{ "1402/1402.6308_arXiv.txt": { "abstract": "We present a {\\it Spitzer} MIPS study of the decay of debris disk excesses at 24 and 70 $\\mu$m for 255 stars of types F4 - K2. We have used multiple tests, including consistency between chromospheric and X-ray activity and placement on the HR diagram, to assign accurate stellar ages. Within this spectral type range, at 24 $\\mu$m, $13.6 \\pm 2.8 \\%$ of the stars younger than 5 Gyr have excesses at the 3$\\sigma$ level or more, while none of the older stars do, confirming previous work. At 70 $\\mu$m, $22.5 \\pm 3.6\\%$ of the younger stars have excesses at $ \\ge$ 3 $\\sigma$ significance, while only $4.7^{+3.7}_{-2.2}$\\% of the older stars do. To characterize the far infrared behavior of debris disks more robustly, we double the sample by including stars from the DEBRIS and DUNES surveys. For the F4 - K4 stars in this combined sample, there is only a weak (statistically not significant) trend in the incidence of far infrared excess with spectral type (detected fractions of 21.9$^{+4.8}_{-4.3}\\%$, late F; 16.5$^{+3.9}_{-3.3}\\%$, G; and 16.9$^{+6.3}_{-5.0}\\%$, early K). Taking this spectral type range together, there is a significant decline between 3 and 4.5 Gyr in the incidence of excesses with fractional luminosities just under $10^{-5}$. There is an indication that the timescale for decay of infrared excesses varies roughly inversely with the fractional lumnosity. This behavior is consistent with theoretical expectations for passive evolution. However, more excesses are detected around the oldest stars than is expected from passive evolution, suggesting that there is late-phase dynamical activity around these stars. ", "introduction": "Understanding planetary system formation and evolution is one of the major initiatives in astronomy. Stars form surrounded by protoplanetary disks of primordial gas and dust where planets grow. The material in these disks that does not fall into the star either collects into planets or is dissipated by processes such as photo-evaporation \\citep[e.g.,][]{clarke2001} and tidal forces from planets \\citep[e.g.,][]{bryden1999}, typically in less than 10 Myr \\citep{williams2011}. The evolution of the systems is not complete, as shown by the events that led to the formation of our Moon and to the Late Heavy Bombardment \\citep{tera1974}, long after the protoplanetary disk cleared from the Sun. It is very difficult to observe these later stages of evolution directly. However, after dissipation of the protoplanetary disks, a relatively low level of dust production can be sustained through debris produced in planetesimal collisions to form planetary debris disks \\citep{wyatt2008}, which can be detected readily in the infrared over the entire range of stellar lifetime (to 10 Gyr). Debris disks are our best current means of studying planet system evolution over its entire duration. About 20\\% of the nearby stars harbor debris disks above current detection limits \\citep{habing2001,trilling2008,carpenter2009,gaspar2013,eiroa2013}. The likelihood of a detectable debris disk at 24 $\\mu$m depends on age, with a higher percentage around young stars than older ones \\citep[e.g.,][]{habing1999,rieke2005,gaspar2013}. The expected timescale for evolution of the disk components that dominate in the far infrared indicates that, depending on disk location and density, there might also be a detectable decay in disk incidence at 70 $\\mu$m \\citep{wyatt2008}. However, it has proven difficult to confirm this prediction definitively \\citep{wyatt2008, bryden2006, trilling2008,carpenter2009}. \\citet{gaspar2013} have recently demonstrated a drop in infrared excess on the basis of comparison with the predictions of a theoretical model. If such a decay can be confirmed by an alternative analysis, it would help substantially to constrain models and particle properties of debris disks. In this paper, we explore debris disk evolution in the far infrared (70 - 100 $\\mu$m) using a large sample of stars to allow reaching statistically robust conclusions. The paper presents high-quality homogeneous data reductions for a large sample of stars observed with {\\it Spitzer}, analyzes their behavior, and then combines this sample with the {\\it Herschel}-observed DEBRIS \\citep{matthews2010} and DUNES \\citep{eiroa2013} samples. We take care in determining stellar ages, since the apparent rate of decay can be strongly influenced by mis-classification of the ages (a small number of young stars mistakenly identified as old ones might dominate the detections among the \"old\" sample). Section 2 describes the selection of the {\\it Spitzer} sample, the determination of stellar ages, and the reduction of the infrared measurements. We determined debris-disk-emitted excesses as described in Section 3. Section 4 presents the analysis of the behavior of these excesses with stellar age, and Section 5 merges this sample and additional {\\it Spitzer} data with {\\it Herschel} observations to study debris disk evolution in a total sample of about 470 stars. Our conclusions are summarized in Section 6. ", "conclusions": "We present a sample of 238 F4 - K2 stars with high-quality {\\it Spitzer} observations through 70 $\\mu$m, and for which we have been able to determine accurate ages from the chromospheric activity index, X-ray emission, and placement on an HR-diagram. Consistent with previous work, we find that the incidence of excesses at 24 $\\mu$m decays dramatically from $13.6 \\pm 2.8$ \\% for stars $<$ 5 Gyr in age to none for older stars. At 70 $\\mu$m, we also find a significant reduction in the incidence of excesses, from 38, or $22.5 \\pm 3.6$ \\% to 4, or $4.7^{+3.7}_{-2.2}$ \\%. We have evaluated the significance of this apparent reduction in the incidence of 70 $\\mu$m excesses in two ways, both of which correctly account for the variable detection limits and resulting censoring of the data. The first is the traditional Kaplan-Meier test, which indicates a decay by greater than 2$\\sigma$ significance, nominally by a factor of 4.5. In the second method, we determined the distribution of excess ratios among the young (ages $<$ 5 Gyr) stars and assumed it could be treated as a probability distribution to determine if a specific star is likely to have an excess above its individual detection limit. This model was used to predict the number of excesses that would have been detected in the old ($>$ 5 Gyr) sample if they behaved identically to the young sample. This approach also predicts a decay at greater than 2 $\\sigma$ significance, and nominally by a factor of five. The two methods therefore agree well on the existence and amount of this decay. Within this sample, there are two stars with excesses at 24 $\\mu$m but no strong excess at 70 $\\mu$m, HD 1466 and HD 13246. We then combined the {\\it Spitzer}-only stars with the measurements for the DEBRIS and DUNES programs (for the former using both {\\it Spitzer} and {\\it Herschel} data to improve the weight of the measurements). This combined sample is doubled in size, allowing us to reach robust conclusions about debris disk evolution in the far infrared. Within this larger sample, we examine the relative incidence of excesses around stars of types F4 - K4, finding no statistically significant trend with spectral type. We therefore analyze the evolution of the entire sample together. We find that, at fractional debris disk luminosity of just under $10^{-5}$, the incidence of far infrared excesses declines significantly between 3000 and 5000 Myr, while for fractional luminosity of just under $10^{-4}$, the decline is at an order of magnitude younger ages. This behavior is consistent with a theoretical model for passive disk evolution. However, there appear to be more excesses for the oldest stars than predicted by passive-evolution models. These excesses are distributed over the full range of spectral types in our sample. This behavior suggests that the oldest debris disks are activated by late-phase dynamical activity." }, "1402/1402.0697_arXiv.txt": { "abstract": "{Halo masses and concentrations have been studied extensively, by means of N--body simulations as well as observationally, during the last decade. Nevertheless, the exact form of the mass--concentration relation is still widely debated. One of the most promising method to estimate masses and concentrations relies on gravitational lensing from massive halos. Here we investigate the impact of the mass--concentration relation on halo peak abundance in weak lensing shear maps relying on the aperture mass method for peak detections. After providing a prescription to take into account the concentration dispersion (always neglected in previous works) in peak number counts predictions, we assess their power to constrain the mass--concentration relation by means of Fisher matrix technique. We find that, when combined with different cosmological probes, peak statistics information from near--future weak lensing surveys provides an interesting and complementary alternative method to lessen the long standing controversy about the mass-concentration relation. } ", "introduction": "Weak gravitational lensing (WL) has emerged over the last decade as one of the most promising methods for testing cosmology and gravity, and unveiling the nature of Dark Energy (DE) and Dark Matter (DM). It relies on the accurate measurement of the small shape distortions of background galaxies due to the bending of light by intervening matter distribution. The smallness of these distortions requires high-quality data and a statistical approach so that the lensing shear signal needs to be measured on a large number of sources. A number of surveys (e.g. GaBoDS \\cite{Hett07}, CFHTLS\\footnote{http://www.cfht.hawaii.edu/Science/CFHLS}) have already shown the WL capacity for constraining cosmological models through cosmic shear measurements (see e.g. \\cite{Hoekstra06,Semboloni06,Hett07}). Several other, more ambitious, surveys both ground-- (KIDS\\footnote{http://kids.strw.leidenuniv.nl}, PanSTARRS\\footnote{http://pan-starrs.ifa.hawaii.edu}, DES\\footnote{http://www.darkenergysurvey.org}, LSST\\footnote{http://www.lsst.org}) and space-based (Euclid\\footnote{http://www.euclid-ec.org}, WFIRST\\footnote{http://wfirst.gsfc.nasa.gov}) are being performed or planned. They will map hundreds of millions of galaxy redshifts and billions of galaxy images. For instance, the Euclid mission \\cite{euclid1,euclid2} aims to map half of the sky in imaging as well as in broad-band spectroscopy up to a redshift $z \\sim 2-3$ with a median redshift of the order of unity. The information contained in these data will permit to measure the matter clustering with unprecedented accuracy. Among the WL probes, it has been shown that the abundance of peaks in shear maps is a sensitive probe of cosmology \\cite{kruse99}. This was tested by \\cite{Reblinsky99} and \\cite{Dietrich10} using ray--tracing in N--body simulations. Shear peaks are regions with high signal--to--noise ratios, $snr$, associated to massive halos (or galaxy clusters) or produced by the alignment of smaller mass concentrations along the line--of--sight. Their number and spatial distribution, therefore, carry information about the underlying cosmology through fundamental parameters as the total matter density of the universe $\\Omega_m$, the normalization of the power spectrum $\\sigma_8$, and the evolution of the DE equation of state $w(z)$. Unlike other available techniques for detecting massive halos (e.g. optical and X--ray identification, Sunyaev-Zeldovich effect, etc), WL does not require any assumption on their dynamical or evolutionary state. Its advantage is to be only sensitive to the mass along the line--of--sight offering, in principle, the opportunity to construct mass--selected halo samples, which can be compared directly with theoretical predictions (e.g. N--body simulations) without assuming any mass--observable relation. Nevertheless, that WL provides truly mass--selected halo samples it is not strictly true. Large scale structures (LSS), into which massive halos are embedded, contribute to the lensing signal. Usually, these projection effects add noise to the halo signal or can result in false positive peaks in the shear map \\cite{Hoekstra01,Hoekstra03,Hoekstra11}. Further, intrinsic ellipticities of galaxies introduce an irremovable noise. All these sources of noise then compromise purity and completeness of halo samples selected by WL. Various theoretical aspects of WL halo detection have been widely investigated by a number of authors \\cite{hamana04,Dietrich10,marian06,maturi10,maturi11,cardone}. The search for shear peaks in data is exemplified in some works \\cite{hett05,dahle06,wittman06,gavazzi07,schirmer07,miyazaki07,berge08,abate09,shan11} although, on the observational side, halo lensing studies have been mainly focused on mass determination and halo properties, e.g. density profile and concentration \\cite{okabe10,israel10}. A basic method for shear peak detection, to which we will refer in the following, is provided by the aperture mass ($M_{ap}$) technique proposed by \\cite{S96}. $M_{ap}$ is a measure of the lensing signal in a shear map smoothed with a suitable filter. The method then relies on the usual approach of searching for points of local maximum (peaks), with $snr$ higher than a given threshold, in the smoothed map. Here, we investigate the impact of the halo concentration on $M_{ap}$ and shear peak counts, as compared to that of some cosmological parameters (i.e. $\\Omega_m$ and $\\sigma_8$). As already noted by \\cite{king11} and detailed in the following, the shear peak abundance, is quite sensitive to halo concentration and its relation with the halo mass. This seems to point out peak counts as a promising and complementary method to delineate the, still debated, relation between halo mass and concentration ($M-c$ relation). Our main interest here is to further deepen this point quantifying the power of peak function to constrain the $M-c$ relation. Halo concentrations have been studied extensively, by means of N--body simulations as well as observationally, during the last decade. The general trend is that the mean halo concentration shows a strong correlation with the halo mass. Typically, it declines with increasing the mass and redshift in accordance with the idea that the central density of halos reflects the mean density of the universe at the time of their formation. Therefore, halos collapsing earlier are expected to be denser than the more massive halos collapsing later. However, there is enough uncertainty on the exact form of the $M-c$ relation. Beside discrepancies between theoretical predictions and observations, differences are found even when comparing simulation results of different groups as well as different sets of observations \\cite{B01,eke01,Comerford07,neto07,gao08,duffy08,maccio08,mandelbaum08,oguri09,okabe10,klypin11,prada11,meneghetti13}. One of the main predictions of N--body simulations is that the density profile of halos assembled hierarchically and close to virial equilibrium, can be well approximated by a Navarro, Frenk \\& White (NFW) profile \\cite{NFW}, regardless of halo mass and the details of the cosmological model, all the cosmological informations being contained in correlation between the parameters of the NFW profile, i.e. in the $M-c$ relation. The median $M-c$ relation predicted by simulations is well described by a power--law \\cite{dolag04,gao08,duffy08,maccio08,zhao09}. However, the variety of individual halo aggregation histories causes a scatter in concentration which can be modeled by a log--normal distribution with a variance ranging from $\\sim 0.15$ up to $\\sim 0.30$ depending on the degree of relaxation of the halo \\cite{Jing00}. On the other hand, the redshift evolution of the concentration is less established. Recent results seem to favor a weaker redshift dependence than previously supposed \\cite{duffy08,maccio08,MC11}. Probably, the best current results from simulations are those by \\cite{klypin11} and \\cite{prada11}. They found larger concentrations than those reported in \\cite{maccio08} and \\cite{duffy08}, with a reasonable agreement for galaxy--size halos ($\\sim 10-15$\\% difference) but substantially larger values ($\\sim 40-50$\\%) for cluster--size halos, more compatible with recent X--ray and kinematic observations. Furthermore, these high--resolution large-volume simulations ({\\it Bolshoi} and {\\it MultiDark}, see http://www.mutidark.org) seem to indicate a $M-c$ relation more complex than previously conceived showing a novel feature: at high redshift, concentration first declines with increasing mass, then flattens and increases slightly at higher masses. On the observational side, many different methods have been used to study masses and concentrations resulting in an even more controversial picture and making the comparison between observations and theory somewhat ambiguous. Gravitational lensing is one of the most promising methods to estimate mass and concentrations. Strong and weak lensing have been used individually or in combination. More classical methods relies on kinematic tracers, e.g. galaxies (\\cite{rines06,WL10} and references therein) and the hydrostatic analysis of X--ray intensity profile of halos. However, all these methods are susceptible to bias effects and are presumably affected by systematics not fully understood and/or modeled, yielding discrepant results. For example, \\cite{mandelbaum08} used the stacked weak lensing signal from galaxies, groups and clusters in the Sloan Digital Sky Survey (SDSS) finding a $M-c$ relation slope consistent with the simulations but a $2\\sigma$ lower normalization. \\cite{oguri09} performed a combined weak and strong lensing analysis of a sample of clusters, reporting concentrations with a $7\\sigma$ excess above the simulation predictions. \\cite{okabe10} performed a weak lensing analysis of X--ray selected clusters reporting a slope somewhat steeper than in simulations although with an error of $\\sim 50$\\%. Other studies based on lensing and X--ray data can be found in \\cite{Comerford07,buote07,SA07,johnston07,broadhurst08,vikhlinin09,ettori11}. Despite the qualitative agreement with simulation predictions, the general picture which emerges from observations can be summarized briefly as follows. In almost all cases, the observed slope of the $M-c$ relation is consistent with or steeper than theoretical predictions. Moreover, strong lensing measures of massive clusters and X--ray analysis give a normalization factor higher than predicted by simulations, strong lensing concentrations being systematically larger than X--ray concentrations. On the other hand, weak lensing methods seem to point out a normalization lower than that found in simulations. For what concerns the lensing measurements, the origin of these discrepancies should be searched in some orientation and shape biases \\cite{oguri05,sereno11,giocoli}. It has been shown that neglecting halo triaxiality can lead to over-- and under--estimates of a factor of $2$ in concentrations and up to $50$\\% in halo mass other than underestimation of statistical uncertainties \\cite{corless09}. In addition, as already outlined above, projection of structures along the line--of--sight can result in apparently high concentration (e.g. \\cite{KC07}). Further complications are due to baryonic processes. Although baryon physics is not expected to drive the structure formation process on very large scales, its effects are likely to be important for low--mass halos or in the central region of larger objects. It has been shown that baryonic feedback and cooling can alter halo profiles. In particular, baryonic cooling (see \\cite{gnedin11} and reference therein) can be responsible for the excess in concentration observed in groups and low-mass clusters \\cite{sereno10,fedeli11}. Dynamical and X--ray techniques assume the halos are in virial and/or hydrostatic equilibrium. Most of real halos are not relaxed and have complex substructures making measurements difficult to interpret. Mergers, active galactic nuclei, cosmic rays, magnetic fields, turbulence and bulk motion of the gas can compromise the hydrostatic equilibrium leading to underestimates of X--ray masses \\cite{evrard96,dolag05,rasia06,nagai07,rasia13}. Although recent hydrodynamical simulations include a number of these non--gravitational processes, they are very difficult to model. Furthermore, many of them take place at scales too small to be resolved by simulations demanding radical simplifications and approximations. It is also worth noticing that the above results were derived in the contest of the standard cosmological model ($\\Lambda$CDM). Discrepancies can also arise if the data--model comparison has been made for the wrong cosmology. A number of studies have shown that alternative DE cosmologies or modified gravity models can lead to a different evolution of matter perturbations affecting the halo concentrations. For example, scalar field DE models with an earlier structure formation resulting in more concentrated halos have been considered in \\cite{dolag04,io1,io2,baldi,baldi2,boni}. The paper is organized as follows: the $M_{ap}$ method for shear peak detection is reviewed in Section \\ref{Map} while in Section \\ref{PF} we investigate the impact of the $M-c$ relation on the probability distribution function of $M_{ap}$ and provide a prescription to analytically calculate it taking into account the dispersion in $M-c$ relation. The impact on peak number counts is then investigated and compared to that of $\\Omega_m$ and $\\sigma_8$. In Section \\ref{fisher} we perform a Fisher matrix analysis in order to assess the capability of upcoming weak lensing survey in constraining the $M-c$ relation by means of peak counts. Section \\ref{concl} is devoted to discussion and conclusions. ", "conclusions": "\\label{concl} Halo masses and concentrations have been studied extensively, by means of N--body simulations as well as observationally, during the last decade. Nevertheless, the exact form of the $M-c$ relation is still widely debated. One of the most promising method to estimate masses and concentrations relies on gravitational lensing, in particular shear peak counts from near--future WL surveys seem a promising and complementary method to delineate the features of the $M-c$ relation, as pointed out in this work. The method for shear peak detection we have considered here relies on the $M_{ap}$ statistics. Hence, we have firstly investigated the impact of the $M-c$ relation on the PDF of $M_{ap}$. In particular, we have provided a prescription to properly take into account the $M-c$ relation dispersion, always disregarded in previous work, in the PDF. We have then shown that such scatter can affect significantly the PDF and, consequently, the $snr$ for peak detection. Secondly, we have investigated the impact of the $M-c$ relation on peaks number counts as compared to that of the cosmological parameters $\\sigma_8$ and $\\Omega_m$. Finally, we have performed a Fisher matrix analysis in order to assess the capability of an Euclid--like survey in constraining the $M-c$ relation and cosmological parameters by using peak number counts. We have found that: i) peak function alone provides constraints on $\\sigma_8$ and $\\Omega_m$ which are competitive with those obtained from different probes. On the other hand, only the normalization $c_0$ of the $M-c$ relation is reasonably (but not so strongly) constrained if peaks with $snr>4$ are considered; ii) adding prior informations on $\\sigma_8$ and $\\Omega_m$ as inferred from current/near--future data, constraints on $M-c$ relation improve significantly so that peak statistics seems helpful in discriminating among the wide variety of $M-c$ relation fits found in literature although no substantial new information on the slope $\\alpha$ is inferred. Although results are quite encouraging, it is worth stressing that the Fisher matrix analysis here presented disregards the effects of the halo sample variance on the error estimates, only Poisson noise being considered. Nevertheless, sample variance might be significant for low mass halos at low redshifts where their abundance is much higher thus reducing the Poisson noise. To yield more accurate predictions, the covariance matrix for number counts $C_{ij}$ \\footnote{it should not be confused with the parameter covariance matrix of eq. (\\ref{sigpiv}) for which we use the same simbol.