the efficiency of that acceleration: we find that a few particles out of a million encountered by the forward shock are accelerated to CR energies. We also notice that this scheme of GCR acceleration does not suffer from problems related to the absence of unstable $^{59}$Ni in observed GCR composition: this heavy nucleus is well inside the SN ejecta and is not reached by the forward shock which accelerates only wind material and ISM. The aforementioned scenario assumes that even the most massive stars, up to 120 \\ms, develop strong forward shocks and accelerate the particles of their WR winds. For non-rotating stars, this is a rather extreme assumption, since it has been argued (Heger et al. 2003) that non-rotating masssive stars of about solar metallicity collapse into black holes, if their mass is in the 30-60 \\ms \\ range (see their Fig. 1). Notice that stars in the 60-120 \\ms \\ range (the most important \\neb \\ producers) end as black holes in their scheme. However, for slightly higher metallicities - such as those encountered in the inner Galactic disk - they find that only neutron stars are formed, because higher stellar mass losses result in a less massive star at explosion. It should be noticed that the details of massive star explosions remain poorly understood at present (see e.g. Hanke et al. 2011) and so is the fate of a massive star above 30 \\ms \\ (see Fryer et al. 2011 for a recent - but certainly not definitive - assessment). The fate of the rotating mass losing stars considered here is even less well known. In view of the aformentioned uncertainties, we feel that the scenario proposed here can be considered as valid at present, although future refinements in our understanding of massive star explosions may change it quantitatively (and even qualitatively, if it turns out that most masive stars above, say, 50 \\ms, end up as black holes). Finally, we explore the idea that CR are accelerated in superbubbles, in which case their composition results from the ejecta of both stellar winds and SN explosions. We first notice that simple nucleosynthesis arguments suggest that the resulting composition, averaged over the stellar IMF, should be very close to solar. We demonstrate this quantitatively, with a simple model for the evolving composition of a superbubble, enriched first by the (\\neb \\ rich) winds of the most massive stars, then by the (\\nea \\ rich) SN ejecta of less massive stars. We find that, after a few Myr the superbubble (\\neo)$_{SB}$ \\ ratio tends to solar, and so does the average (\\neo)$_{CR}$ \\ ratio in accelerated particles. We conclude that superbubbles cannot provide the observed high $R_{Obs}$ value of CR sources and, therefore, are not the main site of CR acceleration. On the contrary, SN remnants - including those expanding in the pre-explosion environment of a stellar wind - appear as suitable sites of GCR acceleration." }, "1112/1112.3878_arXiv.txt": { "abstract": "We assess the viability of successful reconstruction of the evolution of the dark energy equation of state using varying fundamental couplings, such as the fine structure constant or the proton-to-electron mass ratio. We show that the same evolution of the dark energy equation of state parameter with cosmic time may be associated with arbitrary variations of the fundamental couplings. Various examples of models with the same (different) background evolution and different (the same) time variation of fundamental couplings are studied in the letter. Although we demonstrate that, for a broad family of models, it is possible to redefine the scalar field in such a way that its dynamics is that of a standard quintessence scalar field, in general such redefinition leads to the breakdown of the linear relation between the scalar field and the variation of fundamental couplings. This implies that the assumption of a linear coupling is not sufficient to guarantee a successful reconstruction of the dark energy dynamics and consequently additional model dependent assumptions about the scalar field responsible for the dark energy need to be made. ", "introduction": "Introduction} More than one decade ago type Ia supernovae observations suggested, for the first time, that the expansion of the universe is accelerating \\cite{Perlmutter:1998np,Riess:1998cb}. Since then, increasingly precise cosmological observations \\cite{Percival:2009xn,Komatsu:2010fb,Amanullah:2010vv} led to a well tested cosmological model presently dominated by an exotic dark energy form, violating the strong energy condition. In fact, if General Relativity is valid on large cosmological scales then dark energy \\cite{Ratra:1987rm,Frieman:1995pm,Caldwell:1997ii,Ferreira:1997hj,ArmendarizPicon:2000dh,ArmendarizPicon:2000ah} provides the only convincing explanation for the observed acceleration of the universe. Understanding the nature of dark energy is therefore one of the most important challenges of modern cosmology with one of the primary goals being determining whether its energy density is constant or slowing varying (see \\cite{Copeland:2006wr,Frieman:2008sn,Caldwell:2009ix,Silvestri:2009hh,Li:2011sd} for recent dark energy reviews). A fundamental problem associated to a cosmological constant is that its magnitude is constrained to be much smaller than particle physics predictions. On the other hand, it is also not clear if there is a deep physical reason which explains why it became the dominant component of the Universe just around the present day \\cite{Barreira:2011qi}. An arguably better motivated alternative to the cosmological constant is the possibility that dark energy might be described by a dynamical scalar field. One important parameter characterizing dynamical dark energy is its equation of state, the ratio $w$ between the dark energy pressure and energy density. Constant $w$ models are unrealistic unless $w=-1$, which corresponds to the cosmological constant case \\cite{Avelino:2009ze,Avelino:2011ey}. Hence, a measure of $w \\neq -1$ at any redshift or redshift band should be indicative of dynamical dark energy. Considerable efforts are being put forward to constrain the dynamics of $w$ at low redshifts (see \\cite{EditorialTeam:2011mu} for expected future developments with the Euclid mission) using type Ia supernova, galaxy clustering or weak lensing. These are indirect probes which rely on the impact of dark energy on the overall dynamics of the universe. However, dark energy is expected to become subdominant at early times and, consequently, it is not possible to strongly constrain its dynamics at high redshift using standard methods. Still, in realistic models, dynamical scalar fields may couple to other fields, possibly leading to measurable variations of nature's fundamental \"constants\" \\cite{Carroll:1998zi}. The coupling between a quintessence field and fundamental couplings such as $\\alpha$ or $\\mu$ has been investigated by several authors \\cite{Carroll:1998zi,Chiba:2001er,Wetterich:2002ic,Nunes:2003ff,Anchordoqui:2003ij,Copeland:2003cv,Avelino:2004hu,Parkinson,Doran:2004ek,Marra:2005yt,Avelino:2006gc,Avelino:2008dc,Avelino:2009fd,Dent:2009,Calabrese:2011nf,Amendola:2011qp}. The dynamics of $\\alpha$ over the redshift range $z=0-10^{10}$ is severely constrained using using both cosmological and laboratory experiments (see \\cite{Uzan:2010pm} for a recent review). At low redshifts laboratory experiments \\cite{Peik:2006xy,Rosenband:2008} and the Oklo natural nuclear reactor \\cite{Gould:2006,Petrov:2006,Onegin:2010kq} provide very stringent limits on the time-variation of $\\alpha$ and $\\mu$, while at high redshift cosmic microwave background temperature and polarization anisotropies \\cite{Avelino:2000ea,Avelino:2001nr,Martins:2003pe,Rocha:2003gc, Stefanescu:2007aa,Nakashima:2008cb} and light element abundances \\cite{Bergstrom:1999wm,Avelino:2001nr,Nollett:2002da} constrain the value of $\\alpha$ at $z \\sim 10^{10}$ and $z \\sim 10^3$ to be within a few percent of its present day value. Despite a few positive claims for a detection of a variation of the fine-structure constant $\\alpha$ \\cite{Webb:1998cq,Murphy:2006vs} or the proton-to-electron mass ratio $\\mu$ \\cite{Ivanchik:2005ws,Reinhold:2006zn} in the redshift range $z=1-4$, and the more recent claims for a significant spatial variation of $\\alpha$ \\cite{Webb:2011,King:2012}, there is presently no unambiguous evidence for such variation (see, for example, \\cite{Chand2004,Chand2007,Murphy2008} for some strong negative results). Nevertheless, it has been shown that varying couplings may be used to determine the evolution of the dark energy equation of state over a larger redshift range than standard methods, if a number of conditions are verified \\cite{Avelino:2006gc,Avelino:2009fd}. These are: i) that the dark energy can be described by a standard quintessence field; ii) that the relation between the quintessence field and varying fundamental couplings is linear; iii) that such variations are within reach of forthcoming experiments. In this letter we shall relax assumptions i) and ii) and consider more general k-essence models for dark energy, thus testing the robustness of the varying fundamental couplings method for the reconstruction of the evolution of the dark energy equation of state. Throughout this letter we shall use units with $c=8\\pi G/3=H_0=1$ and a metric signature $(+,-,-,-)$. ", "conclusions": "" }, "1112/1112.4175_arXiv.txt": { "abstract": "{We introduce a new technique for imaging the polarized radio sky using interferometric data. The new approach, which we call Faraday synthesis, combines aperture and rotation measure synthesis imaging and deconvolution into a single algorithm. This has several inherent advantages over the traditional two-step technique, including improved sky plane resolution, fidelity, and dynamic range. In addition, the direct visibility- to Faraday-space imaging approach is a more sound foundation on which to build more sophisticated deconvolution or inference algorithms. For testing purposes, we have implemented a basic Faraday synthesis imaging software package including a three-dimensional CLEAN deconvolution algorithm. We compare the results of this new technique to those of the traditional approach using mock data. We find many artifacts in the images made using the traditional approach that are not present in the Faraday synthesis results. In all, we achieve a higher spatial resolution, an improvement in dynamic range of about 20\\%, and a more accurate reconstruction of low signal to noise source fluxes when using the Faraday synthesis technique.} ", "introduction": "Rotation measure (RM) synthesis, introduced by \\citet{brentjens_faraday_2005}, is a technique that makes use of the Faraday effect to improve the sensitivity of polarimetric observations by combining data over wide ranges in frequency. RM synthesis allows for the separation of polarized sources along the line of sight (LOS) by decomposing the observed polarized emission into parts originating from different Faraday depths (in the simplest case, Faraday depth is equivalent to RM), allowing one to generate a 3D representation of the polarized sky. While the Faraday depth axis cannot be mapped to a physical depth, the Faraday depth information can be of significant scientific value since the Faraday depth traces the projected strength and orientation of magnetic fields along the LOS. A more detailed introduction to RM synthesis is provided later. The RM synthesis technique has only recently become viable, owing to the availability of broadband receivers in the next generation of radio telescopes such as the Expanded Very Large Array (EVLA), the upgraded Westerbork Synthesis Radio Telescope (WSRT), and the pathfinder projects leading to the Square Kilometer Array (SKA) such as the Low-Frequency Array (LOFAR). Recently, there have been many successful applications of RM synthesis, and interest is rapidly increasing as new radio telescopes are being commissioned. Applications have included studies of the diffuse polarized emission in the Perseus field \\citep{de_bruyn_diffuse_2005,brentjens_wide_2010} and the Abell 2255 field \\citep{pizzo_deep_2010}, analysis of the polarized emission in nearby galaxies in the WSRT-SINGS survey \\citep{heald_westerbork_2009}, and the detection of a shell of compressed magnetic fields surrounding a local HI bubble \\citep{wolleben_antisymmetry_2010}. The RM synthesis technique will play a critical role in several upcoming polarization surveys, e.g. POSSUM \\citep{gaensler_possum_2010}, GMIMS \\citep{wolleben_gmims_2009}, and future surveys with LOFAR. RM synthesis is very useful for studying magnetism. For instance, \\citet{bell_caustics_2011} have shown that prominent asymmetric features in RM synthesis images known as Faraday caustics, which are related to LOS magnetic field reversals, can be used to study the structural and statistical properties of magnetic fields. In addition to the considerable interest in applications, there has been recent interest in improving RM synthesis imaging techniques. The RMCLEAN deconvolution algorithm was introduced by \\citet{heald_westerbork_2009} and was quickly adopted, owing to its simplicity and similarity to techniques used in aperture synthesis imaging. \\citet{frick_wavelet-based_2010} have proposed a wavelet-based RM synthesis technique. With this approach, one obtains a decomposition of the size scale of structures in addition to their Faraday depth location. There has also been growing interest in applying compressed sensing to RM synthesis \\citep{li_compressed-sensing-rmsynth_2011,andrecut_sparse-rmsynth_2011}. Compressed sensing is a method of reconstructing signals that are sparse in some set of basis functions. If the signal is sufficiently sparse, it can be reconstructed using fewer measurements than indicated by the Nyquist-Shannon sampling theorem. Since compressed sensing techniques make use of wavelet bases, but are implemented in a noise-aware fashion, they can be expected to be superior to pure wavelet based methods. These new imaging techniques have thus far focused on the problem of one-dimensional (1D) reconstruction. However, the product of RM synthesis is a function not only of Faraday depth, but also of position on the sky, making its reconstruction an inherently three-dimensional (3D) problem. In many cases, RM synthesis is performed on sky brightness images that have been produced from radio interferometric data. Observations performed with a radio interferometer sample the aperture plane rather than the image plane, and this is done at many different frequencies. This data space is the 3D Fourier space representation of the polarized sky brightness as a function of Faraday depth. Imaging algorithms should ideally make use of the entire data space to inform the reconstruction at each pixel, since the information about the sky brightness at each pixel is spread throughout Fourier space. However, the current approach is to perform the imaging in a piecewise fashion, first reconstructing the 2D sky plane images at each frequency before doing 1D RM synthesis imaging along each LOS. Therefore, each step of the traditional imaging approach is done with a limited subset of the data, which will reduce the overall sensitivity and degrade fidelity in the final image. In this paper, we introduce a new technique for imaging the Faraday spectrum directly from radio interferometric data. We call this technique \\emph{Faraday synthesis}. Using this approach, one images the polarized emission as a function of sky position and Faraday depth from the visibility data itself, rather than using the traditional piecewise prescription. Faraday synthesis is a natural extension of the aperture synthesis plus RM synthesis techniques that provides improvements in image fidelity and sensitivity. We note that a similar technique was briefly discussed in \\citet{pen_GMRT-eor_2009}, although it was not considered in any detail, nor was it compared to the traditional approach to RM synthesis imaging. Furthermore, deconvolution was not considered. With the advent of RM synthesis the concept of rotation measure, defined to be the amount that the observed polarization angle changes as a function of frequency, has become somewhat outdated. With RM synthesis, one does not measure RMs, but instead reconstructs the polarized intensity as a function of Faraday depth. In the simplest case, where a single, discrete source of polarized emission is positioned behind a Faraday rotating medium, the RM is equal to the Faraday depth. In all other cases this is not true. In general, RM cannot be used as a proxy for Faraday depth, and the full distribution of polarized brightness as a function of Faraday depth is the most appropriate quantity to study. Therefore, we avoid use of the term RM to describe this new method, and instead call it Faraday synthesis. Throughout the remainder of this paper, RM synthesis will refer to the LOS imaging method developed by \\citet{brentjens_faraday_2005}. The traditional practical approach of first imaging individual frequencies using 2D aperture synthesis techniques and then reconstructing the LOS brightness distribution on a pixel-by-pixel basis will be referred to as 2+1D imaging, in contrast to Faraday synthesis, which we will often refer to as 3D imaging. In Sec. \\ref{sec:Synthesis-imaging} we briefly review the theories of aperture and RM synthesis imaging. In Sec. \\ref{sec:Faraday-synthesis} we introduce the Faraday synthesis imaging technique. In Sec. \\ref{sec:Proof-of-concept} we describe the proof of concept software that we have implemented, and in Sec. \\ref{sec:Tests} we compare test results obtained by imaging mock data using both the 3D and 2+1D techniques. We conclude in Sec. \\ref{sec:Discussions-and-conclusions} with a summary and discussion of our results. ", "conclusions": "} We have described a new approach to imaging linearly polarized visibility data that we call Faraday synthesis. With this approach, one directly reconstructs the Faraday spectrum, or the polarized intensity as a function of Faraday depth, from the visibility data. This takes the place of the usual approach of first performing aperture synthesis imaging on the visibility data at each frequency, then performing RM synthesis along each line of sight independently. These two approaches would be equivalent if deconvolution were not required. With Faraday synthesis, the deconvolution is done in one step using the entirety of the broad-band data. In contrast, the traditional approach requires deconvolving the images individually at each frequency. The sensitivity in the narrow-band images is significantly limited, and residual artifacts remain in these images that limit the dynamic range and image fidelity achieved during RM synthesis. Moreover, another deconvolution procedure is necessary to account for the point spread function due to the incomplete sampling of the wavelength space. Artifacts introduced by the first deconvolution algorithm will be compounded during this procedure, further reducing image fidelity. Indeed, we found in our proof-of-concept testing that artifacts were significantly higher in the traditional imaging method than with Faraday synthesis. The noise was roughly 20\\% lower when using the Faraday synthesis technique, and the strongest artifact was about half as bright. We found that one is able to achieve a better resolution in the final image using the new approach. The main lobe of the 3D dirty beam was 30\\% smaller in the sky plane than that of the traditional method. With the 2+1D method, stacking of the images at each frequency requires tapering the visibility data such that the higher frequency images have the same resolution as those at the lower frequencies. We find that the inaccuracies inherent in the stacking process are a significant source of artifacts in the traditional RM synthesis technique. Furthermore, this procedure requires severely down-weighting a large portion of the available data, again limiting sensitivity. With the Faraday synthesis approach, no such down-weighting is required. The Faraday synthesis approach of working directly between visibility and Faraday space is a much better foundation on which to build new image reconstruction algorithms because with it one is able to accurately describe the instrument response. The effects of the intermediate, non-linear deconvolution procedure can not be easily understood or modeled. Many signal inference techniques, like those derived using Information Field Theory \\citep{ensslin_ift_2009,ensslin_gibbs_2010}, depend on a complete description of the instrument response and would need to be built on the framework of Faraday synthesis. While the CLEAN algorithm has worked quite well in radio astronomy for decades, the implicit assumption of a sky sparsely populated by point sources is not well suited for the kinds of diffuse polarized signals that one finds, for example, in the polarized emission from the Milky Way. Novel image reconstruction algorithms built using more appropriate constriants or assumptions are likely justified. Such algorithms could make use of statistical correlations inferred from the data, similar to the extended critical filter algorithm developed by \\citet{oppermann_extcritfilt_2011}." }, "1112/1112.1420_arXiv.txt": { "abstract": "We carry out a comprehensive smooth particle hydrodynamics simulation survey of double-degenerate white dwarf binary mergers of varying mass combinations in order to establish correspondence between initial conditions and remnant configurations. We find that all but one of our simulation remnants share general properties such as a cold, degenerate core surrounded by a hot disk, while our least massive pair of stars forms only a hot disk. We characterize our remnant configurations by the core mass, the rotational velocity of the core, and the half-mass radius of the disk. We also find that some of our simulations with very massive constituent stars exhibit helium detonations on the surface of the primary star before complete disruption of the secondary. However, these helium detonations are insufficiently energetic to ignite carbon, and so do not lead to prompt carbon detonations. ", "introduction": "Type Ia supernovae are commonly accepted to be the observed transient produced after a thermonuclear detonation inside a white dwarf star. While the preferred mechanism for producing SNeIa involves accretion from an evolved main sequence star onto a white dwarf (Whelan \\& Iben 1973; Nomoto 1982; Hillebrandt \\& Niemeyer 2000), the observed SNeIa rate is incompatible with the narrow range of helium accretion rates that initiate a carbon detonation as opposed to accretion induced collapse or classical novae (Nomoto \\& Kondo 1991; Hardin \\etal 2000; Pain \\etal 2002; Ruiter \\etal 2009). Moreover, many recent observations of abnormally luminous SNeIa have been interpreted as having derived from double-degenerate systems involving two white dwarfs. For example, photometric observations of SN~2007if suggest 1.6$\\pm$0.1\\msol of \\nickel[56] was formed, implying a progenitor mass of 2.4$\\pm$0.2\\msol (Scalzo \\etal 2010), which is well above the Chandrasekhar limit (Chandrasekhar 1931). Spectroscopic observations of SN~2009dc suggest $\\apprge1.2$\\msol of \\nickel[56] (Tanaka \\etal 2010), depending on the assumed dust absorption. Since 0.92\\msol of \\nickel[56] is the greatest yield a Chandrasekhar mass can produce (Khokhlov \\etal 1993), this yield also implies a super-Chandrasekhar progenitor mass. And observations of SN~2003fg by Howell \\etal (2006) and of SN~2006gz by Hicken \\etal (2007) infer $\\sim1.3$\\msol of \\nickel[56] each. Generally, for the purposes of cosmological measurements, obvious double-degenerate candidates are excluded from SNeIa surveys. The Phillips relation, or the width-luminosity relation (WLR), which established SNeIa as standard candles (Phillips 1993), relates the peak luminosity of a SNIa to the change in magnitude after 15 days. The WLR for standard SNeIa indicates that SNeIa with bright peak magnitudes also decay at a slower rate than dimmer SNeIa. This is often thought to be the result of a relationship between the \\nickel[56] yield and the opacity of the ejecta material, assuming a total mass not exceeding the Chandrasekhar mass. However, a double-degenerate system may have up to two times the Chandrasekhar mass, and so the relationship between the \\nickel[56] yield and the ejecta opacity need not be similar to single-degenerate scenarios, and the WLR may not be applicable to these SNeIa. In fact, it is more likely that for a given \\nickel[56] production and energy deposition history, an increased ejecta mass results in an increased opacity, reducing the peak magnitude and broadening the lightcurve in a fashion that is the inverse of the standard WLR (Pinto \\& Eastman 2000; Mazzali \\etal 2001; Mazzali \\& Podsiadlowski 2006; Kasen, R\\\"opke, \\& Woosley 2009). Complicating matters is the possibility that SNeIa deriving from double-degenerate progenitors can have ordinary \\nickel[56] yields, depending on the progenitor mechanism (see \\eg collisional mechanisms in Raskin \\etal 2009, 2010 and Rosswog \\etal 2009) and the final, central densities of the degenerate material before ignition. Thus, double-degenerate SNeIa may be masquerading as typical SNeIa. If they do not conform to the WLR, they may introduce systematic errors into cosmological surveys. In order to reduce the scatter in the Hubble diagram, we must first establish whether double-degenerate SNeIa are standardizable, and if not, we must identify the tell-tale signatures of a double-degenerate progenitor mechanism. The most probable double-degenerate progenitor scenario involves two white dwarfs in a tight binary, though other progenitor systems have been considered (Benz \\etal 1989a; Raskin \\etal 2009; Rosswog \\etal 2009). Binary white dwarf systems were first seriously explored as plausible SNeIa progenitors by Iben \\& Tutukov (1984) and Webbink (1984). In such a system, tidal dissipation and gravitational radiation drive the binary pair into an ever closer orbit. Eventually, the least massive white dwarf, being physically larger as $R\\propto M^{-1/3}$, overflows its Roche lobe and begins to accrete material onto the primary, or more massive companion star. For many mass combinations, this is a fundamentally unstable process in which the loss of mass from the secondary causes it to outgrow its Roche lobe faster than its orbit widens due to conservation of angular momentum (see \\eg Marsh \\etal 2004). Benz \\etal (1990) performed one of the first simulations of double-degenerate mergers, examining a binary system consisting of a 1.2\\msol white dwarf primary and a 0.9\\msol white dwarf secondary. They used a smooth particle hydrodynamics code with 3000 particles per star and found that the merger remnant consisted of a 1.7\\msol core surrounded by a rotationally supported disk. This is more massive than the Chandrasekhar limit, but they concluded that the central object was not entirely degenerate, having been considerably heated, and thus, much of the object's support against gravitational collapse came simply from thermal pressure. Since this pioneering work, others have revisited the white dwarf merger scenario with up-to-date simulation codes and higher resolutions than were possible in the past (see \\eg Yoon \\etal 2007; Lor\\'{e}n-Aguilar \\etal 2009). In this paper, we revisit white dwarf mergers simulations, examining a wide range of possible mass combinations with high resolution, more accurate initial conditions, and up-to-date physics for the equation of state and nuclear reaction network. The structure of this paper is as follows. In \\S2, we outline our methods and initial conditions. We discuss the results of our simulations in \\S3, and compare our work to previous studies of white dwarf mergers in \\S4. Finally, in \\S5, we summarize our results and conclusions and discuss possible avenues for advancing remnant evolution in future studies. ", "conclusions": "Double degenerate progenitor scenarios are drawing new interest among the supernova community. They have the potential to explain some of the confounding mysteries that remain about SNeIa, and to enhance their usefulness as cosmological probes. Here, we have conducted a large survey of white dwarf binary merger models in order to begin to understand how these systems evolve, and to piece together some of the observational signatures of double degenerate SNeIa. % In each of our simulations, the merger remnant consisted of a cold, degenerate core surrounded by a hot accretion disk, with $h/r_d\\approx0.1$. The only exception to this remnant configuration was the simulation of two 0.64\\msol white dwarfs, where the cores of both stars merged at the center of mass, heating the remnant core considerably and lifting most of its degeneracy. Since it is unlikely that two identical mass white dwarfs would form in a binary in nature, this scenario might seem of trivial importance, but the merging of the cores was less the result of their masses being identical than it was of the stars being highly susceptible to tidal disruption. Our other equal mass simulations with more massive constituent stars featured merger scenarios wherein one of the stars was completely disrupted into an accretion disk around the core of the other, provided there was a non-trivial asymmetry between the two. Therefore, it is likely that merger scenarios involving slightly unequal but low-mass white dwarfs might also exhibit core merging, though more simulations are certainly needed to confirm this conjecture. Evolving these meta-stable remnants further will require the implementation of new physics beyond hydrodynamics and nuclear burning. Since the disks are optically thick, radiative losses from the surface are the dominant evolutionary mechanism, and efforts are underway to incorporate this process. As the surface layers lose energy through radiation, the material will sink to lower latitudes, driving convective currents that will increase the disk viscosity. We expect this viscosity will mimic an $\\alpha$-disk prescription, and thus, the accretion times will be of order 100 orbital times. However, the three most massive combinations simulated have disk configurations that suggest the cores might grow to become super-Chandrasekhar masses on shorter timescales, after cooling has removed some of the thermal support for the inner portions of the disks. While none of our simulations exhibited prompt carbon detonations, all of the simulations that included a 1.06\\msol primary did exhibit prompt helium detonations on the surface of the primary. These helium detonations were not sufficiently energetic to significantly burn much of the carbon or to unbind either of the constituent stars, but the detonation shocks they produced did alter the structure of the primary before complete merger. Moreover, at $\\sim10^{49}$ ergs, the energy these detonations released is likely sufficient to be observable by many of the upcoming transient surveys, such as LSST. Whether white dwarf mergers produce SNeIa is still an open question. Our simulations and others' represent only the beginnings of our exploration of this progenitor mechanism. Preliminary results look very promising, and it is doubtless that the viability of mergers as SNeIa progenitors will be established in the near term. Meanwhile, it remains an exciting time to be exploring these dynamical scenarios that continue to surprise us." }, "1112/1112.5035_arXiv.txt": { "abstract": "{The largest uncertainty for cosmological studies using clusters of galaxies is introduced by our limited knowledge of the statistics of galaxy cluster structure, and of the scaling relations between observables and cluster mass. A large effort is therefore undertaken to compile global galaxy cluster properties in particular obtained through X-ray observations and to study their scaling relations. However, the scaling schemes used in the literature differ.