}, entering the FM expression, should then be written as: $C_{ij} = P_{ij}+S_{ij}$ where $P_{ij}=\\delta_{ij}N_i^{fid}$ and $S_{ij}=N_i^{fid}N_j^{fid}b_ib_j \\sigma_{ij}^{DM}$ are the Poissonian and sample variance contributions respectively (here $N_i^{fid}$ and $b_i$ are the expected number of peaks and the mean halo bias in the $i$--th redshift bin, while $\\sigma_{ij}^{DM}$ is the covariance of the dark matter fluctuations between the bins $i$ and $j$, se e.g. \\cite{hukra,takada,valageas} for more details). Sample variance effects on halo number counts have been considered in previous works, e.g. \\cite{hukra,takada,valageas,lima,masamune}. In particular, \\cite{hukra,valageas} conclude that sample variance is generally comparable or greater than Poisson noise for number counts above mass threshold $M_{th} \\lesssim 1-4 \\cdot 10^{14} h^{-1} M_\\odot$ depending on the survey width and depth. In our case, the mass thresholds corresponding to given $snr^*$ and $z$ can be deduced from the middle panel of fig. \\ref{snzM}. According to \\cite{hukra,valageas}, for $snr^* \\geq 4$ we expect sample variance to have only moderate effects in the redshift range $\\Delta z=0.15-0.45$ where the counts signal is maximum, becoming negligible at the increasing of $z$. For a raw estimate, we can assume $C_{ij}$ to be diagonal (i.e. halo correlation length smaller than the bin width) and $S_{ij}=P_{ij}$ ($S_{ij}=2P_{ij}$) independently of the redshift. This yields a worsening of the parameter constraints by a factor $\\sqrt 2$ ($\\sqrt 3$). However, \\cite{hukra,takada,valageas} neglect the information arising from the cosmological dependence of the sample variance. When this information is included and considered as ``signal'', it could cause a smaller degradation in the constraints or, in case, even improvements \\cite{lima,masamune}. In this case, the full FM can be well approximated as: $$ F_{ij} = {\\bf N}^T_{,i} {\\bf C}^{-1} {\\bf N}_{,j} + \\frac{1}{2} {\\textrm {Tr}} \\left[{\\bf C}^{-1}{\\bf S}_{,i}{\\bf C}^{-1}{\\bf S}_{,j} \\right]+ \\frac{\\delta_{ij}}{\\sigma_p^2(p_i)} $$ where ${\\bf N}=(N^{fid}_1,...,N^{fid}_{n_{bin}})$ and $_{,i}$ denotes the derivative with respect to the parameter $p_i$. It is worth stressing that the second term, neglected in \\cite{hukra,takada,valageas}, beside the information about the cosmological dependence of the sample variance, also contains information about its dependence on $M-c$ relation trough the derivative of $N_i^{fid}$ and possibly of $b_i$. Indeed, it was pointed out in \\cite{wechsler,jingsu} that halo bias could depend not only on the halo mass but also on additional halo properties, among which the concentration. Adding sample variance information could then, in principle, provide an opportunity to improve measurements of halo concentration (and cosmological parameters) with WL peak statistics through self--calibration techniques for $M-c$ relation similar to those for mass--observable relation (see e.g. \\cite{lima,masamune}). This intriguing point, which deserve a detailed analysis, is left for future works. A second issue concerns the assumption that detected peaks can be properly assigned to redshift bins. The redshift information, however, can not be extracted from WL data only, but additional data need to be included. In order to split the peak sample in redshift bins, optical finders (see e.g. \\cite{postman,koester,milke,bella}) could serve the purpose since they return reasonable redshift estimates, especially because they can be applied on the same optical data retrieved for lensing. Although the correlation between optical detections and WL peaks is not trivial, in a Euclid--like survey, one can look at the position of the peaks in the optical images, and then deduce the peak redshift $z$ from the redshifts of the galaxies closest to the peak position. In a first approximation, we can suppose that this method provides a Gaussian probability distribution function for $z$ with negligible bias and variance $\\sigma_z = \\sigma_0(1+z)$, and then assume a peak to be correctly assigned to a redshift bin roughly asking that the $3-\\sigma$ uncertainty on $z$ is smaller than the bin width. For our assumed value $\\Delta z = 0.1$, this translates in $\\sigma_0 \\leq 0.03$, a precision which could be likely achieved if $z$ is spectroscopically measured ($\\sigma_0 \\simeq 0.001$ according to the Euclid red book), but could be too demanding if one relies on photometric redshift methods ($\\sigma_0 \\simeq 0.05$ for Euclid). In this second case, one should add a non Poissonian uncertainty on $N_{halo}(snr > snr^*, z, \\Delta z)$ in the above Fisher matrix analysis and resort to the pull statistics \\cite{campanelli}. On the other hand, it is likely that the precision of the inferred redshift also depends on the $snr$ of peak so that the net effect should be included in the analysis by convolving the theoretically computed $N_{halo}$ with an empirically determined selection function. A further issue concerns the assumption of NFW spherical halos. Real halos are not spherical and not all relaxed so that deviations from the universal NFW profile are expected. The effects of the diversity of dark matter distributions in individual halos on peak counts has been investigated in \\cite{gruen,hamana12} by means of numerical simulations and analytic methods. The noise originated from halo shape is found to be comparable to the statistical noise discussed in section \\ref{pdf}. Furthermore, halo orientations cause a systematic bias in the peak heights. In order to include these effects in the theoretical predictions, one should consider a triaxial halo model (see \\cite{jingsuto}) which is shown to work quite well \\cite{hamana12}. It is also worth mentioning that, in the FM analysis, we have only investigated a subset of the full parameter space. Here, we were mainly interested in how peak statistics constrain the $M-c$ relation so that those parameters, as $H_0$ and $n_s$, which have a minor impact on peak number counts and are expected to be scarcely constrained, were held fixed to their fiducial values. Nevertheless, when dealing with real data, the full parameter space should be considered. Allowing for a large number of parameters to be varied introduces further degeneracies widening the confidence regions. This degradation in the constraining power, can then be compensated by complementing peak statistics with further datasets, e.g. SNeIa and CMB. Investigating the above issue is however outside our aims here." }, "1402/1402.2976_arXiv.txt": { "abstract": "The massive exploitation of cosmic voids for precision cosmology in the upcoming dark energy experiments, requires a robust understanding of their internal structure, particularly of their density profile. We show that the void density profile is insensitive to the void radius both in a catalogue of observed voids and in voids from a large cosmological simulation. However, the observed and simulated voids display remarkably different profile shapes, with the former having much steeper profiles than the latter. Sparsity can not be the main reason for this discrepancy, as we demonstrate that the profile can be recovered with reasonable accuracy even with very sparse samples of tracers. On the other hand, the observed profile shows a significant dependence on the galaxy sample used to trace the matter distribution. Samples including low-mass galaxies lead to shallower profiles with respect to the samples where only massive galaxies are used, as faint galaxies live closer to the void centre. We argue that galaxies are biased tracers when used to probe the matter distribution within voids. ", "introduction": "Large redshift surveys \\citep{York00, Colless01} and cosmological simulations \\citep{Bond96} have revealed that galaxies are distributed inside a cosmic web of walls, filaments and compact clusters. Such a web encloses large underdense regions, referred to as cosmic voids. Voids were first recognized in the earliest redshift surveys \\citep{Gregory78, Kirshner81} as huge empty holes in the galaxy distribution. Nowadays, there is a general consensus in that voids occupy most of the volume of the Universe \\citep{Sheth04, Vandew11, Pan12}, although there is not yet an agreement on how a genuine void should be defined. Several void finders, which are based on different principles, have been developed. Voids can be identified as spherical regions devoid of galaxies/haloes \\citep{Gottlober03, Patiri06b, Varela12} or underdense regions, relying on the continuous density field \\citep{Plionis02, Colberg05}. More complex algorithms able to capture the complex morphology of voids also exist \\citep{Platen07, Neyrinck08, AragonCalvo13}. Despite their different definition of voids, all these void finders agree in that voids ere extremely empty in the centre and show a sharp increase in the density towards the voids edges (e.g. \\citealt{Colberg08}). Voids are believed to originate from negative density fluctuations in the primordial density field. As a result of their underdensity, they are subject to an effective repulsive peculiar gravity, causing their expansion. As a consequence of such an expansion, the matter within the voids evacuates from the interior and accumulates to the boundaries. This leads to void density profiles that evolve towards a reverse top-hat shape \\citep{Sheth04}. A considerable appeal of cosmic voids is their potential in probing cosmological parameters. In particular, being almost devoid of matter, they are extremely sensitive to the nature of dark energy. Indeed, the void ellipticity and its evolution through cosmic time are intimately connected with the local tidal tensor, which, in turn, depends on the dark energy content \\citep{Park07, Lavaux10, Bos12}. Voids are also the ideal candidate for probing the expansion history of the Universe through the Alcock-Paczynski test \\citep{AP79}, using the average shape of stacked voids \\citep{Lavaux12, Sutter12a}. The application of such a test to the voids that will be identified in the future Euclid survey \\citep{Laureijs11} promises to outperform Baryonic Acoustic Oscillation by an order of magnitude in accuracy. The huge potentiality of voids for precision cosmology requires a robust knowledge of their internal structure, particularly of the density profiles. Works based on cosmological simulations \\citep{Colberg05, Ricciardelli13} indicate that the void density profile is universal. As such, it does not depend on void size. On the observational side, the ideal approach to directly constrain the void density profile is through the weak lensing signal of stacked voids \\citep{Krause13}. However, the number of voids available from spectroscopic catalogues is still limited to provide a robust measurement of the signal \\citep{Melchior13}. At present, we can only rely on the galaxy distribution to trace the density within voids \\citep{Sutter12b}. Thus, to robustly assess a void model to describe the universal density profile, one also needs to assess the systematic effects arising from the use of the sparse galaxy sampling. In a previous work (\\citealt{Ricciardelli13}, hereafter RQP13), we have shown, by means of a cosmological simulation, that a two parameters law can be used to fit the density profile of voids of any size, density, morphology and redshift. The best-fit parameters show some dependence on redshift, density, and, on a less degree, morphology, but they are almost independent on the void size, although the limited statistics prevented us to draw robust conclusions. In this work, we want to test this model and its dependence on void radius, against an observed catalogue of voids and a larger simulation, thus dramatically increasing the statistics. In doing so, we provide a robust determination of the systematic effects arising when using the sparse distribution of void galaxies as density tracers. The structure of the paper is as follows. In Section \\ref{sdss} we introduce our catalogue of observed voids, in Section \\ref{sim} we describe the simulation used and our void identification procedure. The results on the void density profiles are discussed in Section \\ref{profiles}. We conclude in Section \\ref{conclu}. ", "conclusions": "We have robustly assessed the universality of void density profiles, by means of a catalogue of observed voids and a large cosmological simulation. The observed void catalogue has been drawn from the SDSS database, and includes spherical voids whose radius is larger that $7 \\, h^{-1}\\, Mpc$ \\citep{Varela12}. To measure the density profiles in these voids, we rely on the luminous galaxies. As a matter of comparison, we have performed a large cosmological simulation with the code MASCLET, devoted to follow the formation and evolution of the low-density regions. This simulation has been designed to target, with sufficient statistics, voids spanning a wide range of radius. To this aim, we have simulated a large volume, having a comoving side length of $512 \\, h^{-1}\\, Mpc$, with only one level of refinement in the AMR grid, reaching the spatial resolution of $0.5 \\, h^{-1}\\, Mpc$. Since this modest resolution does not allow to follow the formation of structures in the simulated box, void galaxies in our simulation do not form. Therefore, we adopt as density tracer the continuous density field or, where a sparse distribution of tracers is needed, the dark matter particles within the void regions. The void density profiles recovered by means of the observed and simulated voids share the same qualitative shape, showing a significant underdensity in the centre and a sharp density increase approaching the void edges. Both profiles can be well described by the functional form proposed in RQP13. However, the observed profile is significantly steeper than the simulated one. To figure out the reasons for the steepness of the observed profiles, we have assessed the impact of the number and type of tracers on the resulting density profile. The sparsity of the density tracers has been investigated by means of subsamplings of the simulated voids, populated with an increasing number of particles. We have shown that even in the less populated void samples, the original density profile can be recovered with reasonable accuracy. Stacks built with a limited number of voids or sparsely populated present a significant noise at small radii, but no systematic effect with the number of voids/tracers is observed. The low impact of the sparsity of the tracers on the internal void density profiles has been pointed out also by \\citet{Sutter13}, using both dark matter particles and haloes as density tracers. Nevertheless, we observe that the profile shape can have a significant dependence on the type of galaxies used to trace the matter distribution. Within the observed voids, the density profiles recovered by means of faint samples of galaxies are shallower than those determined through the brighter galaxies. The reason for that lies in the galaxy mass segregation within voids. In fact, faint galaxies are those living closer to the void centre and, thus, allow to probe the matter distribution even in the innermost part of the voids. The strong impact of the type of galaxies chosen to trace the density, forces us to use an homogenous sample of galaxies and voids, limited in volume and magnitude, to assess the dependence of the void density profile on the void radius. With such a sample, we have demonstrated the insensitivity of the observed void profile on void radius. Likewise, by using our simulated sample of voids, we do not observe any dependence of the profile shape on the void size, and the same best-fit can correctly describe voids whose size ranges from 7 to $\\sim 50 \\, h^{-1}\\, Mpc$. Finally, we note that the difference in profile between the observed and simulated voids can not be driven by the different algorithms used to identify voids. Indeed, the density profile of our SDSS stack is very similar to the profile published in Pan et al. (2012), using the same SDSS DR7 dataset, albeit with a completely different void finder. Moreover, our simulated void density profiles are in remarkable agreement with the simulations of \\citet{Colberg05}, where voids are identified through spherical underdensities. Therefore, we argue that the difference between observed and simulated void density profiles is a robust result and is due to the biased tracers used, when relying on the observed galaxies. To corroborate this hypothesis, we definitely need high resolution simulations, capable to follow structure formation in the most rarefied regions of the Universe." }, "1402/1402.1849_arXiv.txt": { "abstract": "Accretion disk winds are thought to produce many of the characteristic features seen in the spectra of active galactic nuclei (AGN) and quasi-stellar objects (QSOs). These outflows also represent a natural form of feedback between the central supermassive black hole and its host galaxy. The mechanism for driving this mass loss remains unknown, although radiation pressure mediated by spectral lines is a leading candidate. Here, we calculate the ionization state of, and emergent spectra for, the hydrodynamic simulation of a line-driven disk wind previously presented by \\cite{pk04}. To achieve this, we carry out a comprehensive Monte Carlo simulation of the radiative transfer through, and energy exchange within, the predicted outflow. We find that the wind is much more ionized than originally estimated. This is in part because it is much more difficult to shield any wind regions effectively when the outflow itself is allowed to reprocess and redirect ionizing photons. As a result, the calculated spectrum that would be observed from this particular outflow solution would not contain the ultraviolet spectral lines that are observed in many AGN/QSOs. Furthermore, the wind is so highly ionized that line-driving would not actually be efficient. This does not necessarily mean that line-driven winds are not viable. However, our work does illustrate that in order to arrive at a self-consistent model of line-driven disk winds in AGN/QSO, it will be critical to include a more detailed treatment of radiative transfer and ionization in the next generation of hydrodynamic simulations. ", "introduction": "\\label{introduction} \\begin{figure*} \\includegraphics{fig1_colour.eps} \\caption{The density (colours) and poloidal velocity (arrows) structure of the PK04 model. Radial lines delineate the three zones described in the text. } \\label{lin_den} \\end{figure*} Outflows are found in a vast range of accreting astrophysical objects, from protostars to active galactic nuclei (AGN). In all of these settings, quantifying the mass and energy flows involved is key to understanding how the accreting objects evolve and interact with their local environment. In the case of AGN, the observed outflows can be split into two main classes: highly collimated, relativistic jets, and slower moving ($v \\ltappeq 0.2c$), but more massive `disk winds' driven from the surface of the accretion disk surrounding the central supermassive black hole. These disk winds have been proposed as the underlying structure responsible for many observed AGN spectral features, including the broad absorption lines seen in a significant proportion of QSOs \\citep[e.g.][]{knigge08}, the so called broad absorption line quasars (BALQSOs). However, the geometry of these winds, and even the mechanism by which they are launched, are still not known. Several mechanisms have been proposed to produce the accelerating force for disk winds in AGN, including gas/thermal pressure \\citep[e.g.][]{weymann_82, begelman_91, begelman_83,krolik_kriss}, magnetocentrifugal forces \\citep[e.g.][]{blandford_payne_82, pelletier_pudritz} and radiation pressure acting on spectral lines (``line-driving'') \\citep[e.g.][]{shlosman_85,murray_95}. All of these mechanisms have been shown to produce outflows in suitable conditions. However, line driving is particularly attractive, since the requisite force in this case arises naturally when (primarily ultraviolet [UV]) photons produced by the central engine are scattered by strong resonance lines. Many such lines are observed directly in the UV spectra of (BAL)QSOs, so line driving must certainly be acting in some measure. Moreover, if line driving is the dominant acceleration mechanism, it can produce a unique signature in the profiles of absorption lines produced in the wind (the so-called ``ghost of $\\rm{Ly\\alpha}$'', e.g. \\citealt{arav_95,arav_96}), and this may already have been seen in several BALQSOs (\\citealt{north_knigge_goad}, but also see \\citealt{cottis_10}). Despite these circumstantial reasons for favouring line-driving as the mechanism for the production of disk winds in AGN, there is one main challenge to the model. For line-driving to be efficient, the accelerated material has to be in a moderately low ionization state, despite its proximity to the intense X-ray source at the center of the accretion disk. If the gas becomes over-ionized, the ionic species that can most effectively tap into the momentum of the radiation field are simply not present. \\cite{murray_95} proposed one solution, suggesting that `hitchhiking gas', material interior to the main outflow and itself accelerated by pressure differences can shield the line-driven wind from the central X-rays. Simulations such as those presented in \\cite{proga_stone_kallman} and \\citet[][hereafter PK04]{pk04} also produce a shield, but in this case via a `failed wind'. This failed wind arises close to the X-ray-emitting region and quickly becomes over-ionized as it rises above the disk surface. As it falls back, it produces a dense shield that prevents the intense X-ray radiation from reaching the outer parts of the outflow in its shadow. These low-ionization regions are then able to interact with UV radiation and give rise to a strong line-driven wind. The same type of shielding structure has also been seen in other models of disk winds \\citep[e.g][]{ris_elv,nomura_13}. PK04 used hydrodynamical simulations to investigate whether a line-driven outflow could be accelerated to high velocities purely by the radiation field produced by the accretion disk. They estimated the ionization state and temperature of the wind, taking into account only the central source of ionizing radiation, attenuated by electron scattering. Their main result was that the ionization state of the wind behind the failed wind region remained low enough to permit line driving, so that a fast, dense outflow was generated. Absorption line profiles calculated from their wind model resembled those observed in BALQSOs (e.g. see Fig. 2 in in \\citealt{prog_kuro_10}). It is worth noting that PK04's wind, and especially its base, are virialized systems. This is because a line-driven wind accelerates slowly, so the rotational velocity is dominant over a relatively large distance while the wind base is Keplerian and very dense. This means that, if the broad line region (BLR) is associated with such a wind, black hole (BH) mass estimates based upon the assumption that the BLR is virialized remain correct \\citep{kashi_13}. The PK04 wind model has already been subjected to two more detailed `post-processing' radiative transfer calculations. Both of these were concerned with the impact of the predicted outflow on the X-ray spectra of AGN/QSO. First, \\citet{schurch_09} performed a 1-D simulation which found that the wind produced observable X-ray spectral features. Second, \\citet[][hereafter SP10]{sim_proga_10} carried out a multi-dimensional radiative transfer simulation in which the ionization state and temperature structure of the wind were self-consistently computed. They confirmed that the wind was able to imprint a variety of characteristic features into the X-ray spectra of AGN. However, they also showed that scattering in the (failed) outflow is critically important in setting the ionization state of the wind in regions that would otherwise be shielded from the central engine. PK04 used a simplified treatment of radiative transfer and ionization in their hydrodynamical simulations. These simplifications were essential to make the simulations computationally feasible. However, the work of SP10 implies that a careful treatment of radiative transfer may be required in order to obtain a reliable estimate of the ionization state of the wind. Since the ionization state, in turn, determines the efficiency of the line-driving mechanism itself, it is clearly important to check whether the wind model calculated by PK04 could actually maintain an ionization state that is consistent with efficient line driving and with the production of broad UV absorption lines. Here, we therefore carry out a full, multi-dimensional radiative transfer calculation for the PK04 model that allows us to predict the ionization state, temperature structure and emergent UV spectra for this outflow. This extends the work done by SP10 to the longer wavelengths and lower ionization stages that are critical to allow effective line-driving and the formation of BALs. In particular, we account for the low-energy ($\\rm{< 0.1~keV}$) photons that affect the abundances of key ions, such as C~\\textsc{iv}, N~\\textsc{v} and O~\\textsc{vi}. We also calculate emergent spectra throughout the UV band, where the BAL features associated with the resonance transitions of these species are found. ", "conclusions": "We have presented results from comprehensive radiative transfer and photoionization calculations for the line-driven AGN disk wind predicted by the hydrodynamic simulations carried out in PK04. These simulations were computationally expensive and therefore treated the radiative transfer and ionization in a simplified manner. Here, we have focused on one snapshot from the hydrodynamic calculation and carried out a much more detailed Monte-Carlo simulation of the interaction between the AGN radiation field and the outflow. Our main result is that the ionization state of the outflow is much higher than estimated by PK04, to the extent that line-driving would become inefficient. The over-ionized flow also no longer produces the broad UV absorption lines that are the key observational tracers of disk winds in AGN/QSOs. The main reason for this change in the predicted ionization state of the flow is that self-shielding becomes much less effective when radiative transfer effects are fully accounted for. More specifically, the failed wind region that protects the outflow from over-ionization in the simulations of PK04 can actually be ``circumnavigated'' by ionizing photons via scattering and reprocessing. We conclude that hydrodynamic disk wind simulations need to take account of scattering and reprocessing in order to robustly assess the viability of line driving as an acceleration mechanism. This should ideally take the form of a self-consistent treatment, in which the radiative transfer and hydrodynamics are calculated simultaneously. However, the kind of `post-processing' approach we have used here is at least useful in validating purely hydrodynamic models. For outflows driven non-radiatively, it already provides a means to self-consistently predict the observational characteristics of the flow." }, "1402/1402.4188_arXiv.txt": { "abstract": "Although the mechanisms responsible for heating the Sun's corona and accelerating the solar wind are still being actively investigated, it is largely accepted that photospheric motions provide the energy source and that the magnetic field must play a key role in the process. \\citet{2010ApJ...708L.116V} presented a model for heating and accelerating the solar wind based on the turbulent dissipation of Alfv\\'en waves. We first use a time-dependent model of the solar wind to reproduce one of \\citeauthor{2010ApJ...708L.116V}'s solutions; then we extend its application to the case when the energy equation includes thermal conduction and radiation losses, and the upper chromosphere is part of the computational domain. Using this model, we explore parameter space and describe the characteristics of a fast-solar-wind solution. We discuss how this formulation may be applied to a 3D MHD model of the corona and solar wind \\citep{2009ApJ...690..902L}. ", "introduction": "The identification of the physical processes responsible for the heating of the solar corona and the acceleration of the solar wind still represents an unsolved problem in solar physics. However, there is a general consensus that photospheric motions provide the energy source and that the magnetic field must play a key role in the process. Since it is clear that the measured speeds of fast streams require an extended heating deposition \\citep{1977ARA&A..15..363W,1980JGR....85.4665H,1988ApJ...325..442W}, previous one-dimensional (1D) models generally relied on a parametric heating function exponentially decaying with height \\citep{1982ApJ...259..779H,1982ApJ...259..767H,1988ApJ...325..442W, 1995JGR...10021577H,1995GeoRL..22.1465R,1997ApJ...482..498H}. At the same time, several investigations were based on low-frequency broadband fluctuations on magnetohydrodynamic scales as the mechanism that heats and accelerates the solar wind \\citep{1968ApJ...153..371C,1971JGR....76.3534B,1986JGR....91.4111H, 1988JGR....93.9547H,1994AdSpR..14..123V,1999ApJ...523L..93M,2007ApJ...662..669V,2012ApJ...745...35Z}. Connecting the macroscopic heating of the plasma and acceleration of the wind, as formulated in coronal and inner heliospheric MHD models, with the underlying physical mechanisms is complicated due to the temporal and spatial dynamic ranges involved. In the past decade, turbulent dissipation mechanisms have been progressively incorporated with various degrees of self-consistency into 1D models of the solar wind \\citep{2005ApJ...632L..49S,2005ApJS..156..265C,2007ApJS..171..520C,2010ApJ...710..676C, 2010ApJ...708L.116V,2011ApJ...743..197C}. There are also efforts to replace empirical heating functions in three-dimensional (3D) MHD models with some form of turbulence dissipation mechanism. This is particularly challenging because such 3D models would have to resolve time-scales extending from a millisecond (dissipative time-scale in the solar corona) up to many days (large scale solar wind stream structure). \\citet{2010ApJ...725.1373V} introduced, beside the acceleration of the solar wind through Alfv\\'en waves, the heating of the protons by Kolmogorov dissipation in open field-line regions. \\citet{2011ApJ...727...84U} developed a large-scale MHD heliospheric model with small-scale transport equations for the turbulence energy, normalized cross helicity, and correlation scale, applicable where the solar wind is already supersonic and superalfv\\'enic. The model of \\citet{2013AIPC.1539...30L} included the effect of outwardly propagating Alfv\\'enic turbulence in the solar wind and a phenomenological term to describe nonlinear interactions associated with wave reflection by density gradients in the chromosphere and corona. \\citet{2013ApJ...764...23S} and introduced Alfv\\'en wave turbulence, assuming that this turbulence and its nonlinear dissipation are the only momentum and energy source for heating the coronal plasma and driving the solar wind {\\citep[see][for additional details]{2013arXiv1311.4093V}. } The present work illustrates the integration of the turbulence dissipation heating and acceleration mechanism of \\citet{2010ApJ...708L.116V} in a time-dependent, 1D, hydrodynamic (HD) model of the solar wind, which includes also thermal conduction and radiation losses. In this early, explorative phase, it is expedient to conduct an investigation using 1D models so that we may later apply our gained experience to 3D models. The model of \\citet{2010ApJ...708L.116V} employs strong turbulence closure to treat nonlinear effects, and does not rely on electron heat conduction for radial energy transport, but rather computes the internal energy associated with protons only. After presenting the characteristics of our {turbulence-driven HD} model, we show how it can match one of the solutions derived by \\citet{2010ApJ...708L.116V} without thermal conduction and radiation losses. Subsequently, we extend the application of our model to include transport mechanisms in the energy equation, which are necessary to reproduce plasma emission in agreement with observations \\citep{2009ApJ...690..902L}. Our exploration of parameter space yields solutions compatible with the solar wind properties obtained from \\textit{in situ} measurements and observations. In the future we plan to introduce this formulation into the 3D MHD thermodynamic model of the solar corona and solar wind of \\citet{2009ApJ...690..902L}. This paper is organized as follows: the equations and the solution technique are described in Sec.~\\ref{sec-model}. In Sec.~\\ref{sec-results} we present a solution in the configuration of \\citet{2010ApJ...708L.116V}, we conduct a parameter study that includes thermal conduction and radiative losses, and we describe the details of one of the solutions. We conclude with a discussion. ", "conclusions": "We have implemented a time-dependent model of the solar wind with acceleration and heating through turbulence dissipation. Our model uses the self-consistent formulation of \\citet{2010ApJ...708L.116V} within a 1D HD code with thermal conduction and radiation losses. For the case of a simplified energy equation, such as that used by \\citet{2010ApJ...708L.116V}, we accurately reproduce a solar wind solution. When the model is extended by introducing thermal conduction and radiative losses, our model produces fast solar wind solutions whose characteristics are compatible with \\emph{in situ} measurements. The model of \\citet{2010ApJ...708L.116V} includes a compressional heating in the lower corona; we have found that it is not necessary to include such phenomenological heating when a more realistic energy equation is used. Our model, in comparison with that of \\citet{2013AIPC.1539...30L}, is certainly more complicated, since it advances the amplitude of the perturbations rather than the energies as the latter does. Moreover, the model of \\citet{2013AIPC.1539...30L} neglects the contribution of the Reynolds stress, assuming that outwardly propagating wave is dominant. However, with the present method, we have found that the contribution of the Reynolds stress can be sizable; in one particular case it lowered the wind speed by approximately $25\\%$. Furthermore, if we consider the integration of turbulence dissipation heating and acceleration in 3D MHD models, there are several advantages in the present formulation in respect of that of \\citet{2013AIPC.1539...30L}. First, it is not much more computationally demanding to advance the amplitudes rather the energies. Second, this formulation in terms of $z_\\pm$ does not require us to calculate for each mesh point the reflection coefficient along each field line passing through it. {Third, in our experience the inclusion of time-dependent transport equations for $z_\\pm$ does not increase the physical convergence time to steady state.} Finally, as shown by \\citet{2011ApJ...727...84U}, the Reynolds stress term can be written for a case when only perturbations perpendicular to the large scale magnetic field are considered and whose actual directions do not play a significant role. Therefore we believe that the present formulation can be readily extended to the 3D model of \\cite{2009ApJ...690..902L}." }, "1402/1402.1078_arXiv.txt": { "abstract": "We report the timing and spatial resolution from the Muon Telescope Detector (MTD) installed in the STAR experiment at RHIC. Cosmic ray muons traversing the STAR detector have an average transverse momentum of 6 GeV/$c$. Due to their very small multiple scattering, these cosmic muons provide an ideal tool to calibrate the detectors and measure their timing and spatial resolution. The values obtained were $\\sim$100 ps and $\\sim$1-2 cm, respectively. These values are comparable to those obtained from cosmic-ray bench tests and test beams. ", "introduction": "Data taken over the last decade have demonstrated that RHIC has created a hot, dense medium with partonic degrees of freedom called the Quark-Gluon Plasma (QGP). One of the physics goals for the next decade is to study the fundamental properties of the QGP such as the temperature, density profile, and color-screening length via electro-magnetic probes such as di-leptons~\\cite{starwhitepaper,rhicwhitepaper,dilepton,dileptonII, highptjpsi,satz_0512217,rhicIIQuarkonia,colorscreen,petreczky}. Muons can be measured more precisely because of their relatively reduced Bremsstrahlung radiation in the detector materials. Such an improved measurement is essential for separating the $\\Upsilon$ meson ground state (1S) from its excited states (2S+3S), each of which is predicted to melt at very different temperatures. The Muon Telescope Detector (MTD) in the Solenoidal Tracker at RHIC (STAR) will allow the measurement of the $\\Upsilon$ mesons and $J/\\psi$ mesons, over a broad transverse momentum range through di-muon decays to study color screening features, and $\\mu$-e correlations to distinguish heavy flavor correlations from initial lepton pair production~\\cite{starmtdproposal}. The MTD will thus provide direct information on the temperature and the characteristics of color screening in the QGP created in RHIC collisions. The MTD is based on the Multi-gap Resistive Plate Chambers (MRPC) technology. A similar technology was used for the recently installed STAR TOF system~\\cite{startofproposal,startof}. Unlike the TOF MRPCs~\\cite{startofmrpcs}, however, the MTD MRPCs are much larger, and have long double-ended read-out strips. The MTD detectors are positioned behind the iron return bars of the STAR magnet, and cover 45\\% of the full azimuth within a pseudo-rapidity range of $|\\eta|$$<$0.5. The construction and installation of the MTD system was begun in 2011 and will be completed in 2014. Prototype MTD detectors were built and studied from 2007 to 2011. These detectors were tested in the laboratory with cosmic rays, in test beams, and in STAR experiment during RHIC runs. The cosmic-ray and beam tests indicated a timing ($\\le$100 ps) and the spatial resolution ($\\sim$1 cm) that would be sufficient to achieve the physics goals~\\cite{MTDNIMA}. The operation of the prototype MTD detectors in STAR in 2007-2008 demonstrated that clean muon identification could be achieved for muon transverse momenta above a few GeV/$c$. The bench and test-beam results, as well as detailed simulations, thus indicated that the MTD would provide important physics information on quarkonia and primordial di-lepton measurements at RHIC~\\cite{MTDPerformanceAtSTAR}. However, during the tests of the MTD prototypes in STAR in 2007-2008, the timing resolution was 200-300 ps, which was much worse than that observed in the previous cosmic-ray and test-beam studies. This poorer resolution resulted from the particular digitization electronics~\\cite{startrigger} that were used at that time and the long cables between the MTD detectors and the digitizers. In 2009, new prototype MTD detectors were installed in STAR. For these prototypes, the simple on-board front-end electronics used previously \\cite{mtdelectronics} to drive long cables to the digitizers were replaced with the same electronics as are used in the STAR TOF system~\\cite{startof,tofelectronics}. The TOF electronics are based on the CERN HPTDC~\\cite{HPTDC} chip. In 2010, a cosmic-ray trigger based on the information from the MTD prototypes was implemented in the STAR trigger system. High energy ($\\sim$6 GeV) cosmic rays provide an excellent means to study the timing and spatial resolution as the smearing from multiple scattering in the detector materials is a relatively small effect. In this paper, the timing and spatial resolution of the MTD MRPCs read-out by STAR TOF electronics for cosmic muons reconstructed in the STAR experiment in 2010-2011 will be reported. The performance observed during the operation of 10\\% of the full MTD system during the 2012 RHIC run will also be presented. This paper is arranged as follows. Section~\\ref{mrpc} describes the MTD MRPCs, and Section~\\ref{expr} describes the experimental aspects. The details of the data analysis and the MRPC performance are reported in Section~\\ref{results}. The summary and conclusions are then presented in Section~\\ref{concl}. ", "conclusions": "The time and spatial resolution of the STAR MTD system were obtained using cosmic-ray muons traversing the STAR detector. The relatively high momentum muons can be cleanly triggered upon and selected in the data, and allow studies of the time and spatial resolutions with relatively small contributions from multiple scattering in the STAR detector materials. The MTD resolution values observed in data sets spanning several years were 100~ps for the timing resolution and 1-2 cm for the spatial resolution." }, "1402/1402.1764_arXiv.txt": { "abstract": "\\noindent One of the principal discoveries in modern cosmology is that standard model particles (including baryons, leptons and photons) together comprise only 5\\% of the mass-energy budget of the Universe\\cite{planck}. The remaining 95\\% consists of dark energy and dark matter (DM). Consequently our picture of the universe is known as $\\Lambda$CDM, with $\\Lambda$ denoting dark energy and CDM cold dark matter. $\\Lambda$CDM is being challenged by its apparent inability to explain the low density of DM measured at the centre of cosmological systems, ranging from faint dwarf galaxies to massive clusters containing tens of galaxies the size of the Milky Way. But before making conclusions one should carefully include the effect of gas and stars, which were historically seen as merely a passive component during the assembly of galaxies. We now understand that these can in fact significantly alter the DM component, through a coupling based on rapid gravitational potential fluctuations. ", "introduction": "\\begin{figure*} \\includegraphics[width=0.49\\textwidth]{heic0604a.jpg} \\includegraphics[width=0.49\\textwidth]{martin_fig14.pdf} \\caption{The left panel (by J. Gallagher) shows a composite image of M82 taken by the Hubble Space Telescope. Purple colours correspond to narrow-band H$\\alpha$ emission, allowing us to see recombining hydrogen in outflowing gas. The right panel, from Martin et al\\protect\\cite{martin12}, shows a compilation of measured absorption line blue-shifts for cool gas as a function of the galaxy's star formation rate. Even dwarf galaxies with star formation rates under $1\\,\\Msol\\,\\yr^{-1}$ are able to support winds exceeding $100\\,\\mathrm{km\\,s}^{-1}$. The outflow rate of these winds is typically several times the instantaneous star formation rate of the parent galaxy.} \\label{fig:outflows} \\end{figure*} The viability of the $\\Lambda$CDM picture of structure formation was first evaluated using computer simulations (allowing, for instance, neutrinos to be ruled out as the dominant component of dark matter\\cite{frenk85}). Gas cooling and star formation within DM halos is now the standard paradigm for the origin of galaxies\\cite{whiterees78}. The behaviour of DM can be simulated on computers by chunking a portion of the universe into ``particles'' and evolving. Since the particles interact only through gravity, these simulations are called collisionless. Early attempts used just $30\\,000$ particles to follow large regions of the Universe. Consequently one particle had the mass of a large galaxy -- even so, such simulations were expensive, taking 70 CPU hours on state-of-the-art 3 MHz facilities. Such calculations would now take a few minutes on a cellphone. The growth of computing power and parallel capabilities meant that, by the 1990s, simulations became sufficiently powerful to make detailed predictions of the internal structure of halos in different cosmological scenarios. These simulations highlighted the universal nature of DM halos formed through collisionless collapse. The spherically-averaged density of halos is `cusped' at the centre (scaling approximately as $\\rho \\propto r^{-1}$), rolling to a steeper slope at larger radius (reaching $\\rho \\propto r^{-3}$); such behaviour is known as ``NFW'' after the authors of a pivotal paper\\cite{navarro96}. At the same time, simulations started highlighting a number of deficiencies in the CDM scenarios. The most evident was the overabundance, by more than an order of magnitude, of small satellites\\cite{moore99,klypin99} compared to the number observed orbiting the Milky Way\\cite{mateo98} at the time. Worse, the simulations significantly over-predicted the density of DM at the centre of galaxies\\cite{moore98}. Increasingly precise observations of the rotation curves of field galaxies have confirmed this discrepancy\\cite{deblok08} (see \\S3). Collisionless DM simulations have since reached maturity, with modern simulations using several billion resolution elements for just one Milky Way sized halo\\cite{aquarius08,ghalo}. However to make predictions which are testable against observations of the real Universe, baryon physics must be introduced. (Here we are adopting the astronomical convention of referring to baryons and leptons collectively as `baryons'.) Because baryons dissipate energy and so collapse to smaller scales than DM, they constitute a sizeable fraction of the mass in the central regions of all but the faintest galaxies \\cite{bell01}. Moreover observational constraints on galaxy formation ultimately come from photons, which can only be sourced by baryons. Accordingly much effort has recently been devoted to implementing gas hydrodynamics and a description of star formation within simulations\\cite{gnedin09,enzo11,ramses,keres12}. The energy released by young stellar populations and active galactic nuclei into the surrounding intergalactic medium is critical for regulating star formation\\cite{whiterees78}. Without this energy, most of the gas becomes cold and dense, rapidly collapsing to form stars, contradicting observations. Processes providing the energy to halt collapse are collectively named `feedback' and include supernova winds, radiation from young stars, and radiation and heat from black hole accretion \\cite{croton06,bower06,Stinson06,hopkins12}. Including these effects has led to strides forward in forming realistic disk galaxies, reproducing the efficiency of star formation as a function of galaxy mass, and linking gas accretion and mergers to galaxy morphology\\cite{robertson06,keres05,dekel09}. However, until recently any direct effect of the baryonic component on the DM was limited to a minor `adiabatic' correction\\cite{blumenthal86} (see box A). In other words, star formation (SF) processes resulted in `passive' changes to the galaxy population -- modulating the star formation rate without significant changes to the underlying cosmic DM scaffolding. This picture has recently been subverted. Spectroscopic observations reveal the ubiquity of massive galaxy outflows driven by feedback, carrying significant gas mass away from star forming galaxies throughout cosmic history\\cite{shapley03,weiner09,martin12} (see Section \\ref{sec:evid-galaxy-outfl}). It has slowly been realised that these directly observed processes have a non-adiabatic impact on the associated dark matter halos. The effect is to relieve discrepancies between baseline CDM simulations and the real Universe (discussed in Section \\ref{sec:evidence-cusp-core}). The emerging understanding of these processes constitute the central part of this review (Section \\ref{sec:one-process-rule}). \\begin{figure*} \\vchb{\\includegraphics[width=0.49\\textwidth]{oh_2010_figure.pdf}} \\vchb{\\includegraphics[width=0.49\\textwidth]{weighted_hist_gnfw.pdf}} \\caption{The left panel is a compilation\\protect\\cite{oh11} of observed innermost dark matter density profile slopes ($\\alpha$ where $\\rho_{\\mathrm{DM}}(r) \\propto r^{\\alpha}$) for field dwarf galaxies, plotted at the innermost point where a robust determination has been achieved. Where the slope $\\alpha$ can be measured interior to around one kiloparsec, it is typically much shallower ($\\alpha>-1$) than the simulated ``NFW'' result. The right panel\\protect\\cite{newman13} shows the probability distribution function on the parameter $\\beta=-\\alpha$ for a selection of galaxy clusters ($M\\sim 10^{15}\\Msol$). While the constraints on individual clusters are quite broad, the combined constraints (thick line) again indicate a shallower-than-NFW slope. }\\label{fig:cores} \\end{figure*} ", "conclusions": "The $\\Lambda$CDM cosmology underlies a highly successful paradigm for explaining the formation of visible structure in the universe. Until recently, the key ingredients were passive processes which controlled the association of observable matter with the dark matter (for instance suppressing over-efficient star formation) while having little explicit effect on the underlying dark matter. There is, however, a new, rich literature of processes which violate this basic assumption and lead to fundamental modifications to the observable properties of galaxies. In the last few years these have come into sharp focus as increasingly sophisticated computer simulations have begun to follow the effects of star formation, and many relevant observational techniques have matured to the point that they can be regarded as robust. Direct evidence of precisely which `baryonic processes' are in play and their relative importance in the real Universe at different scales should be our next priority. Because these baryonic processes simultaneously modify a number of observational diagnostics (outflows, dark matter cores, stellar morphology and star formation regulation), they weave into a coherent, testable framework. It remains a possibility that tensions between observation and theory at the scale of faint dwarfs and clusters may point to exotic particle physics. Ultimately we expect that a concerted effort from theorists and observers can achieve the goal of pointing to unique predictions of non-minimal DM models. Of particular interest in the coming years will be ({\\it i}) improved understanding of the dark matter in dwarf spheroidals and faint field galaxies; if cores persist at the faintest end, it is a generic conclusion that baryonic physics cannot account for them\\cite{G12,penarrubia12}; ({\\it ii}) study of the stellar population ages and, separately, metallicity distributions of these objects to determine as far as possible whether the required bursty star formation histories are consistent propositions\\cite{mcquinn10}; ({\\it iii}) better predictions of the scalings of cores in massive galaxies and clusters for different scenarios; ({\\it iv}) observations that constrain the star formation histories of dwarfs\\cite{pacucci13} and the behaviour of gas at high redshift, especially through absorption line studies which are sensitive to internal kinematics and outflows\\cite{viel13}; ({\\it v}) renewed effort to understand how non-minimal dark matter scenarios (such as WDM or SIDM) interact with the revised, more complex baryonic physics of galaxy formation. \\subsection{Acknowledgements} We would like to thank Se-Heon Oh, Simon White, Max Pettini, Crystal Martin, Matt Walker, Jorge Pe{\\~n}arrubia, Alyson Brooks, Tommaso Treu, Richard Ellis, James Wadsley and Lisa Randall for helpful discussions and comments on an early draft." }, "1402/1402.6691_arXiv.txt": { "abstract": "{Stellar differential rotation is important for understanding hydromagnetic stellar dynamos, instabilities, and transport processes in stellar interiors, as well as for a better treatment of tides in close binary and star-planet systems.}{We introduce a method of measuring a lower limit to the amplitude of surface differential rotation from high-precision, evenly sampled photometric time series, such as those obtained by space-borne telescopes. It is designed to be applied to main-sequence late-type stars whose optical flux modulation is dominated by starspots.}{An autocorrelation of the time series was used to select stars that allow an accurate determination of starspot rotation periods. A simple two-spot model was applied together with a Bayesian information criterion to preliminarily select intervals of the time series showing evidence of differential rotation with starspots of almost constant area. Finally, the significance of the differential rotation detection and a measurement of its amplitude and uncertainty were obtained by an a posteriori Bayesian analysis based on a Monte Carlo Markov Chain approach. We applied our method to the Sun and eight other stars for which previous spot modelling had been performed to compare our results with previous ones.