} {The present paper aims to clarify this situation by providing a thorough review of the scaling laws within the standard model of large-scale structure growth and to discus various steps of practical approximations.} {We derive the scaling laws for X-ray observables and cluster mass within the pure gravitational structure growth scenario. Using N-body simulations we test the recent formation approximation used in earlier analytic approaches which involves a redshift dependent overdensity parameter. We find this approximation less precise than the use of a fiducial radius based on a fixed overdensity with respect to critical density.} {Inspired by the comparison of the predicted scaling relations with observations we propose a first order modification of the scaling scheme to include the observed effects of hydrodynamics in structure formation. This modification involves a cluster mass dependent gas mass fraction. We also discuss the observational results of the reshift evolution of the most important scaling relations and find that also a redshift dependence of the gas mass to total mass relation has to be invoked within our modification scheme.} {We find that the current observational data are within their uncertainties consistent with the proposed modified scaling laws.} ", "introduction": "Galaxy clusters form from overdense regions in the large-scale matter distribution, which have small amplitudes at early epochs and only collapsed to objects very recently. In the standard cosmological model the large-scale matter distribution is described by a random Gaussian field characterized by a power spectrum with a smoothly changing power law index over relevant length scales. This implies that the structure evolution will feature a large degree of self-similarity in scale and time (e.g. Peebles 1980). Consequently galaxy clusters, which form an integral part of this large-scale structure, also show an imprint of this general self-similarity. This connection between the framework of the evolution of the gravitating matter on large scales and galaxy cluster formation and their observed structure was realized in early studies, e.g. by Gunn \\& Gott (1972), Fillmore \\& Goldreich (1984), Bertschinger (1985), Hoffman \\& Shaham (1985) and in simulations e.g. Frenk et al. (1985), Zureck, Quinn \\& Solomon 1988, Efstathiou et al. 1988, West, Dekel \\& Oemler 1987). It resulted in a comprehensive description of the structure of dark matter halos, of which galaxy clusters are the most massive representatives, in a series of papers by Navarro, Frenk \\& White (1995, 1996, 1997) and follow-up literature. In this picture of purely dark matter structure growth, dark matter halos form a nearly self-similar, two-parameter family, with the two parameters being mass and a concentration or time-of-formation parameter. This structural model of clusters describes an average behavior of the cluster population, where the different statistical realizations of mass distributions in the protoclusters produce a significant scatter in the observed structural parameters around this mean. Deviations from the equilibrium state after merger events further contribute to this scatter. The addition of baryons to this model leads to a modification of this picture, which for clusters can be seen as a perturbation of the dark matter structure evolution. In this sense galaxy clusters mark the very interesting transition region, where a first order description involving only the large-scale structure gravitational physics provides a very effective guideline and the more complicated hydrodynamics, including radiative cooling and feedback from star formation as well as AGN activity, constitutes a perturbative refinement. At smaller scales, for galaxies the gaseous astrophysics acting on small scales becomes dominant for the appearance of the visible objects and the observed evolution of the large-scale structure on galaxy scales becomes very non-linear. It is therefore on galaxy cluster scales where we can still very successfully apply analytical descriptions as a useful guideline for the understanding of structure evolution. The paper builds on this property of galaxy clusters. X-ray observations are currently providing the most detailed account of galaxy cluster structure and are consequently used extensively to test the predictions of the large-scale structure growth models. However, they do not directly provide a picture of the dark matter halo distribution, but the distribution of the hot intracluster medium (ICM) that fills the entire cluster volume and radiates in X-rays. Therefore the X-ray appearance of clusters includes aspects of the hydrodynamics how the gas reacts to the dark matter density distribution and how the ICM evolves in its thermodynamic properties (e.g. Voit 2005). One can use it as a tracer of the dark matter distribution, for example through the assumption that it is located in the dark matter potentials in hydrostatic equilibrium. We can thus expect that the X-ray appearance of galaxy clusters is featuring some hydrodynamic modification compared to the more readily described dark matter distribution. Because the galaxy cluster formation is so tightly connected to the large-scale structure evolution and the fact that there is a well described self-similar structure statistics in the purely gravitational cluster formation model (Navarro et al. 1997) we can expect that there are simple analytically derivable scaling relations for the global X-ray observables as a function of cluster mass. These relation have been studied already early by e.g. Kaiser (1986), and Evrard \\& Henry (1991) and this theoretical work has been supported by simulations (e.g. Bryan \\& Norman 1998, Borgani 2004, Kravtsov et al. 2006, Evrard et al. 2008, Stanek et al. 2010, Short et al. 2010, Borgani \\& Kravtsov 2010). With the event of detailed observational studies of cluster structure in X-rays by means of the advanced X-ray observatories {\\sl Chandra} and {\\sl XMM-Newton}, large sets of observational data on cluster structure and scaling relations have become available now and the detailed testing of the theoretical predictions for the scaling laws is in full swing (e.g. Markevitch et al. 1998, Arnaud \\& Evrard 1999, Mohr \\& Evrard 1997, Finoguenov et al. 2001, Ikebe et al. 2002, Reiprich \\& B\\\"ohringer 2002, Ponman et al. 2003, Ettori et al. 2004, Vikhlinin et al. 2005, Pointecouteau et al. 2005, Arnaud et al. 2005, Pratt et al. 2006, Kotov \\& Vikhlinin 2006, Zhang et al. 2006, Maughan et al. 2006, Maughan 2007, Arnaud et al. 2007, Pratt et al. 2009, Mantz et al. 2010, Arnaud et al. 2010, Sun et al. 2011, Reichert et al. 2011). An investigation of the relevant literature shows, however, that a number different methods are used for the scaling of the data at different redshifts. The aim of this paper is therefore to critically review these methods and to determine the best approach based on comparison with simulations and observations. In the above mentioned literature mostly analytical formulations of the scaling relations have been used, based on general considerations of structure formation. In order to provide the ground for higher precision in the analysis of the evolution of cluster structure, the simplifications made in the analytical models should be replaced by tests and calibrations with N-body simulations. This situation can be compared to that of the theoretical prediction of the dark matter halo (galaxy cluster) mass function, where the analytical model by Press and Schechter (1974) has paved the way for the general formulation of the solution for the mass function, but the actual formulae now applied, are the result of careful calibration with N-body simulations (e.g. Jenkins et al. 2001, Evrard et al. 2002, Warren et al. 2006, Tinker et al. 2008). Here we adopt a similar approach. We first present the theoretical framework for the description of the evolution of the scaling relations in the classical form based on the assumption of the recent formation approximation and compare it to an alternative scheme used in the literature. We then resort to the results of N-body simulations to test the predictions of the analytical approaches and discuss which of the presently used methods in the literature is best. In the second part of the paper we compare the theoretical scaling relations to observations, inspect the deviations of the observed scaling relations from the predictions based on dark matter structure evolution (often called ''gravitational scaling relations''), and discuss these deviations in the context of the influence of hydrodynamical processes. We then explore a simply empirical modification scheme of the scaling relation using a mass dependent depletion factor for the ICM gas to account for the hydrodynamical scaling effects and compare the so obtained set of scaling relations to observations. In a last step we consider how this modification should depend on redshift to be consistent with the observational data. The paper is structured as follows. In section 2 we derive the ''gravitational scaling relations'' based on dark matter structure evolution with the assumption that the baryonic matter follows the dark matter. In section 3 we investigate the dependence of these scaling relations on the redshift dependent overdensity parameter in a $\\Lambda$CDM cosmology and test this model against the method using a fixed overdensity parameter in section 4 using numerical simulations. In section 5 we discuss the redshift dependence of the overdensity parameter, which defines the proper fiducial radius of the clusters, in the context of numerical studies of the redshift dependence of the concentration parameter of galaxy clusters. Section 6 then starts the second part of the paper where we discuss the modification of the scaling relations to include hydrodynamical effects. After a comparison of the description of the evolution of the scaling relations in terms of the parameters E(z) and (1+z) in section 7 and a brief comparison with some recent simulations in section 8, we provide a comprehensive comparison of scaling relation results in the literature with the model predictions in section 9. Finally section 10 contains a discussion and conclusions. ", "conclusions": "Studying in detail the evolution of the dark matter mass density profiles of simulated galaxy clusters, we have shown that the older model of self-similar scaling relations based on the recent formation approximation which uses a scenario where the fiducial overdensity radius, $r_{\\Delta(z)}$, is taken to be redshift dependent, is not accurate and a scaling with a fixed overdensity provides a better and currently sufficiently precise description of self-similar evolution. For a precise analysis of future cosmological surveys of the galaxy cluster population, we should improve this description further. We plan to do this with larger N-body/hydrodynamical simulations which are performed at present and therefore give no detailed recipes for the scaling corrections in this paper. The corrections as shown in Fig. 5 depend on the cosmological model used and therefore these corrections will have to be worked separately for each model case studied. Studying the different types of scaling relations involving parameters derived from X-rays, we can distinguish two types of scaling behavior: the mass - temperature relation is mostly dependent on the scaling of the dark matter potentials and is therefore very close to the gravitational scaling prediction. Most other relations involving gas density or gas mass are affected by the non-constancy of the gas mass to total mass ratio. With the introduction of modified scaling relations to take this hydrodynamical effects into account, we can describe the currently available data sets within the given uncertainties. Looking at the evolution of the scaling relations with redshift, we find an analogous situation: the $M-T$ relation corresponds within the current uncertainties to the prediction of the gravitational self-similar scenario. All other relations involving parameters which depend on the gas density show deviations, which implies that the gas mass fraction is not constant for given cluster mass with redshift. This is just the consequence of the following effects. At higher redshifts clusters of given mass are more compact and the ICM has to be squeezed into a deeper and narrower potential. Since in preheating models, which seem to explain the data best, the gas starts out with an elevated entropy before cluster formation is complete, the gas is less tightly squeezed into the earlier, narrower potentials than into the later wider potentials. Our modified scaling relation model does not describe all observational effects. As shown in Pratt et al. (2010), the entropy scaling depends on the radius at which the entropy is measured. These results imply that the ICM depletion is larger in the center of groups and clusters than in the outer parts. More data are required that extend out to large cluster radii (to $r_{500}$ and beyond) to substantiate this result. The redshift evolution of the $L-T$ scaling relation, which is predicted to be $\\propto E(z)$ in the simple self-similar scenario, is now found in recent studies to be much less positive or even negative (e.g. O'Hara et al. 2007, Reichert et al. 2011). This has the important consequence that one will find less high redshift galaxy clusters in future X-ray and SZE surveys, than predicted based on the simple scaling models (Reichert et al. 2011). More importantly, a precise measurement of this evolution effect is crucial for using the future X-ray survey data on galaxy clusters for the test of cosmological models. As shown by the comparison of the evolution of the scaling relations compiled by Reichert et al. (2011) and the simulations by Short et al. (2010), the study of the evolution of the scaling relations also provides important insight into the astrophysics of the ICM. Currently the observational data strongly favour a model with ealry preheating of the ICM. In the coming years both the observational data as well as the simulation results will experience further strong improvements. Therefore we see the importance of this paper more in elucidating the way how the scaling relations should be analysed and applied, rather than already providing the best parameterization of the results." }, "1112/1112.0756_arXiv.txt": { "abstract": "Following the discovery of a new radio component right before the GeV $\\gamma$-ray detection since 2008 August by {\\it Fermi} Gamma-ray Space Telescope, we present a detailed study of the kinematics and lightcurve on the central sub-pc scale of 3C~84 using the archival VLBA 43-GHz data covering the period between 2002 January to 2008 November. We find that the new component ``C3'', previously reported by the observations with the VLBI Exploration of Radio Astrometry (VERA), was already formed in 2003. The flux density of C3 increases moderately until 2008, and then it becomes brighter rapidly after 2008. The radio core, C1, also shows a similar trend. The apparent speed of C3 with reference to the core C1 shows moderate acceleration from $0.10c$ to $0.47c$ between 2003 November to 2008 November, but is still sub-relativistic. We further try to fit the observed broadband spectrum by the one-zone synchrotron self-Compton (SSC) model using the measured apparent speed of C3. The fit can reproduce the observed $\\gamma$-ray emission, but does not agree with the observed radio spectral index between 22 and 43~GHz. ", "introduction": "\\label{sec:intro} The radio source 3C~84 is associated with the giant elliptical galaxy NGC~1275 ($z=0.0176$). It is well known that 3C~84 has a pair of compact radio lobe structures on the central 10-pc scale \\citep{1994ApJ...430L..41V, 1994ApJ...430L..45W, 2000ApJ...530..233W, 2006PASJ...58..261A}. Because of its proximity, it allows us to study the region within the central sub-pc scale using a Very Long Baseline Interferometer (VLBI). It is an ideal laboratory to investigate the formation mechanism of relativistic jet ultimately powered by super-massive black holes and the interaction between the jets and ambient matter in the galactic central regions. Thanks to the recent observations with {\\it Fermi}/LAT, we have a new opportunity to explore GeV $\\gamma$-ray production mechanism in misaligned radio-loud AGNs \\citep{2010ApJ...720..912A}. With CGRO/EGRET, a few extragalactic radio galaxies, such as Centaurus A \\citep{1999APh....11..221S}, 3C~111 \\citep{2008ApJ...688..852H}, and NGC~6251 \\citep{2002ApJ...574..693M} had been already detected. However, the detection of NGC~1275 \\citep{2009ApJ...699...31A} by {\\it Fermi}/LAT in 2008 August is particularly noteworthy because NGC~1275 was not detected by CGRO/EGRET \\citep{2003ApJ...588..155R}. The flux density detected by {\\it Fermi}/LAT is about 7 times higher than the upper limit of EGRET sensitivity. Intriguingly, the radio monitoring also shows the flux increase starting in 2005. The time variation of $\\gamma$-ray flux density shows a similar trend with the radio flux density on the timescale of decades, implying the possible connection between the $\\gamma$-ray emission and variable radio component. Overall spectral energy distribution (SED) of NGC~1275 from radio to $\\gamma$-ray can be explained by the synchrotron-self Compton (SSC) model and the deceleration jet model \\citep{2009ApJ...699...31A} adopting sub-pc for the size of emitting region. The authors derived Lorentz factors of the emitting region $\\Gamma=1.8$ ($\\theta=25$ deg, $\\delta=2.3$) for the case of SSC model and $\\Gamma$ varying from $10$ to $2$ ($\\theta=20$ deg, $\\delta$ varying 1.6 to 2.7) for the decelerating model, where $\\theta$ and $\\delta$ are the jet angle to the line of sight and the beaming factor of the emitting region, respectively. Here we should stress that the size of the $\\gamma$-ray emitting region adopted in the SED models is comparable to a VLBI component size. Therefore, it is essential to test the scenario of co-spatiality of GeV $\\gamma$-ray and radio emitting regions by VLBI observations. In order to find the radio counterpart of the GeV $\\gamma$-ray emitting region, we have conducted VLBI observations using the VLBI Exploration of Radio Astrometry (VERA) at 22~GHz between 2006 June 14 and 2009 April 24 (\\citet{2010PASJ...62L..11N}, hereafter Paper I). Surprisingly, we found that the monotonic increase of radio flux density mainly originated in the newly born bright component C3 (Paper I). The measured projected speed of C3, however, was $(0.23 \\pm 0.01)c$ on average between 2007 October 24 and 2009 April 24 in the central sub-pc to pc scale. Given previously estimated jet angle to the line of sight ($11^{\\circ}$ - $55^{\\circ}$: \\citet{2006PASJ...58..261A, 2009AJ....138.1874L, 2009ApJ...699...31A}), the de-projected speed corresponds to 0.24$c$ - 0.55$c$, i.e. slower than the speed of jet derived from the SED modeling \\citep{2009ApJ...699...31A}. The result of VERA observation implies $\\delta < 2$, and there seems to be a discrepancy between the SED models and VLBI observations. A possible idea to explain this discrepancy is that a jet component with relativistic speed other than C3, located much closer to the central black hole (within the central sub-pc), is responsible for the GeV $\\gamma $-ray emission. If such a relativistic motion in the jet of 3C~84 occurs only in vicinity of the black hole, the speed of C3 could have been relativistic soon after the ejection from the core, and then it underwent to a rapid deceleration. \\citet{2010A&A...516A...1L} also suggested that jets are relativistic in bright cluster galaxies but they decelerate very soon (sub-parsec scale) because of a strong interaction with the ISM. In Paper I, we argued possible detection of the relativistic speed at the early stage of emergence of C3. However, this was not conclusive because of (1) the lack of spatial resolution of VERA at 22~GHz and (2) possible screening by free-free absorption. In order to overcome the above potential problems, higher resolution observations at higher frequency are required. We explore the possible relativistic flow of C3 at the early stage of its emergence from the core with the archival VLBA data at 43~GHz. We selected only data obtained from 2002 to 2008, i.e. considering earlier epochs than those discussed in Paper I. We also discuss the SED fit using one-zone SSC model involving the actual measured apparent motion of C3 and possible ideas for incorporating the VLBI observation with GeV $\\gamma $-ray emission. Throughout this paper, we adopt the following cosmological parameters; $H_{0}=71$~km sec$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\mathrm{M}}=0.27$, and $\\Omega_{\\mathrm{\\Lambda}}=0.73$ (1~mas = 0.353~pc, and 0.1~mas~yr$^{-1}$ = $0.113c$). ", "conclusions": "In this section, we try to reproduce the broadband flux from $\\gamma$-ray emitting region of 3C~84 based on the quantities of C3 obtained as in \\S \\ref{sec:results}. In Figure \\ref{fig:SED}, we show the observed SED of 3C~84 and one-zone SSC model fit to the observed SED using the size, flux, and $\\beta_{\\rm app}$ of C3 component measured by the VLBA observations. The flux data are the same as the ones in \\citet{2009ApJ...699...31A} but we add the 43-GHz flux of C3 measured on 2008 August 27. We assume that $\\gamma$-ray and VLBA 43-GHz flux are emitted from the same sub-pc region. Optical data (after the extraction of host galaxy contribution; c.f., \\citet{2009ApJ...699...31A}) is also used as flux upper limit of C3. The $\\beta _{\\rm app}$ at 2008 August is $0.44$ (in Figure \\ref{fig:gamma}). Assuming the jet viewing angle as $25^{\\circ}$ (\\citet{2009ApJ...699...31A}), the intrinsic jet speed $\\beta$ can be estimated as $0.54$, which corresponds to $\\Gamma=1.19$ ($\\delta=1.65$). The Gaussian-fitted model size of C3 is $\\sim 0.2$-$0.3~{\\rm mas}$. To fit the 43-GHz flux, the size of $\\gamma$-ray emitting region is adopted as $1.45 \\times 10^{17}~{\\rm cm}~{\\rm(0.133~mas)}$ which is slightly smaller than the C3 model size. The detailed numerical treatment of synchrotron and inverse Compton scattering processes is shown in \\citet{2002ApJ...564...97K} and references therein. If we adopt $\\gamma_{\\rm min} \\sim 10^{3}$ as is the case in \\citet{2009ApJ...699...31A}, 43~GHz lies below the frequency of low energy cutoff and therefore the spectral index should be $\\alpha _{_K}^{_Q} = 1/3$ ($\\nu f_{\\nu} \\propto 4/3$). This is inconsistent with the observed spectral index between 22 and 43~GHz $\\alpha _{_K}^{_Q} = -0.87\\pm0.3$ as shown in Figure \\ref{fig:spectrum}. Then, we adopt $\\gamma_{\\rm min} = 1$ to raise the cutoff frequency. This fit is indicated by a solid line in Figure \\ref{fig:SED}. The model spectrum requires an extremely hard electron energy index $s=1.2$ occasionally seen in TeV blazars (e.g., \\citet{2000ApJ...528..243K}) if we try to reproduce the flux data at other wavelengths. The synchrotron opacity becomes unity at $\\sim 30$~GHz, and the resultant model spectrum shows $\\alpha _{_K}^{_Q} \\sim -0.05$, which is still inconsistent with the observed spectral index. Hence it seems difficult to attribute the broadband SED to the one-zone SSC emission from C3. This indicates that some other components C1 and/or C2 also should be taken in account for the candidates for $\\gamma$-ray emitting region. Following the detection of 3C~84 by {\\it Fermi}/LAT in 2008, we discussed the kinematics of the newly formed jet component C3 using our VERA radio observations for the period 2006 June - 2009 April in Paper I. In this paper, we further explored the kinematics and lightcurves of C1, C2, and C3 using archival 43-GHz VLBA