}{We find that autocorrelation is a simple method for selecting stars with a coherent rotational signal that is a prerequisite for successfully measuring differential rotation through spot modelling. For a proper Monte Carlo Markov Chain analysis, it is necessary to take the strong correlations among different parameters that exist in spot modelling into account. For the planet-hosting star Kepler-30, we derive a lower limit to the relative amplitude of the differential rotation of $\\Delta P / P = 0.0523 \\pm 0.0016$. We confirm that the Sun as a star in the optical passband is not suitable for measuring differential rotation owing to the rapid evolution of its photospheric active regions. In general, our method performs well in comparison to more sophisticated and time-consuming approaches.}{} ", "introduction": "\\label{intro} The Sun and other stars do not rotate as rigid bodies owing to latitudinal and radial transport of angular momentum induced by anisotropic turbulent Reynolds stresses, meridional flows, and magnetic fields \\citep[e.g., ][]{Rudiger89}. Differential rotation (hereinafter DR) plays a fundamental role in hydromagnetic dynamo \\citep{BrandenburgSubramanian05} and as a source of hydrodynamic and magnetohydrodynamic instabilities in stellar interiors \\citep{KnoblochSpruit82}. Moreover, it plays a role in tidal interaction in close binary systems \\citep{Scharlemann81,Scharlemann82} and is thus expected to affect the tidal interaction between a close-in planet and its host star \\citep{Mathisetal13}. The measurement of the mean rotation period of a main-sequence late-type star, from which its age is estimated by the method of gyrochronology, is also affected by the amplitude of latitudinal DR \\citep[see, e.g., ][]{EpsteinPinsonneault12}. In the Sun, we can study DR in detail in the photosphere by measuring the rotation rate at different latitudes by Doppler shifts of the plasma spectral lines as well as by using sunspots as tracers for the motion of the surrounding plasma. The interior DR is accessible by helioseismic techniques that reveal a time dependence in some of the layers, probably related to the feedback of the Lorentz force associated with hydromagnetic dynamo action \\citep[e.g., ][]{Howeetal00,Lanza07,Howe09}. In distant stars, we have much more limited information because only spatially unresolved data can be acquired. Recently, asteroseismic techniques have provided the first hints on radial DR in red giants \\citep[e.g., ][]{Deheuvelsetal12,Mosseretal12,Goupiletal13}, while for main-sequence stars information on surface DR has been extracted through spectroscopic or photometric techniques. The advent of space-borne high-precision photometry with MOST \\citep[Microvariability and Oscillations of STars experiment, ][]{Rucinskietal03}, CoRoT \\citep[Convection, Rotation and Transit experiment, ][]{Auvergneetal09}, and Kepler \\citep{Boruckietal10} has made available large and homogeneous datasets of photometric measurements of late-type stars that represent a treasure trove to study stellar rotation and DR, in particular. Therefore, we introduce in the present work a technique to measure stellar DR from high-precision and evenly-sampled photometric time series of late-type main-sequence stars and discuss its advantages and limitations in the context of previously proposed approaches. Main-sequence stars of the A and F spectral types are generally quite fast rotators and do not show brightness inhomogeneities in their photospheres { \\citep[see, however, ][]{Balona13}}, thus making the effect of surface DR on the rotational broadening of spectral line profiles directly measurable by means of deconvolution techniques based on Fourier analysis \\citep[e.g., ][]{ReinersSchmitt03}. Applying this approach, \\citet{Reiners06} and \\citet{AmmlervonEiffReiners12} measured DR in a sample of A and F stars and found a remarkable increase in its amplitude with increasing effective temperature \\citep[see ][ for theoretical models that could explain such a dependence]{KukerRudiger05}. Stars of the spectral types G, K, and M generally show brightness inhomogeneities in their photospheres that are analogous to sunspots; i.e., they are associated with surface magnetic fields. Those having a sufficiently fast rotation ($ v \\sin i \\ga 15$~km~s$^{-1}$) can be mapped through the Doppler Imaging techniques \\citep[e.g., ][]{Donatietal97,DonatiCollierCameron97,Strassmeier09} allowing their surface DR to be measured and its dependence on temperature and rotation rate to be studied. \\citet{Barnesetal05} find that the amplitude $\\Delta \\Omega$ of the DR has a weak dependence on the angular velocity of rotation $\\Omega$, i.e., $\\Delta \\Omega \\propto \\Omega^{\\alpha}$ with $\\alpha = 0.15 \\pm 0.10$, while a remarkably stronger correlation is found with stellar effective temperature, i.e., $\\Delta \\Omega$ increases with increasing effective temperature from M to G-type stars, thus extending the dependence found by \\citet{Reiners06} to lower temperatures. { In main-sequence stars, the detected differential rotation is solar-like; i.e., the equator rotates faster than the pole. Nevertheless, an anti-solar differential rotation has been suggested in some late-type giants \\citep[see, e.g., ][ and references therein]{Kovarietal07}}. For late-type stars that are slowly rotating ($v \\sin i \\la 12-15$~km~s$^{-1}$), Doppler Imaging cannot be applied and information on surface DR can be extracted solely by photometric techniques. The chromospheric fluxes in the cores of the Ca~II~H\\&K lines have been monitored along several decades for a sample of late-type main-sequence stars in the framework of the Mt.~Wilson project to study rotation and stellar activity cycles. It has provided information on the dependence of DR on stellar rotation rate \\citep{Donahueetal96} thanks to the quite long lifetime of chromospheric plages in comparison with the stellar rotation period that makes them good tracers for pointing out the differences in rotation rate at different latitudes \\citep{Donahueetal97a,Donahueetal97b}. In the Sun, the activity belts where active regions form migrate with the phase of the activity cycle. This is interpreted as a migration of the latitude of maximum toroidal magnetic field close to the base of the convection zone \\citep[e.g., ][]{DikpatiCharbonneau99}. A similar migration is expected in late-type stars that have a solar-like dynamo, producing a systematic variation in the period of the photometric modulation with the phase of the cycle. Such a variation has indeed been observed in the rotational modulation of the Sun-as-a-star chromospheric flux and provides an estimate of the amplitude of solar DR \\citep{DonahueKeil95}. A key parameter is the length of the time interval used to determine the seasonal solar rotation period. It is calibrated by trying to match two contrasting requirements: a) avoid remarkable variations of the large scale pattern of chromospheric inhomogeneities that would imply using as short an interval as possible; b) attain sufficient time resolution and low false-alarm probability in determining the period of the rotational modulation that would benefit from a time interval that is as long as possible. In the Sun, the optimal extension of the seasonal time interval is found to be $150-200$ days that makes a compromise between the two opposite requirements. This is allowed because chromospheric active regions and activity complexes are remarkably long-lived in comparison with photospheric spots having a mean lifetime of $50-80$ days vs. $\\sim 10-15$ days, respectively \\citep[cf. ][]{Donahueetal97a,Lanzaetal03}. As a matter of fact, a similar approach based on photospheric sunspots was not successful because of the random longitude appearance and short lifetime of individual sunspot groups \\citep{LaBonte82}. The situation is different in the case of very active, young solar-like stars whose rotation period is shorter than the Sun's and whose photospheric starspots have lifetimes of several months. Therefore, the method was successful in that case \\citep[see, e.g., ][]{MessinaGuinan02,MessinaGuinan03}. For the highly active and fast-rotating subgiant component stars in close binary systems such as RS Canum Venaticorum binaries, the persistence of active longitudes for decades allows us to measure a low-amplitude DR using photospheric starspots as tracers \\citep[$\\Delta \\Omega /\\Omega \\sim 10^{-3}$ in HR 1099, cf. ][]{Lanzaetal06,BerdyuginaHenry07}. The majority of the stars observed by CoRoT and Kepler in the optical passband are not suitable to this approach because their photospheric active regions have lifetimes that are shorter than the typical timescale of DR shear, i.e., $1/\\Delta \\Omega$, where $\\Delta \\Omega$ is the amplitude of the DR. This limits the precision in the determination of the rotation period attainable with periodogram techniques, even in the case of a uniformly sampled time series \\citep{Lanzaetal93,Lanzaetal94}. On the other hand, if starspot intrinsic evolution is negligible, periodogram techniques coupled with a pre-whitening approach can be successful for estimating DR in solar-like stars by pinpointing the rotation frequencies of spots at different latitudes \\citep{ReinholdReiners13,Reinholdetal13}. To make progress in the measurement of DR in late-type stars having starspots that evolve on a timescale comparable to $1/\\Delta \\Omega$ or possibly shorter, we investigate the potentiality of a simple starspot model to extract DR. Our approach applies a simple autocorrelation technique to estimate the coherence time of the light modulation that provides an estimate of the spot lifetime to be compared with the shear timescale. This allows us to select promising candidates for spot modelling. For a given star, we perform a screening of the time intervals showing variations that likely stem from the effect of DR rather than from intrinsic starspot evolution. Finally, we apply a Monte Carlo Markov Chain (hereafter MCMC) method to estimate the most probable value of DR and its standard deviation following an approach introduced by \\citet{Croll06}. We compare the proposed method with previous ones by analysing a sample of stars observed by MOST, CoRoT, and Kepler whose DR has been extracted with different spot modelling approaches. Moveover, we also consider the case of the Sun as a star to show the limitation of the method in the case of a slowly rotating star. ", "conclusions": "We have introduced a method of searching for differential rotation (DR) signals in high-precision photometric time series of the late-type stars acquired by space-borne telescopes (see Sect.~\\ref{overview}). The even sampling of the time series allowed us to apply an autocorrelation to measure the timescale of the intrinsic evolution of the spot pattern that limits the possibility of measuring DR. We selected a sample of eight stars for which previous determinations of DR based on photometry were available along with some information on the inclination of the stellar rotation axis. We added the Sun as a star to this sample to study the rotation of solar analogues. We found that a significant DR can be detected when the relative height of the second maximum in the autocorrelation function is at least $0.6-0.7$. In this case, we subdivided the time series into intervals the duration of which was shorter than the typical timescale of starspot evolution, although long enough to reveal the drift of the spot longitudes produced by DR. The best interval was then selected by comparing the best fits obtained with a simple two-spot model with and without DR. For that interval, a fully Bayesian analysis was undertaken by means of an MCMC approach to assess the significance of DR and determine its most probable value and uncertainty. We found that the amplitude of the DR derived with our method is generally consistent with previous results when the different assumptions of the other approaches are considered. The advantage of our approach is the simplicity of the spot model that allows us to run MCMCs with tens or hundreds of million steps to study the a posteriori distribution of the parameters. This is particularly useful in spot modelling given the strong correlations among different parameters. We took advantage of the available information on the inclination of the stellar rotation axis to fix the a priori distribution of the inclination that is strongly correlated with the colatitudes and the unprojected areas of starspots in our model. Other correlations may become important when the signal-to-noise ratio of the photometry is low, spots are rapidly evolving ($\\tau_{\\rm s} \\la (2.5-3.0) P $), or the DR relative amplitude is small ($\\Delta P / P \\la 0.01$). They severely hamper the proper mixing and convergence of the MCMC procedure, although their effect can be controlled as described in Sect.~\\ref{param_corr}, provided that a clear signal of DR is present in the light curve. In the case of stars more active than the Sun, we find that the measured amplitude of DR depends on the specific time interval considered. In general, our approach only provides a lower limit to the amplitude of surface DR. We did not attempt to extract the amplitude and sign of the pole-equator shear as in other works \\citep[e.g., ][]{Froehlichetal12} because spot colatitude and inclination of the stellar spin axis are highly correlated in our simple spot model. On the other hand, an independent estimate of the inclination is generally not available, except in the case of close eclipsing binaries or transiting star-planet systems." }, "1402/1402.4836_arXiv.txt": { "abstract": "We present updated analyses of pulse profiles and their arrival-times from PSR B1534+12, a 37.9-ms radio pulsar in orbit with another neutron star. A high-precision timing model is derived from twenty-two years of timing data, and accounts for all astrophysical processes that systematically affect pulse arrival-times. Five ``post-Keplerian\" parameters are measured that represent relativistic corrections to the standard Keplerian quantities of the pulsar's binary orbit. These relativistic parameters are then used to test general relativity by comparing the measurements with their predicted values. We conclude that relativity theory is confirmed to within 0.17\\% of its predictions. Furthermore, we derive the following astrophysical results from our timing analysis: a distance of $d_{\\textrm{GR}} = 1.051 \\pm 0.005$ kpc to the pulsar-binary system, by relating the ``excess\" orbital decay to Galactic parameters; evidence for pulse ``jitter\" in PSR B1534+12 due to short-term magnetospheric activity; and evolution in pulse-dispersion properties. As a secondary study, we also present several analyses on pulse-structure evolution and its connection to relativistic precession of the pulsar's spin axis. The precession-rate measurement yields a value of $\\Omega_1^{\\textrm{spin}}$ = $0.59^{+0.12}_{-0.08}$ $^{\\circ}$/year (68\\% confidence) that is consistent with expectations, and represents an additional test of relativistic gravity. ", "introduction": "Pulsars in relativistic binary systems have provided the most rigorous tests of gravitational theory in strong fields to date. High-precision timing of such an object produces a timing model that describes ``post-Keplerian\" (PK) effects that characterize relativistic corrections to the standard orbital elements \\citep{dd85,dd86}, as well as its nominal spin, astrometric, and environmental properties. Comparisons between measured and expected PK parameters produce tests of the gravitational theory in question. The ``Hulse-Taylor\" pulsar \\citep{ht75a} provided the first such positive case for general relativity, and still serves as an excellent laboratory for strong-field gravity \\citep{wnt10}. The recent discovery of a massive pulsar in a highly relativistic orbit with a white dwarf yields a testing ground for tensor-scalar extensions of gravitational theory \\citep{afw+13}. An extensive analysis of the ``double-pulsar\" system \\citep{bdp+03} constrains general relativity to within 0.05\\% of its predictions and remains the most stringent pulsar-timing test so far \\citep{ksm+06}. PSR B1534+12 was discovered by \\citet{wol91a} to be in a highly inclined, 10.1-hour binary orbit with another neutron star. Follow-up timing studies on this pulsar \\citep{sac+98,sttw02} produced a timing solution that yielded measurements of five PK parameters: \\(\\dot{P}_b\\), the orbital decay of the binary system; \\(\\dot{\\omega}\\), the advance of periastron longitude; \\(\\gamma\\), the time-averaged gravitational-redshift and time-dilation parameter; \\(r\\textrm{ and }s\\), the ``range\" and ``shape\" of the Shapiro time delay. Simultaneous measurement of these parameters produced a self-consistent set of tests that complemented the Hulse-Taylor results by including tests based only on quasi-stationary, non-radiative PK parameters \\citep{twdw92}. Additional results were derived using the fitted timing model, including a precise estimate of the pulsar's distance using the measured excess of orbital decay due to relative motion \\citep{bb96}. \\begin{deluxetable*}{ccccccccc} \\tabletypesize{\\scriptsize} \\tablecaption{Timing Parameters for Each Backend and Frequency} \\tablewidth{0in} \\tablehead{ \\colhead{Parameter} & \\colhead{Mark III} & \\colhead{Mark IV} & \\colhead{Mark IV} & \\colhead{ASP} & \\colhead{ASP} & \\colhead{ASP} & \\colhead{ASP} & \\colhead{ASP} } \\startdata Frequency (MHz) \\dotfill & 1400 & 430 & 1400 & 424 & 428 & 432 & 436 & 1400 \\\\ Bandwidth (MHz) \\dotfill & 40 & 5 & 5 & 64 & 64 & 64 & 64 & 64 \\\\ Spectral Channels\\dotfill & 32 & 1\\tablenotemark{a} & 1\\tablenotemark{b} & 1 & 1 & 1 & 1 & 16\\tablenotemark{c} \\\\ Number of TOAs \\dotfill & 1185 & 3102 & 664 & 1204 & 1197 & 1190 & 1124 & 231 \\\\ Dedispersion type \\dotfill & Incoh. & Coh. & Coh. & Coh. & Coh. & Coh. & Coh. & Coh. \\\\ Integration time (s) \\dotfill & 300 & 190 & 190 & 180 & 180 & 180 & 180 & 180 \\\\ Date span (years) \\dotfill & 1990-94 & 1998-2005 & 1998-2005 & 2004-12 & 2004-12 & 2004-12 & 2004-12 & 2004-12 \\\\ RMS residual, \\(\\sigma_{rms}\\) & 5.31 & 4.21 & 6.73 & 4.48 & 4.34 & 4.65 & 4.93 & 8.27 \\enddata \\label{tab:backend} \\tablenotetext{a}{Four sub-bands centered at 430 MHz were taken when the Mark IV data were originally recorded, but were averaged together to build signal strength.} \\tablenotetext{b}{Two sub-bands centered at 1400 MHz were taken when the Mark IV data were originally recorded, but were also averaged together to build signal strength.} \\tablenotetext{c}{The number of actual channels recorded sometimes varied due to computational limitations, so this value represents a typical number of channels used.} \\end{deluxetable*} The time-averaged pulse profile of PSR B1534+12 is undergoing a secular change in observed radiation pattern at a rate of 1\\% per year \\citep{arz95}. Such changes can be linked to spin-orbit coupling in a strong gravitational field, which results in a precession of the pulsar's spin axis \\citep[``relativistic spin precession\";][]{ds16} and an evolving view of the two-dimensional beam structure \\citep{kra98}. \\citet*{sta04} (hereafter STA04) developed a general technique to characterize the overall profile shape at a given epoch and derive a precession rate by measuring and comparing spin-precession and orbital-aberration effects that produce the observed shape evolution. The results of this study yielded a direct measurement of the precession rate that was consistent with the rate predicted by general relativity, albeit with considerably limited precision. Furthermore, the geometry of PSR B1534+12 and its binary system was derived by combining these results with a rotating-vector-model \\citep[RVM,][]{rc69a} analysis of the evolving polarization properties. PSR B1534+12 currently remains the only pulsar for which special-relativistic orbital aberration is observed. {\\bfref The effects of relativistic spin precession on pulse structure have also been observed in PSR B1913+16 \\citep{wrt89}, the double-pulsar system \\citep{bkk+08}, and most dramatically in PSR J1141-6545 \\citep{mks+10}.} In this work, we report on updated timing and profile-evolution analyses of PSR B1534+12, using data sets that collectively span 22 years since its discovery. Results from the analyses described below include improvements in tests of general relativity, an improved measurement of the pulsar's precession rate, and additional findings extracted from our time series. A full discussion of all current results is provided in Section \\ref{sec:discuss}. ", "conclusions": "\\label{sec:discuss} \\subsection{Pulse Jitter in PSR B1534+12} \\label{sec:jitter} The measurability of $\\ddot{\\nu}$ and $\\dddot{\\nu}$ in PSR B1534+12 strongly suggests a significant amount of timing noise across our data set. This polynomial whitening in our best-fit model removes most of the long-term timing noise, which is usually attributed to rotational instabilities and variations in magnetospheric torque. However, TOA residuals generally exhibit scatter on shorter, pulse-period timescales in excess of standard measurement uncertainties. This residual ``jitter\" is manifested from slight changes in the shape, amplitude and pulse phase of recorded profiles between successive pulses, and is likely due to variable activity within the pulsar magnetosphere. Recent studies of pulse jitter suggest that timing precision can be improved when averaging consecutive TOA residuals together \\citep{sc12}. We believe that such pulse jitter is evident in our timing analysis of PSR B1534+12. Figure \\ref{fig:jitter} displays TOA residuals as a function of orbital phase from a global fit using unweighted TOA uncertainties recorded simultaneously on MJD 53545. The dark-blue data points were recorded with the Mark IV backend, while additional points represent the four channelized ASP data sets. While residual variations are visibly uncorrelated within the shown timescale, there is a visible correlation between the two overall data sets despite significant differences in backend specifications. We therefore associate this backend-correlated scatter as pulse jitter due to irregular activity within the pulsar's magnetosphere. \\begin{figure}[h] \\begin{center} \\includegraphics[scale=0.43]{./f4.pdf} \\caption{Pulse jitter in PSR B1534+12. The above figure shows global-fit residuals of TOAs recorded on MJD 53545 as a function of orbital phase. The dark-blue points were recorded with the Mark IV pulsar backend, and the remaining colors represent the four channels of ASP 430-MHz data recorded simultaneously. } \\label{fig:jitter} \\end{center} \\end{figure} Several implications arise from the observed pulse jitter. First, pulse jitter will become a significant source of timing error in future timing studies of PSR B1534+12. The recent installation of the PUPPI signal processor\\footnote{\\url{http://www.naic.edu/~astro/guide/node11.html}} is expected to produce high-precision residuals with scatter that strongly reflects time-dependent inhomogeneities in the magnetosphere. Second, a long-term solution to jitter with PSR B1534+12 cannot involve averaging a large number of consecutive TOAs. The main objective of strong-gravity tests with pulsars is to monitor time-dependent changes to orbital elements and quasi-static PK parameters, which requires full coverage of the orbit over long periods of time. We therefore use this jitter as a means to justify the TOA-uncertainty compensation for the global timing solution described in Section \\ref{sec:obsred}. Lastly, further instrumental upgrades will only improve measurements made at 1400 MHz, where PSR B1534+12 is intrinsically weaker and signal-to-noise is currently limited. However, the overall timing solution is still expected to improve with additional observations over time. \\begin{deluxetable*}{lcc} \\tabletypesize{\\scriptsize} \\tablecaption{Orbital Elements for PSR B1534+12} \\tablewidth{0in} \\tablehead{\\colhead{Parameter} & \\colhead{DD Model} & \\colhead{DDGR Model}} \\startdata Projected semimajor axis, \\(x\\) (s) \\dotfill & 3.7294636(6) & 3.72946417(13) \\\\ Eccentricity, \\(e\\) \\dotfill & 0.27367752(7) & 0.27367740(4) \\\\ Epoch of periastron, \\(T_0\\) (MJD) \\dotfill & 52076.827113263(11) & 52076.827113271(9) \\\\ Orbital Period, \\(P_b\\) (days) \\dotfill & 0.420737298879(2) & 0.420737298881(2) \\\\ Argument of periastron, \\(\\omega\\) (deg) \\dotfill & 283.306012(12) & 283.306029(10) \\\\ & & \\\\ Rate of periastron advance, \\(\\dot{\\omega}\\) (deg yr\\(^{-1}\\)) \\dotfill & 1.7557950(19) & 1.755795\\tablenotemark{a} \\\\ Time-averaged gravitational redshift, \\(\\gamma\\) (ms) \\dotfill & 2.0708(5) & 2.0701\\tablenotemark{a} \\\\ Orbital decay, \\((\\dot{P}_b)^{\\textrm{obs}}\\) (\\(10^{-12}\\)) \\dotfill & -0.1366(3) & -0.19244\\tablenotemark{a} \\\\ Shape of Shapiro delay, \\(s = \\textrm{sin}i\\) \\dotfill & 0.9772(16) & 0.97496\\tablenotemark{a} \\\\ Range of Shapiro delay, \\(r = T_{\\odot}m_2\\) (\\(\\mu\\)s) \\dotfill & 6.6(2) & 6.627\\tablenotemark{a} \\\\ & & \\\\ Companion mass, \\(m_2\\textrm{ }(M_{\\odot})\\) \\dotfill & 1.35(5) & 1.3455(2) \\\\ Pulsar mass, \\(m_1\\textrm{ }(M_{\\odot})\\) \\dotfill & $\\cdots$ & 1.3330(2)\\tablenotemark{a} \\\\ Total mass, \\(M = m_1+m_2 \\textrm{ }(M_{\\odot})\\) \\dotfill & $\\cdots$ & 2.678463(4) \\\\ Excess \\(\\dot{P}_b\\textrm{ }(10^{-12})\\) \\dotfill & $\\cdots$ & 0.0559(3) \\\\ \\enddata \\label{tab:binarypar} \\tablecomments{Values in parentheses denote the uncertainty in the preceding digit(s).} \\tablenotetext{a}{derived quantity} \\end{deluxetable*} \\begin{figure}[h] \\begin{center} \\includegraphics[scale=0.43]{./f5.pdf} \\caption{Phase structure function $D_{\\phi}$ as a function of time lag $\\tau$. The solid line is a best-fit model of Equation \\ref{eq:dmstructpred} for data with lags between 70 and 900 days.} \\label{fig:dmstruct} \\end{center} \\end{figure} \\subsection{Long-term Variations in Dispersion Measure} \\label{sec:vardm} Figure \\ref{fig:dm} illustrates an irregular evolution in DM over time for PSR B1534+12. Five large bins, each with a fitted gradient, are used in our timing model to fully describe the observed changes across different timespans. Recent studies have shown that several objects also exhibit nonlinear evolution in DM along different directions and distances \\citep[e.g.][]{kcs+13}. As with these studies, we believe that the dominant source of such evolution is the inhomogeneity of the interstellar medium, which is traced by the pulsar's signal as the line of sight sweeps through different regions due to a significant relative motion. {\\bfref These long-term DM measurements are useful for a statistical analysis of turbulence within the interstellar medium \\citep*[e.g.][]{ktr94}, which usually assumes that the power spectrum of spatial variations in electron density is a power law within a range of length scales \\citep{ric90}, \\begin{equation} P(q) \\propto q^{-\\beta}, \\hspace{20pt} q_o < q < q_i \\label{eq:pq} \\end{equation} \\noindent where $q = 2\\pi/l$ is a spatial frequency and $l$ is a scattering length. The frequency range in Equation \\ref{eq:pq} corresponds to a range between an ``inner\" ($l_i$) and ``outer\" ($l_o$) length scale where the power-law form is valid. The observed spatial fluctuations due to a relative transverse velocity $v$ are related to a time lag $\\tau$ by $l = v\\tau$. The power spectrum $P(q)$ can therefore be estimated by computing a pulse-phase structure function $D_{\\phi}(\\tau) = \\langle[\\phi(t+\\tau)-\\phi(t)]^2\\rangle$, where the angle brackets represent an ensemble average over observing epoch $t$. The pulse phase $\\phi$ is linearly related to DM, which therefore relates $D_{\\phi}(\\tau)$ to a DM structure function $D_{\\textrm{DM}}(\\tau) = \\langle[\\textrm{DM}(t+\\tau)- \\textrm{DM}(t)]^2\\rangle$, \\begin{equation} D_{\\phi}(\\tau) = \\bigg(\\frac{2\\pi C}{f}\\bigg)^2D_{\\textrm{DM}}(\\tau) \\end{equation} \\noindent where $C = 4.148\\times10^3 \\textrm{ MHz}^2\\textrm{ pc}^{-1}\\textrm{ cm}^3$ s, and $f$ is the observing frequency in MHz. Moreover, $D_{\\phi}(\\tau)$ is a power law in $\\tau$ within the inner length scales defined in Equation \\ref{eq:pq}, which finally requires that \\begin{equation} D_{\\phi}(\\tau) = \\bigg(\\frac{\\tau}{\\tau_0}\\bigg)^{\\beta-2} \\label{eq:dmstructpred} \\end{equation} \\noindent where $\\tau_0$ is a logarithmic intercept. Scintillation theory requires that $\\tau_0 = \\tau_\\textrm{d}$, where $\\tau_\\textrm{d}$ is the diffractive timescale, if the inner length-scale $l_i \\leq v\\tau_{\\textrm{d}}$. We computed values of $D_{\\phi}(\\tau)$ at $f$ = 430 MHz using the Mark IV and ASP small-bin measurements of DM shown in Figure \\ref{fig:dm}. The Mark III DM point was measured using all Mark III TOAs collected over several years, which were generated with a different standard profile than the one used for the Mark IV and ASP data; we therefore chose to ignore this measurement in order to avoid incorporating bias in the structure function. Uncertainties in $D_{\\phi}(\\tau)$ were determined by propagating errors from our DM($t$) measurements. Our estimate of $D_{\\phi}(\\tau)$ is shown in Figure \\ref{fig:dmstruct}, and illustrates a power-law evolution between time lags of roughly 70 and 900 days. We fitted Equation \\ref{eq:dmstructpred} to this segment of data, and found that \\begin{align} \\beta &= 3.70 \\pm 0.04 \\nonumber \\\\ \\tau_0 &= 3.0 \\pm 0.8 \\textrm{ minutes} \\label{eq:n8} \\end{align} \\noindent which is shown as a solid black in in Figure \\ref{fig:dmstruct}. The measured spectral index $\\beta$ is consistent with the value for a ``Kolmogorov\" medium, $\\beta_{\\rm Kol}$ = 11/3. Furthermore, $\\beta$ and $\\tau_0$ in Equation \\ref{eq:n8} are consistent with the structure-function estimates reported by \\citet{sw12}. Our estimate of $\\tau_0$ is also consistent with the value of $\\tau_{\\textrm{d}}$ measured from the autocorrelation function of a dynamic spectrum of PSR B1534+12 \\citep{bplw02}. At large timescales, the structure function departs from the fitted model at a lag $\\tau_o \\approx 900$ days, which suggests that \\begin{equation} l_o \\approx 52 \\bigg(\\frac{v}{100\\textrm{ km/s}}\\bigg)\\textrm{ AU} \\label{eq:l_o} \\end{equation} \\noindent \\citet{bplw02} derived an interstellar scintillation (ISS) velocity of 192 km/s. They noted in their study that ISS velocities of pulsars are dominated by the systemic transverse component, which means that $v \\approx 192$ km/s for PSR B1534+12, and $l_o \\sim 100$ AU $\\sim 10^{15}$ cm from Equation \\ref{eq:l_o}. This estimate is consistent with the upper limit of $l_o$ observed for several pulsars by \\citet{pw91}. By contrast, there is no evidence for a significant inner scale from our data set, since bins with mean values less than 70 days contain only one or two pairs of DM$(t)$ and were therefore ignored from the analysis. We did not apply any correction for the solar-wind contribution of our DM($t$) measurements, due to a covariance between the TEMPO solar-wind DM model and a fitted timing parameter that is discussed at the end of Section \\ref{sec:dist}.} \\subsection{High-precision Distance to PSR B1534+12} \\label{sec:dist} Relative acceleration between the observatory and pulsar systems in the Galactic potential causes significant Doppler-factor biases in PK parameters. \\citet{sac+98,sttw02} noted such behavior in earlier data sets of PSR B1534+12 with the observed orbital decay and applied a distance-dependent kinematic correction derived by \\citet{dt92} in order to include it as a consistent, radiative test of general relativity (see Section \\ref{sec:tests} below). However, the corrected value yielded a large uncertainty due to the imprecise distance to the pulsar derived from DM measurements using the \\citet{tc93} model of free electrons in the Galaxy. Previous studies of PSR B1534+12 therefore solved the inverse problem suggested by \\citet{bb96}, where general relativity is assumed to be correct; the distance is then derived using the measured ``excess\" orbital decay, a model of the Galactic acceleration \\citep{kg89}, and the expression for kinematic correction derived by \\citet{dt92}. Using this procedure, \\citet{sttw02} were able to derive a distance with a relative uncertainty of 4.9\\% when doubling their TEMPO uncertainties. We used the same approach in this study to update the derived distance with a substantially longer timespan of Arecibo data. We also corrected the expression for the kinematic bias presented in \\citet{sac+98,sttw02} for missing factors of the cosine of the pulsar's galactic latitude; the correct equation is given by \\citet[Equation 5]{nt95}. Our derived distance to PSR B1534+12, using our timing results, the corrected kinematic equation and updated Galactic parameters from \\citet{rmb+14}, is {\\bfref \\begin{equation} d_{\\textrm{GR}} = 1.051 \\pm 0.005 \\textrm{ kpc} \\label{eq:distance} \\end{equation} \\noindent where the value and its uncertainty (68\\% confidence level) were estimated using a Monte-Carlo method: all uncertain parameters were randomly sampled from a normal distribution with mean and standard deviation equal to their fitted values and uncertainties, respectively, which were then used to derive a value of $d_{\\textrm{GR}}$ by applying Newton-Raphson's method \\citep[e.g.][]{pftv86}. This process was repeated $10^5$ times, and resulted in a distribution of $d_{\\textrm{GR}}$ that is shown in Figure \\ref{fig:histdist}.} This new distance is consistent with the previous estimate of $1.02 \\pm 0.05$ kpc made by \\citet{sttw02}. The relative uncertainty of this result (0.48\\%) is slightly lower than that of the derived distance to PSR J0437-4715 estimated by \\citet{vbv+08}. We attribute this improvement in precision to the updated $(\\dot{P}_b)^{\\textrm{obs}}$ listed in Table \\ref{tab:binarypar}. The uncertainty in $d_{\\textrm{GR}}$ is dominated by uncertainties in Galactic acceleration and rotation parameters used to derive the estimate. \\begin{figure}[h] \\begin{center} \\includegraphics[scale=0.43]{./f6.pdf} \\caption{Distribution of $d_{\\textrm{GR}}$ obtained by using a Monte-Carlo method described in Section \\ref{sec:dist}.} \\label{fig:histdist} \\end{center} \\end{figure} Despite its high level of precision, the derived distance presented in Equation \\ref{eq:distance} is a model-dependent quantity. An ideal measure of distance can be obtained from the geometric, model-independent parallax through low-frequency interferometry. The recent inclusion of PSR B1534+12 into an extension of the PSR$\\pi$ interferometry program\\footnote{\\url{https://safe.nrao.edu/vlba/psrpi}} will likely provide such an estimate in the next 2-3 years. Another independent distance measure can be derived from a timing parallax estimated in the global-fit timing solution. However, our measured timing parallax was found to be significantly covariant with an input DM parameter associated with free electrons from the solar wind. Such a covariance is unexpected since the solar-wind component of DM is strongest for pulsars close to the ecliptic plane, while PSR B1534+12 is $\\sim 30^{\\circ}$ above the plane. Since the expected solar contribution is much smaller than the scatter of the 80-day DM bins in Figure \\ref{fig:dm}, we chose to set the solar DM component to zero for the global timing fit while acknowledging that the timing parallax is unreliable as a fitted parameter. \\begin{figure}[h] \\begin{center} \\includegraphics[scale=0.43]{./f7.pdf} \\caption{\\emph{Top}: MCMC posterior distribution of $\\Omega_1^{\\textrm{spin}}$ obtained from the profile-shape analysis of Mark IV and ASP data discussed in Section \\ref{sec:precres}. \\emph{Bottom}: Markov chain for $\\Omega_1^{\\textrm{spin}}$ determined from the MCMC algorithm.} \\label{fig:mcmcpos} \\end{center} \\end{figure} \\subsection{Precession and Geometry} \\label{sec:precres} Results from the MCMC fit on several data sets can be found in Table \\ref{tab:profdata}, and the posterior distribution for $\\Omega_1^{\\textrm{spin}}$ derived from our Mark IV and ASP data sets is shown in Figure \\ref{fig:mcmcpos}. {\\bfref We generated $3\\times10^5$ samples for each application of the algorithm, after burning the first 5000 samples in order to remove non-convergent iterations.} We provided the original results obtained by STA04, as well as a reproduced set of results from the STA04 data set using the MCMC algorithm, for comparison with our extended Mark IV and ASP profiles. We assumed that values of $F'$ must be negative while using the MCMC algorithm, since the simultaneous-linear-fit technique used and described by STA04, which avoids any consideration of $F'$, estimates that $\\cos\\eta < 0$. These results agree well with predictions from general relativity, where $\\Omega_1^{\\textrm{spin}} = 0.51^{\\circ}$/yr using the derived masses in Table \\ref{tab:binarypar}, and previous measurements made by STA04. General improvements in precision come from the new fitting procedure, which permitted direct sampling of the precession rate and other free parameters, as well as the addition of the ASP 2005 campaign and several strong bi-monthly observations. \\begin{deluxetable}{lcccc} \\tabletypesize{\\scriptsize} \\tablewidth{0pt} \\tablecaption{Profile-evolution MCMC Parameters} \\tablehead{\\colhead{Parameter} & \\colhead{STA04\\tablenotemark{a}} & \\colhead{STA04} & \\colhead{Mark IV} & \\colhead{All}} \\startdata $\\Omega_1^{\\textrm{spin}}$ $(^{\\circ}/\\textrm{yr})$ \\dotfill & $0.44^{+0.48}_{-0.16}$ & $0.51^{+0.10}_{-0.08}$ & $0.48^{+0.09}_{-0.07}$ & $0.59^{+0.12}_{-0.08}$ \\\\ [5pt] $\\eta$ $(^{\\circ})$ \\dotfill & $\\pm103^{+10}_{-10}$ & $\\pm99^{+2}_{-2}$ & $\\pm118^{+10}_{-15}$ & $\\pm139^{+16}_{-25}$ \\\\ [5pt] $F'$ \\dotfill & n/a & $-5.9^{+0.9}_{-1.0}$ & $-2.2^{+0.6}_{-0.7}$ & $-1.3^{+0.3}_{-0.5}$ \\\\ [5pt] $\\epsilon$ ($10^{-3}$) \\dotfill & $-1.5^{+0.3}_{-0.3}$ & $-1.90^{+0.08}_{-0.09}$ & $-1.21^{+0.08}_{-0.08}$ & $6.67^{+0.07}_{-0.07}$ \\\\ [5pt] \\enddata \\label{tab:profdata} \\tablenotetext{a}{Original, non-MCMC results from STA04.} \\tablecomments{Uncertainties reflect 68\\% confidence intervals of posterior distributions.} \\end{deluxetable} The RVM analysis yielded values of $\\alpha$ and $\\beta$ at different times using the Mark III (reticon), Mark IV and ASP campaign profiles. The values of $\\beta$ measured for each campaign are shown in Figure \\ref{fig:betavtime}. Measurements of $\\alpha = 103.5(3)^{\\circ}$ are consistent with no evolution in time, while the values of $\\beta$ are found to change significantly, where $d\\beta/dt$ = -0.23 $\\pm$ 0.02 $^{\\circ}$/yr. This is consistent with the STA04 result of -0.21 $\\pm$ 0.03 $^{\\circ}$/yr. The assumption that general relativity is correct requires that $d\\beta/dt = \\Omega_1^{\\textrm{spin}} \\sin i\\cos\\eta$, and therefore yields $\\eta = \\pm 117 \\pm 3^{\\circ}$ (68\\% confidence), which agrees with the value determined from the MCMC analysis described above. With these values, the misalignment angle $\\delta$ between the spin and orbital angular momentum axes can be derived through spherical trigonometry by $\\cos\\delta = -\\sin i\\sin\\lambda\\sin\\eta+\\cos\\lambda\\cos i$. The sign ambiguity in $\\eta$ and $i$, as well as the requirement that $\\cos i\\tan\\eta > 0$ pointed out by STA04, gives an expected value of $\\delta = 27.0 \\pm 3.0^{\\circ}$ or $\\delta = 153.0 \\pm 3.0^{\\circ}$. Physical arguments based on alignment of angular momenta prior to the second supernova suggest that the smaller angle is correct \\citep{bai88}, and therefore requires that $\\eta = -117 \\pm 3^{\\circ}$ and $i = 77.7 \\pm 0.9^{\\circ}$. The consistency between the MCMC and RVM analyses serves as an improved, independent check of precession within this relativistic binary system. These results also confirm the geometric picture of this pulsar-binary system derived in STA04. \\citet{kws03} pointed out that relativistic spin precession can eventually imprint a timing signature as a second-derivative in spin frequency. However, they predicted that it would take an additional 25 years of timing observations in order to measure the predicted signature with reasonable accuracy. Moreover, our best-fit timing model of PSR B1534+12 data indicates that the measured $\\ddot{\\nu}$ and $\\dddot{\\nu}$ are dominated by timing noise. \\begin{figure}[h] \\begin{center} \\includegraphics[scale=0.43]{./f8.pdf} \\caption{Impact angle $\\beta$ between the magnetic axis and line of sight as a function of time. The black line is a best-fit slope of -0.23 $\\pm$ 0.02 $^{\\circ}$/yr.} \\label{fig:betavtime} \\end{center} \\end{figure} \\begin{figure}[h] \\begin{center} \\includegraphics[scale=0.43]{./f9.pdf} \\caption{Mass-mass plot for PSR B1534+12. Each set of black curves represents the 68\\%-confidence region delimited by the labeled PK (timing) parameter. The dot-dashed red curves represent the 68\\% confidence region determined from the spin-precession rate. } \\label{fig:m1m2} \\end{center} \\end{figure} \\subsection{Tests of General Relativity} \\label{sec:tests} In general relativity, each PK parameter is expressed as a function of at least one of the two binary-component masses; one can therefore define a ``mass-mass\" space where each PK parameter corresponds to a curve in this plane. \\citet{sac+98} presented the first mass-mass plot that incorporated up to 5 PK curves, as well as the first ``non-mixed\" test of quasi-static parameters using the $\\dot{\\omega}-\\gamma-s$ combination. Figure \\ref{fig:m1m2} presents our evaluation of PK parameters and tests of general relativity using PSR B1534+12. Each curve corresponds to a PK value measured either with the best-fit DD timing model or profile-shape model, while the filled circle represents the best-fit DDGR solution of the pulsar and companion masses of $m_1 = 1.3330\\textrm{ }M_{\\odot}$, $m_2 = 1.3455\\textrm{ }M_{\\odot}$, respectively. The $\\dot{P}_b$ curve was corrected for kinematic bias assuming a pulsar distance of $d = 0.7 \\pm 0.2$ kpc, which was estimated using the electron number-density model developed by \\citet{tc93}, as discussed in Section \\ref{sec:dist}. The discrepancy between this distance and the expected distance (Equation \\ref{eq:distance}) prevents the curve from intersecting the other PK curves, while its large uncertainty dominates the corrected orbital-decay error estimate. The $\\Omega_1^{\\textrm{spin}}$ curves intersect the best-fit DDGR point well at the 68\\% confidence level. The Shapiro $r$ parameter remains a slightly weaker constraint than the PK timing parameters, but its relative uncertainty has improved by nearly a factor of two since the last measurement by \\citet{sttw02}. The $\\dot{\\omega}-\\gamma-s$ combination is the strongest test from this pulsar and confirms general relativity to within 0.17\\% of its predictions. This test quality is slightly larger than the 0.05\\% test from the double-pulsar system, which uses the mass ratio as determined by the projected semi-major axes of both pulsars \\citep{ksm+06}. The high-precision DDGR masses of PSR B1534+12 and its companion are consistent with previous estimates. As noted in \\citet{sttw02}, the significant difference between $m_1$ and $m_2$ presents a conundrum where the spun-up pulsar is actually less massive than its companion. This suggests that a period of ``mass inversion\" -- mass transfer from the pulsar's progenitor to its companion -- took place during the system's evolution, though a more thorough understanding of mass-transfer processes and its effects on stellar structure is needed. The effectiveness of mass estimates from Shapiro-delay measurements will continue to provide a better view of the general pulsar-companion mass population \\citep{kkdt13}.\\\\ The Arecibo Observatory is operated by SRI International under a cooperative agreement with the National Science Foundation (AST-1100968), and in alliance with Ana G. M\\'{e}ndez-Universidad Metropolitana, and the Universities Space Research Association. Pulsar research at UBC is supported by an NSERC Discovery Grant. We thank Z. Arzoumanian, F. Camilo, A. Lyne, D. Nice, J. H. Taylor, and A. Wolszczan for their earlier contributions to this project. We also thank P. Freire, I. Hoffman, A. Lommen, D. Lorimer, D. Nice, E. Splaver, and K. Xilouris for previous assistance with Mark-IV observations. We also thank R. Ferdman, M. Gonzalez, other NANOGrav observers for help with some of the ASP observations reported here. We thank W. Zhu for providing a TEMPO MCMC program. We are grateful for useful comments provided by an anonymous referee." }, "1402/1402.6002_arXiv.txt": { "abstract": "The Canadian Automated Meteor Observatory (CAMO) detects occasional meteors with two maxima in the image intensified CCD based light curves. We report early results from an analysis of 21 of these events. Most of these events show qualitatively similar light curves, with a rounded first luminous peak, followed by an almost linear sharp rise in the second peak, and a relatively rapid curved decay of the second peak. While a number of mechanisms could explain two maxima in the light curves, numerical modelling shows that most of these events can be matched by a simple dustball model in which some grains have been released well before intensive ablation begins, followed by a later release of core grains at a single time. Best fits to observations are obtained with the core grains being larger than the pre-released outer grains, with the core grains typically $10^{-6}$~kg while the early release grains are of the order of $10^{-9}$~kg. ", "introduction": "Light curves provide one of the best indicators of the mode of meteor ablation and fragmentation, and implied meteoroid structure. The intensity of the light produced is indicative of the almost instantaneous mass loss rate. In this paper we consider implications for meteoroid structure and ablation from dual peak meteor light curves, in particular looking at whether these observations support a dustball model. \\cite{Jacchia1955} suggested that photographic observations of shortened trails, flares and meteor wake supported a dustball structure for at least some meteors. However the fragile dustball structure proposed was inconsistent with bright meteors that survived high pressures to low observed heights. \\cite{HawkesJones1975} developed a two component dustball model that would fit both faint and brighter meteor light curves. The key idea of this model was that meteoroids have two components, a grain component with a silicate metallic composition that is responsible for the light production, and a glue component (possibly organic in nature) which has a lower boiling point and does not produce significant luminosity. With this model meteoroids are not necessarily fragile. Under this model some grains will be released early (as the glue reaches its boiling point) while other grains will be separated during intensive light production by the meteor. \\cite{FisherEt2000} reviewed the observational evidence in favour of this dual component dustball model (e.g. shorter more symmetric light curves, relative independence of heights of meteors past a certain mass cutoff, good prediction of ablation profiles across a wide mass range). ", "conclusions": "Perhaps the most obvious question is why some meteors, but not all, show a dual peak structure. A related question is why these events have not been more frequently reported in the past. One possibility is that the relatively high response in the near infrared of the CAMO detection systems may play a factor. One of the most striking aspects of the study was the similarity of different light curves. In almost all cases the first peak is more rounded and symmetric, while the second peak has a sharp almost linear (in a plot using the logarithmic astronomical magnitude as the vertical axis), followed by a fairly sharp rounded decline. This suggests to us that a similar mechanism is producing most of the double light curve events. While other mechanisms can probably also match the results, the simple dustball model used here seems adequate for the events studied. \\cite{MalhotraMathews2010} have conducted a statistical study of smaller meteoroids observed with large aperture radar, and find that only about one-quarter are consistent with simple single body ablation. They find that 48\\% seem to show fragmentation, while 20\\% imply differential ablation is important. If a more detailed future analysis continues to suggest that the core grains are larger than the early release grains, that is an important and somewhat surprising result. This work suggests that core grains are typically~$10^{-6}$~kg while early release grains are of the order of ~$10^{-9}$~kg. A number of previous studies have sought to establish the size of meteor constituent grains. \\cite{Simonenko1968} found from an analysis of rapid onset flares on bright meteors a mean grain mass of about $2\\times10^{-9}$~kg. \\cite{BorovickaEt2007} have modelled six electro-optical Draconid meteors, using an erosion model in which grains are released over generally the first half of the light curve, although for the brightest meteor a number of grains were resistant to release and separated later. They generally found agreement using grains from $10^{-11}$ to $10^{-9}$ kg. They find a model with total pre-release prior to intensive ablation inconsistent with the meteors in this small sample. Of course it should be kept in mind that these were Draconid meteors, known to have a porous low density structure (\\cite{BorovickaEt2007}). \\cite{CampbellKoschny2004} have modelled Leonid light curves using a thermal erosion two component dustball model. The find that two of the meteors are well matched with Gaussian mass distribution of grains, while the third requires a power law distribution. Grain sizes generally range from $1\\times10^{-11}$ to $4\\times10^{-7}$~kg, in general agreement with this study. \\cite{BeechMurray2003} used a power-law mass distribution to match Leonid light curves, generally finding the need for grains in the mass range from $10^{-10}$ to $10^{-7}$~kg. In an analysis of four Leonid fireball bursts \\cite{HawkesEt2002} found that relatively large grains $10^{-5}$ to $10^{-4}$~kg were needed to match the observations. It is also interesting to compare the grain size distribution with that obtained by Stardust during encounter with the coma of Comet 81P/Wild~2. \\cite{Green2007} show that most grains are smaller than those reported here, although the total range is about $10^{-15}$ to $2\\times10^{-5}$~kg. It is possible that the outer grains released prior to ablation are produced in a similar manner to the less spatially and temporally constrained clusters of meteoroids occasionally observed (\\cite{WatanabeEt2003, PiersHawkes1993}). While \\cite{WatanabeEt2003} discusses possible production mechanisms, there is not yet clarity as to how these clusters occur. \\cite{BroschEt2004} point out the many parameters of meteor light curve analysis, and how they can help to constrain meteoroid structure and ablation. \\cite{StokanEt2013} have studied optical trail widths using the CAMO system and this information will help constrain the possible meteoroid models. We plan to in a later paper incorporate the tracking data for a small number of events in a more in depth investigation of dual light curve events which may help to further constrain possible models." }, "1402/1402.3627_arXiv.txt": { "abstract": "First evidence of high-energy astrophysical neutrino observation with the IceCube detector from May 2010 to May 2012 is presented. Selecting for high-energy neutrino events with vertices well contained in the detector volume, the analysis has sensitivity starting at approximately 50 TeV, and