data spanning a time interval from 2002 January to 2008 November, i.e. before the VERA observations. Summary and discussions are as follows: \\begin{enumerate} \\item In the multi-epoch high resolution images obtained by VLBA at 43~GHz (Figure \\ref{fig:3C84_Q_all}), we find that C3 has been visible since 2003 November. We succeed in obtaining a strong constraint on the time of the C3 birth before 2003 November. \\item The lightcurve of VLBA at 43~GHz shows a trend similar to that of Mets\\\"{a}hovi at 37~GHz, indicating that the increase of radio flux density arises from the region within the central $\\sim1$~pc. In particular, the rapid increase of flux density has been seen since the middle of 2008. Both C1 and C3 show the largest change of flux density for this flare. This indicates that the radio flare in the central $\\sim1$~pc region originates in both C1 and C3. The flux density increase of C1 was not prominent in Paper I, possibly because of large optical depth at 22~GHz. \\item The apparent speed of C3 with respect to C1 changes from $(0.10 \\pm 0.05)c$ to $(0.47 \\pm 0.08)c$ between 2003 November 20 and 2008 November 27 with acceleration rate $\\dot{\\beta }_{\\rm app} = 0.08 \\pm 0.02 $ yr$^{-1}$. No superluminal motion is detected with VLBA observations at 43~GHz before 2008 November. \\item The one-zone SSC model using measured apparent speed of C3 is fitted to the observed broadband spectrum, in particular, the 43-GHz and GeV $\\gamma $-ray fluxes. Our model fit is difficult to reproduce the optically-thin radio spectrum $\\alpha _{_K}^{_Q} \\sim -0.9$ that is measured by the VLBA observations. This indicates that it seems difficult to attribute the broadband SED to the one-zone SSC emission from C3. Therefore, the other components should not be excluded as candidates for $\\gamma$-ray emitting regions. \\end{enumerate}" }, "1112/1112.6084.txt": { "abstract": "We study the effect of primordial magnetic fields (PMFs) on the anisotropies of the cosmic microwave background (CMB). We assume the spectrum of PMFs is described by log-normal distribution which has a characteristic scale, rather than power-law spectrum. This scale is expected to reflect the generation mechanisms and our analysis is complementary to previous studies with power-law spectrum. We calculate power spectra of energy density and Lorentz force of the log-normal PMFs, and then calculate CMB temperature and polarization angular power spectra from scalar, vector, and tensor modes of perturbations generated from such PMFs. By comparing these spectra with WMAP7, QUaD, CBI, Boomerang, and ACBAR data sets, we find that the current CMB data set places the strongest constraint at $k\\simeq 10^{-2.5}$ Mpc$^{-1}$ with the upper limit $B\\lesssim 3$ nG. ", "introduction": "Introduction} %_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/ %_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/ %_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/ Many researchers have studied magnetic fields for wider ranges in the early Universe. Recently, magnetic fields with strength $\\sim 1 \\mu$ G have been observed in clusters of galaxies, and primordial magnetic fields (PMFs) have been studied by many authors to explain their origins. The strength of PMFs is constrained as $B\\lesssim 1$ nG from cosmological observations, such as temperature and polarization anisotropies of cosmic microwave background (CMB) and matter power spectra \\cite{Subramanian:1998fn, Mack:2001gc, Subramanian:2002nh, Lewis:2004ef, Yamazaki:2004vq, Kahniashvili:2005xe, Challinor:2005ye, Dolgov:2005ti, Gopal:2005sg, Yamazaki:2005yd, Kahniashvili:2006hy, Yamazaki:2006mi, Yamazaki:2006bq, Yamazaki:2006ah, Giovannini:2006kc, Yamazaki:2007oc, Paoletti:2008ck, Finelli:2008xh, Yamazaki:2008bb, 2008nuco.confE.239Y, Sethi:2008eq, Kojima:2008rf, 2008PhRvD..78f3012K, Giovannini:2008aa, Yamazaki:2009na, Shaw:2009nf, 2010PhRvD..81b3008Y, 2010PhRvD..81j3519Y, Yamazaki2010aa}. Many theoretical models have been proposed for generating PMFs of cosmological scales. Generation models of PMFs from inflation can create strong fields, whose amplitude is of the order of $1$ nG at redshift $z\\sim 0$, depending on assumed hypothetical fields and couplings \\cite{Turner:1987bw,Ratra:1991bn,Bamba:2004cu,2004PhRvD..69d3507B, 2007JCAP...02..030B}. Furthermore, PMFs can be generated by primordial perturbations of density fields \\cite{Takahashi:2005nd,ichiki:2006sc}, the Beiermann battery mechanism in primordial supernovae remnants \\cite{Hanayama:2005hd} or the Weibel instabilities \\cite{2005MNRAS.364..247F}. These PMFs can evolve into the magnetic fields observed in galaxies and/or galaxy clusters directly or through the dynamo process. While magnetogenesis during inflation can produce PMFs beyond the horizon scale, PMFs generated by the other mechanism are expected to have the characteristic scale because they are based on causal processes. Here it should be noted that the coherence length of PMFs can grow after generation due to inverse cascade \\cite{2001PhRvE..64e6405C,Banerjee:2004df}, although the efficiency is still under study. In the previous studies to put constraints on PMFs from CMB anisotropies, the spectrum of PMFs was assumed to be power-law shape. Although power-law spectrum is natural for magnetogenesis during inflation, this is not the case for the other mechanisms based on causal processes. Furthermore, in previous studies with power-law PMFs, it was not so clear which scale of PMFs mainly contributes to the constraints. Since inflationary mechanisms tend to generate power-law magnetic fields, the magnetic fields which have a characteristic scale at cosmological scales are difficult to be produced and seem somewhat artificial. However, it is still possible that such fields can be constrained by observation, and if observed it would give us useful information about generation mechanism of cosmological magnetic fields. Hence, in this article, as a toy example we use a log-normal distribution (LND) for the PMF spectrum $f_\\mathrm{LND}(k;k_\\mathrm{M},\\sigma_\\mathrm{M})$ which has a characteristic scale expressed by $k_\\mathrm{M}$, and aim to constrain PMFs scale by scale instead of the power-law PMF spectrum. These quantities reflect the generation mechanism but here we regard them just as parameters. We study features of angular power spectra of the CMB from the PMFs with log-normal distribution. Finally we constrain the strength of the PMFs for fixed sets of the parameters. %_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/ %_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/ ", "conclusions": "%_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/ %_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/ %_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/_/ It is natural to assume that the power spectrum of the PMF from inflation is the power law. However, when we consider PMF generated by causal mechanisms, it would be appropriate to assume a spectrum with a characteristic scale. In this paper, we assumed the log-normal distribution as the PMF spectrum and studied the effects of the PMFs on the temperature and polarization anisotropies of the CMB. We have revealed the features of the LND-PMF effects on the CMB as follows: (1) For larger $k_\\mathrm{M}$ and/or $\\sigma_\\mathrm{M}$, the CMB spectra of TT, TE and EE from the LND-PMF are dominated by the vector mode. Meanwhile, in the opposite case, these spectra are dominated by the scalar mode. (2) The CMB spectrum of the BB mode for all scales and other modes for smaller scales is dominated by the vector modes. Because three parameters which characterize the LND-PMF affect the CMB power spectra differently at small and large scales we expect that tight constraints can be placed on these parameters without degeneracy. We found that the LND-PMF which generates CMB anisotropies among 1000 $<\\ell<$ 2000 is most effectively constrained by the current CMB data sets. For example, $B_\\mathrm{LND}$ for $10^{-3}\\mathrm{Mpc}^{-1}~\\le~k/h~\\le 10^{-2}\\mathrm{Mpc}^{-1}$ and $\\sigma_\\mathrm{M} = 1.0$ is limited most strongly as shown in Fig.\\ref{fig6}. In the near future, the tighter constraints on $B_\\mathrm{LND}$ at $k_\\mathrm{M}/h>10^{-2}$Mpc$^{-1}$ will be expected from the observations, such as the Planck, QUIET, and PolarBear missions. \\begin{figure} \\includegraphics[width=1.0\\textwidth]{fig2}% Here is how to import EPS art \\caption{\\label{fig2} CMB spectra from the LND-PMF at $\\sigma_\\mathrm{M}=1.0$ and $B_\\mathrm{LND} = $10nG. TT, EE, BB and TE in each panel indicate the temperature auto-correlation, the E-mode auto-correlation, the B-mode auto-correlation and the temperature E-mode cross correlation, respectively Curves in all panels are theoretical lines as indicated in the legend box at the left bottom. Curves of all TE modes are plotted in the absolute value. } \\end{figure} \\begin{figure} \\includegraphics[width=1.0\\textwidth]{fig3}% Here is how to import EPS art \\caption{\\label{fig3} CMB spectra from the LND-PMF at $k_\\mathrm{M}=10^{-3}$ and $B_\\mathrm{LND} = $10nG. Curves in all panels are theoretical lines as indicated in the legend box at the left bottom. Curves of all TE modes are plotted in the absolute value. } \\end{figure} \\begin{figure} \\includegraphics[width=0.8\\textwidth]{fig1}% Here is how to import EPS art \\caption{\\label{fig1} Power spectrum of the energy density of the PMF. Panels (a) and (c) on this figure are for the fixed $\\sigma_\\mathrm{M}$ value at $\\sigma_\\mathrm{M}=1.0$, and panels (b) and (d) are for the fixed $k_\\mathrm{M}$ value at $k_\\mathrm{M}=10^{-3}$. Curves in different colors in all panels correspond to the different PMF parameters as indicated in the legend box in the figure. Note that from Eqs.(\\ref{p:se})-(\\ref{p:t}) we see that the other spectra, such as magnetic tension, anisotropic stress and so on, have very similar shape to that of the energy density so we only show the spectrum of the energy density in the figure.} \\end{figure} \\begin{figure} \\includegraphics[width=1.0\\textwidth]{fig4}% Here is how to import EPS art \\caption{\\label{fig4} Comparison of total CMB spectra from the LND-PMF at $\\sigma_\\mathrm{M}=1.0$ and $B_\\mathrm{LND} = $10nG with the best-fitting CMB power spectra without the LND-PMF (dashed line). Curves in all panels are the theoretical lines as indicated in the legend box on the figure. Curves of the TE mode are plotted in the absolute value. } \\end{figure} \\begin{figure} \\includegraphics[width=1.0\\textwidth]{fig5}% Here is how to import EPS art \\caption{\\label{fig5} Comparison of total CMB spectra from the LND-PMF at $k_\\mathrm{M}=10^{-3}$ and $B_\\mathrm{LND} = $10nG with the best-fitting CMB power spectra without the LND-PMF (dashed line). Curves in all panels are the theoretical lines as indicated in the legend box on the figure. Curves of the TE mode are plotted in the absolute value. } \\end{figure} \\begin{figure} \\includegraphics[width=1.0\\textwidth]{fig6}% Here is how to import EPS art \\caption{\\label{fig6} Constraint on the strengths of LND-PMF for $10^{-5}< k_\\mathrm{M} <10^{-1}$. The bold curve is the 2 $\\sigma$ upper limits of $B_\\mathrm{LND}$ [nG]. We fix the standard cosmological parameters and use the best-fitted value from WMAP 7th + tensor mode\\cite{2011ApJS..192...16L}. } \\end{figure}" }, "1112/1112.3033_arXiv.txt": { "abstract": "Mapping one-dimensional stellar profiles onto multidimensional grids as initial conditions for hydrodynamics calculations can lead to numerical artifacts, one of the most severe of which is the violation of conservation laws for physical quantities such as energy and mass. Here we introduce a numerical scheme for mapping one-dimensional spherically-symmetric data onto multidimensional meshes so that these physical quantities are conserved. We validate our scheme by porting a realistic 1D Lagrangian stellar profile to the new multidimensional Eulerian hydro code \\CASTRO. Our results show that all important features in the profiles are reproduced on the new grid and that conservation laws are enforced at all resolutions after mapping. ", "introduction": "Multidimensional simulations shed light on how fluid instabilities arising in supernova explosions mix ejecta \\cite{herant1994,candace2009,candace2010,candace2011}. Unfortunately, computing the full self-consistent three-dimensional (3D) stellar evolution initial models for the explosion setup is still beyond the realm of contemporary computational power. One alternative is to first evolve the main sequence star in a 1D stellar evolution code in which the equations of momentum, energy and mass are solved on a spherically symmetric Lagrangian grid, such as \\KEPLER{} \\cite{kepler} or \\MESA{} \\cite{mesa}. Once the star reaches the pre-supernova phase, its 1D profiles can then be mapped into multidimensional hydro codes such as \\CASTRO{} \\cite{zhang2011,ann2010} or \\FLASH{} \\cite{flash} and continue to be evolved until the star explodes. Differences between codes in dimensionality and coordinate mesh can lead to numerical issues such as violation of conservation of mass and energy when profiles are mapped from one code to another. A first, simple approach could be to initialize multidimensional grids by linear interpolation from corresponding mesh points on the 1D profiles. However, linear interpolation becomes invalid when the new grid fails to resolve critical features in the original profile such as the inner core of a star. This is especially true when porting profiles from 1D Lagrangian codes, which can easily resolve very small spatial features in mass coordinate, to a fixed or adaptive Eulerian grid. Besides conservation laws, some physical processes such as nuclear burning are very sensitive to temperature, so slight errors in mapping can lead to very different outcomes for the simulation. Only a few studies have examined mapping 1D profiles to 2D or 3D meshes \\cite{zingale2002}, and none address the conservation of physical quantities by such procedures. We investigate these issues and introduce a new scheme for mapping 1D data sets to multidimensional grids. We first describe our mapping algorithm in \\Sect{method} and then present results of porting a massive star model from \\KEPLER{} to \\CASTRO{} in \\Sect{result}. Finally, we conclude in \\Sect{conclusion}. ", "conclusions": "\\lSect{conclusion} Multidimensional stellar evolution and supernova simulations are numerically challenging because multiple physical processes (hydrodynamics, gravity, burning) occur on many scales in space and time. For computational efficiency, 1D stellar models are often used as initial conditions in 2D and 3D calculations. Mapping 1D profiles onto multidimensional grids can introduce serious numerical artifacts, one of the most severe of which is the violation of conservation of physical quantities. We have developed a new mapping algorithm that guarantees that conserved quantities are preserved at any resolution and reproduces the most important features in the original profiles. Our method is practical for 1D and 2D calculations, and we are now developing integral methods (an explicit integral approach instead of using volume subsampling) that are numerically tractable in 3D." }, "1112/1112.5565_arXiv.txt": { "abstract": "We investigate baryogenesis in the $\\nu$MSM (neutrino Minimal Standard Model), which is the MSM extended by three right-handed neutrinos with masses below the electroweak scale. The baryon asymmetry of the universe can be generated by the mechanism via flavor oscillation of right-handed (sterile) neutrinos which are responsible to masses of active neutrinos confirmed by various experiments. We present the kinetic equations for the matrix of densities of leptons which describe the generation of asymmetries. Especially, the momentum dependence of the matrix of densities is taken into account. By solving these equations numerically, it is found that the momentum distribution is significantly distorted from the equilibrium one, since the production for the modes with lower momenta $k \\ll T$ ($T$ is the temperature of the universe) is enhanced, while suppressed for higher modes. As a result, the most important mode for the yields of sterile neutrinos as well as the baryon asymmetry is $k \\simeq 2 T$, which is smaller than $\\langle k \\rangle$ inferred from the thermal average. The comparison with the previous works is also discussed. ", "introduction": "\\label{sec:Introduction} The baryon asymmetry of the universe (BAU) is one of the most puzzling problems in particle physics and cosmology. It is certain that the Minimal Standard Model (MSM) cannot account for its origin. There have so far been proposed various scenarios of baryogenesis by considering physics beyond the MSM. (See, for example, a recent review \\cite{Riotto:1999yt}.) Among them leptogenesis~\\cite{Fukugita:1986hr} by superheavy right-handed neutrinos is one of the most motivated scenarios. This is because these new particles allow us to give non-zero masses to neutrinos which have been confirmed in many oscillation experiments. Being singlet under the MSM gauge group they can obtain Majorana masses which are completely independent on the electroweak scale. Right-handed neutrinos having superheavy masses then can generate the lepton asymmetry by their decays which can be a source of the BAU. In addition, the observed smallness of active neutrino masses is naturally explained by such fermions through the seesaw mechanism~\\cite{Seesaw}. The required masses are so heavy that it is impossible to directly test these new particles in near future experiments. Akhmedov, Rubakov and Smirnov have proposed~\\cite{Akhmedov:1998qx} another attractive scenario of baryogenesis by using right-handed neutrinos (mechanism via neutrino oscillation). One of the most important features is that the mechanism works when their masses are smaller than the electroweak scale, which are within the reach of current experiments. The flavor oscillation among right-handed neutrinos in the early universe leads to the separation of the lepton number into right-handed neutrinos and left-handed leptons. The sphaleron process then converts the asymmetry of the left-handed leptons into the baryon asymmetry for high temperatures~\\cite{Kuzmin:1985mm}. This mechanism can be realized in a simple and attractive framework called as the neutrino MSM ($\\nu$MSM)~\\cite{Asaka:2005an,Asaka:2005pn} in which three right-handed neutrinos are introduced with Majorana masses below the electroweak scale. Because of the very suppressed neutrino Yukawa interaction the seesaw mechanism is still effective, which leads to three active neutrinos $\\nu_i$ ($i=1,2,3$) and three sterile neutrinos $N_I$ ($I=1,2,3$) as mass eigenstates. The former ones are responsible to the neutrino oscillations observed in experiments. The latter ones give the solutions to the cosmological problems. The lightest sterile neutrino $N_1$ with mass $\\sim 10$ keV can be a candidate of dark matter, while the rest two $N_2$ and $N_3$ with masses $\\sim 1$ GeV can generate the BAU through the mechanism above. (For a review see Ref.~\\cite{Boyarsky:2009ix}.) Furthermore, the non-minimal coupling of Higgs field to gravity allows to realize the cosmic inflation~\\cite{Bezrukov:2007ep}. Various aspects of baryogenesis in the $\\nu$MSM has been studied until now~\\cite{Asaka:2005pn,Shaposhnikov:2008pf,Asaka:2010kk,Canetti:2010aw}. The generation of the asymmetry is described by the matrix of densities~\\cite{Dolgov:1980cq,Barbieri:1990vx,Sigl:1992fn} of sterile (right-handed) neutrinos and active (left-handed) leptons. The previous works were based on the kinetic equations for the matrix of densities in Ref.~\\cite{Asaka:2005pn}, where the terms describing the transfer of asymmetries between sterile neutrinos and active leptons are added to the original ones in \\cite{Akhmedov:1998qx}. As shown in \\cite{Asaka:2005pn}, such terms are essential to generate enough amount of the BAU in the $\\nu$MSM. In spite of their significance such terms are introduced by a heuristic approach. The main motivation of this article is to improve the estimation of the BAU in the $\\nu$MSM. One of the most unsatisfactory points in the previous works is that the evolution of the matrix of densities has been traced only by an approximate way. Namely, the kinetic equations analyzed previously are for the typical, single mode with momentum $k \\sim T$ ($T$ is the temperature of the universe), which is expected to give a dominant contribution to the asymmetry. By extrapolating the result of such a mode to other modes the asymmetry in the number density is estimated. In this article, thus, we derive the kinetic equations in which the momentum dependence of the matrix of densities is fully taken into account. For this purpose we shall re-evaluate the destruction and production rates of sterile neutrinos as well as active leptons paying attention to the two points. One is the momentum dependence of the particle of interest. The other is the corrections coming from the fact that the states in the destruction and production processes may differ from the thermal equilibrium states (\\ie, the deviations from the thermal equilibrium for sterile neutrinos and the chemical potentials for active leptons). As we will show later, the kinetic equations are written as simultaneous integrodifferential equations for the matrices of densities of sterile neutrinos and active leptons. Interestingly, our kinetic equations include the terms connecting between sterile and active sectors, which is crucial for the baryogenesis in the $\\nu$MSM as mentioned before. It will be shown that such terms arise automatically as the corrections to the destruction and production rates owing to the deviations from the equilibrium states in the scattering processes. It turns out that the coefficients of these terms are different from~\\cite{Asaka:2005pn} and, in addition, there appears a new type of terms which couples sterile neutrinos to active anti-leptons (rather than active leptons). This paper is organized as follows: In Sec.~\\ref{sec:BGinNuMSM} we briefly review the framework of the analysis and the baryogenesis scenario in the $\\nu$MSM. In the Sec.~\\ref{sec:KE}, we derive the kinetic equations which describe the generation of the baryon asymmetry paying attention to the momentum dependence in the matrices of densities. Sec.~\\ref{sec:NS} is devoted to study the numerical solutions to the obtained equations. We show the mode by mode evolution of the matrices of densities and their momentum distributions. The comparison with the previous works is done in Sec.~\\ref{sec:Comparison}. We perform the quantitative comparison of the yields of the BAU, but also clarify the differences between the kinetic equations in the literature. We conclude in Sec.~\\ref{sec:Conc}. ", "conclusions": "\\label{sec:Conc} In this paper we have investigated baryogenesis induced by the flavor oscillation between sterile (right-handed) neutrinos $N_2$ and $N_3$ in the $\\nu$MSM. We have presented the kinetic equations for the matrices of densities of sterile neutrinos $\\rho_N$ and $\\rho_{\\bar N}$ and chemical potentials of active (left-handed) leptons $\\mu_{\\nu_\\alpha}$ shown in Eqs.~(\\ref{eq:KE1}) and (\\ref{eq:KE3}). By using these equations the time evolution for the matrices of densities can be traced for each mode, and the momentum distributions can be found at any moments relevant for baryogenesis. As a result they allow us to estimate the yield of the baryon asymmetry more precisely. We have evaluated the destruction and production rates of sterile and active leptons for the kinetic equations. Especially, we have paid a special attention to the following two respects. The first is to include the momentum dependence of the rate correctly, which is mandatory to obtain the equations of $\\rho_{N, \\bar N}$ for a given momentum. The second is to include the effect from the deviation of the initial or final state from the thermal equilibrium in the scattering process. As explained in Sec.~\\ref{sec:KE}, the second point is crucially important in the construction of the kinetic equations for baryogenesis in the $\\nu$MSM. We have shown that the terms connecting between sterile and active sectors, which are required for the successful baryogenesis, arise in a consistent way from the second effect to the rate. To be concrete, the terms connecting the matrices of densities $\\rho_N$ and $\\rho_L$, which introduced in Ref.~\\cite{Asaka:2005pn}, appear as the corrections to the production rates $\\delta \\Gamma_N^p$ and $\\delta \\Gamma_\\nu^p$ via the scattering processes (A) and (B). As shown in Sec.~\\ref{sec:Comparison} the prefactor is different by 2/3 to the previous result in Ref.~\\cite{Asaka:2010kk}. Moreover, we have found the new terms which link $\\rho_N$ to $\\rho_{\\bar L}$ (rather than $\\rho_L$). They are induced as the corrections $\\delta \\Gamma_N^d$ and $\\delta \\Gamma_\\nu^d$ to the destruction rates via the process (C). Importantly, such a term is proportional to the product of $\\rho_{N , {\\bar N}}$ and $\\mu_{\\nu_{\\alpha}}$. Thus, our kinetic equations are no longer linear in $\\rho_N , \\rho_{\\bar N}$ and $\\mu_{\\nu_{\\alpha}}$. It should also be stressed that both types of these terms are required to ensure the conservation of the total lepton number. The impacts of these terms on the generation of the baryon asymmetry have been described in detail. We have then investigated the numerical solutions of the kinetic equations. Because of the momentum dependence in the rates the production of sterile neutrinos is enhanced for the modes with $k/T \\lesssim 1$ while diminished for the higher momentum modes. Consequently, the momentum distributions of $\\rho_N$ and $\\rho_{\\bar N}$ are significantly distorted from the equilibrium one. This clearly shows the importance to utilize the momentum-dependent kinetic equations. It has also been found that the low modes get in the thermal equilibrium and the asymmetries carried by them receive the wash-out effect at higher temperatures. In our choice of the parameters, the mode with $k/T \\simeq 2$ is the most significant for the occupation numbers as well as the asymmetries including the BAU. The comparison with the previous works has also been discussed. We have first explained what kind of approximations are needed to reproduce the previous kinetic equations staring from the present ones. It is shown that the use of the destruction rate~\\cite{Akhmedov:1998qx} obtained by the thermal average is crucial to ensure the lepton number conservation when we treat the momentum dependence of $\\rho_N$ approximately. We have then compare quantitatively the yields of the BAU. Apart from the correct treatment of the momentum dependence, the present equations gives the smaller BAU by a factor of $\\simeq 2/3$ compared with Ref.~\\cite{Asaka:2010kk}. Within the specific choice of parameters, the inclusion of the momentum dependence enhances the BAU by about 40\\% and 10\\% for $M_N=100$ MeV and 10 GeV, respectively. This is because the enhancement in the generation of the asymmetry from the lower modes overcomes the suppression induced by the higher modes. On the other hand, such an amplification may be disturbed by the wash-out effect for large neutrino Yukawa coupling constants. The present analysis shows that the momentum distribution of the matrices of densities differs from the equilibrium one and the inclusion of their momentum dependence modifies the yield of the BAU in non-trivial manner depending on the parameters of neutrino Yukawa coupling constants and masses of sterile neutrinos. Therefore, it is important to explore the full parameter space of the $\\nu$MSM accounting for the observed BAU by using the kinetic equations presented in this paper. This issue will be performed elsewhere." }, "1112/1112.2942_arXiv.txt": { "abstract": "We use the photoionisation and dust radiative transfer code MOCASSIN to create a model of the dwarf irregular galaxy NGC 4449. The best-matching model reproduces the global optical emission line fluxes and the observed spectral energy distribution (SED) spanning wavelengths from the UV to sub-mm, and requires the bolometric luminosity of 6.25$\\,\\times\\,$10$^9$ $L_\\odot$ for the underlying stellar component, $M_{\\mathrm{dust}}$/$M_{\\mathrm{gas}}$ of 1/680 and $M_{\\mathrm{dust}}$ of 2.2$\\,\\times\\,$10$^6$ $M_\\odot$. ", "introduction": "The spectral energy distribution (SED) of a galaxy emerges as a combination of the underlying stellar emission, the degree of reprocessing taking place in the interstellar medium (ISM) and thermal or non-thermal emission due to dynamical or evolutionary processes. As a consequence, galaxies emit across the entire range of the electromagnetic spectrum and the distribution of the energy output as a function of wavelength provide clues to the galaxy's past and present. Numerical models that reproduce the observed global properties of a galaxy over a wide range of wavelengths can be useful in understanding the physical and chemical nature of the underlying stellar and dust components. Here we present a MOCASSIN model of the dwarf irregular galaxy NGC~4449. MOCASSIN \\cite[(Ercolano \\etal\\ 2003, 2005, 2008)]{Ercolano2003,Ercolano2005,Ercolano2008} is a fully 3-dimensional photoionisation and dust radiative transfer code. Recent applications of the code include modelling the planetary nebula NGC~6302 \\cite[(Wright \\etal\\ 2011)]{Wright2011} and studying dust formation by SN~2008S \\cite[(Wesson \\etal\\ 2010)]{Wesson2010}. Modelling entire galaxies is technically challenging due to their large physical sizes and the high level of complexity of the galactic environments. Dwarf galaxies are ideal candidates due to their intrinsically small sizes, relatively simple star formation histories and the low degree of processing of their interstellar material. NGC~4449 (Fig.\\,\\ref{fig1}) is a metal-poor magellanic-type irregular galaxy. It is not a typical dwarf irregular: at the distance of 4.21 Mpc its H~{\\sc i} envelope extends to a radius of 14.2 kpc (11$.\\!\\!^{\\prime}$6), equal to approximately 6$R_{25}$ \\cite[(Swaters \\etal\\ 2002)]{Swaters2002}, and exhibits a counter-rotating core. It was selected for this study for its extensive coverage of photometric measurements for wavelengths ranging from the UV to sub-mm. Recently, new FIR observations were acquired by the {\\it Herschel Space Observatory} \\cite[(Pilbratt \\etal\\ 2010)]{Herschel} as part of the Guaranteed Time Key Project: Dwarf Galaxy Survey (PI: Suzanne Madden). The emission line fluxes, used to constrain the models, were derived from global spectra, acquired with the drift scanning technique, and are representative of the entire galaxy \\cite[(Kobulnicky \\etal\\ 1999)]{Kobulnicky1999}. \\begin{figure} \\vspace{2mm} \\begin{center} \\includegraphics[width=5.4cm]{./n4449_bw_sdssr.jpg} \\caption{SDSS \\textit{r}-band image of NGC~4449. North is up, east is to the left. The image is centred at 12$^{\\rm h}$28$^{\\rm m}$11$.\\!\\!^{\\rm s}$1, +44$^{\\circ}$05$^{\\prime}$37$^{\\prime\\prime}$ (J2000). The bar is 2$^{\\prime}$ in length.} \\label{fig1} \\end{center} \\end{figure} \\vspace{-2mm} ", "conclusions": "The gas and dust masses derived from our MOCASSIN model of NGC~4449 are consistent with earlier estimates (\\cite[Hunter \\etal\\ 1999]{Hunter1999}; \\cite[Engelbracht \\etal\\ 2008]{Engelbracht2008}). Our best-matching star formation scenario broadly agrees with the presence of an old (3--5 Gyr) and a young population of stars, and continuous star formation over the last 1~Gyr (\\cite[Bothun 1986]{Bothun1986}; \\cite[Martin \\& Kennicutt 1997]{Martin1997}; \\cite[Annibali \\etal\\ 2008]{Annibali2008}). However, the predicted present-day star formation rate is only one-fourth of the estimate based on the H$\\alpha$ luminosity \\cite[(Hunter \\etal\\ 1999)]{Hunter1999}. The discrepancy between these two independent estimates highlights the differences in the adopted initial mass function (IMF) as well as the limitations of the two-episode model. The presented technique assumes spherical symmetry for NGC~4449 and a two-episode star formation history, both of which are great simplifications. Also, at present PAHs are not included in the models and, as a consequence, the 4--20~$\\mu$m range is not correctly predicted. Nevertheless, most of the global parameters are predicted to within the observational uncertainties, and therefore this technique may prove useful in studying the characteristics of dust and the underlying stellar components in other dwarf galaxies. \\begin{table} \\vspace{0.2cm} \\begin{center} \\begin{tabular}{l@{\\hspace{8mm}}cc} \\cline{1-3}\\noalign{\\medskip} parameter & \\multicolumn{2}{c}{MOCASSIN model of NGC~4449}\\tabularnewline \\hline\\noalign{\\smallskip} $L_\\star$ & \\multicolumn{2}{c}{6.25$\\,\\pm\\,$0.25$\\,\\times\\,$10$^9$ $L_\\odot$} \\tabularnewline $M_\\star$ & \\multicolumn{2}{c}{1.2$\\,\\pm\\,$0.1$\\,\\times\\,$10$^9$ $M_\\odot$} \\tabularnewline $M_\\mathrm{gas}$ & \\multicolumn{2}{c}{1.5$\\,\\pm\\,$0.2$\\,\\times\\,$10$^9$ $M_\\odot$} \\tabularnewline $M_\\mathrm{dust}$ & \\multicolumn{2}{c}{2.2$\\,\\pm\\,$0.2$\\,\\times\\,$10$^6$ $M_\\odot$} \\tabularnewline $M_{\\mathrm{dust}}$/$M_{\\mathrm{gas}}$ & \\multicolumn{2}{c}{ 1/680 (1/850 to 1/540)} \\tabularnewline dust composition & \\multicolumn{2}{c}{100\\% amorphous carbon} \\tabularnewline dust grain sizes & \\multicolumn{2}{c}{0.005--1~$\\mu$m, $n \\propto a^{-3.5}$} \\tabularnewline \\noalign{\\medskip} SF episodes & 6$\\,\\pm\\,$2~Gyr to 120$\\,\\pm\\,$40~Myr & 120$\\,\\pm\\,$40~Myr to present \\tabularnewline relative $M_\\star$ & 120 & 1 \\tabularnewline $<$SFR$>$ & 0.2 to 0.3 $M_\\odot$\\,yr$^{-1}$ & 0.06 to 0.12 $M_\\odot$\\,yr$^{-1}$ \\tabularnewline \\noalign{\\smallskip}\\cline{1-3} \\end{tabular} \\vspace{0.2cm} \\caption{Summary of global parameters derived from the best-matching MOCASSIN model of NGC~4449 assuming two continuous episodes of star formation. The global emission line fluxes are matched to within 20 per~cent or better, with the exception of the sulphur lines [S~{\\sc ii}] $\\lambda\\lambda$6717,31, which are overpredicted by a factor of four.} \\label{tab1} \\end{center} \\end{table} \\vspace{-4mm}" }, "1112/1112.0176_arXiv.txt": { "abstract": "{The Small Magellanic Cloud (SMC) hosts a large number of Be/X-ray binaries, however no Be/white dwarf system is known so far, although population synthesis calculations predict that they might be more frequent than Be/neutron star systems.} {\\xmms was found as a new faint super-soft X-ray source (SSS) with a likely Be star optical counterpart. We investigate the nature of this system and search for further high-absorbed candidates in the SMC.} {We analysed the \\xmm X-ray spectrum and light curve, optical photometry, and the $I$-band OGLE III light curve.} {The X-ray spectrum is well represented by black-body and white dwarf atmosphere models with highly model-dependent temperature between 20 and 100 eV. The likely optical counterpart AzV\\,281 showed low near infrared emission during X-ray activity, followed by a brightening in the $I$-band afterwards. We find further candidates for high-absorbed SSSs with a blue star as counterpart.} {We discuss \\xmms as the first candidate for a Be/white dwarf binary system in the SMC.} ", "introduction": "\\label{sec:introduction} Super-soft X-ray sources \\citep[SSSs, e.g. ][]{2006AdSpR..38.2836K} are a phenomenological class of X-ray sources, defined by a very soft X-ray spectrum without significant emission above 1 keV. They were discovered first in the Magellanic Clouds due to their close distance and low foreground absorption. The general scenario for a SSS is thermonuclear burning on the surface of an accreting white dwarf (WD), which can be stable for certain accretion rates \\citep{2007ApJ...663.1269N}. SSSs are associated with cataclysmic variables (CVs), symbiotic stars, and post-outburst optical novae. Super-soft X-ray emission at lower luminosity can e.g. originate from some CVs, cooling neutron stars (NSs), active galactic nuclei (AGN), PG 1159 stars (hot cooling isolated WDs), and planetary nebulae. Also Be stars \\citep[e.g. ][]{2003PASP..115.1153P} are believed to harbour accreting WDs. By a mechanism not well understood and in combination with high rotation, Be stars eject material in the equatorial plane causing the build up of a circumstellar decretion disc \\citep{2001PASJ...53..119O,2002MmSAI..73.1038Z}. This disc causes line emission and a near infrared (NIR) excess, both showing high variability pointing to the instability of these discs. How far the generation of the Be phenomenon is related to close binary evolution or another mechanism is still under debate \\citep{2000ASPC..214..668G}. In the case of close binary evolution, matter transfer causes a spin up of the gainer, becoming a Be star, whereas the donor turns into a He star, WD or NS. Hereafter, accretion from the decretion disc onto the compact object causes X-ray emission. A high recent star formation \\citep{2010ApJ...716L.140A} and low metalicity \\citep{2006MNRAS.370.2079D} are probably responsible for the remarkably high number of $\\sim$90 known Be/X-ray binaries in the SMC. In these systems a NS accretes matter from the decretion disc of the Be star \\citep[for a recent review see ][]{2011Ap&SS.332....1R}. Binary system evolution models predict, that Be/WD systems might be more frequent than Be/NS systems \\citep{1991A&A...241..419P,2001A&A...367..848R}. \\citet{2001A&A...367..848R} obtain an abundance ratio for He star/NS/WD of 2/1/7, but no Be/WD system is known so far in the SMC in contrast to the large number of Be/NS systems. Similar to an accreting NS, WDs are expected to show hard X-ray emission powered by accretion, but at lower luminosity of $10^{29-33}$ erg s$^{-1}$, compared to $10^{34-38}$ erg s$^{-1}$ for NSs \\citep{1989A&A...220L...1W}. Suggested candidates are $\\gamma$\\,Cas like objects \\citep{1995A&A...296..685H,2006A&A...454..265L}. In addition, nuclear burning on the WD surface can produce super-soft X-ray emission at luminosities of $10^{35-38}$ erg s$^{-1}$, which might be absorbed by the decretion disc in most cases, complicating the discovery of such systems \\citep{1991A&A...248..139A}. Only one Be/WD SSS system has been proposed to date, \\wdbelmc \\citep{2006A&A...458..285K} in the Large Magellanic Cloud (LMC). To investigate the observational boundary conditions for Be/WD systems and to constrain the evolutionary models in general, it is of high importance to find more such systems. The \\xmm \\citep{2001A&A...365L...1J} large programme survey of the SMC (Haberl et al. 2012, in preparation), in combination with archival observations, provides a flux limited sample of X-ray sources of the central field of the SMC. This enables comprehensive studies of SSSs \\citep[e.g.][]{2010A&A...519A..42M,2011A&A...529A.152S} as well as Be/X-ray binaries \\citep[e.g.][]{2011A&A...527A.131S,2011MNRAS.414.3281C}. In a search for new faint SSSs in our point source catalogue (Sturm et al. 2012, in preparation), we found one candidate, correlating with an early type SMC star. In this study, we present the results of the analysis of the \\xmm EPIC X-ray data, together with optical information. This led to the discovery of the first candidate Be/WD system in the SMC -- \\xmms\\ -- and illustrates, that \\wdbelmc in the LMC is not a unique case. ", "conclusions": "\\label{sec:discussion} We analysed \\xmm X-ray data and OGLE III photometry of \\xmms. In the following we will discuss possible classifications of the source: {\\it A Galactic star} can produce soft coronal X-ray emission. In this case the fainter object could be the true counterpart. But the emission of stars extends to higher energies than we see from our SSS causing a positive $HR_1$ in general \\citep{2004A&A...426...11P,2011A&A...534A..55S}. Also, the derived $\\chi_{\\rm red}^2$ of the apec model suggests formally that a description of the spectrum by a thermal plasma is less likely than by the other models. {\\it A cooling isolated compact object}, i.e. a NS or PG 1159 star, can emit super-soft X-rays. We would not expect to see an optical counterpart. In both cases, we cannot explain the variability, seen in the X-ray light curve. For an {\\it AGN} with very soft X-ray emission and with a faint undetected optical counterpart, we expect to see the Galactic and total SMC absorption in the line of sight. The total SMC column density in the direction of the source is $N_{\\rm H, SMC} = 5.1 \\times 10^{21}$ cm$^{-2}$ \\citep{1999MNRAS.302..417S}. We derive a $N_{\\rm H, SMC} < 1.5 \\times 10^{21}$ cm$^{-2}$ for the power-law model, which seems to contradict the AGN assumption. {\\it A CV in the SMC} which by chance correlates with AzV\\,281 can account for the super-soft emission. However, there is no accretion powered system known with super-soft emission exceeding $10^{33}$ erg s$^{-1}$, which is below the observed luminosity of \\xmms for the distance of the SMC. In the case of nuclear surface burning, higher luminosities can be reached (e.g. CAL 83 in the LMC). This would need a highly obscured WD, as discussed below. Another possibility for variable super-soft X-ray emission would be the rare case of a nova explosion. Here, super-soft emission is observed up to 2--5 years \\citep{2011A&A...533A..52H} after outburst. However, a nova explosion in this system is not known, which of course could have been missed. Due to the low nova rate in the SMC (only 3 novae were discovered during the last 10 years) this is an unlikely scenario. {\\it A Galactic CV} cannot be excluded with the available data. E.g. the Galactic star could contain a WD companion. In this case, a soft intermediate polar \\citep{1995A&A...297L..37H} could account for the spectral properties, but only a handful of these systems is known in the Galaxy so far. During the \\xmm SMC survey data analysis, we did not find any other SSS candidate correlating with an early-type emission line star. To estimate the chance for a random coincidence of a SSS or SSS candidate (14 in total) with a blue emission line star, we used a subset of early type SMC stars ($V<17$ mag, $-0.5$ mag $$50 mas yr$^{-1}$ (134780 sources) results in 773 chance coincidences. This further affirms AzV\\,281 as the most likely optical counterpart. Also the coincidence of the X-ray emission with the NIR flux minimum further supports the identification of the super-soft X-ray source with AzV\\,281. From a {\\it Be/NS system} we would expect outbursts with hard X-ray emission during accretion, which has not been observed for this system, although it was monitored regularly during the last 11 years. Also, observing only supersoft X-ray emission is atypical for a Be/NS system, as dominant hard emission is expected in additon to a soft excess \\citep{2004ApJ...614..881H}. Even in the extreme case of an outburst of RX\\,J0103.6-7201 \\citep{2008A&A...491..841E}, a hard X-ray component is present and the soft component extends to higher energies. From the bb+po model we obtain a limit for $L_{\\rm bb}/L_{\\rm X}$ of $>$0.5 in the (0.15-10.0) keV band, proofing the dominance of the thermal component. Therefore, we argue, that the system is less likely a classical Be/X-ray binary. In contrast to the former possibilities, a {\\it WD/Be system} can account for both the super-soft X-ray emission and the variability. Black-body emission is a crude approximation to the spectra of SSSs \\citep{2010AN....331..146R}. Physically more meaningful are nlte models, but higher statistics and high resolution spectra would be necessary to constrain the parameters. To demonstrate the parameter dependence on elemental abundances we list two nlte models in Table~\\ref{tab:spec}. Solar abundances result in similar parameters as the black-body model with an emission radius of (73--106) km compared to (19--75) km from the bb model. Halo abundances result in a higher absorption, higher luminosity and an emission radius of (90000--190000) km. For an object in the SMC we expect abundances between solar and halo values \\citep{1992ApJ...384..508R,2010A&A...520A..85D}. Therefore the black-body emission radius is rather a lower limit. This further demonstrates, that luminosity and absorption can not be determined uniquely for \\xmms and that the spectra might be compatible with absorbed emission from a WD and luminosities sufficiently high for surface burning. The presumable optical counterpart shows properties typical for a Be star. This makes the system the second candidate for a Be/WD binary after \\wdbelmc in the LMC \\citep{2006A&A...458..285K}. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics[angle=0,clip=]{opt_lc.ps}} \\caption{ {\\it Top:} $I$-band light curve from OGLE III. {\\it Bottom:} X-ray flux in the (0.2--1.0) keV band. } \\label{fig:lc} \\end{figure} \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics[clip=]{opt_spec.ps}} \\caption{ Dereddened optical \\citep{2002AJ....123..855Z} and NIR \\citep{2007PASJ...59..615K} fluxes (squares) compared to a Kurucz model atmosphere for a B1V (upper line) and an O7V (lower line) star both normalised to the $U$-band flux. } \\label{fig:spec_opt} \\end{figure} The X-ray turn-off around MJD 52500 can be caused by an exhaustion of the nuclear burning on the WD, as it occurs for post-nova SSSs, if the accretion rate is too low for stable nuclear burning. In the case of ongoing mass transfer, reignition of nuclear burning on the WD is possible in a nova explosion, followed by another super-soft X-ray emission state. More likely, the decrease in soft X-rays in the later observations might be explained by increasing photo-electric absorption due to a build-up of the decretion disc around the Be star. An additional absorbing column density of \\nh$\\gtrsim 10^{22}$ cm$^{-2}$ would be required for a non-detection of \\xmms\\ in the merged images. Be stars can loose and rebuild their discs on a time scale of some years including intervals without a disc of some hundred days \\citep{2011Ap&SS.332....1R,2010ApJ...709.1306W}. The decrease of the NIR flux before or during the X-ray active phase and the low photo-electric absorption during the three X-ray detections suggest this scenario. The X-ray detections have only a small overlap with the optical light curve, but since X-ray turn-off, the $I$-band shows an increase in flux, which can be caused by the recreation of a disc. We note, that the optical light curve of \\wdbelmc \\citep[cf. Fig. 4 in ][]{2006A&A...458..285K} has similar characteristics of both a long-term increase in the $I$-band and a periodicity around one thousand days. The low X-ray absorption suggests a location of the system on the near side of the SMC bar. Also, the system intrinsic absorption along the line of sight must have been low during the X-ray detections. If the 1264 d period is caused by binarity, we get a relatively large separation of 5.7 AU ($\\sim$$140 R_{\\ast}$) for $M_{\\rm WD}=1\\msun$ and $M_{\\rm \\ast}=15\\msun$. But we note, that the realness and nature of the periodicity are uncertain, since the OGLE light curve covers only about two cycles. While \\citet{2001A&A...367..848R} argue that most Be/WD systems are not detectable in X-rays due to absorption by the decretion disc, truncation of the disc, as it is known from Be/NS systems with low excentricity \\citep{2001A&A...377..161O} might only allow accretion during disc instabilities. Also, the absorption by the SMC has to be low, to be able to observe such systems in X-rays. We estimate the completeness in our SMC catalogue in the (0.2--1.0) keV band to a flux limit of 2.5\\ergcm{-15}, which would require a SMC absorption of less than 3 $\\times 10^{21}$ cm$^{-2}$ for the best-fit black-body model. Therefore, \\xmms might be the tip of the iceberg of the SMC Be/WD system population. We report the discovery and analysis of a new SSS in the SMC, having a presumable early type star companion similar to \\wdbelmc in the LMC. This star has a variable NIR excess and emission lines, typical for a Be star. This makes the system the first candidate for a Be/WD system in the SMC. The X-ray turn off of the source around MJD 52400 can be caused by an increasing photo-electric absorption, possibly caused by the build up of a decretion disc around the Be star. We list further candidates for brighter but more absorbed Be/WD systems." }, "1112/1112.0495_arXiv.txt": { "abstract": "Extending previous studies, we derive generic predictions for lower order cumulants and their correlators for individual tomographic bins as well as between two different bins. We derive the corresponding one- and two-point joint probability distribution function for the tomographic convergence maps from different bins as a function of angular smoothing scale. The modelling of weak lensing statistics is obtained by adopting a detailed prescription for the underlying density contrast. In this paper we concentrate on the convergence field $\\kappa$ and use top-hat filter; though the techniques presented can readily be extended to model the PDF of shear components or to include other windows such as the compensated filter. The functional form for the underlying PDF and bias is modelled in terms of the non-linear or the quasilinear form depending on the smoothing angular scale. Results from other semi-analytical models e.g. the lognormal distribution are also presented. Introducing a {\\em reduced} convergence for individual bins, we are able to show that the tomographic PDFs and bias for each bin sample the same functional form of the underlying PDF of density contrast but with varying variance. The joint probability distribution of the convergence maps that correspond to two different tomographic bins can be constructed from individual tomographic PDF and {\\em bias}. We study their dependence on cosmological parameters for source distributions corresponding to the realistic surveys such as LSST and DES. We briefly outline how photometric redshift information can be incorporated in our computation of cumulants, cumulant correlators and the PDFs. Various approximate results for cumulants and their correlators are presented. Connection of our results to the full 3D calculations is elucidated. Analytical results for inclusion of realistic noise and finite survey size are presented in detail. ", "introduction": "Following the first weak lensing measurements \\citep{BRE00,Wittman00,KWL00,Waerbeke00} , the field of weak lensing has witnessed a tremendous progress in all fronts (see \\citet{MuPhysRep08} for a review). Currently, in terms of cosmological observations, weak lensing plays a role complementary to both Cosmic Microwave Background (CMB) studies and studies involving large scale structure (LSS) surveys. The ability of weak gravitational lensing to reveal cosmological information, particularly the dark energy equation of state is considerably enhanced by the inclusion of tomographic information. The impotance of weak lensing has spurred tremendous progress on the technical front in terms of specification and control of systematics. There are many ongoing and future weak lensing surveys such as the CFHT{\\footnote{http://www.cfht.hawai.edu/Sciences/CFHLS/}} legacy survey, the Pan-STARRS{\\footnote{http://pan-starrs.ifa.hawai.edu/}} and the Dark Energy survey{\\footnote{https://www.darkenergysurvey.org/}}, and further in the future, the Large Synoptic Survey Telescope{\\footnote{http://www.lsst.org/llst\\_home.shtml}}, Joint Dark Energy Mission or JDEM{\\footnote{http://jdem.gsfc.nasa.gov/}} that will map the dark matter and dark energy distribution of the entire sky in unprecedented details. In particular, owing to the large fraction of the sky coverage and tighter control on systematics as well as dense sampling of source galaxy populations it will be soon possible to study gravity induced non-Gaussianity with extreme accuracy. The gravity induced non-Gaussianity is typically probed using real space correlation functions as well as in the harmonic domain using their harmonic counterparts i.e. the multispectra (see e.g. \\cite{Pen03}). These correlation functions provide a set of tools to go beyond the usual power spectrum analysis. The higher-order correlation functions are important not only to break the parameter degeneracy inherent in power spectrum analysis (e.g. between the amplitude of the matter power spectrum $\\sigma_8$ and the matter density parameter $\\Omega_{\\rm M}$) but also to understand error-estimates of lower-order correlations functions. Starting with the study of the three-point correlation function \\citep{Vil96,JainSeljak97} higher order statistics of weak lensing shear, convergence or flexions are now well understood from a theoretical point of view. The power spectrum of density perturbations remains the most commonly used statistic in many cosmological studies. Weak lensing surveys probe the non-linear regime and are sensitive to non-Gaussianity which can not be probed using only the two-point correlation function or its harmonic analog the power spectrum. The statistics of shear or convergence probe the statistics of underlying mass distribution in an unbiased way \\citep{JSW00,MuJai01,Mu00,MuJai00,Valageas00, VaMuBa05,TakadaWhite03,TakadaJain04}, sensitive to nonlinear evolution due to gravitational clustering. Various analytical schemes from perturbative calculations to halo models have been employed to model the weak lensing statistics \\citet{Fry84,Schaeffer84, BerSch92,SzaSza93, SzaSza97, MuBaMeSch99, MuCoMe99a, MuCoMe99b, MuMeCo99, MuCo00, MuCo02, MuCo03, CooSeth02}). In addition to studying the statistics in projection they have also been studied in 3D using photometric redshifts. This approach can further tighten the constraints on e.g. the neutrino mass as well as the dark energy equation of state \\citep{Heav03,HRH00, HKT06, HKV07, Castro05, Kit08}. Tomographic techniques have also been employed as an intermediate strategy between projected surveys and 3D mapping \\citep{Hu99,TakadaJain04,TakadaJain03,Semboloni08}. In this paper we extend previous results \\citep{JSW00,MuJai01,Mu00,MuJai00,Valageas00} on projected surveys by analysing the entire one-point PDF and the two-point PDF with tomographic information.The PDF contains information about the correlation hierarchy to an arbitrary order; the correlation hierarchy of the convergence field is directly related to that of the underlying mass distribution. We employ a generating function formalism that relies on {\\em hierarchical ansatz} on smaller angular smoothing scales and on perturbative results on larger scales. We define a reduced convergence for each bin and show that the different bins sample the same underlying PDF and bias functions (to be defined later) for the density contrast. The entire joint two-point PDFs for different pairs of redshift bins and individual PDF for each bins can be constructed from the PDF and the bias associated with individual bins because the joint PDF is factorisable in terms of the individual PDFs, bias and cross-correlations among various bins and different angular scales. We will show that individual redshift-resolved tomographic maps can be used to map out the PDF of the underlying mass distribution for a wide range of variance. This underlying PDF of the density contrast can be used to recover the tomographic PDF with the use of just two individual variables $\\kappa_{min}$ and the reduced variance for each bin; both of these variables are uniquely determined by the geometry and matter content of the Universe. The results are applicable not only to the PDFs as determined under hierarchical ansatz but also for other well motivated approximations for PDF such as the lognormal distribution. Recent cosmological observations favour an accelerating Universe. This implies existence of energy of unknown nature (dark energy) which has negative pressure \\citep{Amen10,Wang10}. Current data continues to be consistent with dark energy being a non-zero cosmological constant. Though many other alternative dark energy candidates have been consider which are consistent with data as well, e.g. quinessence, k-essence, spintessence. Different dark energy models can be classified according to the equation of state of of the dark energy component $w_{\\rm X}$. For quintessence model $dw_{\\rm X}/dz>0$ while for k-essence models $dw_{\\rm X}/dz<0$. There are many complimentary probes for dark energy, the distance-redshift relation of cosmological standard candles; Cosmic Microwave Background Anisotropy; volume redshift relations using galaxy counts; the evolution of galaxy clustering; weak lensing, etc. The different methods to probe dark energy are complementary to other and can provide important consistency check. Weak lensing surveys are particularly suitable for dark energy studies. All major weak lensing surveys has dark energy as their one of prime science driver. We will use the techniques developed in this paper to study two different dark energy model and compare the predictions against those of standard $\\Lambda$CDM model. The methods presented here are complementary to the usual Fisher matrix based approach that rely on two-point correlation functions or the power spectrum as it includes non-Gaussian information upto order . This paper is organised as follows. In \\textsection2 we introduce our notation and present some standard results. In \\textsection3 we link the lower order statistics of weak lensing convergence to that of the underlying density distribution. In \\textsection4 we briefly review the hierarchical ansatz in the context of generating function formalism. In \\textsection6 we discuss the lognormal model in the context of weak lensing statistics. In \\textsection7 we derive the PDF and bias for various tomographic bins. The results are quite generic and can be used for arbitrary source redshift distribution. Finally the \\textsection8 is left for discussion of our results. In an appendix we outline how in the context of tomographic binning the evolution topological estimators such as Minkowski Functionals can be studied using the lognormal distribution. ", "conclusions": "\\label{sec:conclu} Previous tomographic studies of weak lensing have typically worked with the lower order cumulants; we have generalized here these results to the case of the entire one- and two-point PDF, which contain information about the cumulants to an arbitrary order. The construction was performed using a generating function formalism based on hierarchical ansatz and a lognormal model. Our analysis generalizes previously obtained results derived for projected or 2D surveys. Though we have considered a top-hat filer convergence maps due to their simplicity, similar results can be obtained for related statistics such as the shear components or aperture mass $M_{ap}$ \\citep{BV00}. The PDFs for the individual bins are constructed by generalization of the previously introduced global variable $\\kappa^{\\rm min}$, for individual bins i.e.