the highest energy events are a few PeV. A significant excess in flux is observed above expected backgrounds from atmospheric muons and neutrinos. The sample of 28 events includes the highest energy neutrinos ever observed, and has properties consistent in flavor, arrival direction, and energy with generic expectations for neutrinos of extraterrestrial origin. Using the new found astrophysical neutrinos, spatial searches of clusters are performed to look for exact sources, marking the birth of neutirno astronomy. ", "introduction": "Observation of high-energy neutrinos are thought to provide insight into the origins and acceleration mechanisms of high-energy cosmic rays. Cosmic-ray protons and nuclei produce neutrinos in interactions with gas in their source environment and interstellar dust through decay of charged pions and kaons. These neutrinos have energies proportional to the cosmic rays that produced them and point back to their sources since they are neither affected by magnetic fields nor absorbed by matter opaque to radiation. Neutrinos can also provide insight into the creation of gamma-rays at astronomical sources. Gamma-rays can be produced by mesons from cosmic ray interactions in a similar process to neutrinos. This is known as the hadronic process, but they can also be produced by leptonic processes such as inverse compton scattering of electrons in high-energy source environments. Neutrino observation of gamma-ray sources will confirm their hadronic origin. Large-volume Cherenkov detectors like IceCube \\cite{daqpaper} can detect these neutrinos through production of secondary leptons and hadronic showers when they interact with the detector material. This analysis has an event selection with an energy threshold at about 50 TeV and is sensitive to energies up to 10 PeV and beyond, depending on the flux being measured. By selecting for events well contained in the detector volume, it is sensitive to all neutrino flavors from all directions. Hints of astrophysical sources are seeked by characterizing the astrophysical flux measured and looking for spatial clustering among these events. The analysis uses a data-taking period that started in May 2010 using 79 strings and continuing with the completed detector (86 strings) from May 2011 to May 2012 for a total livetime of 662 days. ", "conclusions": "An analysis of two years of IceCube data from 2010 to 2012 has observed a flux incompatible with expectations from terrestrial processes. It contains a mixture of neutrino flavors compatible with a flux proportional to $E^{-2}$, a spectrum expected for neutrinos associated with primary cosmic ray acceleration. The sample is thus consistent with generic expectations for a neutrino population with origins outside the solar system. We did not observe significant spatial clustering of the events, although this study is currently limited by low statistics and poor angular resolution for the majority of the observed events. Future observations with IceCube will provide improved measurements of the energy spectrum and origins of this flux, providing insight into the underlying processes responsible for these events." }, "1402/1402.5976_arXiv.txt": { "abstract": "We present spectral energy distributions (SEDs) of 69 QSOs at $z>5$, covering a rest frame wavelength range of 0.1\\,$\\mu$m to $\\sim$80\\,$\\mu$m, and centered on new {\\it Spitzer} and {\\it Herschel} observations. The detection rate of the QSOs with {\\it Spitzer} is very high (97\\% at $\\lambda_{\\rm rest}\\,\\lesssim\\,4\\,\\mu$m), but drops towards the {\\it Herschel} bands with 30\\% detected in PACS (rest frame mid-infrared) and 15\\% additionally in the SPIRE (rest frame far-infrared; FIR). We perform multi-component SED fits for {\\it Herschel}-detected objects and confirm that to match the observed SEDs, a clumpy torus model needs to be complemented by a hot ($\\sim$1300K) component and, in cases with prominent FIR emission, also by a cold ($\\sim$50K) component. In the FIR detected cases the luminosity of the cold component is on the order of $10^{13}$ $L_{\\odot}$ which is likely heated by star formation. From the SED fits we also determine that the AGN dust-to-accretion disk luminosity ratio declines with UV/optical luminosity. Emission from hot ($\\sim$1300K) dust is common in our sample, showing that nuclear dust is ubiquitous in luminous QSOs out to redshift 6. However, about 15\\% of the objects appear under-luminous in the near infrared compared to their optical emission and seem to be deficient in (but not devoid of) hot dust. Within our full sample, the QSOs detected with {\\it Herschel} are found at the high luminosity end in $L_{\\rm UV/opt}$ and $L_{\\rm NIR}$ and show low equivalent widths (EWs) in H$\\alpha$ and in Ly$\\alpha$. In the distribution of H$\\alpha$ EWs, as determined from the {\\it Spitzer} photometry, the high-redshift QSOs show little difference to low redshift AGN. ", "introduction": " ", "conclusions": "" }, "1402/1402.2290_arXiv.txt": { "abstract": "We explore the use of different radio galaxy populations as tracers of different mass halos and therefore, with different bias properties, to constrain primordial non-Gaussianity of the local type. We perform a Fisher matrix analysis based on the predicted auto and cross angular power spectra of these populations, using simulated redshift distributions as a function of detection flux and the evolution of the bias for the different galaxy types (Star forming galaxies, Starburst galaxies, Radio-Quiet Quasars, FRI and FRII AGN galaxies). We show that such a multi-tracer analysis greatly improves the information on non-Gaussianity by drastically reducing the cosmic variance contribution to the overall error budget. By using this method applied to future surveys, we predict a constraint of $\\sigma_{f_{nl}}=3.6$ on the local non-Gaussian parameter for a galaxy detection flux limit of 10$ \\mu$Jy and $\\sigma_{f_{nl}}=2.2$ for 1 $\\mu$Jy. We show that this significantly improves on the constraints obtained when using the whole undifferentiated populations ($\\sigma_{f_{nl}}=48$ for 10 $\\mu$Jy and $\\sigma_{f_{nl}}=12$ for 1 $\\mu$Jy). We conclude that continuum radio surveys alone have the potential to constrain primordial non-Gaussianity to an accuracy at least a factor of two better than the present constraints obtained with Planck data on the CMB bispectrum, opening a window to obtain $\\sigma_{f_{nl}}\\sim1$ with the Square Kilometer Array. ", "introduction": "In the current standard cosmological model, the large scale structures observed in the Universe originated from small fluctuations present in the matter density field arising after the inflationary phase shortly after the Big Bang. Although slow-roll inflation models predict this random field to be essentially Gaussian \\citep{Maldacena, Acquaviva}, other evolutionary models after inflation predict a non-vanishing non-Gaussian component in the primordial matter density field \\citep{Verde, Liguori}. The detection of non-Gaussianity could open a new window in the knowledge of early Universe physics. The most widely used method so far to constrain primordial non-Gaussianity is to measure the bi-spectrum in the Cosmic Microwave Background (CMB) temperature anisotropy maps. Recently, this method was applied to Planck data \\citep{Planck2} to provide the best measurement of local non-Gaussianity up to date ($f_{nl}=2.7 \\pm 5.8$)\\footnote{Note that in our definition of $f_{nl}$, with the growth factor normalized to unity today, this value should be multiplied by a factor of $\\approx 1.3$ \\citep[see e.g.][]{Dalal,Afshordi}.}. A complementary way to access non-Gaussianity is to measure its impact on Large Scale Structure (LSS) at lower redshifts \\citep{Dalal,b2, carbone} which affects the bias of dark matter tracers. Large galaxy surveys have been the probe of choice to constrain the clustering properties of dark matter on large scales and measure this non-Gaussian effect \\citep{Xia2010, Xia2011, Bernardis}. Given the fact that this non-Gaussian signal is specially sensitive on large scales, other type of surveys, based on intensity mapping techniques, have been suggested to go after this effect \\citep{Joudaki, Camera2013}. Although observing large scales has the advantage of probing the regime where the non-Gaussian effect on the bias of the dark matter tracers is stronger, it has the problem that it is nonetheless limited by cosmic variance, e.g., the lack of enough independent measurements for the scales we are trying to probe, given the limited size of the volume that is observed. \\citet{Seljak} proposed a way to get around this cosmic variance limitation by using different biased tracers of the underlying dark matter distribution. With at least two tracers with different bias, we can make a measurement of the ratio of these two biases that is only limited by shot noise and hence beats cosmic variance. This is specially sensitive when the bias of one object is much larger than the other. In order to understand this, let us assume that we measure two density maps at a given redshift, one for a biased tracer and another for the dark matter itself. Due to cosmic variance, there will be several power spectra (e.g. several cosmologies) that are consistent with the dark matter map. However, the ratio of the two maps should give a direct measurement of the bias, with an uncertainty just given by the shot noise of the tracer. In this paper we propose to use the mass of the haloes hosting the dark matter tracers as an extra source of information in order to constrain the bias. This will have 3 advantages: 1) it will allow us to select objects with large bias factors (without any mixture with low mass objects) making them more sensitive to the non-Gaussian effect; 2) it will allow us to compare directly the bias of different tracers in order to cancel cosmic variance and 3) it will allow us to have a more physical description of the bias parameters used in the analysis. Although this idea can be applied to different surveys, it can be particularly relevant to radio galaxy surveys, given that they usually lack redshift information but on the other hand should provide a more direct relation between bias and mass. In here we concentrate on surveys that can be achieved with the SKA (http://www.skatelescope.org) and its pathfinders. The outline for this work is the following: In Sec. 2 we describe how to model non-Gaussian features in LSS bias, and their mass dependence. In Sec. 3 we present the forecasts for future radio surveys, focusing on the improvement of the constraints on $f_{nl}$ when using tracers of different masses with standard angular power spectrum measurements. In Sec. 4, we show how to improve these constraints by eliminating cosmic variance using multi tracer bias. Finally, in Sec. 5 we discuss the results and present our conclusions. \\section[]{Primordial non-Gaussianity in LSS clustering} \\label{section2} \\subsection{The effect on bias} In most standard inflationary scenarios, non-Gaussianity in the primordial fluctuation field is characterized by a local feature of the Bardeen gauge invariant potential {$\\Phi$}, which is expressed as: \\begin{equation} \\Phi=\\phi+f_{nl}(\\phi^2-\\langle \\phi \\rangle^2) \\label{eq:1} \\end{equation} where $\\phi$ is a random Gaussian field and $f_{nl}$ is a constant that defines the amplitude of non-Gaussianity. On sub-horizon scales, the Bardeen potential simply becomes the opposite of the gravitational potential. In terms of large scale structure clustering in the Universe, the main consequence of such deviation from a Gaussian field is to increase power on large scales in 2-point statistics such as the tridimensional power spectrum of clustered objects. A scale dependent correction appears in the total halo bias: \\citep{Dalal,b2} \\begin{equation} b_h(M,z)=b_L(M,z)+f_{nl}\\delta_c \\left[(b_L(M,z)-1\\right]\\frac{3 \\Omega_m H_0^2}{c^2 k^2 T(k) D(z)} \\label{eq:2} \\end{equation} The total halo bias thus depends on the non-Gaussianity amplitude $f_{nl}$, the critical overdensity for spherical collapse $\\delta_c$, at redshift z=0, as well as the linear transfer ($T(k)$) and growth ($D(z)$) functions (with $D(0)=1$). This correction is placed upon the usual Gaussian linear bias, which assumes the expression: \\begin{equation} b_L(M,z)=1+\\frac{q \\nu -1}{\\delta_{c}(z)}+\\frac{1}{\\delta_c(z)}\\frac{2p}{1+(q\\nu)^p} \\label{eq:3} \\end{equation} where $\\nu=\\delta_c^2(z)/\\sigma_0^2(M)$ and $\\delta_c(z)= D(z) \\delta_c$. This model was first proposed in \\cite{MoWhite} and \\cite{SethTormen} obtained the values for the parameters which better fit numerical simulations of dark matter collapse and galaxy formation, finding $p=0.3$ and $q=0.75$. \\subsection{Mass dependence of non-Gaussian effects in the power spectra} For a given tracer with bias $b_h(z)$, the total 3D halo power spectrum is then simply $P^{3D}_h=b_h^2 P_{\\delta}$ (on large scales), where $P_{\\delta}$ is the underlying dark matter power spectrum. If no redshift information is available to compute the 3D power spectrum, the adequate estimator to use is the angular power spectrum, where the distribution of halos along the line of sight is projected over the 2D field of view. If several tracers are considered, the total statistical information is described by the auto and cross correlation power spectra. The result consists in a set of multipole values $C_l^{i,j}$ and the full computation reads \\citep{Huterer}: \\begin{equation} C_l^{i,j}=\\frac{2}{\\pi}\\int_{k_{min}}^{k_{max}} k^2 P_\\delta(k) W^i_l(k)W^j_l(k)dk \\label{eq:4} \\end{equation} where $W^i_l$ is a window function which accounts for the clustering properties of a given biased tracer $i$ and the angular geometry for a given multipole $l$: \\begin{equation} W^i_l=\\int \\frac{dn}{dz}^iD(z)b_h^i(z)j_l(k r)dz\\;. \\label{eq:5} \\end{equation} Here $dn/dz$ is the angular redshift distribution of sources normalized to unity, $r$ is the comoving radial distance to redshift $z$ and $j_l$ are the spherical Bessel functions of order $l$. \\begin{figure*} \\hspace{-9pt} \\begin{minipage}[b]{8.6cm} \\includegraphics[width=1\\textwidth]{fig1_a.eps} \\end{minipage} \\begin{minipage}[b]{8.6cm} \\includegraphics[width=1\\textwidth]{fig1_b.eps} \\end{minipage} \\captionof{figure}{\\emph{Left:} 3D halo power spectrum at z$=$1 for different values of $f_{nl}$ with an effective halo bias (eq. \\ref{eq:6}) computed in the mass range from $10^{11}h^{-1}M_{\\odot}$ to $10^{12}h^{-1}M_{\\odot}$(lower curves) and from $10^{13}h^{-1}M_{\\odot}$ to $10^{14}h^{-1}M_{\\odot}$ (Upper curves). \\emph{Right:} Ratio between the $P^{3D}_h$ curves presented on the left panel for the same values of $f_{nl}$.} \\label{fig1} \\end{figure*} As shown in the previous section, the total scale dependent correction introduced by non-Gaussianity in halo clustering ($\\delta_h=b_h\\delta_m$) depends on the redshift and also on the mass of the halo. In the absence of non-Gaussianity, this mass dependence simply translates into an increased amplitude in the power spectrum or correlation function for more massive halos/galaxies when compared to lower mass objects. However, if non-Gaussianity is present, the mass dependence of the linear bias in eq. \\ref{eq:2} will introduce different scale dependence features in clustering 2-point statistics. More massive halos will then be more sensitive to non-Gaussianity than lower mass ones. Fig. 1 shows this effect for different values of $f_{nl}$ and halo mass for the 3D halo power spectrum. One can clearly see the differential effect of using mass bins, with the effect being significant only for $k$ less than 0.02 $h$Mpc$^{-1}$, and having a relative amplitude up to order 1 order of magnitude. Such features offers the possibility to obtain additional information on $f_{nl}$ by constructing observational 2-point statistics from objects corresponding to different halo mass ranges. In order to exploit this, we need to address the issue of differentiating halo masses from an observational point of view. ", "conclusions": "" }, "1402/1402.5824_arXiv.txt": { "abstract": "Giant flares on soft gamma-ray repeaters that are thought to take place on magnetars release enormous energy in a short time interval. Their power can be explained by catastrophic instabilities occurring in the magnetic field configuration and the subsequent magnetic reconnection. By analogy with the coronal mass ejection (CME) events on the Sun, we develop a theoretical model via an analytic approach for magnetar giant flares. In this model, the rotation and/or displacement of the crust causes the field to twist and deform, leading to flux rope formation in the magnetosphere and energy accumulation in the related configuration. When the energy and helicity stored in the configuration reach a threshold, the system loses its equilibrium, the flux rope is ejected outward in a catastrophic way, and magnetic reconnection helps the catastrophe develop to a plausible eruption. By taking SGR 1806 - 20 as an example, we calculate the free magnetic energy released in such an eruptive process and find that it is more than $10^{47}$ ergs, which is enough to power a giant flare. The released free magnetic energy is converted into radiative energy, kinetic energy and gravitational energy of the flux rope. We calculated the light curves of the eruptive processes for the giant flares of SGR 1806 - 20, SGR 0526-66 and SGR 1900+14, and compared them with the observational data. The calculated light curves are in good agreement with the observed light curves of giant flares. ", "introduction": "Soft gamma repeaters (SGRs) and anomalous X-ray pulsars (AXPs) are believed to be magnetars - a small class of spinning neutron stars with ultra-strong magnetic fields ($\\emph{B}$ $\\geq$ $10^{15}$ G ), which are thought to result from dynamo action during supernova collapse \\citep{DT92,Kel98,TM01,Lyu03,HL06}. The emission from a magnetar is powered by the dissipation of non-potential (current-carrying) magnetic fields in the magnetosphere \\citep{DT92,TD96,TLK02,Lyu06}. Both SGRs and AXPs show quiescent persistent X-ray and repeated soft gamma-ray emissions \\citep{Mel04}. Extremely rarely, an SGR produces a giant flare with enormous energy ($\\backsimeq 10^{44}-10^{47}$ erg) and long burst duration. These exceptionally powerful outbursts begin with a very short ($\\sim$ 0.2~s) spike of $\\gamma$-rays containing most of the flare energy and the spike is followed by a pulsating tail lasting a few hundreds of seconds \\citep{Hur05}. So far, three SGRs have been reported to produce giant flares \\citep{Maz79,Hur99,Hur05}. They include SGR 0526-66 on 5 March 1979 \\citep{Maz79}, SGR 1900+ 14 on 27 August 1998 \\citep{Hur99,Kou99,Vrb00}, and SGR 1806-20 on 27 December 2004 \\citep{Hur05,Pal05}. The giant flare from SGR 1806- 20 was much more luminous than the other two events \\citep{Hur05,Pal05}. Its initial $\\gamma$-ray spike released an energy of $\\sim 10^{46}$ erg within $\\sim 0.2$~s, and its rising and falling times were $\\tau_{rise}\\leq 1$ ms and $\\tau_{fall}\\approx 65$ ms, respectively. The main spike was followed by a tail with $\\sim 50$ pulsations of high-amplitude at the rotation period (7.56s) of SGR 1806 - 20 \\citep{Hur05,Pal05}. Although the energy of a magnetar outburst is widely believed to come from the star's magnetic field, details of the physical process in which the magnetic energy is stored and released remain unknown. So far, two models of giant flares of SGRs exist, which depend on the location where the magnetic energy is stored prior to the eruption: one assumes that the energy is stored in the crust of the neutron star (crust model) and the other one assumes that the storage occurs in the magnetosphere (magnetosphere model). In the crust model, a giant flare is caused by a sudden untwisting of the internal magnetic field \\citep{TD95,TD01,Lyu06}. Subsequently, a large and quick rotational displacement on the time-scale of a flare leads to the giant flare. Alternatively, in the magnetosphere model, the magnetic energy is slowly stored in the magnetosphere on time scales much longer than that of the giant flare itself, until the system reaches a critical state at which the equilibrium becomes unstable. Then further evolution in the system occurs and leads to flares, in analogy with solar flares and coronal mass ejections (CMEs) taking place in the solar atmosphere \\citep{Lyu06}. Observations of the giant flare from SGR 1806-20 on 27 December 2004 showed that it lasted a very short rise time, $\\sim$ 0.25~ms \\citep{Pal05}. This time interval of the eruption is short compared to the time-scale required for the crust model \\citep{TD95,Lyu03,Lyu06}. Therefore, at least for the SGR 1806 - 20, the time-scale of the crust model is too long to account for the triggering and early stage evolution of the event. The magnetosphere model based on an analogy with solar CMEs was proposed by \\citet{Lyu06}. In this model, the magnetic energy released during a giant flare is built up slowly in the magnetosphere, not in the crust of the neutron star. Although the energy storage process is gradual and long, the energy release takes place within a very short timescale in a dynamical fashion \\citep{Lyu06}. Transition from slow to fast evolution constitutes the catastrophe that is similar to what happens in solar flares and CMEs \\citep[see][] {FI91,Iea93}. Moreover, the profile of the light curve of solar-flare/ CME events resembles that of magnetar giant flares. Both of them have an impulsive phase and a tail emission. Similar morphology and characteristics between magnetar giant flares and solar-flare/ CME events indicate the operation of a common physical mechanism. Therefore, solar flares and CMEs give an important prototype context for magnetar giant flares. \\citet{Mas10} constructed a theoretical model for a magnetar giant flares based on the solar-flare/ CMEs model. They described magnetar giant flare using a magnetic reconnection model of a solar flare proposed by \\citet{SY99} while taking account of chromospheric evaporation. In their work, the preflare activity produces a baryon-rich prominence. Then the prominence erupts as a result of magnetic reconnection, and the eruption constitutes the origin of the observed radio-emitting ejecta associated with the giant flare from SGR 1806- 20 \\citep{Tay05,Cam05,Gae05,Mas10}. A giant flare should be induced as the final outcome of prominence eruption accompanied by large-scale field reconfigurations \\citep{Mas10}. Numerical simulations have also been performed in order to construct an MHD model for a magnetar giant flare. \\citet{Pfr12a,Pfr12b,Pfr13} presented the simulations of evolving strongly twisted magnetic fields in the magnetar magnetosphere. Their results showed that slow shearing of the magnetar crust leads to a series of magnetospheric expansion and reconnection events, corresponding to X-ray flares and bursts. They studied the relationship between the increasing twist and the spindown rate of the star, and concluded that the observed giant flares could be caused by the sudden opening of large amounts of overtwisted magnetic flux, resulting in an abrupt increase in spin period. Motivated by CME studies, \\citet{Yu11} constructed a general relativistic model of non-rotating magentars, which simulated how the magnetic field could possess enough energy to overcome the Aly-Sturrock energy constraint and open up. Furthermore, by taking into account the possible flux injections and crust motions, \\citet{Yu12} built a force-free magnetosphere model with a flux rope suspended in the magnetosphere and investigated the catastrophic behavior of the flux rope in a background with multi-polar magnetic field. In this model, a gradual process leads to a sudden release of the magnetosphere energy on a dynamical timescale \\citep{Yu12}. Therefore, the existing catastrophe model for solar eruptions could be a good template for constructing a theoretical model for magnetar giant flares. The initiation and development of magnetar giant flares have been extensively studied and several models have been suggested \\citep{TD95,TD01,Lyu06, Mas10,Yu11,Yu12}. But the origin and development of these giant flares remains unclear. The detailed physical process of the magnetic energy storage and release is still an open question. In this work, we consider relativistic effects and construct a magnetohydrodynamical (MHD) model for magnetar giant flares in the framework of the CME catastrophe model \\citep{LF00}, then duplicate the dynamical process of the giant flares produced by SGR 1806- 20. We describe our model in next section. Results of calculations and comparisons with observations are in Section 3. Finally, we discuss these results and summarize this work in Section 4. ", "conclusions": "We develop a theoretical model via an analytic approach for a magnetar giant flare in the framework of the solar CME catastrophe model \\citep{LF00}. We considered the physical process that causes the catastrophic loss of equilibrium of a twisted flux rope in the magnetar magnetosphere. The model was constructed according to manifestations of the eruption from the magnetar SGR 1806-20. As happens on the Sun, the free magnetic energy that drives the eruption on the magnetar is slowly stored in the magnetosphere of the magnetar long before the outburst until the system loses mechanical equilibrium, and then is quickly released in the consequent eruptive process, also known as a magnetar giant flare. In our model, the energy driving the eruption comes from the magnetosphere. The motion of footpoints causes the magnetic configuration to lose its equilibrium and release magnetic energy eventually, which is in principle the same as the results of numerical simulations \\citep{Pfr12a,Pfr12b,Pfr13,Yu11,Yu12}. In the present model, the eruption starts with a loss of equilibrium in the magnetic configuration, which could be triggered either by the change in the background magnetic field, or by the sudden break of a crust piece on the surface of the star where the disrupting magnetic configuration roots in. The evolution in this stage could be purely mechanical, namely no dissipation or magnetic reconnection needs to take place in the magnetosphere. But the dissipation is required in the consequent progress, otherwise the conversion of magnetic energy into radiative and kinetic energy would be stopped, and the evolution in the system ceases without plausible eruptive phenomenon happening \\citep[e.g., see detailed discussions by][]{ FL00,LF00,Lin02,PF02,Lel03,Fel06}. Our main results deduced from the present analysis are summarized as follow: 1. The system could store free energy of more than $10^{47}$~ergs prior to the eruption, which is enough to drive a giant flare like the one from magnetar SGR 1806-20 observed on 27 December 2004. 