$\\kappa^{\\rm min}$, that was used in the context of 2D projected maps. Next, using $\\kappa_{(i)}^{\\rm min}$, reduced variable $\\eta^{(i)}$ is defined for each individual bins whose statistics can directly be linked to that of underlying density contrast $\\delta$. The convergence in individual bins can then be mapped to unique values of $\\eta=1+\\delta$ for a given smoothing angular scales $\\theta_0$. For modelling the statistics of underlying density contrast $\\delta$ we have assumed two completely different model: the hierarchical ansatz along with its perturbative counterpart as well as the lognormal distribution. Both these approximations have been used successfully in various cosmological contexts. There are a wide class of models that are available under the general category of hierarchical ansatz. The main motivation behind our choice of a specific hierarchy is simplicity. In recent years more sophisticated models of hierarchical clustering have been proposed which rely more on inputs from numerical simulations. The generic results we have derived here indeed can be improved using such modelling though the fundamental characteristics will remain unchanged. \\begin{figure} \\begin{center} {\\epsfxsize=6. cm \\epsfysize=6. cm {\\epsfbox[27 426 315 709]{noise_diff.eps}}} \\end{center} \\caption{The difference of noisy $\\Lambda$CDM PDF and dark energy models $\\Delta_p(\\kappa)=p(\\kappa)-p_{\\Lambda \\rm CDM}(\\kappa)$ is plotted as a function of $\\kappa$. The smoothing angular scale, bin size and galaxy number density is as depicted. The scatter in estimation is smaller compared to the difference in the PDFs considered. The comsological parameters considered are the same as the ones in Figure-(\\ref{fig:dark}). The two survey configurations that we have considered both produces near identical results.} \\label{fig:10} \\end{figure} In our treatment we find, in agreement with \\cite{MuWa03}, the dynamical and geometrical contribution can be treated separately. The geometrical effects are completely encoded in a parameter $\\kappa^{min}$. The reduced convergence as defined is independent of the background geometry of the universe and essentially probe the evolution of gravitational clustering. We showed that a set of $\\kappa^{\\rm min}_{(i)}$ defined for a given set of redshift slices are adequate to characterize not only individual PDFs for each bin but it is also sufficient to study the joint two-point PDF among two different bins. The PDF of the reduced convergence $\\eta^{(i)}$ for individual bins or joint PDFs for a pair of bins generalizes the earlier studies where the projected or 2D maps were considered in a straight forward manner. We also note that the construction of convergence maps is difficult compared to the direct evaluation of non-Gaussian statistics from shear maps. On the other hand convergence statistics can directly be modelled at arbitrary order whereas for shear field the computation is done mostly order by order manner. An independent analysis of convergence maps constructed from shear maps should therefore be useful in constraining various errors which might get introduced during various stages of data reduction. In our analysis we have ignored the noise from intrinsic ellipticity distribution of galaxies as well as from shot noise resulting from finite number of galaxies that are used to trace the underlying mass distribution. These issues have been dealt with in great detail in \\cite{MuCo03,VaBaMu04}. Dividing the source population into bins reduced the number-density of sources. This in turn will increase the level of noise or the scatter in the estimator. In our analysis we have considered two different survey configurations, i.e. LSST and DES and found that for our choice of tomographic bins the one- and two-point PDFs are very similar in nature. The lognormal distribution has already been used to model the statistics of weak lensing observables \\citep{Mu00,TTHF02} and the clustering of Lyman alpha absorption systems e.g. \\citep{BD97}. One-to-one mapping of initial density fields to evolved density fields using maps that are consistent with lognormal distribution function was not found to be very successful and the success of a lognormal distribution function in reproducing the statistics of gravitational clustering still remains somewhat unclear. Tomographic weak lensing surveys can be cross-correlated with external data sets including frequency cleaned maps of secondaries from ongoing CMB surveys; e.g. the thermal Sunyaev-Zeldovic (tSZ) maps or $y$-maps that will be available from CMB surveys such as Planck. The cross-correlation with tomographic information can help to understand the evolution of cosmological pressure fluctuations responsible for tSZ effect with redshift. The formalism presented here is perfectly suitable for such analysis. Detailed results of such analysis will be presented elsewhere. In addition to the weak lensing surveys the Supernova pencil beam surveys might also benefit for the results presented here. To summarize, we have extended results derived in three different previous papers \\citep{MuCo03,VaMuBa05,VaBaMu04} to take into account tomographic bins within which the photometric redshift are available. The results obtained previously for one-point PDF are now extended to two-point PDF. These results can provide an alternative to usual Fisher-matrix analysis that is employed to optimize survey strategies. We have concentrated mainly on analytical results in this paper. The numerical results regarding optimization of survey strategy using these results will be considered elsewhere." }, "1112/1112.5279_arXiv.txt": { "abstract": "{In the last few years Fermi and AGILE observations have indicated the existence of a possible population of transient MeV-GeV sources located on the Galactic plane and characterized by fast flares lasting only a very few days. Notably, no blazar-like counterparts are known within their error boxes so they could represent a completely new class of Galactic transient high energy emitters. The task of identifying their counterparts at lower energies remains very challenging. Despite this difficulty, INTEGRAL observations have provided intriguing hints that reliable candidate counterparts for these unidentified MeV-GeV transients could be found among the members of the recently discovered class of Supergiant Fast X-ray Transients (SFXTs). In this context, to date the best test case is represented by the association between the two sources IGR J17354$-$3255 and AGL J1734$-$3310. We will discuss their possible physical link and implications stemming from this association.} \\FullConference{The Extreme and Variable High Energy Sky - extremesky2011,\\\\ September 19-23, 2011\\\\ Chia Laguna (Cagliari), Italy} \\begin{document} ", "introduction": "The field of high energy astronomy is relatively young. Breakthrough results have been obtained only in the last twenty years thanks to satellites carrying instruments whose survey capabilities unveiled the extreme richness of objects at hard X-rays (E$>$20 keV, e.g. INTEGRAL/IBIS, Swift/BAT) as well as gamma-rays (E$>$100 MeV, e.g CGRO/EGRET, AGILE/GRID, Fermi/LAT). Interestingly, the great majority of such objects are still unidentified, with no firmly established counterparts at lower energies. Their identification is one of the great challenges of current high energy astronomy, it could leave some room for novel and unexpected discoveries. In this context, recent AGILE/GRID and Fermi/LAT observations have indicated the existence of a possible population of fast transient MeV-GeV sources located on the Galactic plane and characterized by flares lasting only a very few days (e.g. Hays et al. 2009, Bulgarelli et al. 2009, Chen et al. 2007). Notably, no blazar-like counterparts are known within their error boxes so they could represent a completely new class of Galactic fast high energy transients. The task of identifying their counterparts at lower energies remains very challenging, mainly because of their often large error circles (e.g radii typically from 10 arcmin to 0.5 degrees). INTEGRAL/IBIS is particularly suited to search for reliable best candidate counterparts thanks to i) a large field of view (FoV) which ensure a total coverage of the gamma-ray error box ii) a good angular resolution which is crucial to disentangle the hard X-ray emission of different sources in crowded fields such as those on the Galactic Plane iii) a good sensitivity above 20 keV. In particular, recent INTEGRAL results provided intriguing hints that reliable best candidate counterparts could be found among the members of the recently discovered class of Supergiant Fast X-ray Transients (Sguera et al. 2009, Sguera et al 2011, Sguera 2009). Supergiant Fast X-ray Transients (SFXTs) are a new class of High Mass X-ray Binaries (HMXBs) mainly unveiled thanks to INTEGRAL observations of the Galactic plane. They host a massive OB supergiant star as companion donor (Negueruela et al. 2006) and display X-ray flares lasting from a few hours to a few days (Sguera et al. 2005, 2006). It is generally assumed that all SFXTs host a neutron star as compact object because their broad band X-ray spectra (0.2--100 keV) strongly resemble those of accreting X-ray pulsars in HMXBs, i.e. absorbed cutoff power law shape (e.g. Sidoli et al. 2009). This idea is supported by the detections of X-ray pulsations in some systems (e.g. Sguera et al. 2007). The typical dynamic range of classical SFXT is 10$^{3}$--10$^{5}$, however some systems show a lower value of $\\sim$10$^{2}$ and so they have been named as intermediate SFXTs (Sidoli et al. 2011, Clark et al. 2010, Walter \\& Zurita 2007). The physical mechanism driving the peculiar X-ray behaviour of SFXTs is still unclear and highly debated. Several different models have been proposed in the literature (see Sidoli 2009 for a review). We note that in principle SFXTs have all the \"ingredients\" to possibly be MeV-GeV emitters since they host a compact object (i.e. neutron star) as well as a bright and massive OB star which could act as source of seed photons (for the Inverse Compton emission) and target nuclei (for hadronic interactions). In this respect it is important to point out that in the last few years a few classical HMXBs, having the same \"ingredients\" of SFXTs in terms of compact object and companion stellar donor, have been firmly detected at MeV-TeV energies as persistent and variable sources (e.g. LS 5039 and LS I +61 303, Paredes 2008, Hill et al. 2010) or as fast flaring transients (e.g. Cyg X-3 and Cyg X-1, Tavani et al. 2009, Sabatini et al. 2010), providing evidence that particles can be efficiently accelerated to very high energies in HMXBs. At odds with such gamma-ray binaries, the eventual MeV-GeV emission from SFXTs must be in the form of fast flares and should be expected only for a very small fraction of time (i.e. from few hours to few days), making very difficult their detection. Despite this drawback, some observational evidences have been recently reported in the literature on SFXTs as best candidate counterparts of unidentified transient MeV-GeV sources located on the Galactic Plane (Sguera et al. 2009, Sguera et al 2011, Sguera 2009). These evidences are merely based on intriguing hints such as a spatial correlation and a common transient behaviour on similar, though as yet not simultaneous, short time scales. This scenario is also supported from an energetic standpoint by a theoretical model based in the microquasar accretion/jet framework (Sguera et al. 2009). The so far proposed associations represent an important first step towards obtaining reliable candidates on which to concentrate further efforts to obtain quantitative proofs for a real physical association. In this respect, so far, the best test case is represented by the association between the two sources IGR J17354$-$3255 and AGL J1734$-$3310. ", "conclusions": "" }, "1112/1112.0340_arXiv.txt": { "abstract": "Radio observations of galaxy clusters show that the intra cluster medium is permeated by $\\rm{\\mu G}$ magnetic fields. The origin and evolution of these cosmological magnetic fields is currently not well understood and so their impact on the dynamics of structure formation is not known. Numerical simulations are required to gain a greater understanding and produce predictions for the next generation of radio telescopes. We present the galactic chemodynamics smoothed particle magnetohydrodynamic (SPMHD) code (GCMHD+), which is an MHD implementation for the cosmological smoothed particle hydrodynamic code GCD+. The results of 1, 2 and 3 dimensional tests are presented and the performance of the code is shown relative to the ATHENA grid code. GCMHD+ shows good agreement with the reference solutions produced by ATHENA. The code is then used to simulate the formation of a galaxy cluster with a simple primordial magnetic field embedded in the gas. A homogeneous seed field of $10^{-11} \\rm{G}$ is amplified by a factor of $10^3$ during the formation of the cluster. The results show good agreement with the profiles found in other magnetic cluster simulations of similar resolution. \\vspace{2mm} \\noindent\\textbf{Key words:} magnetic fields - MHD - methods: numerical - galaxies: clusters: general \\vspace{0.1cm} ", "introduction": "Cosmological magnetic fields are thought to be ubiquitous through out the universe and have been detected by radio observations on scales as large as galaxy clusters. On the galactic cluster scale, magnetic fields have been detected via Faraday rotation measurements of distant quasars/AGN, by the observation of radio galaxies within the cluster and by the detection of diffuse synchrotron emission from radio haloes \\citep[e.g.][]{obs94,obs02,obs04,obs11,obs01}. The current observations find magnetic fields of $\\rm{\\mu G}$ strength permeates the intra-cluster medium (ICM) of most galaxy clusters. Beyond the cluster scale, measurements of the magnetic field are far less certain. Cosmological fields can be generated via Weibel's instability \\citep{sch03,med06}, Biermann's battery \\citep{bbm50,sub94,kul97,gne00}, structure formation \\citep{har70,ici06}, galactic winds \\citep{bec96,ber06,zfe09}, relativistic charged particles \\citep{min11} and various processes in the early universe \\citep{the02}. Current observations of the structural detail of these fields is limited, but the next generation of radio telescopes, such as the Square Kilometer Array (SKA) \\citep[e.g.][]{ska06,ska08}, will produce a wealth of observational data on the strength and structure of cosmological magnetic fields. To compare these observations with our knowledge of cosmological magnetic fields we require numerical simulations. The predictions made by simulations allow for a comparison of the theory with the observation. This will produce a more detailed understanding of cosmological magnetic fields. The evolution of primordial magnetic fields in a cosmological setting has been studied using both smoothed particle magnetohydrodynamics \\citep[SPMHD, e.g.][]{cp09,niMHD11} and adaptive mesh refinement codes \\citep[e.g.][]{pmc08,pmc11}. The different techniques produce good agreement on the predicted strength and profile of a magnetic field in a galaxy cluster. They show that the initial structure of the field has little influence on the final field. The central cluster field strength predicted by these simulations agrees with the values inferred from observation. A few simulations have followed the creation of a cosmological magnetic field from either galactic wind pollution of the ICM \\citep{zfe09}, AGN pollution of the cluster \\citep{zfe10} or the Biermann battery mechanism during structure formation \\citep{zfe98} to generate a field from zero field initial conditions. The magnetic field strength can also be predicted a posteriori using the velocity fields from a purely hydrodynamical cosmological simulations \\citep{zfe08}. These simulations produce different field strengths than the values found from following the evolution of a primordial magnetic field, especially in filament structures. Further simulations are required to examine the different origins of a magnetic field in the ICM and to test the validity of the predictions made. We present the implementation of magnetohydrodynamics (MHD) for the existing galactic chemodynamics code (GCD+) \\citep{gcd03,gcd09,gcd11}. GCD+ is a three-dimensional tree N-body/Smoothed Particle Hydrodynamics (SPH) code which incorporates self-gravity, hydrodynamics, radiative cooling, star formation, supernova feedback and metal enrichment. The following MHD implementation is fully compatible with all of the original features of the code and allows for a complete cosmological simulation to be run. We discuss the choice of method to ensure tensile instability is avoided to an appropriate level. We also discuss the effect of an applied time dependant dissipation scheme \\citep{tdd97} for the magnetic field and the effect of the Balsara switch \\citep{cp95} on the test simulations, including its need in a cosmological simulation. For the tests shown in this paper we ignore the additional processes and concentrate on the effect of introducing MHD to the simulations and the accuracy of the non-radiative solutions produced. As such, radiative cooling, star formation, supernovae feedback and metal enrichment are switched off in all results presented. In Section 2 we present the numerical implementation of the MHD equations used in GCMHD+. Section 3 shows the performance of the code in various test simulations. Section 4 shows the results of a cosmological simulation for the formation of a galaxy cluster with a simple primordial magnetic field embedded in the gas. Our conclusions are then given in Section 5. ", "conclusions": "We have introduced the MHD component to the N-body/SPH code GCD+. We discussed the addition of the equations of ideal MHD, the choice of instability correction and the addition of dissipative terms to treat discontinuities in the magnetic field. We implement schemes to remove the tensile instability, suggested by \\citet{ic01}, and for artificial magnetic dissipation, following \\citet{cp04a}. The code's ability to vary this dissipation in the simulation and allow each particle to evolve it own dissipation constant is presented. We put the code through a set of standard 1D shocktube tests, the fast rotor test, the Orszag-Tang vortex test and the MHD Point like explosion test. The numerical parameters were varied for all tests and the best compromise between noise reduction and minimised smoothing was found. The code with the best compromised parameter set performs very well in these tests and agrees with the reference solutions provided by the ATHENA mesh code, where they are available. The code shows no sign of the tensile instability and the magnetic dissipation scheme produces very little smoothing, while allowing the code to accurately capture the features. We then applied the code to a cosmological simulation for the formation of the Santa Barbara galaxy cluster. The code produces the expected hydrodynamic parameters of the cluster. The magnetohydrodynamic parameters are also well captured and show a similar level of magnetic field to the literatures with similar resolution \\citep{cp09}. We demonstrate that no minimum limit for the parameter of the dissipation of magnetic field, $\\alpha^B_{min}=0.0$, is necessary to minimise the artificial dissipation for the cluster scale magnetic field. This requirement is significantly lower than the previous SPMHD implementations, except for \\citet{niMHD11}. Our extensive test simulations in Section 3 demonstrate that $\\alpha^B_{min}=0.0$ still leads to satisfactory results. Encouraged by the success of our new MHD code, GCMHD+, we will apply it to higher resolution cosmological simulations, and study how magnetic fields developed in the evolving universe." }, "1112/1112.4506_arXiv.txt": { "abstract": "We examine the {\\it Kepler} light curves of V1504 Cyg and V344 Lyr, encompassing $\\sim$736 d at 1 min cadence. During this span each system exhibited $\\sim$$64-65$ outbursts, including six superoutbursts. We find that, in both systems, the normal outbursts lying between two superoutbursts increase in duration over time by a factor $\\sim$1.2$-$1.9, and then reset to a small value after the following superoutburst. In both systems the trend of quiescent intervals between normal outbursts is to increase to a local maximum about half way through the supercycle $-$ the interval from one superoutburst to the next $-$ and then to decrease back to a small value by the time of the next superoutburst. This is inconsistent with Osaki's thermal-tidal model, which predicts a monotonic increase in the quiescent intervals between normal outbursts during a supercycle. Also, most of the normal outbursts have an asymmetric, fast-rise/slower-decline shape, which would be consistent with outbursts triggered at large radii. The exponential rate of decay of the plateau phase of the superoutbursts is 8 d mag$^{-1}$ for V1504 Cyg and 12 d mag$^{-1}$ for V344 Lyr. This time scale gives a direct measure of the viscous time scale in the outer accretion disk given the expectation that the entire disk is in the hot, viscous state during superoutburst. The resulting constraint on the Shakura-Sunyaev parameter, $\\alpha_{\\rm hot}\\simeq0.1$, is consistent with the value inferred from the fast dwarf nova decays. By looking at the slow decay rate for superoutbursts, which occur in systems below the period gap, in combination with the slow decay rate in one long outburst above the period gap (in U Gem), we infer a steep dependence of the decay rate on orbital period for long outbursts. We argue that this relation implies a steep dependence of $\\alpha_{\\rm cold}$ on orbital period, which may be consistent with recent findings of Patterson, and is consistent with tidal torquing as being the dominant angular momentum transport mechanism in quiescent disks in interacting binary systems. ", "introduction": "Cataclysmic variables (CVs) are semi-detached interacting binaries in which a Roche lobe filling K or M secondary transfers matter to a white dwarf (WD). CVs evolve to shorter orbital periods and show a ``gap'' between $P_h=2$ and 3 (where $P_h = P_{\\rm orbital}/1$ hr) during which time the secondary star loses contact with its Roche lobe and mass transfer ceases. Thus the binary becomes fully detached. At $P_h=2$ the secondary comes back into contact with its Roche lobe and mass transfer resumes. For $P_h < 2$ angular momentum loss from the binary is thought to be due solely to gravitational radiation. The CV subclass of dwarf novae (DNe) are characterized by their semi-periodic outbursts. SU UMa stars are DNe lying below the period gap that exhibit short, normal outbursts (NOs) and superoutbursts (SOs). We refer the time from one SO to the next as the supercycle. SOs show superhumps which are modulations in the light curve at periods slightly exceeding the orbital period. There are two further subdivisions within the SU UMa grouping: (i) the VW Hyi stars at long orbital periods, near $P_h=2$, for which the decay rate is fairly constant during a SO, and (ii) the WZ Sge stars at short orbital periods, a little greater than $P_h=1$, which have less frequent, larger amplitude SOs, for which the decay rate decreases during a SO. DNe outbursts are thought to be due to a limit cycle accretion disk instability (Lasota 2001) in which material is accumulated in quiescence and then dumped onto the WD during outburst. During short outbursts in longer period DNe, a few percent of the stored mass is accreted, and during long outbursts a significant fraction $\\sim$0.2 of the stored mass is accreted. For the SU UMa stars, a SO is thought to accrete $\\ga0.7-0.8$ of the stored mass. Although the accretion disk is never in steady state during the limit cycle, it is close to steady state during SO, with the local rate of accretion inside the disk $|{\\dot M}(r)|$ decreasing linearly from a maximum value at the disk inner edge to zero at the outer edge. The accretion disk modeling has traditionally been done within the Shakura \\& Sunyaev (1973, hereafter SS) formalism, using two values for the $\\alpha$ viscosity parameter, $\\alpha_{\\rm hot}$ for gas in the hot, ionized disk, and $\\alpha_{\\rm cold}$ for gas in the quiescent disk. \\begin{figure*} \\centering \\epsscale{1.0} \\includegraphics[scale=0.75]{figure1.jpg} \\vskip .35cm \\figcaption{ {\\it Kepler} light curve for V1504 Cyg at 1 min cadence, spanning $\\sim$736 d. The light curve covers six superoutbursts and $59$ normal