2. A combination of the following three factors determines that the disrupting magnetic configuration is highly stretched, and a long magnetically neutral current sheet forms separating two magnetic fields of opposite polarity. These factors are the mechanical property of the loss of equilibrium, the inertia of the magnetic field that tends to keep the original topological features in the configuration unchanged without diffusion in the system \\citep[see detailed discussions by][]{FI91,Iea93}, and the fact that the reconnection time-scale is long compared to that of the loss of equilibrium. 3. A long current sheet is usually unstable to perturbations due to various plasma instabilities, such as the tearing mode instability, and the consequent turbulence quickly dissipates the magnetic field such that the stored magnetic energy is rapidly converted into radiative and kinetic energy of the plasma inside the current sheet, as well as the kinetic energy and the gravitational potential energy of the mass in the ejected flux rope. Most of the released magnetic energy turns into radiative energy in the case of the giant flare from SRG 1806-20. 4. Conversion of energy by reconnection mainly occurs at the lower part of the current sheet, and the hottest part of the current sheet is then expected close to the star surface. A large amount of energetic particles and a heat conduction front are also created by reconnection and propagate downward along magnetic field lines. They eventually reach the star surface and may heat the relevant region significantly, which then leads to the formation of a ``fireball\" near the neutron star's surface. Therefore, a ``fireball\" is a straightforward and natural consequence of our model. 5. We calculated the light curve on the basis of our model and compare it with the observational one obtained by the RHESSI $\\gamma$- ray detectors. We note that the calculated light curve consists of a hard spike lasting a half ms and a tail emission last a few $10^{2}$~s, which are consistent with the observations. 6. Our calculations indicate that the magnetic flux preexisting in the flux rope before the eruption was about $7.9\\times 10^{26}$~Mx, and that the magnetic flux brought from the environment around the disrupting magnetic field into the ejecta bubble was around $6.8\\times 10^{26}$~Mx. Therefore, total magnetic flux of more than $1.5\\times 10^{27}$~Mx was sent into interstellar space from the central star by the super eruption from SGR 1806-20. 7. We duplicated our calculations for the giant flares from SGR 0526-66 and SGR 1900+14, respectively, and also found good agreement of our model with observations. Furthermore, the the total magnetic fluxes ejected into interstellar space by the eruption were $6.9\\times 10^{26}$~Mx for SGR 1900+14 and $6.7\\times 10^{26}$~Mx for SGR 0526-66, respectively." }, "1402/1402.1770_arXiv.txt": { "abstract": "\\edit{We carry out a systematic study of the X-ray emission from the active nuclei of the $0.02{{(\\rm B/T)}_{\\rm thresh}}$, where the parameter ${(\\rm B/T)}_{\\rm thresh}$ may adopt values in the range $\\sim 0.7-0.85$. Thus, the formation of elliptical galaxies is directly connected with bulge formation, which in \\sag~takes place through global disk instabilities and two different kinds of galactic mergers. When a galaxy merger occurs, the stellar mass ratio between the satellite galaxy and the central galaxy, $M_{\\rm sat}/M_{\\rm cen}$, is evaluated. If $M_{\\rm sat}/M_{\\rm cen}>0.3$, then the merger is considered as a major one. In this case, all the gas in the remnant galaxy is consumed in a starburst contributing to the bulge formation, and the stellar disc is completely relaxed and transferred to the bulge. The triggering of a starburst in a minor merger ($M_{\\rm sat}/M_{\\rm cen}\\leq 0.3$) will depend on the fraction of cold gas present in the disc of the central galaxy, as implemented by \\citet{lcp08} following the work of \\citet{malbon2007}. If the ratio between the cold gas and the disc mass of the central galaxy, $M_\\text{cold,cen} / M_\\text{disc,cen}$, is larger than a fixed parameter $f_\\text{burst}=0.6$, then the perturbation introduced by the merging satellite drives all the cold gas from both galaxies into the bulge component, where it is consumed in a starburst; the disc mass is given by the sum of the cold gas and the stars formed through quiescent SF. However, if the satellite is much less massive than the central galaxy ($M_{\\rm sat}/M_{\\rm cen}\\leq$~0.05), no burst occurs. In minor mergers, only the stars of the merging satellite are transferred to the bulge component of the central galaxy. The other channel that contributes to the bulge formation is the global disk instability. Some configurations of galactic discs do not remain stable with time. When a galactic disc is sufficiently massive that its self-gravity is dominant, it becomes unstable. This condition is expressed in the model through the Efstathiou-Lake-Negroponte \\citep{Efstathiou+1982} criterium, that is, the stability to bar formation is lost when \\begin{equation} \\epsilon_{\\rm d} \\equiv \\frac{V_{\\rm disc} }{ (G M_{\\rm disc} / R_{\\rm disc})^{1/2}} \\le \\epsilon_{\\rm thresh}, \\label{eq:diskinstab} \\end{equation} where $M_{\\rm disc}$ is the mass of the disc (cold gas plus stars), $R_{\\rm disc}$ is the disc scale radius, and $V_{\\rm disc}$ is the circular velocity of the disc. For the latter, we use the velocity where the rotation curve flattens, which we approximate by the velocity calculated at $\\sim 3\\,R_{\\rm disc}$ (see \\citet{Tecce2010} for details concerning disc features). The free parameter $\\epsilon_{\\rm thresh}$ has values close to unity; in our model, we adopt $\\epsilon_{\\rm thresh}=1$. We consider an additional free parameter that takes into account the influence of a perturbing galaxy that effectively triggers the disc instability; this is modelled by computing the mean separation between galaxies in a \\fof~group. When the mean separation is smaller than $f_\\text{pert}\\,R_{\\rm disc}$, being $f_\\text{pert}$ a free parameter, we consider that a neighbouring galaxy perturbs the unstable disc. As a consequence of this, all the stars and cold gas in that disc are transferred to the bulge component, and all the gas present is consumed in a starburst. \\subsubsection{Chemical enrichment model} \\label{sec:chem_model} In our chemical model, we follow the production of ten chemical elements (H, $^4$He, $^{12}$C, $^{14}$N, $^{16}$O, $^{20}$Ne, $^{24}$Mg, $^{28}$Si, $^{32}$S, $^{40}$Ca) generated by stars in different mass ranges, from low- and intermediate-mass stars to quasi massive and massive stars. This version of \\sag~is characterized by a new set of stellar yields. We use the best combination of stellar yields reported by \\citet{Romano2010}, selected to be in accordance with the large number of constraints for the Milky Way. For low and intermediate-mass stars (LIMS), in the mass interval $1-8 \\Msun$, we use the yields of \\citet{Karakas2010}. For the mass loss of pre-supernova stars (He and CNO elements), we use the yields computed by the Geneva group \\citep{Hirschi2005}, and for the yields of the explosive nucleosynthesis due to SNe CC, we use the results of \\citet{Kobayashi2006}. In all cases, we have adopted the total ejected mass of solar metallicity models, for which the solar abundances of \\citet{AndersGrevesse89} are assumed, with a solar composition $Z_{\\odot} = 0.02$. Ejecta from SNe Ia are also included, which are characterized by high iron production ($\\sim 0.6\\, {\\rm M}_{\\odot}$); we consider the nucleosynthesis prescriptions from the updated model W7 by \\citet{Iwamoto99}. The SNe Ia rates are estimated using the single degenerate model, in which a SN Ia occurs by carbon deflagration in C-O white dwarfs in binary systems whose components have masses between $0.8$ and $8\\,\\Msun$ \\citep{Greggio83}; we implement the formalism presented by \\citet{Lia2002} (see their Section 4.2.2). The fraction of these binary systems, $A_{\\rm bin}$, is one of the free parameters of \\sag. We take into account the return time-scale of mass losses and ejecta from all sources considered. For this purpose, we use the stellar lifetimes given by \\citet{Padovani93} who use results from \\citet{Matteucci86} for masses over $6.6\\,\\Msun$, and from \\citet{Renzini86} for masses below this limit. This aspect becomes especially relevant for this work because of the different delay times that characterize different types of SNe, which affect the abundance of $\\alpha$-elements relative to iron. For a detailed description of the implementation in \\sag~of the chemical enrichment of the different baryonic components, we refer the reader to \\citet{Cora06}. \\subsubsection{Extended starbursts} \\label{sec:ExtendedBurstsInSAG} In previous versions of the model, the formation of stars in the bulge occurs through starbursts that consume the available cold gas in a single step \\footnote{Differential equations in \\sag~are integrated in time-steps of equal size used to subdivide the intervals between simulation outputs.}. We now consider that starbursts are characterized by a given time-scale during which the cold gas driven to the galactic center is consumed gradually in several steps. Thus, this cold gas has the possibility of being progressively contaminated by the stars formed in the bulge. This way, the $[\\alpha/{\\rm Fe}]$ abundance ratio of successive generations of stars in the bulge is not only determined by the abundances of the disc cold gas at the moment in which the starburst is triggered, but is also modified by the relative contribution of different types of SNe, which depends on the relation between the time-scale of the duration of the starburst and the lifetime of SNe progenitors. The cold gas that will be eventually converted in bulge stars is referred to as bulge cold gas, in order to differenciate it from the disc cold gas. The amount of bulge cold gas available allows us to make a first estimation of the mass of stars that will be formed in the bulge. Then, we compute the time-scale in which they would form, which is chosen to be the dynamical time considered in quiescent SF as a first approximation, that is, the one defined as the ratio between the scale radius of the exponential profile that characterizes the galactic disc and its circular velocity. This choice seems reasonable since a burst normally takes place via a bar embedded in the galaxy disc. Thus, from the ratio of the bulge cold gas mass and the time-scale estimated for the duration of the starburst, we know the mass of stars that will be formed in each time step. For simplicity, SF throughout the extended starburst process occurs at a constant rate. Other works in the literature that have taken into account the starbursts time-scales consider that the SFR in a burst decays exponentially with time after the burst is triggered \\citep{Granato2000, Lacey2008}. We justify our assumption considering that the bar formed during the instability, triggered either by a galaxy merger or a disk instability, favours the gradual feeding of the bulge cold gas reservoir by the disc cold gas that is being driven to the galactic center \\citep[e.g.][]{KormendyKennicutt2004}, keeping constant the conditions that give rise to SF. Note that the presence of a bar during the process of bulge formation is only implicit in the assumption of constant SFR during starbursts. Modelling bar formation is not an easy task \\citep[see][]{KormendyKennicutt2004, Athanassoula2013}, and is beyond the scope of this paper. \\subsection{Variable IMF} \\label{sec:varIMF} The stellar IMF has a great influence on the chemical enrichment of galaxies since it defines the amount of stars formed in each stellar mass interval for each star formation event and, consequently, impacts on the amount of metals returned to the ISM, where new stars will eventually form. The IMF also determines the number of SNe CC involved in the estimation of the reheated mass that is transferred from the cold to the hot phase during SN feedback. Today, there is some consensus that in a simple stellar population the slope of the IMF above $1\\,\\Msun$ is not strictly different from that of Salpeter IMF ($\\alpha \\equiv 1+x = 2.35$); stars with masses over $1\\,\\Msun$ are the major contributors to chemical enrichment \\citep{Portinari98b}. Several attempts have been made in measuring this important distribution function in the solar neighbourhood \\citep{Salpeter55, Chabrier2003} or directly in star clusters \\citep{Kroupa2001, Kroupa2002}. The latter is represented by a multi-component power-law IMF given by \\begin{equation} \\xi(m) = k \\left\\{\\begin{array}{ll} k'\\,\\left(\\,\\frac{m}{m_{\\rm H}} \\right)^{-\\alpha_{0}}&\\hspace{-0.25cm},{m_{\\rm low}} \\le m/{M}_\\odot < {m_{\\rm H}}\\\\ \\left(\\frac{m}{m_{\\rm H}} \\right)^{-\\alpha_{1}}&\\hspace{-0.25cm},m_{\\rm H} \\le m/{M}_\\odot < m_0,\\\\ \\left(\\frac{m_{0}}{m_{\\rm H}} \\right)^{-\\alpha_{1}} \\left(\\frac{m}{m_{0}} \\right)^{-\\alpha_{2}}&\\hspace{-0.25cm},m_0 \\le m/{M}_\\odot < m_1,\\\\ \\left(\\frac{m_{0}}{m_{\\rm H}} \\right)^{-\\alpha_{1}} \\left(\\frac{m_{1}}{m_{0}} \\right)^{-\\alpha_{2}} \\left(\\frac{m}{m_{1}} \\right)^{-\\alpha_{3}}&\\hspace{-0.25cm},m_1 \\le m/{M}_\\odot \\le m_\\mathrm{max},\\\\ \\end{array} \\right. \\label{eq:4pow} \\end{equation} \\noindent with $m_{\\rm low}=0.01$, $m_{\\rm H}=0.08$, $m_{0}=0.5$ and $m_{1}=1.0$, and exponents $\\alpha_0=0.3$, $\\alpha_1=1.3$, $\\alpha_2=\\alpha_3=2.35$ . The normalization constant $k$ contains the desired scaling of the IMF (for example, to the total mass of the system). Note that brown dwarfs are a separate population and contribute about 1.5 per cent by mass only. Therefore, they have a different normalisation factor, $k'$ \\citep[$k' \\sim 1/3$,][]{ThiesKroupa2007, ThiesKroupa2008}. In the case of stellar systems with composite stellar populations, other aspects must be considered. A galaxy-wide IMF may be represented by an integrated galactic initial mass function (IGIMF) proposed by \\citet{Kroupa2003}, whose theory was developed in the last decade \\citep[see][ for a review]{Kroupa2012}. This formalism states that star formation takes place exclusively in star clusters \\citep{LadaLada2003}. This way, the stellar populations of each galaxy are composed by stars in surviving and dissolved star clusters. Within each star cluster, stars form with the stellar IMF given by eq. \\ref{eq:4pow}. The embedded star clusters in the galaxy also follow an initial distribution of masses of the form $\\xi_{\\rm ecl}(M_{\\rm ecl})\\, dM_{\\rm ecl}\\propto M_{\\rm ecl}^{-\\beta}\\,dM_{\\rm ecl}$, where $M_{\\rm ecl}$ is the mass of the embedded star cluster. Moreover, the most massive star in a cluster, $m_{\\rm max}$, is related in a non-trivial way with the cluster mass \\citep{Weidner2013b}; the mass of the most massive star is higher in more massive clusters. Furthermore, observations indicate that higher SFRs lead to the formation of brighter clusters \\citep{Larsen2002}. \\citet{Weidner2004} find that this empirical correlation can be transformed into a relation between the SFR of the galaxy and the maximum embedded cluster mass of the form \\begin{equation} M_{\\rm ecl}^{\\rm max}(SFR)=8.5 \\times 10^4 \\, SFR^{0.75}\\,\\Msun. \\label{eq:MeclSFR} \\end{equation} The IGIMF is the sum of all the new born stars in all of the star clusters considering the above ingredients , that is, \\begin{equation} \\xi_{\\rm IGIMF}(m,t)=\\int_{M_{\\rm ecl}^{\\rm min}}^{M_{\\rm ecl}^{\\rm max(SFR(t))}}\\xi(m\\leq m_{\\rm max(M_{\\rm ecl})})\\,\\xi_{\\rm ecl}(M_{\\rm ecl})\\,dM_{\\rm ecl}, \\label{eq:IGIMF} \\end{equation} where $\\xi(m\\leq m_{\\rm max(M_{\\rm ecl})})$ is the stellar IMF given by eq. \\ref{eq:4pow}, and $M_{\\rm ecl}^{\\rm min}$ defines the minimum mass of cluster that can be formed in a galaxy, which is a free parameter of the IGIMF model. Now, for this formulation of the IGIMF, it is assumed that star formation in all the star clusters at all epochs is produced with a canonical IMF. However, violent star formation conditions could drive crowding in massive star clusters \\citep{Elmegreen2004, Shadmehri2004}, and the formation of massive stars could be favoured under such conditions, given that the low-mass limit of star formation would be higher. Following preliminary work of \\citet{Marks2012}, \\citet[][WKP11 hereafter]{Weidner2011} consider that, for embedded clusters $M_{\\rm ecl}<2\\times 10^5\\,\\Msun$, the slope of the canonical IMF (eq. \\ref{eq:4pow}) for stars more massive than $1.3\\,\\Msun$ is $\\alpha_3 = 2.35$. For larger cluster masses, the dependence of the stellar IMF slope for stars in this mass range is given by \\begin{equation} \\alpha_{3}(M_{\\rm ecl})=-1.67 \\times {\\rm log}_{\\rm 10}\\left(\\frac{M_{\\rm ecl}}{10^6\\,\\Msun}\\right)+1.05, \\label{eq:alpha3Ecl} \\end{equation} The effect of this assumption in the IGIMF is analysed in WKP11. The high-mass end of the IGIMF is, in general, steeper than the stellar IMF since the formation of low mass stars is favoured. However, if one includes the effect of crowding in massive clusters, as we have seen, the slope of the canonical IMF for massive stars can become more top heavy than the Salpeter IMF for these massive clusters. When a starburst event with high SFR occurs in a galaxy, the formation of massive clusters is favoured and the formation of massive stars is highly enhanced, so the IGIMF can become also more top heavy than the Salpeter IMF for high mass stars . This formulation leads to the top heavy IGIMF (TH-IGIMF; WKP11). The slope of this TH-IGIMF above $1.3\\,\\Msun$ can then be computed by a least-squares fit to the calculated IGIMF. Hence, the TH-IGIMF can be translated into a power law of a form similar to the canonical stellar IMF used to derive the IGIMF , but with an exponent $\\alpha_{\\rm TH}$ for high mass stars, that is, \\begin{equation} \\xi_{\\rm TH}(m) = k \\left\\{\\begin{array}{ll} \\left(\\frac{m}{m_{\\rm H}} \\right)^{-\\alpha_{1}}&\\hspace{-0.25cm},m_{\\rm H} \\le \\frac{m}{{M}_\\odot} < m_0,\\\\ \\left(\\frac{m_{0}}{m_{\\rm H}} \\right)^{-\\alpha_{1}} \\left(\\frac{m}{m_{0}} \\right)^{-\\alpha_{2}}&\\hspace{-0.25cm},m_0 \\le \\frac{m}{{M}_\\odot} < m_1,\\\\ \\left(\\frac{m_{0}}{m_{\\rm H}} \\right)^{-\\alpha_{1}} \\left(\\frac{m_{1}}{m_{0}} \\right)^{-\\alpha_{2}} \\left(\\frac{m}{m_{1}} \\right)^{-\\alpha_{\\rm TH}}&\\hspace{-0.25cm},m_1 \\le \\frac{m}{{M}_\\odot} \\le m_\\mathrm{max}.\\\\ \\end{array} \\right. \\label{eq:THIGIMF-FIT} \\end{equation} where $m_{\\rm H} = 0.1$, $m_{\\rm 0} = 0.5$ and $m_{\\rm 1} = 1.3$. After having introduced the basic aspects of the TH-IGIMF theory, we describe the way in which it is implemented in \\sag. Although the reference to ``starbursts condition'' in the work of WKP11 implies a high level of star formation rate in a starburst, we use the resulting dependence of the $\\alpha_{\\rm TH}$ with SFR for any event of star formation in our model, that is, without making any distinction between quiescent star formation mode and starbursts. For each star formation event, we assign to it the TH-IGIMF given by eq. \\ref{eq:THIGIMF-FIT}, according to the corresponding SFR. The model considers small variations of log(SFR) so that the slope of the TH-IGIMF is binned in steps of $0.05$. We adopt $m_{\\rm H}=0.1 M_{\\odot}$ for the lower limit in the TH-IGIMF according to our chemical implementation. We consider embedded cluster mass functions characterized by different values of the exponent $\\beta$, and of the minimum embedded cluster masses $M_{\\rm ecl}^{\\rm min}$. We explore cases with (i) $\\beta=2$ and $M_{\\rm ecl}^{\\rm min}=5 \\, \\Msun$, (ii) $\\beta=2.1$ and $M_{\\rm ecl}^{\\rm min}=5 \\, \\Msun$, and (iii) $\\beta=2$ and $M_{\\rm ecl}^{\\rm min}=100 \\, \\Msun$. It is worth noting that some stellar associations with masses as low as $5\\,\\Msun$ are found in the galaxy \\citep[e.g. Taurus-Auriga star-forming regions,][]{Kirk2011} , thus favouring the lowest value of $M_{\\rm ecl}^{\\rm min}$ considered here. Therefore, at least in quiescent star forming conditions, the $M_{\\rm ecl}^{\\rm min}$ is supported observationally. However, this constraint could not be valid in more extreme star-forming conditions. \\begin{figure} \\includegraphics[width=0.49\\textwidth]{fig1.eps} \\vspace{1pt} \\caption{Dependence of the slope of the TH-IGIMF ($\\alpha_{\\rm TH}$) on the SFR for different minimum embedded cluster masses and slopes of the embedded star cluster mass function: (i) $\\beta=2$, $M_{\\rm ecl}^{\\rm min}=5 \\Msun$ (solid line), (ii) $\\beta=2.1$, $M_{\\rm ecl}^{\\rm min}=5 \\Msun$ (dashed-dotted line), and (iii) $\\beta=2$, $M_{\\rm ecl}^{\\rm min}=100 \\Msun$ (dashed line). } \\label{fig:alphaTHIGIMF} \\end{figure} Fig. \\ref{fig:alphaTHIGIMF} shows the slope $\\alpha_{\\rm TH}$ as a function of the SFR for different values of $M_{\\rm ecl}^{\\rm min}$ and $\\beta$. It is clear from this plot that the slope $\\alpha_{\\rm TH}$ of the TH-IGIMF is larger than the slope of the Salpeter IMF (characterized by a fixed value of $\\alpha$ for all SFR) in lower star formation regimes. The values of SFR at which the TH-IGIMF becomes more top heavy than Salpeter depend on the value of $\\beta$ and $M_{\\rm ecl}^{\\rm min}$, being of the order of $1,\\,10$ and $1000\\ \\Msun\\,{\\rm yr^{-1}}$ for $\\beta=2$ and $M_{\\rm ecl}^{\\rm min}=100 \\,\\Msun$, $\\beta=2$ and $M_{\\rm ecl}^{\\rm min}=5 \\,\\Msun$, and $\\beta=2.1$ and $M_{\\rm ecl}^{\\rm min}=5 \\,\\Msun$, respectively. During periods of high star formation activity, the formation of massive stars is favoured. For higher SFRs, the number of high mass stars in a TH-IGIMF becomes larger than in the Salpeter IMF, with the corresponding reduction of the number of low mass stars required by the IMF normalization. This aspect is crucial to understand the build-up of the $[\\alpha/{\\rm Fe}]$-stellar mass relation, as we will see in Subsection \\ref{sec:AlphaFeTHIGIMF}. ", "conclusions": "" }, "1402/1402.4964_arXiv.txt": { "abstract": "We present a setup that provides a partial UV-completion of the ghost inflation model up to a scale which can be almost as high as the Planck mass. This is achieved by coupling the inflaton to the Lorentz-violating sector described by the Einstein-aether theory or its khronometric version. Compared to previous works on ghost inflation our setup allows to go beyond the study of small perturbations and include the background dynamics in a unified framework. In the specific regime when the expansion of the Universe is dominated by the kinetic energy of the inflaton we find that the model predicts rather high tensor-to-scalar ratio $r \\sim 0.02\\div 0.2$ and non-Gaussianity of equilateral type with $f_{NL}$ in the range from $-50$ to $-5$. ", "introduction": "\\label{sec:intro} All structures in the observed Universe, such as galaxy clusters, galaxies and stars are believed to arise from tiny quantum fluctuations, amplified during a primordial stage in the expansion of the Universe. The most successful model of this stage is the inflationary theory which fits very well all current cosmological data. Generally, inflation is supposed to occur at very high energies (probably a few orders below the Planck mass), where the effects of new unknown physics may show up. Thus the inflationary period provides a unique opportunity to probe these scales through cosmological observations. An intriguing possibility is that the known space-time symmetries become invalid at the inflationary energies. One of such symmetries is Lorentz invariance (LI) which is at the basis of the highly successful Standard Model of particle physics and General Relativity (GR). However, beautiful as it is, GR suffers from the problem of non-renormalizability precluding it from being a consistent quantum theory at energies above the Planck scale. It has been suggested by Ho\\v rava \\cite{Horava:2009uw} that the situation can be improved by allowing LI to be violated at high energy. A consistent implementation of this proposal \\cite{Blas:2009qj} involves additional light degrees of freedom in the gravitational sector that introduce departures from LI even at energies well below Planckian. The description at such energies is provided by the so-called ``khronometric'' model \\cite{Blas:2010hb}, which can be considered as a variant of the phenomenological ``Einstein-aether'' theory \\cite{Jacobson:2000xp,Jacobson:2008aj} for the study of Lorentz-violating (LV) effects in gravity (see \\cite{Jacobson:2010mx,Jacobson:2013xta} for the precise relationship between the two models). The existing astrophysical and cosmological data significantly constrain the parameters of the model \\cite{Elliott:2005va, Blas:2010hb,Blas:2011zd,Audren:2013dwa,Shao:2013wga,Yagi:2013qpa, Yagi:2013ava}, but still leave open a theoretically motivated portion of the parameter space. It is worth mentioning that any application of these ideas to realistic model building must include a mechanism that would prevent significant percolation of LI breaking from gravity into the Standard Model sector where LI has been tested with an outstanding precision; several options have been discussed in \\cite{GrootNibbelink:2004za,Pospelov:2010mp,Pujolas:2011sk,Bednik:2013nxa}. LV in gravity, though tightly constrained at low energies accessible to current experiments, may be significantly stronger during inflation. An interesting alternative to the standard slow-roll inflation involving LV in gravity is provided by the ghost inflation model \\cite{ArkaniHamed:2003uz} and its ``tilted'' extension \\cite{Senatore:2004rj}. The important feature of these models is the modified dispersion relation for the excitations of the inflaton\\footnote{By inflaton we loosely understand the field generating the primordial perturbations.