outbursts. The vertical red lines indicate the local maxima for the normal outbursts in which coverage permits a reliable determination, and with well-sampled decays down to $\\sim$2 mag below maximum. \\label{fig1}} \\smallskip \\end{figure*} There are two bright SU UMa stars in the {\\it Kepler} field exhibiting a variety of temporal behavior that make them worthy of a detailed statistical study of their outbursts, V1504 Cyg ({\\it Kepler} ID 7446357; $P_h=1.67$) and V344 Lyr ({\\it Kepler} ID 7659570; $P_h=2.1$). These are members of the VW Hyi subdivision. To date the two light curves have amassed 736.4 d at 1$-$min cadence. Excluding gaps and bad data points, the light curves contain 1000431 and 1000345 data entries, respectively. Previous studies of the {\\it Kepler} data on SU UMa stars have found quiescent superhumps in V344 Lyr (Still et al. 2010), presented numerical models of the long term light curve of V344 Lyr (Cannizzo et al. 2010; hereafter C10), and studied superhumps, both positive and negative, in the long term V344 Lyr light curve (Wood et al. 2011). \\begin{figure*} \\centering \\epsscale{1.0} \\includegraphics[scale=0.75]{figure2.jpg} \\vskip .35cm \\figcaption{ {\\it Kepler} light curve for V344 Lyr at 1 min cadence, spanning $\\sim$736 d. The light curve encompasses six superoutbursts and $58$ normal outbursts. The conventions are the same as in Figure 1. \\label{fig2}} \\smallskip \\end{figure*} Statistical studies of DNe have been useful in delineating the long-term properties of outbursts, e.g.., the outburst duration, recurrence times, and quiescent intervals, and placing physical constraints on models of outbursts (e.g., Campbell \\& Shapley 1940, Sterne, Campbell \\& Shapley 1940, Bath \\& van Paradijs 1983, van Paradijs 1983, Szkody \\& Mattei 1984, Cannizzo \\& Mattei 1992, 1998, Ak, et al. 2002, Simon 2004). Several interesting studies of the SU UMa stars have been carried out. For instance, van Paradijs (1983) studied the variation in outburst duration $t_b$ with orbital period for 15 DNe spanning the $2-3$ hr CV period gap. He found that short outburst durations increase with orbital period, whereas long outburst durations are relatively constant with orbital period. Therefore for the SU UMa systems, which lie at short orbital period, the ratio $t_b({\\rm SO})/t_b({\\rm NO})$ is large. The relation of superoutbursts to normal outbursts for DNe below the period gap and the relation of long outbursts to normal outbursts for DNe above the period gap are part of a general trend; superoutbursts are just long outbursts in short orbital period DNe. This finding was amplified by Ak et al. (2002) using a larger sample. In this work we seek to extend previous studies by analyzing the high fidelity {\\it Kepler} light curves of two SU UMa stars to investigate the properties of the outbursts. In Section 2 we examine the outburst properties of the NOs and SOs, in Section 3 we discuss the results, in particular the scaling of the SO decay rate with orbital period, and in Section 4 we summarize our results. ", "conclusions": "We have analyzed long term {\\it Kepler} light curves of V1504 Cyg, containing 65 outbursts, and V344 Lyr, containing 64 outbursts. Each system showed six superoutbursts. The findings are: \\smallskip (1) The NO decays are faster-than-exponential, with most of the deviation occurring near maximum light. Near quiescence the decays are close to exponential, $\\sim$0.7 d mag$^{-1}$ for V1504 Cyg and $\\sim$0.6 d mag$^{-1}$ for V344 Lyr. These rates are in line with that expected from the Bailey relation. Our values are based on frequency histogram distributions of only $\\sim$50 NOs and should improve as more outbursts are accumulated. \\smallskip (2) The NO outburst durations $t_b$ increase by a factor $\\sim$1.2$-$1.9 over the time interval between consecutive superoutbursts, resetting to a small value after a superoutburst. This is consistent with a monotonic increase of the accretion disk mass with each successive NO. \\smallskip (3) The quiescent intervals $t_q$ between normal outbursts show a general trend in both systems of increasing to a local maximum about mid-way between superoutbursts. (There was also an anomalously long quiescence interval at the start of the V344 Lyr light curve, potentially the result of a tilted disk.) This behavior for $t_q$ is inconsistent with Osaki's thermal tidal instability model for SOs, in which one expects a monotonic increase in $t_q$ values accompanying the monotonic build-up in disk mass and angular momentum leading to a SO. If $t_q$ is correlated with the NO triggering radius $r_{\\rm trig}$, then $r_{\\rm trig}$ moves outward with each successive NO only through roughly the first half of a supercycle, and then recedes. \\smallskip (4) The NOs in both systems are asymmetric, i.e., $t_{\\rm rise}/t_b < 0.5$, consistent with ``Type A'' outbursts (Smak 1984). \\smallskip (5) The inference of the steep relation $\\alpha_{\\rm cold} \\simpropto P_h^{4}$ gives strength to the notion of tidal torquing as the dominant angular momentum transport mechanism operating in quiescent accretion disks in interacting binaries. \\smallskip If one considers the slow decay rate of the long, viscous outbursts spanning DNe from WZ Sge to U Gem, one infers an overall variation $\\simpropto P_h^{1.6}$. It is unfortunate that only one long outburst in one DN longward of the $2-3$ hr period gap exhibited enough dynamic variation in flux during its plateau phase that a decay rate can be reliably extracted for the slow decay portion. It is fortuitous, given the general scatter in DN outburst properties at a given orbital period, that the decay rate between V1504 Cyg and V344 Lyr shows a similar law $\\simpropto P_h^{1.8}$. These dependencies are much steeper than the $P_h^{0.25}$ expected from the previous theoretical expression, due to Warner (1995a, 1995b). By starting with the definition for the viscous time scale in the outer disk during superoutburst, and taking into account the scaling for material accumulated during quiescence, we derive a more physically motivated formula, and find that a steep dependence of the quiescent state value of $\\alpha$ is required, $\\alpha_{\\rm cold}\\simpropto P_h^{4.2}$. In simple terms, the dramatic increase in stored quiescent mass with decreasing orbital period brings about a decrease in the viscous time scale of matter in the superoutburst disk, which is formed out of redistributed gas from the quiescent state. Recent work by Patterson (2011) on the recurrence time scale in SU UMa stars for SOs also supports a steep dependence of $\\alpha_{\\rm cold}$ on $P_h$. The long outbursts in systems above the period gap and superoutbursts in systems below the gap both represent viscous decays, with the absence of transition waves in the disk. It is somewhat counterintuitive that the recurrence time for these long, viscous outbursts varies as $P_h^{-2.6}$, and yet their decay time varies as $P_h^{1.8}$. The steepness of the $\\alpha_{\\rm cold}(P_h)$ relation explains why accretion disk modelers have had to adopt such small $\\alpha_{\\rm cold}$ values for the SU UMa stars, compared to DNe above the period gap, and also probably accounts for the fact that the extreme mass ratio black hole transient systems undergoing limit cycle accretion disk outbursts, like A0620-00, have such long recurrence times for outbursts. The {\\it Kepler} data have already provided important constraints on the physics of accretion disks, in particular the Shakura-Sunyaev $\\alpha$ parameter: C10 showed that an interpolation between $\\alpha_{\\rm cold}$ and $\\alpha_{\\rm hot}$ based on the degree of partial ionization of gas was needed to account for the shoulders in the SOs of V344 Lyr, and in this work we have shown that the scaling of SO decay rate with orbital period mandates a steep dependence of $\\alpha_{\\rm cold}$ on orbital period, enforcing the notion of tidal torquing as being the dominant viscosity mechanism in quiescent accretion disks. \\smallskip We acknowledge the contributions of the entire {\\it Kepler} team. \\def\\mnras{MNRAS} \\def\\apj{ApJ} \\def\\apjs{ApJS} \\def\\apjl{ApJL} \\def\\aj{AJ} \\def\\araa{ARA\\&A} \\def\\aap{A\\&A} \\def\\aapl{A\\&AL} \\def\\pasj{PASJ}" }, "1112/1112.3260_arXiv.txt": { "abstract": "We present a multifrequency approach which optimizes the constraints on cosmological parameters with respect to extragalactic sources and secondary anisotropies contamination on small scales. We model with a minimal number of parameters the expected dominant contaminations in intensity, such as unresolved point sources and the thermal Sunyaev-Zeldovich effect. The model for unresolved point sources, either Poisson distributed or clustered, uses data from {\\it Planck} early results. The overall amplitude of these contributions are included in a Markov Chain Monte Carlo analysis for the estimate of cosmological parameters. We show that our method is robust: as long as the main contaminants are taken into account the constraints on the cosmological parameters are unbiased regardless of the realistic uncertainties on the contaminants. We show also that the two parameters modelling unresolved points sources are not prior dominated. ", "introduction": "The CMB anisotropy data constitute a fundamental tool to test the standard cosmological model and its extensions. In particular small scale CMB anisotropies have a great importance in cosmology and are one of the current frontiers in CMB observations. Many experiments have been dedicated to the observations of small scale anisotropies. The {\\it Planck} satellite \\citep{Bluebook} will cover all the scales up to $\\ell\\sim 2500$ but there are also ground based experiments observing smaller regions of the sky but with higher angular resolutions, such as: ACBAR \\citep{ACBAR}, CBI \\citep{CBI}, QUaD \\citep{QUAD}, SPT \\citep{SPT}, ACT \\citep{ACT}. At these scales, the Silk damping suppresses the primordial CMB contribution with respect to extragalactic contamination and secondary anisotropies. These contaminations affect the constraints on cosmological parameters derived from small scale data introducing biases and therefore it is necessary to properly account for them in the data analysis. We present an approach which has been developed to optimize, with respect to the contamination by point source contributions, the constraints on cosmological parameters which could be obtained from high resolution multi-frequency data.\\\\ The problem of extragalactic sources and secondary anisotropies contamination of the angular power spectrum and consequentely of the cosmological parameters has already been addressed in several recent articles \\citep{Douspis2006,TaburetBias,TaburetSZ,EfstathiouGratton,Millea,IRCTemplate,SZGeorge}, together with the specific tools developed by the data analysis groups of the most recent small scale experiments \\citep{SPT2011,ACT2010}. \\\\ The approach we present in this article consists in modelling the astrophysical contributions to the angular power spectrum with minimal parametrizations, where minimal relates to the number of parameters necessary to describe the signals, and considering the frequency dependences of the different contributions. This is the most important difference with respect to previous approaches \\citep{EfstathiouGratton,Millea,IRCTemplate,SPT2011,ACT2010}. These parametrizations are included as additional contributions to the primary CMB in the Markov Chain MonteCarlo analysis to constrain cosmological parameters. The parameters characterizing the astrophysical contributions are included in the analysis together with the cosmological ones. As described in the following, we focus on the frequencies: $70,100,143,217,353$ GHz. We use the available data from {\\it Planck} early results \\citep{PlanckRadio,PlanckCIB} to derive the parametrizations to describe the astrophysical signals. When no data are available, we rely on predictions from theoretical models and empirical simulations.\\\\ In the following, we use a fiducial cosmological model with the parameters from \\cite{WMAP7} which are reported in the third column of Table 4. The polarization from point sources and SZ effect is negligible we thus do not take it into account. We also do not account for residual signal, in intensity nor in polarization, from our Galaxy. \\\\ The article is organized as follows. In Section 2 we describe the astrophysical signals and we derive the parametrizations of their contributions to the angular power spectrum. In Sect 3 we describe the analysis method and the frequency channel combinations. In Section 4 we present the results of the Markov Chain MonteCarlo (MCMC) analysis for the cosmological and astrophysical parameters. In section 5 we present the discussion and conclusions. ", "conclusions": "In the present study, we developed a multi-frequency approach of the analysis of the CMB power spectrum aiming at obtaining unbiased cosmological parameters in the presence of unresolved extragalactic point sources and TSZ signal. We have considered the three main extra-galactic contributions to the CMB signal at small scales, such as radio sources, infrared sources and SZ clusters. We have modelled these three signals in a minimal way using the observational constraints from {\\it Planck}'s early results \\citep{PlanckRadio,PlanckCIB}. In particular, the radio and IR sources were modelled through a unique Poissonian term and a unique clustering term. In total, three amplitude terms suffice to account for and characterize the extra-galactic contributions. This minimal number of parameters was obtained by considering the frequency dependences within our parametrizations of the TSZ signal, and IR and radio point sources. The use of a data-based physical model including the frequency dependence of the astrophysical signals under consideration to build the parametrizations of the small-scale contribution to the CMB is one of the original points of the present work. Previous studies \\citep{EfstathiouGratton,Millea} chose blind parametrizations of the signals which consider generic shapes and parametrize each frequency with a different set of parameters. Such an approach applied for a frequency channels, leads to either a rapid increase of the number of parameters necessary to describe the astrophysical signals, like in \\cite{Millea}, or it requires the use of a limited number of frequencies, like in \\cite{EfstathiouGratton}. Other studies focused only on one single contribution using more detailed and physical based parametrizations which involve a higher number of parameters with respect to our approach (see for example \\cite{SZGeorge} for the TSZ) or considered the frequency dependence on a much simpler level like for the clustering contribution in \\cite{IRCTemplate}. A different approach, relying directly on the data, has been used by the SPT team \\citep{SPT2011,SPT}. They took advantage of the high resolution to acquire data on scales where the CMB is suppressed. They used the data on smallest scales to measure the amplitudes of their blind templates for the astrophysical signals and consequently to remove the associated contamination from the CMB data. We have applied our multi-frequency minimal parametrization of the extra-galactic contamination to estimate the cosmological parameters for a standard six-parameter $\\Lambda$CDM model and for a model with tensor perturbations and a running of the spectral index, without including residuals from our Galaxy and without using the $B$-mode information. We show that the best combination of frequency is the one which considers the frequencies from 70 to 353 GHz. It allows to estimate the cosmological parameters and at the same to put the tightest constraints on the amplitude of the IR clustering term. We show, both in the standard $\\Lambda$CDM model and in the \"extended-parameter\" model, that the input cosmological parameters together with the astrophysical ones (amplitudes of the TSZ, Poisson and clustering terms) are recovered as unbiased and with slighlty enlarged error bars. We show that the Poisson and clustering terms are degenerate. Our approach instead of blind parametrizations prefers to use the whole available information on point sources and theoretical knowledge of the TSZ to have models of the astrophysical signals which are the most similar as possible to the expected signal. The use of the frequency dependence and the technique used to fit the data allowed a very restricted number of parameters to characterize the signals, in particular only a total of three parameters for all the signals. We have investigated for both cosmological models, standard and extended, the dependence of our approach and of the derived parameters on the accuracy of our model of extra-galactic contributions. In this way we tested its robustness against our theoretical priors on frequency dependence. We tested different Poissonian and clustering contributions by varying the associated terms both in amplitude and in frequency dependence. We showed that for both cosmological models the cosmological parameters are recovered unbiased even in situations where the mock data differ by several sigmas from the model. The astrophysical parameters, amplitudes $A_{\\rm PS}$ and $A_{\\rm CL}$, are recovered well in the case of a ``real'' Poissonian term different from that of the parametrisation. In the case of a different clustering term, the degeneracy with the Poissonian term induces a shift also in the Poissonian term. We have provided a minimal parametrisation to take into account the extragalactic contaminations on small scales. Our minimal analysis can be extended to include additional contributions like residuals from the Galaxy contamination. However, it is ideal in the sense that systematic effects, beam uncertainties etc are not accounted for. \\begin{table} \\centering \\small\\addtolength{\\tabcolsep}{-4pt} \\begin{tabular}{|l|c|c|c|c|c|c|} \\hline Channel & $70$ & $100$ & $143$ & $217$ & $353$\\\\ (GHz) & & & & & \\\\ \\hline $A^{\\mathrm Low}$& 8.06 &7.76 & 7.47 & 7.16 & 6.81 \\\\ $B^{\\mathrm Low}$& $10^{14}$ & $10^{14}$ & $10^{14}$ & $10^{14}$ & $ 10^{14}$ \\\\ $C^{\\mathrm Low}$& 1 & 1 & 1 & 1 &1 \\\\ $\\alpha^{\\mathrm{Low}}$& 0.45 & 0.45 & 0.45 & 0.45 & 0.45 \\\\ $\\beta^{\\mathrm{Low}}$& 2.5 & 2.5 & 2.5 & 2.5 & 2.5 \\\\ $w_{\\nu}^{\\mathrm Low}$& 4.4 & 4.0 & 5.0 & 7.5 & 4.5 \\\\ \\hline $A^{\\mathrm{Med}}$& 306.98 &224.99 & 241.08 & 203.69 & 165.85 \\\\ $B^{\\mathrm{Med}}$& 20408 & 40000 & 111111 & 308642 & 501187 \\\\ $C^{\\mathrm Med}$& 1 & 1 & 1 & 1 &1 \\\\ $\\alpha^{\\mathrm{Med}}$& 0.75 & 0.74 & 0.75 & 0.73 & 0.7 \\\\ $\\beta^{\\mathrm{Med}}$& 2. & 2. & 2. & 2 & 1.9 \\\\ $w_{\\nu}^{\\mathrm Med}$&2.3 & 2.3 & 2.7 & 2.7 & 3.8 \\\\ \\hline $A^{\\mathrm High}$& 58.41 & 22.66 & 70.89 & 23.12 &51.17 \\\\ $B^{\\mathrm High}$& 1. & 0.91 & 3.52 & 1.48 & 0.93 \\\\ $C^{\\mathrm High}$& 1 & 0.25 & 1 & 0.29 &1 \\\\ $\\alpha^{\\mathrm High}$& 0.75 & 0.85 & 0.8 & 0.85 & 0.75 \\\\ $\\beta^{\\mathrm High}$& 1.45 & 1.0 & 1.2 & 1.1 & 1.5 \\\\ $w_{\\nu}^{\\mathrm High}$&1.0 & 1.0 & 1.04 & 1.0 & 1.03 \\\\ \\hline \\end{tabular} \\label{FPOI} \\caption{The exponents $\\alpha$, $\\beta$ and the coefficients $A$, $B$, $C$ and $w_{\\nu}$ for each term of the fits of the radio galaxies composite number counts. } \\end{table} \\begin{table} \\centering \\small\\addtolength{\\tabcolsep}{-4pt} \\begin{tabular}{|l|c|c|c|c|c|c|} \\hline Channel & $70$ & $100$ & $143$ & $217$ & $353$\\\\ (GHz) & & & & & \\\\ \\hline $P_1$& $1.1 10^{-4}$ & $3.6\\times 10^{-5}$ & $1.1\\times 10^{-5}$ & $4.3\\times 10^{-6}$ & $1.8\\times 10^{-5}$ \\\\ $\\chi_1$& 0.95 & 0.95 & 0.95 & 0.95 &0.95 \\\\ $a_1$& 0.34 & 0.34 & 0.34 & & \\\\ $b_1$& 1 & 1 & 1 & 1 & 0.34 \\\\ $c_1$& 1.334 & 1.34 & 1.34 & 1.34 & 1\\\\ $d_1$&$-10^{14}$ &$ -10^{14}$ & $ -10^{14}$ & $-10^{14}$ & $ -10^{14}$ \\\\ $e_1$& 2.5 & 2.8 & 2.8 & 2.8 &2.8 \\\\ \\hline $P_2$&$ 5.9\\times 10^{-3}$ & $1.4 \\times 10^{-3}$ & $ 3.6 \\times 10^{-4}$ & $1.5 \\times 10^{-4}$ & $4.1 \\times 10^{-4}$\\\\ $\\chi_2$&1.28 & 1.24 & 1.2 & 1.15 &1.2 \\\\ $a_2$& 0.64 & 0.62 & 0.6 &0.58 & 0.63\\\\ $b_2$& 1 & 1. & 1. & 1. & 1.\\\\ $c_2$& 1.64 & 1.62 & 1.6 & 1.58 &1.63 \\\\ $d_2$& -20408.2 & -40000 & -111111. &-308642. & -501187 \\\\ $e_2$& 2 & 2 & 2 &2 & 1.9 \\\\ \\hline $P_3$&$2.6\\times 10^{-3}$ &$1.2\\times 10^{-3}$ & $3.6 \\times10^{-4}$ & $2.5 \\times 10^{-4}$ & $4.5 \\times 10^{-4}$ \\\\ $\\chi_3$&1.25 & 1.35 & 1.3 & 1.35 &1.25 \\\\ $a_3$& 0.86 & 1. & 1 & 1. & 0.83\\\\ $b_3$& 1. & 1.35 & 1.08 & 1.22727 & 1.\\\\ $c_3$&1.86 & 2.35 & 2.08 & 2.23 & 1.83\\\\ $d_3$& -1 & -3.64 & -3.52 & -5.10 & -0.93\\\\ $e_3$& 1.45 & 1 & 1.2 & 1.1 & 1.47\\\\ \\hline \\end{tabular} \\label{ClPTerms} \\caption{ The coefficients and exponents of the radio galaxy Poissonian contribution to the angular power spectrum Eq.\\ref{RadioP}. } \\end{table} \\begin{table} \\centering \\small\\addtolength{\\tabcolsep}{-5pt} \\begin{tabular}{|l|c|c|c|c|c|} \\hline $a_i^j$&j & 0 & 1 & 2 & 3 \\\\ \\hline i & & & & & \\\\ \\hline 0 & & 14541.6 & -298.31 & 1.84 &$ - 3.26\\times 10^{-3}$ \\\\ \\hline 1 & & -718.25 &15.14 & $- 9.8\\times 10^{-2}$ & $1.9\\times 10^{-4}$ \\\\ \\hline 2 & & 19.52 & - 0.42 & $2.8\\times 10^{-3}$ &$- 5.99 \\times10^-6$ \\\\ \\hline 3 & & -0.19 & $4.2\\times 10^{-3}$ & $ - 2.8\\times 10^{-5}$ & $6.1\\times 10^{-8} $ \\\\ \\hline 4 & & $ -2.\\times 10^{-4} $ & $4.4\\times 10^{-6}$ & $- 3.03\\times 10^{-8}$ & $6.72\\times 10^{-11}$ \\\\ \\hline 5 & & $ 1.83\\times 10^{-8}$ & $- 3.99\\times 10^{-10} $ & $ 2.78\\times 10^{-12}$ & $6.18\\times 10^{-15}$ \\\\ \\hline 6 & & $ 4.58\\times 10^{-13} $ & $- 9.72\\times 10^{-15}$ & $6.50\\times 10^{-17}$ & $-1.36\\times 10^{-19}$ \\\\ \\hline 7 & & -108.25 & 2.40 & $ - 1.67\\times 10^{-2}$ & $ 3.66\\times 10^{-5}$\\\\ \\hline 8 & & 3.81 & $ - 6.03\\times 10^{-2}$ & $ 3.87\\times 10^{-4} $ & $-6.53\\times 10^{-7}$ \\\\ \\hline 9 & & 2.76 & $- 2.08\\times 10^{-3}$ & $- 1.13\\times 10^{-5} $ & $ 4.07\\times 10^{-8}$ \\\\ \\hline \\end{tabular} \\label{ClusteringTable} \\caption{ The coefficients and exponents of the fits for the clustering term Eq.\\ref{clfit}. } \\end{table} \\begin{figure} \\includegraphics[{width=85mm,height=90mm}]{CMBPSSZCL_70353_RUNGW.eps} \\caption{Results of the analysis which considers together with cosmological and astrophysical parameters also the variation of the tensor to scalar ratio and the running of the spectral index. Comparison between the results varying or not the astrophysical contributions: the black (solid) curve considers the astrophysical signals whereas the red (dashed) curve represent the results with astrophysical contributions fixed to the fiducial value. Vertical bars are the input parameters.} \\label{GWRUNFG} \\end{figure} \\begin{figure} \\includegraphics[scale=0.3]{CMBPSSZCL_70353_RUNGW_tri.eps} \\caption{Results of the analysis which considers together with cosmological and astrophysical parameters also the variation of the tensor to scalar ratio and the running of the spectral index. Comparison between the results varying or not the astrophysical contributions: the black (solid) curve considers the astrophysical signals whereas the red (dashed) curve represent the results with astrophysical contributions fixed to the fiducial value. We show the comparison of the two analysis for the running of the spectral index and the scalar spectral index. Curves are the $68\\%$ and $95\\%$ confidence level.} \\label{GWRUNFG_2D} \\end{figure} Acknowledgements: We wish to thank Nicolas Taburet for useful discussions and for providing the TSZ power spectrum template and Marco Tucci for provinding the number counts of \\cite{Tucci}.This work is partially supported by ASI contract Planck-LFI activity of Phase E2, by LDAP. NA wishes to thank IASF Bologna for hosting and DP thanks IAS for hosting. The simulations for this work have been carried on IASF Bologna cluster. We wish thank the financial support by INFN IS PD51 for NA visit in Bologna. \\begin{figure} \\includegraphics[{width=85mm,height=100mm}]{CMBPSSZCL_70353_FGVariation_RUNGW.ps}\\\\ \\caption{Robustness test part 4. Cosmological and astrophysical parameters with different astrophysical signals with respect to the fiducial model considering also the tensor perturbations and the running of the spectral index. The black (solid) curve are the results of the analysis which considers a Poissonian contribution double its fiducial value. The red (dashed) curve are the results of the analysis which considers the 100, 143 and 217 GHz clustering terms double their fiducial value. The blue (dot-dashed) curve represent the analysis with a clustering term for the 217 GHz half its fiducial value whereas double its fiducial value for 100 and 143 GHz.} \\label{FGV_RG} \\end{figure}" }, "1112/1112.1786_arXiv.txt": { "abstract": "In this paper we investigate gravitationally bound, spherically symmetric equilibrium configurations consisting of ordinary (polytropic) matter nonminimally coupled to an external chameleon scalar field. We show that this system has static, regular, asymptotically flat general relativistic solutions. The properties of these spherical configurations, such as total mass, distribution of matter, and size, depend strongly on the surrounding scalar field. The mass is found in terms of the parameter $\\sigma$, which is proportional to the central mass density of the ordinary matter. We perform a stability analysis of these spherical solutions and find an upper bound for $\\sigma$ where dynamical instability occurs. ", "introduction": "Scalar fields have been applied to a wide range of model building in cosmology and astrophysics. At the cosmological scale scalar fields have been used to model inflation and the late time acceleration due to dark energy. At the galactic scale scalar fields have been tried as models for dark matter (for a review with extensive references to scalar field models of dark energy and dark matter, see \\cite{sahni:2004}). At the astrophysical scale there are boson stars (see the reviews \\cite{Jetzer:1991jr,Schunck:2003kk}). Boson stars are stellar size or smaller, self-gravitating configurations of scalar fields. The scalar fields may be self-interacting or not, and one can include a fluid which interacts with the scalar field either only gravitationally \\cite{Henriques:1989ar,Henriques:1989ez} or through direct coupling \\cite{Lee:1986tr}. The properties of these boson stars strongly depend on the type of scalar fields of which they are composed as well as how these fields interact with each other and the surrounding matter/fluid. In scalar-tensor gravitational theories there is also the possibility of constructing mixed configurations of a scalar field plus ordinary matter. In such theories, a conformal transformation from the original Jordan frame to the Einstein frame is always accompanied by the appearance of a nonminimal coupling between ordinary matter and a scalar field, which leads to new features not present in general relativity. In particular, this nonminimal coupling between the matter of a neutron star and the scalar field can lead to the effect of spontaneous scalarization of the neutron star \\cite{Damour:1993hw,Damour:1996ke}. In turn, this leads to a redistribution of the scalar field inside the neutron star, which has a significant influence on the process of the neutron star's collapse \\cite{Novak:1997hw}. In cosmology, this ability of scalar fields to interact nongravitationally with other fields has been used within the framework of chameleon cosmology \\cite{Khoury:2003aq,Khoury:2003rn,Brax:2004qh}, where a chameleon scalar field $\\phi$ interacts directly with ordinary matter through a conformal coupling of the form $e^{\\alpha \\phi}$. The name ``chameleon scalar field'' was suggested in \\cite{Khoury:2003aq} since the properties of the scalar field, such as its mass, depend sensitively on the environment in which the scalar field finds itself. The idea that the properties of a scalar field could be influenced by the environment/matter surrounding the scalar field was studied earlier, in particular, in the papers \\cite{Ellis:1989as,Mota:2003tc} where the interaction between matter and the scalar field was used to model a dependence of fundamental coupling constants on the local environment. The possibility of having scalar fields interact directly with ordinary matter has been used in~\\cite{Brax:2004qh} to describe the present accelerated expansion of the Universe. These cosmological models were developed further in \\cite{Farajollahi:2010pk}, where the authors choose a generalized expression for the nonminimal coupling between the scalar field and the matter having the form $f(\\phi) L_m$, with $f(\\phi)$ being an arbitrary coupling function and $L_m$ being the Lagrangian of ordinary matter. In this case, for certain parameters of the model, the numerical calculations of \\cite{Farajollahi:2010pk} showed that it was possible to describe the evolving Universe with a contraction phase, a subsequent bounce phase, and finally a phase of accelerated expansion. One can also find analytical solutions \\cite{Cannata:2010qd} within the framework of the cosmological model of \\cite{Farajollahi:2010pk}. These analytical solutions describe the evolution in time of the chameleon scalar field. It is also possible to consider the effect of a chameleon scalar field at astrophysical scales rather than cosmological scales. In \\cite{Dzhunushaliev:2011ma} the influence of a chameleon field on the structure of a star composed of a polytropic fluid nonminimally coupled to a chameleon scalar field was studied. It was found that for special choices of the scalar field potential energy and the coupling function $f(\\phi)$, one could obtain static, regular, asymptotically flat solutions in both relativistic and nonrelativistic limits. However, a preliminary stability analysis of the solutions found in \\cite{Dzhunushaliev:2011ma} indicated that they were not stable in that the matter of the solutions would either collapse or disperse. In \\cite{Folomeev:2011uj} it was shown that taking an isothermal fluid coupled to a chameleon scalar field led to both stable and unstable static solutions, and for the unstable solutions, the problem of the gravitational collapse of the nonrelativistic configurations has been considered. In this paper we continue the study of the influence of a chameleon scalar field on the structure of polytropic spheres begun in \\cite{Dzhunushaliev:2011ma}. We consider a relativistic, polytropic, spherically symmetric fluid embedded in an external chameleon scalar field, which is distributed homogeneously and isotropically over the Universe. When polytropic matter is placed in such a homogeneous background, it begins to interact with the background scalar field both gravitationally and through the nonminimal coupling. This results in a change of the inner structure of the polytropic star depending on the properties of the surrounding chameleon scalar field. In general, the scalar field will evolve over cosmological time, and this will lead to a time dependence for the term $f(\\phi) L_m$, which in turn can have a strong influence on the structure of the compact fluid configuration which is embedded in the scalar field. In this paper we restrict ourselves to only a simplified version of the problem, when we just study changes in the structure of a {\\it static} polytropic, fluid sphere embedded in an external {\\it dynamic} chameleon scalar field. For the time variation of the scalar field we take the time dependence suggested in \\cite{Cannata:2010qd} (see below in Sec.~\\ref{num_res}). The paper is organized as follows: In Sec.~\\ref{gen_equations_cham_star} the general equations describing an equilibrium configuration consisting of a real scalar field coupled to a perfect polytropic fluid are derived. In Sec.~\\ref{static_sol_exp} these equations are written for a particular case when the coupling function $f(\\phi)$ is chosen to be the exponential one. In Sec.~\\ref{num_res} we give the results of the numerical calculations for this choice of nonminimal coupling function. In Sec.~\\ref{stab_gen_exp} we address the issue of the stability of the obtained equilibrium configurations against small radial perturbations. Finally, in Sec.~\\ref{concl_exp} we summarize the main results. ", "conclusions": "\\label{concl_exp} Based on the assumption that time evolving chameleon scalar fields may exist in the Universe, this paper studies the possible influence that such fields may have on the inner structure and stability of compact gravitating configurations consisting of ordinary matter and embedded in an external, homogeneous chameleon scalar field. The ordinary matter is taken to have a polytropic equation of state which is parametrically given by expressions \\eqref{eqs_cham_star}-\\eqref{pressure_fluid_theta}. We choose a specific form for the nonminimal coupling between the ordinary matter and the scalar field given by \\eqref{fun_f_exp}. This form was used in \\cite{Farajollahi:2010pk} within the framework of chameleon cosmology. Here we study the consequences of the presence of a chameleon scalar field on the inner structure of compact astrophysical objects (i.e. ``stars\") rather than the cosmological consequences of chameleon fields. In Ref.~\\cite{Dzhunushaliev:2011ma} this type of gravitating configuration of a polytropic fluid nonminimally coupled to a scalar field was called a ``chameleon star.'' One of the main conclusions of the present work is that the characteristics of the chameleon stars -- their masses, sizes, matter distribution -- depend strongly on the parameters of the surrounding scalar field $\\phi$. Assuming that this background scalar field changes with cosmological times, we predict that the characteristics of the stars embedded in the scalar field should also change. For example, we considered three different values of the background scalar field, $\\phi_{\\text{ext}}=-2, 0, 2$, which could represent the time evolved value of the field at different times according to the law \\eqref{cham_field_evol}. These three values could correspond to three different periods of time in the evolution of the Universe -- from the relatively early Universe to the future Universe, for example, already near the big rip singularity. The results of the numerical calculations presented in Figs.~\\ref{energ_fig}-\\ref{fig_bin_energy_exp} (which cover the values of the polytropic index, $n=1.5$ and $n=3.0$) indicate that the characteristics of the configurations have a strong dependence on the value of the scalar field $\\phi_{\\text{ext}}$. In particular, from Fig.~\\ref{energ_fig} one finds that, at the center of the configurations considered, there can be either a larger or smaller concentration of the matter density compared to the case of a star without a scalar field. This would have an impact on the rate of fusion reactions inside the stars, and thus on their lifetimes so that the life span of the star could be altered by the properties of the background chameleon scalar field surrounding the star. Some other speculations about possible physical applications of the chameleon star model can be found in Ref.~\\cite{Dzhunushaliev:2011ma}. Another important consequence of the presence of the chameleon scalar field is its influence on the stability of the configurations studied here. From previous investigations of the stability of polytropic gravitating fluids it is known that for equilibrium configurations instabilities occur once one moves to the right of the first peak in the binding energy (or the first peak in the mass) \\cite{Tooper2}. A similar situation also occurs for purely boson stars (see the erratum from Ref.~\\cite{Gleiser:1988rq}). However, the dynamical stability analysis performed in Sec.~\\ref{stab_gen_exp} shows that the configurations considered here have their stable points shifted to the right of the first peak in the binding energy. These shifts are shown by the bold dots in Figs.~\\ref{fig_mass_exp} and \\ref{fig_bin_energy_exp}. Thus, in contrast to the usual polytropic and boson stars, the expansion of the region of stability is in the direction where $\\sigma$ increases. It is obvious that this new effect is caused by the presence in the system of the nonminimal coupling between the ordinary matter and the scalar field. The study of mixed configurations with scalar fields and ordinary matter sources is of interest since various types of scalar fields are believed to exist in the Universe. Most often, such scalar fields play a cosmological role~-- as a mechanism to drive inflation (the inflaton field), as dark matter, or as dark energy. Here we consider the possibility that one type of scalar field (a chameleon scalar field) might also have some influence on astrophysical objects such as stars." }, "1112/1112.1265_arXiv.txt": { "abstract": "In order to mimic the phase changes in the primordial Big Bang, several {\\it cosmological} solid-state experiments have been conceived, during the last decade, to investigate the spontaneous symmetry breaking in superconductors and superfluids cooled through their transition temperature. In one of such experiments the number of magnetic flux quanta spontaneously trapped in a superconducting loop was measured by means of a long Josephson tunnel junction built on top of the loop itself. We have analyzed this system and found a number of interesting features not occurring in the conventional case with simply connected electrodes. In particular, the fluxoid quantization results in a frustration of the Josephson phase, which, in turn, reduces the junction critical current. Further, the possible stable states of the system are obtained by a self-consistent application of the principle of minimum energy. The theoretical findings are supported by measurements on a number of samples having different geometrical configuration. The experiments demonstrate that a very large signal-to-noise ratio can be achieved in the flux quanta detection. ", "introduction": "Long Josephson tunnel junctions (LJTJs) were traditionally used to investigate the physics of non-linear phenomena\\cite{barone}. In the last decade they have been employed to shed light on other fundamental concepts in physics such as the symmetry principles and how they are broken\\cite{PRLS,PRB08,gordeeva}. A recent experiment\\cite{PRB09} has demonstrated spontaneous symmetry breaking during the superconducting phase transition of a metal ring and both fluxoids or antifluxoids can be trapped in the ring while it is cooled rapidly through the superconducting critical temperature. The basic phenomenon of quantization of magnetic flux in a multiply connected superconductor was suggested long time ago as one among several possible condensed matter \\textit{cosmological} experiments\\cite{zurek2} suitable to check the validity of the \\textit{causality} principle in the early Universe\\cite{kibble1}. In the experiment of Ref.\\cite{PRB09} the magnetic flux quanta are spontaneously trapped in the ring during its cooling through the transition temperature. Much later at lower temperature when superconductivity is fully established, the number $0,\\pm 1, \\pm 2$ ... of flux quanta is registered as a function of the quench rate. This can be done in a variety of ways, one of which is the detection of the induced persistent currents by the magnetic field modulation of the critical current of a planar LJTJ \\textit{built} on top of the ring. In these experiments, the quench rate can be varied over four decades. This allows for an accurate check of the theoretical predictions of the involved second-order phase transitions. This is of interest within cosmology and of major importance for the physical understanding of many order-disorder processes. However, the working principles of that experiment had not yet been reported. The general task of this work is to study the static properties of a planar LJTJ for which at least one of the superconducting electrodes is multiply-connected, i.e., not every closed path can be transformed into a point. In the simplest case, one of the superconducting thin-film stripes forming the LJTJ is shaped as a doubly-connected loop. This configuration is illustrated in Fig.\\ref{geometry}(a) in which the ring-shaped base electrode is in black, while the top electrode is in gray and the junction area is in white. The geometry of the loop is not critical to our discussion; however, a ring-shaped bottom electrode simplifies the analysis. \\begin{figure}[tb] \\centering \\subfigure[ ]{\\includegraphics[width=7.0cm]{geometrya.png}} \\subfigure[ ]{\\includegraphics[width=7.0cm]{geometryb.png}} \\caption{ a) Sketch of an in-line Josephson tunnel junction with a single doubly connected ring-shaped base electrode. $L_{loop}=L_1+L_2+L_b$. b) Sketch of an in-line Josephson tunnel junction with two doubly connected electrodes. The base electrode is in black, the top electrode is in gray, and the tunneling insulating layer is white. The dotted withe arrows indicate the direction of the circulating currents.\\label{geometry}} \\end{figure} \\noindent For the sake of generality, in our analysis, we will include an external flux $\\Phi_e$ linked to the loop by some externally applied field and the presence of an integer number $n$ of flux quanta trapped in the loop; altogether they induce a current $I_{cir}=(n\\Phi_0-\\Phi_e)/L_{loop}$ circulating clockwise in the loop and inversely proportional to its inductance, $L_{loop}$; in turns, the circulating current produces at the loop surface a radial magnetic field, $H_{rad} \\propto I_{cir}$, that adds to any external field, $H_{app}$, applied in the loop plane. With no loss of generality, we will assume that the width, $W_t$, of the top film does not exceed the width, $W_b$, of the bottom one, $W_t\\leq W_b$ and, to simplify the analysis, we will also assume that both widths are much smaller than the mean radius, $R$, of the ring; in this narrow ring approximation, the current distribution in the ring and the surface radial field become radially independent\\cite{brandt}. A dc current $I$ is injected into the loop at an arbitrary point $O$ along the ring and is inductively split in the two loop arms before going through the LJTJ; let $\\alpha$ (1-$\\alpha$) be the fraction of the bias current $I$ diverted in the left (right) side of the loop. In principle, $\\alpha$ values outside the $[0,1]$ range are possible if the current $I$ would include also a persistent current $I_{cir}$ circulating in the loop: however, since the two currents are independent, we will treat them separately. Independently of the $\\alpha$ value, the bias current $I$ is extracted at one end of the junction via the top electrode. With the current entering and exiting at the junction extremities we have the well-known case of the so-called \\textit{in-line} configuration treated in the pioneering works on LJTJs soon after the discovery of the Josephson effect\\cite{ferrel,OS,stuehm,basa,radparvar85}. \\noindent Throughout the paper we will limit our interest to LJTJs in the zero-voltage time-independent state; this can be achieved as far as the applied current $I$ is smaller than the junction critical current $I_c$. To further simplify the analysis, we assume that the Josephson current density $J_c$ is uniform over the barrier area and that the junction width $W_{}$ is smaller than the Josephson penetration depth $\\lambda _J\\equiv \\sqrt{\\Phi_0/ 2\\pi \\mu _{0}d_{e} J_{c}}$ setting the length unit of the physical processes occurring in the \\Jos \\jun (here $\\Phi_0$ is the magnetic flux quantum, $\\mu_0$ the vacuum permeability and $d_e$ the junction magnetic thickness). The gauge-invariant phase difference $\\phi$ of the order parameters of the superconductors on each side of the tunnel barrier obeys the Josephson equations\\cite{joseph}: \\begin{equation} \\label{jos} J_Z(X)=J_c \\sin \\phi(X), \\end{equation} \\noindent and \\begin{equation} \\label{gra} \\kappa {\\bf \\nabla} \\phi(X) = {\\bf H}\\times {\\bf \\hat{n}}, \\end{equation} \\noindent in which $-{\\rm{L}}/2\\leq X \\leq {\\rm{L}}/2$ is a curvilinear coordinate and ${\\rm{L}}$ is the long dimension of the junction. The net current crossing the tunnel barrier is $I \\equiv W_{} \\int_{-{\\rm{L}}/2}^{{\\rm{L}}/2} J_Z(X)dX$. The last equation states that the phase gradient is everywhere proportional to the local magnetic field ${\\bf H}$ and parallel to the barrier plane. Therefore, in the case of a curvilinear one-dimensional junction, a uniform external field applied in the junction plane has to be replaced by its radial component\\cite{PRB96}. $\\kappa \\equiv {\\Phi_0}/{2\\pi d_e \\mu _0}=J_c \\lambda_J^2$ has the dimension of a current ($\\kappa\\approx 2.5\\,$mA when $d_e \\approx 100\\,$nm, which is typical of all-Niobium \\Jos junctions) and ${\\bf \\hat{n}}$ is the versor normal to the insulating barrier separating the two superconducting electrodes. It is well known\\cite{ferrel,OS} that combining Eqs.(\\ref{jos}) and (\\ref{gra}) with the static Maxwell's equations, a static sine-Gordon equation is obtained that describe the behavior of a one-dimensional in-line LJTJ: \\begin{equation} \\lambda_J^2 \\frac{d^2 \\phi}{d X^2} = \\sin \\phi(X). \\label{sG} \\end{equation} \\noindent Equation(\\ref{sG}) was first introduced in the analysis of \\textit{asymmetric} in-line LJTJs by Ferrel and Prange\\cite{ferrel} in 1963; few years later, Owen and Scalapino\\cite{OS} reported an extensive study of its analytical solutions for \\textit{symmetric} in-line \\Jos junctions (provided that ${\\rm{L}}\\geq\\pi \\lambda_J /2$). Ampere's law applied along the barrier perimeter requires that the magnetic fields at the two ends of the junctions differ by the amount of the enclosed current: $I=W_{} [H_Y ({\\rm{L}}/2) -H_Y (-{\\rm{L}}/2)]$. We remark that Eqs.(\\ref{jos}), (\\ref{gra}) and (\\ref{sG}) automatically satisfy the Ampere's law. \\noindent As it is usually done in the modeling of Josephson interferometers, it is useful to divide the loop inductance $L_{loop}$ in three inductive paths characterized by positive coefficients $L_1$, $L_2$ and $L_b$ having units of inductance such that $L_{loop}=L_1+L_2+L_b$ so that any current $I_{cir}$ circulating around the loop \\textit{sees} them in series; furthermore, $\\alpha=1$ ($\\alpha=0$) in the limit of $L_2>>L_1+L_b$ ($L_2<5 \\lambda_{Lb,t}$). The magnetic penetration $d_e$ of a tunnel barrier with negligible height $t_{ox}<< d_{b,t}$ is\\cite{wei} $d_e \\simeq \\lambda_{b} + \\lambda_{t}$. Insofar as the width $W_b$ is much larger\\cite{swihart,orlando} than the strip-line magnetic thickness $d_e$, then $\\mathcal{L}_b \\simeq \\mu_0 \\lambda_{b}/W_b$; this expression also takes into account the kinetic inductance due to the motion of the superelectrons. In the wide-strip approximation, most of the magnetic energy is confined in the region between the plates and the fringing field can be ignored; as the strip width becomes narrower, the fringe field effects become more important and may dominate if $W_b$ and $d_e$ are comparable\\cite{chang}. It is worth pointing out that, since, in all practical cases, the width of the loop is much larger than the London penetration depth, then $\\mathcal{L}_b$ is considerably smaller than the inductance per unit length along the ring $L_{ring}/2\\pi R$; this is due to the presence of a counter electrode acting as a superconducting ground plane\\cite{chang,vanDuzer}. Similarly, we introduce $\\mathcal{L}_t \\simeq \\mu_0 \\lambda_{t}/W_t$, the inductance per unit length along the current direction of the top plate. Since the electrodes have different widths and penetration depths, in general, $\\mathcal{L}_t \\neq \\mathcal{L}_b$. In Section III we will show that for our high quality all-Niobium LJTJs, having the base electrode thinner and wider than the top one, we found $\\mathcal{L}_t \\approx 3 \\mathcal{L}_b$. According to the theory of the two-conductor transmission lines\\cite{kraus}, the inductance per unit length, $\\mathcal{L}_J$, of a LJTJ, seen as a transmission line structure, is simply obtained as the sum of the inductances/unit lengths of the bottom and top stripes, i.e., \\begin{equation} \\mathcal{L}_J=\\mathcal{L}_b + \\mathcal{L}_t=\\mu_0 \\left( \\frac{\\lambda_{b}}{W_b}+ \\frac{\\lambda_{t}}{W_t} \\right). \\label{indu} \\end{equation} \\noindent Historically, the boundary conditions for Eq.(\\ref{sG}) were derived under the implicit assumption that $W_b = W_t=W_{}$, so that\\cite{scott76,vanDuzer} $\\mathcal{L}_J=\\mu_0 d_e/W_{}$. However these conditions are not fulfilled in real samples, especially for window-type LJTJs used nowadays whose electrodes have quite different widths; typically, $W_b>W_t>W_{}$. \\noindent The paper is organized in the following way. In Section II we will overcome this limitation by extending the existing theoretical model\\cite{OS} to LJTJs having different electrodes widths, strictly speaking, different inductances per unit length. At the same time we will derive the most general boundary conditions for Eq.(\\ref{sG}) needed to correctly describe any self-field effect in LJTJs. Next we will focus on the specific case of LJTJs with doubly connected electrode(s). Later on we will consider the consequences of the fluxoid quantization and energy minimization principles. In the next section we will describe our experimental setup and our samples; in addition we will present their magnetic diffraction patterns and discuss how the experimental findings can be unambiguously interpreted in term of our modeling. Finally, the conclusions will be drawn in Section IV. ", "conclusions": "\\noindent In this paper we have revisited the theory of the self-field effects that characterize the long Josephson tunnel junctions and made them interesting for the investigation of non-linear phenomena. Our analysis goes beyond the previous works in two ways: (i) it takes into account the different inductances per unit length of the electrodes forming the junction and (ii) it provides the boundary conditions for the most general junction biasing configuration. We applied the theory to the specific case of long Josephson tunnel junctions with not simply connected electrodes. Apart from their intriguing physical properties, the interest for LJTJs built on a superconducting loop stems from the fact that they were successfully used to detect trapped fluxoids in a {\\it cosmological} experiment aimed to study the spontaneous defect production during the fast quenching of a superconducting loop through its normal-to superconducting transition temperature\\cite{PRB09}. We found that the single-valuedness of the phase $\\phi_1$ of the order parameter of the bottom (or top) superconducting electrode (fluxoid quantization) gives raise to a variety of unexpected non-linear phenomena when coupled to the sine-Gordon equation for the phase difference $\\phi_2-\\phi_1$ of the order parameters in the superconductors on each side of the tunnel barrier. The principle of energy minimization was also invoked to determine the possible states of the system. We have focused on static phenomena such as the reduction of the junction critical current and its dependence on magnetic fields applied in and out of the loop plane. Nevertheless, also the dynamic properties, such as the propagation of non-linear waves, are expected to be drastically affected by the doubly connected electrode(s). Our experiments unambiguously corroborate the analytical findings and provide hints to implement the modeling. Future work should go in the direction of investigating the consequences of the fluxoid quantization for small and intermediate length Josephson tunnel junctions and on the dynamic properties of long junctions." }, "1112/1112.1862_arXiv.txt": { "abstract": "The relation between redshift and the CMB temperature, $T_{\\rm CMB}(z)=T_0(1+z)$ is a key prediction of standard cosmology, but is violated in many non-standard models. Constraining possible deviations to this law is an effective way to test the $\\Lambda$CDM paradigm and search for hints of new physics. We present state-of-the-art constraints, using both direct and indirect measurements. In particular, we point out that in models where photons can be created or destroyed, not only does the temperature-redshift relation change, but so does the distance duality relation, and these departures from the standard behaviour are related, providing us with an opportunity to improve constraints. We show that current datasets limit possible deviations of the form $T_{\\rm CMB}(z)=T_0(1+z)^{1-\\beta}$ to be $\\beta=0.004\\pm0.016$ up to a redshift $z\\sim 3$. We also discuss how, with the next generation of space and ground-based experiments, these constraints can be improved by more than one order of magnitude. ", "introduction": "Introduction} Cosmology and particle physics are presently experiencing a truly exciting period. On the one hand, both have remarkably successful standard models, which are in agreement with a plethora of experimental and observational data. On the other hand, there are also strong hints that neither of these models is complete, and that new physics may be there, within the reach of the next generation of probes. There are three compelling and firmly established observational facts that the standard model of particle physics fails to account for: neutrino masses, the existence of dark matter, and the size of the baryon asymmetry of the Universe. For each of these, the model makes very specific statements, failing however to reproduce the experimental evidence. It is precisely our confidence in the model and our ability to calculate its consequences that lead us to the conclusion that it is incomplete, and new phenomena must be anticipated. This is, of course, the reason for the LHC project. Similarly, the last decade saw the emergence of the so-called concordance model of cosmology. This can reproduce all the available observations with only a small number of parameters, but also requires that about $96\\%$ of the content of the universe is in a form that has never been seen in the laboratory (and is only known indirectly from its gravitational properties). It is thought that dark matter is a subdominant part of this, while the dominant one is an even more mysterious component usually called dark energy. In this context, it is important to identify laboratory or astrophysical probes that can give us more information about the nature and properties of this still unknown physics. In this work we will discuss one such probe---the temperature-redshift relation---, and lay the foundations for exploring its cosmological implications. One of the most precise measurements in cosmology is the intensity spectrum of the cosmic microwave background radiation: the COBE-FIRAS experiment revealed a very precise black-body spectrum \\cite{Mather}. However, this measurement tells us nothing about the behaviour of the Cosmic Microwave Background (CMB) at non-zero redshift. If the expansion of the Universe is adiabatic and the CMB spectrum was a black-body at the time it originated, this shape will be preserved with its temperature evolving as $T(z)=T_0(1+z)$. This is a robust prediction of standard cosmology, but it is violated in many non-standard models, including string theory motivated scenarios where photons mix with other particles such as axions (see~\\cite{JaeckRing} for a recent review), and those where dimensionless couplings like the fine-structure constant vary \\cite{RoySoc}. A few measurements of $T(z)$ already exist, but the currently large uncertainties do not allow for strong constraints on the underlying models to be set. However, with future datasets this will become a competitive probe. It is therefore timely to discuss what these measurements can tell us about the underlying cosmological paradigms. At low redshifts, say $z<1$, the $T(z)$ relation can be measured via the Sunyaev-Zeldovich (SZ) effect towards galaxy clusters. This method was applied to ground-based CMB observations \\cite{Battistelli,Luzzi}, which demonstrated its potential. With a new generation of ground experiments becoming operational and a forthcoming all-sky survey of SZ clusters to be carried out by Planck \\cite{Horellou}, the potential of this method will come to fruition. At higher redshifts, $z>1$, $T(z)$ can be evaluated from the analysis of quasar absorption line spectra which show atomic and/or ionic fine structure levels excited by the photon absorption of the CMB radiation \\cite{Srianand}. (The CMB is an important source of excitation for species with transitions in the sub-millimeter range.) Although the suggestion is more than four decades old, measurements (as opposed to upper bounds) were only obtained in the last decade, and the best ones so far still have errors at the ten percent level \\cite{Noterdaeme}. Here we will study these issues in detail, but we will also place them in a wider context. For example, in models where photons can be created or destroyed, not only does the temperature-redshift relation vary, but so does the distance duality relation (also known as the Etherington relation \\cite{Etherington1}), and these two different departures from the standard behaviour are quantitatively related. One issue that has been overlooked so far is that in such models, where photon number is not conserved, this relation between $T(z)$ and distance duality provides us with an opportunity to improve constraints. By combining data from different observations one not only reduces the statistical uncertainties on underlying phenomenological parameters but, given the different nature of both observational datasets, one also has a much better control over possible systematics. We therefore discuss in detail the origin of the above relation, as it can be a unique consistency test for the standard paradigm and, at the same time, a valuable tool for probing new physics beyond the standard model. We also study further imprints of these models in the CMB, and present forecasts for improvements that Planck, as well as planned Baryon Acoustic Oscillations (BAO) missions and spectrographs planned for the VLT and the E-ELT, will soon make possible. Last but not least, we derive the strongest constraints to date on deviations of these relations from their standard behaviour, and quantify the improvements to be expected from the aforementioned forthcoming experiments. ", "conclusions": "" }, "1112/1112.4992_arXiv.txt": { "abstract": "{Through an optical campaign performed at 4 telescopes located in the northern and the southern hemispheres, plus archival data from two on-line sky surveys, we have obtained optical spectroscopy for 28 counterparts of unclassified or poorly studied hard X-ray emitting objects detected with Swift/BAT and listed in the 39 months Palermo Swift/BAT hard X-ray catalogue. We have been able to pinpoint the optical counterpart of these high energy sources by means of X-ray observations taken with Swift/XRT or XMM which allowed us to restrict the positional uncertainty from few arcmin to few arcsec; satellite data also provided information on the X-ray spectra of these objects. We find that 7 sources in our sample are Type 1 AGN while 20 are Type 2 AGN, with their redshifts lying between 0.009 and 0.075; the remaining object is a Galactic cataclysmic variable (CV). In this work we provide optical information for all 28 sources and the results of the soft X-ray analysis of 3 out of 5 AGN observed with XMM/Newton.} \\FullConference{The Extreme and Variable High Energy Sky - extremesky2011,\\\\ September 19-23, 2011\\\\ Chia Laguna (Cagliari), Italy} \\begin{document} ", "introduction": "The {\\it Swift} mission was designed to study cosmic gamma-ray bursts (GRBs) in a multiwavelength context ([7]), but it is also able to perform dedicated X-ray and UV-optical observations as well as surveys of the entire sky. {\\it Swift} carries three instruments, i.e. the burst alert telescope (BAT; [1]), the X-ray telescope (XRT; [2]) and the ultraviolet/optical telescope (UVOT; [13]) and therefore can detect and follow up X-ray emitting objects over a wide range of wavelengths. In particular, BAT, the high energy instrument, is a coded mask detector operating with good sensitivity in the energy range 14--195 keV over a field of view of 1.4 sr with a point source location accuracy of $1^{\\prime}-4^{\\prime}$ ([7]) depending on the source intensity. Its sensitivity is estimated to be $\\sim$1 mCrab at high Galactic latitudes and $\\sim$3 mCrab over the Galactic plane. This instrument is not only able to detect GRBs, but also to perform highly sensitive hard X-ray surveys (e.g. [4], [5], [14]). In particular, the BAT surveys allow the study of the extragalactic X-ray sky, and the observation of many absorbed AGNs which are often missed by lower energy instruments. Quantifying the number of such absorbed objects, especially at low redshifts, is very important if one wants to understand the accretion \\mbox{mechanisms} at work in AGNs and to estimate the contribution of all AGN to the cosmic X--ray background ([3]). However, many of the objects listed in the BAT surveys are still unclassified or poorly studied and hence they need optical follow up work to be fully characterized. For this work we have selected from the 39 months Palermo {\\it Swift}/BAT AGN survey ([4]), a group of objects (28 in total) either without optical identification, or not well studied or without published optical spectra. Following the method applied by [8] and references therein or [11] for the optical spectroscopic follow up work of unidentified {\\it INTEGRAL} and/or BAT sources, we determine the nature of these 28 selected objects by means of X-ray observations (to pinpoint the likely X-ray counterpart) and optical measurements (to provide the source classification). In the following sections we show the results obtained with our optical spectroscopic campaigns and we discuss in detail the results of the X-ray analysis of 3 out of 5 objects observed by XMM/Newton. ", "conclusions": "With this work we have been able to either give or confirm or correct the optical classification of 28 {\\it Swift} sources belonging to the Palermo 39 months {\\it Swift} catalogue (see also [12] for details). This was achieved through a multisite observational campaign in Europe, South Africa, Central and South America. We found that our sample is composed of 27 AGNs (7 of Type 1 and 20 of Type 2), with redshifts between 0.009 and 0.075, and 1 CV. Among these sources we found some peculiar objects, such as 3 likely LINERs and 1 narrow line seyfert 1. The X-ray spectral analysis of 3 out of 5 sources observed with XMM-Newton shows a complex best-fit model with an absorbed power law component as a primary emission model; all 3 require an extra component at low energies to fit an emission excess below few keV, while only two display iron line emission features. This work shows the importance of optical spectroscopic follow up observations for sources discovered by hard X-ray surveys and either unclassified or poorly studied. By increasing the number of the identifications in hard X-ray catalogues, it is possible to perform more reliable statistical studies as multiwavelength characterization of the sources, thus allowing a better understanding of the physical processes that drive the powerful AGN detected." }, "1112/1112.1053_arXiv.txt": { "abstract": "The IceCube Neutrino Observatory is the world's largest high energy neutrino telescope, using the Antarctic ice cap as a Cherenkov detector medium. DeepCore, the low energy extension to IceCube, is an infill array with a fiducial volume of around 30 MTon in the deepest, clearest ice, aiming for an energy threshold as low as 10 GeV and extending IceCube's sensitivity to indirect dark matter searches and atmospheric neutrino oscillation physics. We will discuss the analysis of the first year of DeepCore data, as well as ideas for a further extension of the particle physics program in the ice with a future PINGU detector. ", "introduction": "\\label{sec:intro} The IceCube neutrino telescope, now fully operational at depths of 1450-2450 m below the surface of the Antarctic ice cap, was designed to detect high energy neutrinos from astrophysical accelerators of cosmic rays. Although the energy threshold of a large volume neutrino detector is not a sharp function, the original IceCube design focused on efficiency for neutrinos at TeV energies and above. Recently, the IceCube collaboration decided to augment the response of the detector at lower energies with the addition of DeepCore, a fully contained subarray aimed at improving the sensitivity of IceCube to neutrinos with energies in the range of 10's of GeV to a few hundred GeV. This energy range is of interest for several topics related to particle physics, including measurements of neutrino oscillations and searches for neutrinos produced in the annihilation or decay of dark matter. DeepCore consists of an additional eight strings of photosensors (Digital Optical Modules, or DOMs) comprising 10'' Hamamatsu photomultiplier tubes and associated data acquisition electronics housed in standard IceCube glass pressure vessels. For most of the DeepCore DOMs, the standard IceCube R7081 PMTs were replaced with 7081MOD PMTs with Hamamatsu's new super-bialkali photocathode. These PMTs provide approximately 35\\% higher quantum efficiency (averaged over the detected Cherenkov spectrum) than the standard bialkali PMTs. Sited at the bottom center of the IceCube array, DeepCore benefits from the high optical quality of the ice at depths of 2100-2450 m, with an attenuation length of approximately 50 m in the blue wavelengths at which most Cherenkov photons are detected in ice. DeepCore also benefits from the ability of the standard IceCube sensors to detect atmospheric muons penetrating the ice from cosmic ray air showers above the detector, allowing substantial reduction in the background rate by vetoing events where traces of penetrating muons are seen. Each DeepCore string bears 50 DOMs in the fiducial region, with an additional 10 DOMs deployed at shallower depths to improve the vetoing efficiency for steeply vertical muons. In addition to the new DeepCore strings, the DeepCore fiducial volume for analysis includes 12 standard IceCube strings, chosen so that the fiducial region is shielded on all sides by a veto region consisting of three rows of standard IceCube strings, as shown in Fig.~\\ref{fig:layout}. The random noise rate of IceCube DOMs is quite low (around 500 Hz, on average) due to the low temperatures and radiopurity of the ice cap. This permits DeepCore to be operated with a very low trigger threshold, demanding that 3 DOMs within the DeepCore fiducial region detect light in ``local coincidence'' within a period of no more than 2500 ns. The local coincidence criterion counts DOMs as being hit (i.e., having detected light) only if one of the four neighboring DOMs on a string (two above and two below) also registers a hit within $\\pm1 \\;\\mu$s. Most of the resulting 185 Hz of triggers are due to stray light from muons which simultaneously satisfy the main IceCube trigger condition of 8 DOMs hit in local coincidence within 5 $\\mu$s, but the DeepCore trigger contributes an additional (exclusive) rate of around 10 Hz. \\begin{figure} \\begin{center} \\includegraphics[width=0.8\\columnwidth]{icecube_deepcore_pingu_koskinen_new.pdf} \\caption{Schematic layout of DeepCore within IceCube. The shaded region indicates the fiducal volume of DeepCore, at the bottom center of IceCube, plus the extra veto cap of DOMs deployed at shallower depths to reinforce the veto against vertically-downgoing atmospheric muons. This schematic depicts both the DeepCore configuration used in 2010, when 79 IceCube strings were operational, and the final DeepCore layout and fiducial region used in the 2011 run. \\label{fig:layout}} \\end{center} \\end{figure} The vast majority of the events which trigger DeepCore, irrespective of whether they also trigger IceCube, are due to either penetrating atmospheric muons or random coincidences of dark noise. Immediately after data acquisition, events triggering DeepCore are subjected to an online data rejection algorithm which calculates a characteristic time and location for the activity observed in the DeepCore fiducial region, as an initial estimate of the putative neutrino vertex. The estimated location is the average position of the hit DOMs, and the time is determined by subtracting the time of flight $dn/c$ of an unscattered photon emitted from that location from the observed arrival time of the first photon to hit each DOM. After outliers due to dark noise or scattered light are removed, the average inferred emission time is used as the estimated time of the underlying physics event. Based on this estimated time and location, every locally coincident hit recorded in the veto region prior to the vertex time is examined to determine whether it lies on the light cone connecting it with the estimated event vertex. The distributions of the inferred speed required to connect hits in the veto region to the DeepCore vertex, for both simulated atmospheric muons and simulated neutrinos, is shown in Fig.~\\ref{fig:veto}; positive speeds indicate hits occuring in the veto region prior to the DeepCore vertex time. If any hits are found with inferred speeds between +0.25 and +0.4 m/ns, the event is rejected as being most likely due to an atmospheric muon. This algorithm reduces the event rate by more than two orders of magnitude, to 18 Hz, while retaining over 99\\% of simulated triggered events due to neutrinos interacting within the fiducial volume. Additional background rejection criteria are applied offline, depending on the goals of each physics analysis making use of these data. \\begin{figure} \\includegraphics[width=\\columnwidth]{ParticleSpeedProbabilities_v22.pdf} \\caption{Distribution of probabilities of observing hits leading to a given inferred particle speed, for simulated atmospheric muons (dashed line) and atmospheric neutrinos (solid). Positive speeds indicate activity in the veto region prior to that in the DeepCore volume, and a peak around $c =$ 0.3 m/ns is visible for penetrating muons. The integral of each distribution corresponds to the mean number of hits observed in the veto region for the given class of events. \\label{fig:veto}} \\end{figure} The effective volume of the DeepCore detector for detection low energy muon neutrinos, accounting for this online data filter, is shown in Fig.~\\ref{fig:nuMuVolume}. It should be stressed that this effective volume curve does \\emph{not} include losses due to later background rejection or event quality criteria. The contribution of DeepCore to low energy analysis is evident in the fact that despite its relatively small geometric volume, around 3\\% that of IceCube, the overall sample of neutrino events below 100 GeV consists primarily of those detected by DeepCore. This energy range is of considerable interest for several topics in particle physics, including searches for dark matter and measurements of neutrino oscillations. While DeepCore does not have a sharp energy threshold, it retains around 7 megatons of effective volume at energies as low as 10 GeV. Further details regarding DeepCore's instrumentation and performance are available in Ref.~\\cite{Collaboration:2011ym}. \\begin{figure} \\includegraphics[width=\\columnwidth]{effectiveVolume_IC86_numu_GENIE_effVolumes_logScale_prelim.pdf} \\caption{Effective volume of DeepCore for muon neutrinos at trigger level (solid) and after application of the online veto algorithm described in the text (dot-dashed line). The effective volume of IceCube as originally proposed is shown for comparison. \\label{fig:nuMuVolume}} \\end{figure} ", "conclusions": "The effectiveness of IceCube at energies below 100 GeV has been significantly enhanced by the addition of DeepCore, which extends IceCube's reach to energies of 10's of GeV. This range is of interest for observations of neutrino oscillations, as well as searches for dark matter. As a first step toward these studies, we have observed a significant sample of atmospheric neutrino-induced cascades, enabled by the ability of the IceCube detector to identify and veto atmospheric muons penetrating to the DeepCore volume. We are also investigating the potential for a further reduction in the energy threshold of IceCube with an additional extension known as PINGU, which could extend IceCube's reach to energies as low as a few GeV" }, "1112/1112.5994_arXiv.txt": { "abstract": "Spectroscopic monitoring with Mercator-HERMES over the past two years reveals that MWC~314 is a massive binary system composed of an early B-type primary LBV star and a less-luminous supergiant companion. We determine an orbital period $\\rm P_{\\rm orb}$ of 60.85 d from optical S~{\\sc ii} and Ne~{\\sc i} absorption lines observed in this single-lined spectroscopic binary. We find an orbital eccentricity of $e$=0.26, and a large amplitude of the radial velocity curve of 80.6 $\\rm km\\,s^{-1}$. The ASAS $V$ light-curve during our spectroscopic monitoring reveals two brightness minima ($\\Delta$$V$$\\simeq$$0^{\\rm m}$.1) over the orbital period due to partial eclipses at an orbital inclination angle of $\\sim$70$\\deg$. We find a clear correlation between the orbital phases and the detailed shapes of optical and near-IR P Cygni-type line profiles of He~{\\sc i}, Si~{\\sc ii}, and double- or triple-peaked stationary cores of prominent Fe~{\\sc ii} emission lines. A preliminary 3-D radiative transfer model computed with {\\sc Wind3D} shows that the periodic P Cygni line profile variability results from an asymmetric common-envelope wind with enhanced density (or line opacity) in the vicinity of the LBV primary. The variable orientation of the inner LBV wind region due to the orbital motion produces variable P Cygni line profiles (with wind velocities of $\\sim$200~$\\rm km\\,s^{-1}$) between orbital phases $\\phi$ = 0.65 to 0.85, while weak inverse P Cygni profiles are observed half an orbital period later around $\\phi$ = 0.15 to 0.35. We do not observe optical or near-IR He~{\\sc ii}, C~{\\sc iii}, and Si~{\\sc iii} lines, signaling that the LBV's spectral type is later than B0. Detailed modeling of the asymmetrical wind properties of massive binary MWC~314 provides important new physical information about the most luminous hot (binary) stars such as $\\eta$~Carinae. ", "introduction": "MWC~314 (V1492~Aql; BD$+$14$\\deg$3887; $V$=$9^{\\rm m}$.9) is a candidate Luminous Blue Variable (LBV) previously proposed to be one of the most luminous stars of the Galaxy by \\citet{miro1} with log(${L}_{\\star}/{L}_{\\odot}$)$\\simeq$6.1$\\pm$0.3, $T_{\\rm eff}$$\\simeq$25 to 30 kK, and $\\dot{M}$$\\sim$ $\\rm 3\\,10^{-5}$ $\\rm M_{\\odot}\\,{yr}^{-1}$. More recently, \\citet{mura1} found that the star is a binary system with an orbital period of $\\sim$30 d using optical spectra, however without determining other orbital parameters. We therefore observed 12 high-resolution spectra over the past two years with Mercator-HERMES ($R$=80,000) at La Palma (Spain). HERMES is a high-efficiency {\\'e}chelle spectrograph covering 420 nm to 900 nm \\citep{rask1}. We observed the spectra of MWC~314 with large SNR$\\sim$100 for accurate radial velocity (RV) measurements and detailed line profile studies. On 5 \\& 9 Sep 2009, and on 17 \\& 20 Mar 2011 we also observed two spectra within 5 d to investigate possible short-time spectroscopic variability in MWC~314 \\citep[see also][]{lobe1}. ", "conclusions": "Based on long-term spectroscopic monitoring with Mercator-HERMES we confirm the binarity of candidate LBV MWC 314, first conjectured by Muratorio et al. in 2008. However, we determine an orbital period of 60.85 d (twice longer than $P_{\\rm orb}$$\\sim$1 m they proposed) from an accurate solution of the RV-curve. We also compute an orbital eccentricity of 0.26 with LBV periastron passage oriented almost towards the observer. The visual brightness curve reveals two unequal minima signaling partial eclipses at an orbital inclination angle of $\\sim$70$\\deg$ in the plane of the sky. We also confirm the LBV character of MWC 314 with strong P Cygni-type line profiles observed during the orbital phases of fastest approach around periastron passage. A 3-D radiative transfer model we compute for the wind of MWC~314 shows that the P Cyg profiles result from enhanced LBV wind density inside a circumbinary expanding wind envelope." }, "1112/1112.2785_arXiv.txt": { "abstract": "In this paper we study the non-Gaussian features of the primordial fluctuations in loop quantum cosmology with the inverse volume corrections. The detailed analysis is performed in the single field slow-roll inflationary models. However, our results reflect the universal characteristics of bispectrum in loop quantum cosmology. The main corrections to the scalar bispectrum come from two aspects: one is the modifications to the standard Bunch-Davies vacuum, the other is the corrections to the background dependent variables, such as slow-roll parameters. Our calculations show that the loop quantum corrections make $f_{{\\rm NL}}$ of the inflationary models increase $0.1\\%$. Moreover, we find that two new shapes of non-Gaussian signal arise, which we name $\\mathcal F_{1}$ and $\\mathcal F_{2}$. The former gives a unique loop quantum feature which is less correlated with the local, equilateral and single types, while the latter is highly correlated with the local one. ", "introduction": "As a non-perturbative and background-independent theory, Loop Quantum Gravity (LQG) \\cite{rovelli04,ashtekar04,thiemann07} has achieved great successes in past years: derivations of the quantized area and volume operators \\cite{rovelli95,ashtekar97,ashtekar98,thiemann98a}, calculations of black holes entropy \\cite{rovelli96} and Loop Quantum Cosmology(LQC) \\cite{bojowald08a}, {\\it etc.} And the nonperturbative quantization procedure of LQG is also valid for a more general class of four-dimensional metric theories of gravity~\\cite{zhang11a,zhang11b,zhang11c}. As an example of LQG, LQC gives a quantization scheme of LQG for a symmetry-reduced model in the homogeneous and isotropic Friedmann-Lema\\^ itre-Robertson-Walker universe. The discrete spacetime geometry in LQC scenarios predicts a non-singular bouncing universe in some simplified models, which satisfies most of the astronomical and cosmological observational constraints. Although the quantum correction effects are being diluted with the expansion of our universe, it remains present in a weaker form, especially on/near the super-horizon scales. Recently, the gauge invariant cosmological perturbation theory has been systematically constructed in \\cite{bojowald06,bojowald08,bojowald09} for inverse volume corrections and in \\cite{wu10,Cailleteau11} for holonomy corrections. Some relevant applications have been considered in~\\cite{bojowald11a,bojowald11b,bojowald11c,zhu11}. The inverse volume corrections and the holonomy corrections are two main quantum corrections in LQC. The inverse volume corrections come from the quantization of the inverse of the volume operator in LQG. The inverse volumes exist in the Hamiltonian constraint of gravity and the usual matter Hamiltonian, especially in the kinematic terms. Since the volume can be taken the value zero, there does not exist well defined inverses of the volume operator. Fortunately, with the Thiemann trick~\\cite{thiemann98}, we can construct well defined inverse volume operators, which bring the quantum corrections. While the holonomy corrections arise from the loop quantization based on the holonomies instead of the direct connection. The holonomy corrections become important when the energy scale of our Universe approaches the Planck one. Both the modifications to the scalar and tensor primordial power spectra from the inverse volume corrections are carefully investigated by the authors of \\cite{bojowald11a}. Their results show that the inverse volume corrections could give rise to the enhancement of the power spectra on the large scales, i.e., a red-tilt one. However, some other mechanisms such as the non-commutative geometry or string theory~\\cite{tsujikawa03,piao04,calcagni04}, could also lead to similar features. Therefore to seek for signature of loop quantum cosmology, the study for the non-Gaussianity features in loop quantum cosmology is necessary. Because the primordial non-Gaussianities are quite helpful to distinguish inflationary models, so far a lot of papers have been devoted to studying the non-Gaussianities in different inflation models, see the relevant references in the nice reviews \\cite{Bartolo:2004if,Chen:2010xk}. Inspired by the studies of \\cite{bojowald11a}, in this paper we mainly consider the non-Gaussianities from the inverse volume corrections in LQC. The reason for consider inverse volume corrections only is as following. We denote $\\delta_{\\rm inv}$ as the correction term coming from the inverse volume operator and $\\delta_{{\\rm hol}}$ as the correction term from the holonomy corrections. We can estimate the inverse volume correction as\\cite{bojowald11c} \\begin{equation}\\label{compare} \\delta_{{\\rm inv}}\\sim\\left(\\frac{8\\pi}{3}\\frac{\\rho}{\\rho_{{\\rm Pl}}}\\delta_{{\\rm hol}}^{-1}\\right)^2\\;, \\end{equation} where the Planck density $\\rho_{{\\rm Pl}}$ is assumed as the quantum gravity scale. From the above expression, we can see that the inverse volume corrections behave very differently from what is normally expected for quantum gravity. For low densities, the holonomy corrections is small, but the inverse volume one may still be large because they are magnified by the inverse of $\\delta_{{\\rm hol}}$. For an example, the small holonomy corrections of size $\\delta_{{\\rm hol}}<10^{-6}$ then requires the inverse volume corrections larger than $\\delta_{{\\rm inv}}>10^{-6}$ even at scale $\\rho\\approx10^{-9}\\rho_{\\rm Pl}$. These novel features make the investigations on the inverse volume corrections more interesting than the former at sub-Planckian inflationary scales. So we only consider the inverse volume correction in this work. Explicitly, in the perturbation theory in LQC, the inverse volume operator can be captured by a correction function such as $\\bar{\\alpha}\\simeq1+\\alpha_0\\delta_{\\rm inv}\\simeq 1+\\alpha_0(a_{\\rm inv}/a)^{\\sigma}$, where $a$ is the scale factor of the FLRW universe and $a_{\\rm inv}$ is introduced to describe the characteristic scale of the inverse volume correction, which is not the Planck one in general. When $a_{\\rm inv}/{a}\\ll 1$, we can ignore the correction term. However, if $a_{inv}/a \\lesssim 1$ during inflation, one cannot neglect inverse volume corrections. In this case, the correction approximates $\\alpha_0\\delta_{\\rm inv}(k)\\approx\\delta(k_0)(k_0/k)^{\\sigma}$, where $k$ and $k_0$ are, respectively, the considered perturbation wave number and some characteristic number involved in the inverse volume correction. In addition, many works about LQC imply that $\\sigma\\in[0,6]$ \\cite{bojowald11b}. From the form of the inverse volume correction, we can see that a small sigma corresponds to a small the inverse volume correct, vice versa. Therefore, as an example, following \\cite{bojowald11b}, we take $\\sigma=2$ in this paper. For other $\\sigma$ values the behavior will be similar. Furthermore, in terms of spherical multiples the wave number could be expressed as $k\\approx10^{-4}hl$, with $h$ representing for the reduced Hubble parameter $h=0.7$ and $l$ for the spherical multiples. In the typical linear regime of Cosmic Microwave Background (CMB), the multiples $l$ range in $2