}. The dispersion relation is quadratic, $\\omega^2\\propto k^4$, in the case of the original ghost inflation and linear, $\\omega^2=\\delta^2\\cdot k^2$, with the small ``sound speed'' $\\delta\\ll 1$ in the tilted version. This, combined with the specific form of the interactions, leads to interesting predictions for the amplitude and shape of non-Gaussianity, which have been constrained by the Planck results \\cite{Ade:2013ydc}. The ghost inflation models are formulated as effective field theories (EFT) for cosmological perturbations valid below a certain cutoff. While this approach has the advantage of being very general, it is unable to fully capture the evolution of the inflaton background; in particular, it is problematic to incorporate in it the end of inflation and reheating. Besides, if one assumes that the ghost condensate present at inflation persists till today, stability requires the cutoff to be rather low \\cite{ArkaniHamed:2003uy,ArkaniHamed:2005gu}. This makes a UV-completion of ghost inflation desirable. In this paper we show that ghost inflation can be embedded in the framework of the khronometric or Einstein-aether gravity. The latter being also an EFT, it has a cutoff as well. But this can be as high as just a few orders of magnitude below the Planck mass. In other words, khronometric / Einstein-aether can provide a partial UV-completion (``UV-extension\") for the ghost condensate model almost up to the Planck scale. This allows to describe the evolution of the background and perturbations within a self-contained theory. Interestingly, the UV-extension happens without restoration of the broken Lorentz symmetry, similar to the case recently discussed in \\cite{Endlich:2013vfa}. Further, in the khronometric version of the setup there is a potential UV-completion all the way above the Planck scale in the form of Ho\\v rava gravity. We study the observational signatures of the extended model. For two reasons the analysis somewhat differs from the discussion of the generic ghost inflation present in the literature. First, the ability to keep the background evolution under control allows to consider the situation when the VEV of the inflaton time derivative $\\langle \\dot\\Theta\\rangle$ --- the ``ghost condensate'' --- varies with time. This produces a contribution into the tilt of the power spectrum in addition to that coming from a potential for $\\Theta$, which in some cases can actually be dominant. Second, the ghost inflation is characterized essentially by three energy scales: the scale $\\rho_{inf}^{1/4}$ of the inflationary energy density; the scale $M$ of the ghost condensate $M^2=\\langle\\dot\\Theta\\rangle$ which determines the strength of the inflaton self-interaction producing the leading non-Gaussianity; and the scale $M'$ suppressing terms with higher spatial derivatives in the quadratic effective action for inflaton perturbations. In the previous treatments of the ghost inflation the latter two scales have been commonly assumed to be of the same order with the first scale being much higher, $\\rho_{inf}^{1/4}\\gg M\\sim M'$. We will see that in the extended model where all the above scales are derived quantities, $M$ and $M'$ have different parameter dependence, and the requirement that the EFT description is valid for the inflationary background imposes the hierarchy\\footnote{At first sight, this hierarchy might seem surprising from the viewpoint of the EFT for the inflaton {\\it perturbations}. However, it is straightforward to verify that it is stable under radiative corrections and thus perfectly natural.} $M\\gg M'$. On the other hand, $\\rho_{inf}^{1/4}$ can naturally be of the same order as $M$. As a result of this new hierarchy, the amplitude of non-Gaussianity is somewhat suppressed compared to the original prediction of the ghost inflation; still, it remains large enough to be observationally interesting. The paper is organized as follows. In Sec.~\\ref{sec:fast} we describe the model and identify the inflationary regime that reproduces ghost inflation. In Sec.~\\ref{sec:lin} we study linear cosmological perturbations emphasizing the similarities and differences with the generic ghost inflation treatment. Sec.~\\ref{sec:bispectra} contains calculation of the bispectrum. In Sec.~\\ref{sec:kinetic} we apply our results to the special case when the background dynamics is dominated by the kinetic energy of the inflaton and derive observational constraints on the model parameters in this case. Sec.~\\ref{sec:discussion} is devoted to conclusions. Some details of the analysis are postponed to the Appendices. ", "conclusions": "\\label{sec:discussion} In this paper we presented a setup that provides a partial UV-completion (``UV-extension\") of the well-known ghost inflation model up to a scale which can be almost as high as the Planck mass. This is achieved by coupling the inflaton to the Lorentz-violating sector described by the Einstein-aether theory or its khronometric version. The cutoff of our model coincides with the cutoff of the Einstein-aether sector and is unrelated to the scale of the ``ghost condensate'' --- the dynamically developed expectation value for the time derivative of the inflaton. In the khronometric version the construction can be potentially completed even further in the framework of the Ho\\v rava gravity. Curiously, the UV-extension occurs without restoration of the broken Lorentz symmetry, cf. \\cite{Endlich:2013vfa}. We have studied the inflationary evolution of the Universe and the generation of primordial perturbations in the model. The latter are described by an effective theory containing a single degree of freedom and governed by the same effective action as in the ghost inflation. The novelty of our model compared to the previous works on ghost inflation is that it allows to go beyond the study of small perturbations and incorporates in a unified framework the background dynamics. We have found that the ghost condensate gives positive contribution into the energy budget of the Universe. This makes inflation possible even in the absence of any potential for the inflaton --- the regime that we called kinetically driven inflation. In principle, it is straightforward to incorporate in our model the graceful exit from inflation and reheating which proceeds through vanishing\\footnote{At least up to the tiny present value of the dark energy density. We do not consider a fine-tuned situation when after inflation the ghost condensate stays large and its positive energy is cancelled by a negative cosmological constant.} of the inflaton potential and the ghost condensate. This would be impossible in the original formulation of the ghost inflation as the low-energy EFT because its cutoff goes to zero in this limit. On the phenomenological side, we have calculated the characteristics of the power spectrum and bispectrum predicted by the model. Specifically, in the kinetically driven regime the model predicts a rather high tensor-to-scalar ratio and is already constrained by the bounds from the Planck mission. The non-Gaussianity is predicted to be close to the equilateral type with the amplitude ranging from $f_{NL} \\sim -50$ for $r\\sim 0.02$ to $f_{NL} \\sim -5$ for $r\\sim 0.2$. This is still well within the limits set by Planck. Optimistically, one can expect it to be probed by planned surveys \\cite{Baumann:2008aq,Amendola:2012ys,Andre:2013nfa}. There are two directions in which our work can be extended. First, additional insight will be gained from the study of higher statistics. In the general EFT formulation of ghost inflation the trispectrum depends on an additional free coupling constant standing in front of the quartic interaction of inflaton perturbations $(\\d_i\\pi)^2 (\\d_j\\pi)^2$ \\cite{Izumi:2010wm}. On the other hand, in our model this coupling is unambiguously determined by the specific form of the UV-extended theory. Repeating the analysis presented in the beginning of Sec.~\\ref{sec:bispectra} it is straightforward to find the leading quartic interaction \\[ S_{[\\pi]}^{(4)}=\\int d^4x\\;\\frac{1}{8\\m^4c_\\T^6 a}\\,(\\d_i\\pi)^2 (\\d_j\\pi)^2\\;. \\] Clearly, the quartic coupling is related the cubic one, see Eq.~(\\ref{Hint}), and thus the shape of the trispectrum will be uniquely predicted. Second, it is interesting to investigate if the (partial) UV-completion along the lines presented in this paper can be found for other Lorentz-violating models of modified gravity. From the theory viewpoint, the case of Lorentz-violating massive gravity \\cite{Dubovsky:2004sg} deserves particular attention as its completion above the scale of strong coupling associated to the graviton mass could be considered as the analog of the Higgs mechanism in massive Yang--Mills theory." }, "1402/1402.5545_arXiv.txt": { "abstract": "We present results of optical broad-band and narrow-band H$\\alpha$ observations of a sample of forty nearby early-type galaxies. The majority of sample galaxies are known to have dust in various forms viz. dust lanes, nuclear dust and patchy/filamentary dust. A detailed study of dust was performed for 12 galaxies with prominent dust features. The extinction curves for these galaxies run parallel to the Galactic extinction curve, implying that the properties of dust in these galaxies are similar to those of the Milky-Way. The ratio of total to selective extinction ($R_{V}$) varies between 2.1 to 3.8, with an average of 2.9$\\pm$0.2, fairly close to its canonical value of 3.1 for our Galaxy. The average relative grain size $\\frac{}{a_{Gal}}$ of dust particles in these galaxies turns out to be 1.01$\\pm$0.2, while dust mass estimated using optical extinction lies in the range $\\sim$ $10^{2}$ to $10^{4}$ M$_{\\odot}$. The H$\\alpha$ emission was detected in 23 out of 29 galaxies imaged through narrow-band filters with the H$\\alpha$ luminosities in the range 10$^{38}$ - 10$^{41}$ erg sec$^{-1}$. The mass of the ionized gas is in the range $\\sim$ 10$^{3}$ - 10$^{5}$ M$_{\\odot}$. The morphology and extent of ionized gas is found similar to those of dust, indicating possible coexistence of dust and ionized gas in these galaxies. The absence of any apparent correlation between blue luminosity and normalized IRAS dust mass is suggestive of merger related origin of dust and gas in these galaxies. ", "introduction": "Interstellar matter (ISM), being an important factor governing the formation and subsequent evolution of galaxies, has drawn considerable attention since their detection and is still being continued to gain better understanding in various galactic and extragalactic environments. The availability of space observatories/satellites as well as ground-based telescopes with state of the art detectors has made exploration of different phases of ISM possible. From the first detection of dust lanes in early-type galaxies (\\citealt{Ber78}), the possibility of using its orientation to extract additional information regarding intrinsic shape of galaxies was explored (\\citealt{Gun79, vanA82, Habe85, Habe88}). Importance of dust in understanding three dimensional structure of these galaxies led to search for dust in several of them (\\citealt{Haw81, Ebn85, Ver88, Sad85}). Subsequently, deep optical imaging surveys of large sample of early-type galaxies detected dust in a significant fraction of them (\\citealt{Goud94b, Fer99}). Dust has also been detected in the innermost regions of early-type galaxies in deep {\\it Hubble Space Telescope} (HST) images, which remained unresolved with the ground-based observations (\\citealt{Jaf94, van95, deKoff00, Tran01}). Further, the detection of significant far-infrared (FIR) emission from early-type galaxies, using {\\it Infra Red Astronomical Satellite} (IRAS), {\\it Infrared Space Observatory} (ISO) and {\\it Spitzer Space Telescope} has added a new dimension in the study of dust and its possible role in underlying dynamics of galaxies (\\citealt{Temi04, Xil04}). Optical broad-band imaging of galaxies with prominent dust features allows one to investigate physical properties of dust such as particle size, extinction, reddening, total dust content of galaxies and the processes that govern their evolution in different environments. Investigation of dust extinction in different bands {\\it i.e.} extinction curve has been traditionally a basic tool for studying dust properties. A refined form of this technique has been applied to study the properties of dust in a number of dusty early-type galaxies (\\citealt{Goud94a, Sahu98, Pat07, Fin08, Fin10d}). These studies showed that the extinction curves of dust in early-type galaxies run almost parallel to that of the Milky-Way, with average relative grain size not very much different from that in our Galaxy. Optical spectroscopic observations of early-type galaxies show that $\\sim$55 - 60\\% of them have faint emission line, indicating the presence of ionized gas, with mass $\\sim$ 10$^{3}$ -10$^{4}$ M$\\odot$ (Caldwell 1984, Phillips et al. 1986). Further, the H$\\alpha$ imaging survey of large sample of early-type galaxies confirmed the presence of ionized gas with various morphologies; such as flattened disc, ring or filamentary structures (\\citealt{Kim89, Shi91, Trin91, Bus93, Goud94a, Sing95, Macch96, Mart04}). The possible sources of gas excitation have been explored and the post-asymptotic giant branch (pAGB) stars were identified as the main contributor of ionizing radiation. However, for at least 10\\% of early-type galaxies, the ionized emission is powered by recently formed stellar subcomponent (\\citealt{Sarzi10}). Other forms of ISM, {\\it i.e.} hot and cold gas have also been found in early-type galaxies. Hot gas halos have been detected around early-type galaxies with X-ray observatories (\\citealt{For85, Can87, Fab89, Osul01, Sar01, Kim03, Kim10}). Cold molecular gas is detected through CO emission in $\\sim$ 22\\% of all early-type galaxies (\\citealt{You11} and references therein). In a large fraction of dusty early-type galaxies, the morphology and extent of ionized gas match with that of dust (\\citealt{Goud94b, Fer99, Pat07}) and in some cases with the X-ray emitting region too (\\citealt{Goud98}). This points towards a possible physical connection between hot, warm and cold phases of ISM. The origin and fate of ISM in early-type galaxies is important as it holds clue to the formation and subsequent evolution of early-type galaxies (\\citealt{Goud95}). The recent studies have shown that neither internal nor external origin can explain all the observed properties of ISM in these galaxies. Hence, a good balance between internal and external origin is suggested as the source of ISM in early-type galaxies. In the present paper, we discuss the properties of dust and ionized gas in a sample of forty low redshift early-type galaxies. The paper is organized as follows: Observation and data reduction in Section 2. Section 3 gives the methodology used for analyzing dust and ionized gas, the results have been discussed in Section 4. We summarize our results in Section 5. \\begin{table} \\begin{scriptsize} \\begin{center} \\caption{Global parameters of sample galaxies} \\begin{tabular}{|lllllcc|} \\hline Object & RA & DEC & Morph. &$B^{0}_{T}$& $V_{Helio}$& Size \\\\ &(J2000) &(J2000) &(RC3) & & (km/s) & (arcmin)\\\\ \\hline NGC\\,0383& 01:04:39 & \\,36:09:07 & SA0 & 13.38 & 5098 & 1.6x1.4 \\\\ NGC\\,0708& 01:52:46 & \\,36:09:07 & E & 13.70 & 4855 & 3.0x2.5 \\\\ NGC\\,0720& 01:53:00 & -13:44:19 & E5 & 11.16 & 1745 & 4.7x2.4 \\\\ NGC\\,1052& 02:41:04 & -08:15:21 & E4 & 12.10 & 1510 & 3.2x2.1 \\\\ NGC\\,1167& 03:01:42 & \\,35:12:21 & SA0 & 13.38 & 4945 & 2.8x2.3 \\\\ NGC\\,1199& 03:01:18 & -15:48:29 & E3 & 12.37 & 2570 & 2.4x1.9 \\\\ NGC\\,1395& 03:38:29 & -23:01:40 & E2 & 10.97 & 1717 & 5.9x4.5 \\\\ UGC\\,2783& 03:34:18 & \\,39:21:25 & S0 & 12.99 & 6173 & 1.3x1.2 \\\\ NGC\\,1407& 03:40:11 & -18:34:49 & E0 & 10.70 & 1779 & 4.6x4.3 \\\\ NGC\\,2534& 08:12:54 & \\,55:40:19 & E1 & 13.70 & 3447 & 1.4x1.2 \\\\ NGC\\,2644& 08:41:31 & \\,04:58:49 & S & 13.31 & 1939 & 2.1x0.8 \\\\ NGC\\,2768& 09:11:37 & \\,60:02:14 & S0 & 10.84 & 1373 & 8.1x4.0 \\\\ NGC\\,2851& 09:20:30 & -16:29:43 & E & 15 & 5195 & 1.2x0.5 \\\\ NGC\\,2855& 09:21:27 & -11:54:34 & SA & 12.63 & 1897 & 2.5x2.2 \\\\ NGC\\,3065& 10:01:55 & \\,72:10:13 & SA & 13.5 & 2000 & 1.7x1.7 \\\\ NGC\\,3115& 10:05:14 & -07:43:07 & S0 & 09.87 & 663 & 7.2x2.0 \\\\ NGC\\,3377& 10:47:42 & \\,13:59:08 & E5 & 11.24 & 665 & 5.2x3.0 \\\\ M\\,105 & 10:47:49 & \\,12:34:54 & E1 & 10.24 & 911 & 5.4x4.0 \\\\ NGC\\,3489& 11:00:18 & \\,13:54:04 & S & 11.12 & 677 & 3.5x2.0 \\\\ NGC\\,3607& 11:16:54 & \\,18:03:07 & SA & 10.82 & 960 & 4.9x2.5 \\\\ NGC\\,3801& 11:40:16 & \\,17:43:41 & S0 & 12.96 & 3317 & 3.5x2.1 \\\\ NGC\\,4125& 12:08:06 & \\,65:10:27 & E6 & 10.65 & 1356 & 5.8x3.2 \\\\ NGC\\,4233& 12:17:07 & \\,07:37:28 & S0 & 12.8 & 2371 & 2.3x0.9 \\\\ NGC\\,4278& 12:20:06 & \\,29:16:51 & E1 & 11.20 & 649 & 4.1x3.8 \\\\ NGC\\,4365& 12:24:28 & \\,07:19:03 & E3 & 10.52 & 1243 & 6.9x5.0 \\\\ NGC\\,4494& 12:31:24 & \\,25:46:30 & E1 & 10.71 & 1344 & 4.8x3.0 \\\\ NGC\\,4552& 12:35:39 & \\,12:33:23 & E & 10.73 & 340 & 5.1x4.7 \\\\ NGC\\,4648& 12:41:44 & \\,74:25:15 & E3 & 12.96 & 1414 & 2.1x1.6 \\\\ NGC\\,4649& 12:43:39 & \\,11:33:09 & E2 & 09.81 & 1117 & 7.4x6.0 \\\\ NGC\\,4697& 12:48:35 & -05:48:02 & E6 & 10.10 & 1241 & 7.2x4.7 \\\\ NGC\\,4874& 12:59:35 & \\,27:57:34 & cD & 12.63 & 7224 & 1.9x1.9 \\\\ NGC\\,5322& 13:49:15 & \\,60:11:26 & E3 & 11.14 & 1754 & 5.9x3.9 \\\\ NGC\\,5525& 14:15:39 & \\,14:16:57 & S0 & 13.6 & 5553 & 1.4x0.9 \\\\ NGC\\,5812& 15:00:55 & -07:27:26 & E0 & 12.19 & 1970 & 2.1x1.9 \\\\ NGC\\,5846& 15:06:29 & \\,01:36:20 & E0 & 11.05 & 1714 & 4.1x3.8 \\\\ NGC\\,5866& 15:06:29 & \\,55:45:48 & S0 & 10.74 & 672 & 4.7x1.9 \\\\ NGC\\,6166& 16:28:38 & \\,39:33:06 & E2 & 12.78 & 9100 & 1.9x1.4 \\\\ NGC\\,7052& 21:18:33 & \\,26:26:48 & E & 13.40 & 4672 & 2.5x1.4 \\\\ NGC\\,7454& 23:01:07 & \\,16:22:58 & E4 & 12.78 & 2022 & 2.2x1.6 \\\\ IC\\,2476 & 09:27:52 & \\,29:59:09 & S0 & 13.85 & 8007 & 1.5x1.4\\\\ \\hline \\multicolumn{7}{p{12cm}}{ Cols.(2) and (3) list galaxy co-ordinates, Col.(4) lists morphological classification of the galaxies,blue luminosity of the program galaxies are listed in Col.(5), while Col.(6) lists heliocentric velocity, Col.(7) optical size of the galaxies, all taken from RC3(de Vaucoulers et al. 1991)} \\end{tabular} \\label{basic} \\end{center} \\end{scriptsize} \\end{table} ", "conclusions": "\\textit{Dust}: For the galaxies with obvious features of dust extinction, it is found that morphology and extent of the dust feature in colour-index image is similar to that of extinction map. This indicates that the extinction map approximates the dust obscured region to a high degree. An analysis of extinction curves for 12 galaxies is presented in this work. The extinction curves for eight galaxies are reported for the first time. The value of $R_{V}$ and relative grain size for the sample galaxies obtained by analyzing the extinction maps are given in Table 5. The value of $R_{V}$ for sample galaxies varies from 2.1 to 3.8 with an average of 2.95. It is close to the previously reported values by \\cite{Goud94c} ($R_{V}$= 2.7), \\cite{Pat07} ($R_{V}$= 3.01) and \\cite{Fin10d} ($R_{V}$= 2.82). We find that the galaxies having $R_{V}$ values less than the canonical value of 3.1 have well settled dust morphology in the form of a dust lane or a ring and have relatively smaller grain size than that of the Milky Way. Conversely, the galaxies with larger values of $R_{V}$ show irregular dust morphology and larger relative grain size. This result is in good agreement with results of \\cite{Goud94b}, \\cite{Pat07} and \\cite{Fin10i}. There are four galaxies in common with the samples of \\cite{Goud94c} and \\cite{Pat07}. Table 8 presents the value of $R_{V}$, relative grain size and optical dust mass for the four common galaxies along with the previously reported values in the literature. Except for galaxy \\astrobj{NGC 4125}, which shows complex dust structure (\\citealt{Goud94c}), our results are in good agreement with those reported in the literature. \\cite{Goud94b} explored various dust destruction mechanisms such as sputtering of dust by supernova blast wave, turbulent shocks, thermal ions and hot gas etc. It is shown that for volatile grain of radius $\\sim$ 0.1 $\\mu$m, the lifetime of dust grain varies from 10$^{7}$ to 10$^{9}$ years. The mass of dust and its grain size are expected to drop gradually with time since the dust was acquired by the galaxy. In this scenario, the smooth and regular dust lane morphology associated with smaller grain size is expected, if the galaxy had sufficient time for the dust to settle down in a regular shape. On the other hand if the dust is acquired recently, it will lack regular dust morphology as in the case of \\astrobj{NGC 3801}. Among our sample galaxies \\astrobj{NGC 3801} shows considerably larger relative grain size (1.47) and larger value of $R_{V}$. A multiwavelength study of this galaxy by \\cite{Hota12} revealed that this galaxy has a kinematically decoupled core or an extremely warped gas disc, dust filaments with complex structures, recent star burst with age less than 500 Myr and evidence of ionized gas (as seen in this study too). All these facts led \\cite{Hota12} to suggest that this galaxy is a merger-remnant early-type galaxy. The results of \\cite{Hota12} support that \\astrobj{NGC 3801} acquired dust in the recent past and did not have enough time to settle it into a smooth, regular morphology, which may result in the observed larger grain size and $R_{V}$ value, as suggested by \\cite{Goud94c}. The dust mass reported in this work using optical extinction is in the range $10^2 - 10^4$ $M_{\\odot}$ while, the dust mass estimated using IRAS fluxes is higher by a factor of $\\sim$ 10$^2$. This indicates that a significant fraction of dust is diffusely distributed throughout the galaxy (\\citealt{Goud95, Wise96}) which could not be detected with optical observations, this is in agreement with that of previous studies (\\citealt{Pat07, Fin08, Fin10d}). \\begin{table*}\\tiny{ \\caption{Comparison of dust properties} \\begin{tabular}{|l|ccl|cccl|} \\hline Galaxy Name&\\multicolumn{3}{|c|}{This Paper}& \\multicolumn{4}{|c|}{Reported values for comparison}\\\\ \\hline & $R_{V}$&$\\frac{}{a_{Gal}}$ & log($\\frac{M_{d,opt}}{M_{\\odot}}$) & $R_{V}$&$\\frac{}{a_{Gal}}$&log($\\frac{M_{d,opt}}{M_{\\odot}}$) & Reference\\\\ \\hline NGC\\,2534&2.1$\\pm$0.20&0.83&3.32$\\pm$0.03&2.03$\\pm$0.28&0.80&3.78& \\cite{Pat07} \\\\ NGC\\,3489&3.1$\\pm$0.33&1.04&2.40$\\pm$0.05&3.38$\\pm$0.21&1.09&3.66& \\cite{Pat07} \\\\ NGC\\,4125&3.6$\\pm$0.27&1.36&2.10$\\pm$0.02&2.74$\\pm$0.33&0.96&5.22& \\cite{Goud94c}\\\\ NGC\\,5525&3.2$\\pm$0.28&1.01&4.04$\\pm$0.09&3.15$\\pm$0.17&0.99&5.67& \\cite{Pat07} \\\\ \\hline \\end{tabular} \\label{dust_compare} } \\end{table*} \\textit{Ionized gas}: H$\\alpha$ line emission is detected in $\\sim$ 85\\% of our sample galaxies observed in narrow-band, the results are listed in Table 7. % The H$\\alpha$ luminosity of the sample galaxies lies in the range 10$^{38}$ - 10$^{41}$ {\\it erg sec$^{-1}$} and the mass of ionized gas lies in the range $\\sim$ 10$^{3}$ - 10$^{5}$ M$_{\\odot}$. The ionized gas and dust have identical morphology in almost all the galaxies detected in H$\\alpha$ (refer Figure 2 and 4). % We have several galaxies in common with earlier studies by \\cite{Kim89}, \\cite{Trin91}, \\cite{Shi91}, \\cite{Goud94b}, \\cite{Macch96} and \\cite{Fin10i}. The H$\\alpha$ flux estimated in this work are in good agreement with the flux reported in the literature (refer Table 9).% There are 6 galaxies for which H$\\alpha$ flux are available from more than one source and a significant scatter is noticed among them. The source of scatter may be due to the higher uncertainty involved in the method of determination of the line emission fluxes as discussed in Section 3.3. Similar discrepancy in H$\\alpha$ flux was also reported by \\cite{Fin10i}. \\begin{table}\\tiny{ \\begin{center} \\caption{Comparison of ${H\\alpha}$ fluxes} \\begin{tabular}{|llll|} \\hline Galaxy &$\\textit{f}_{H\\alpha}$& Other obs. & Reference \\\\ name& This paper &&\\\\ &$ergcm^{2}s^{-1}$&$ergcm^{2}s^{-1}$&\\\\ \\hline NGC0708& 39$\\times 10^{-14}$ &49$\\times 10^{-14}$& \\cite{Fin10i}\\\\ NGC0720& 15.7$\\times 10^{-14}$&$<$ 1.49$\\times 10^{-14}$& \\cite{Goud94b}\\\\ &&$< 3.53 \\times 10^{-14}$& \\cite{Shi91}\\\\ &&17.66$\\times 10^{-14}$& \\cite{Macch96}\\\\ NGC1052& 22.4$\\times 10^{-14}$&8.1$\\times 10^{-14}$& \\cite{Kim89}\\\\ NGC1199& 3.5$\\times 10^{-14}$ &4.9$\\times 10^{-14}$& \\cite{Fin10i}\\\\ NGC1395& 3.91$\\times 10^{-14}$&2.2$\\times 10^{-14}$& \\cite{Goud94b}\\\\ &&22.7$\\times 10^{-14}$& \\cite{Macch96}\\\\ &&9.33$\\times 10^{-14}$& \\cite{Trin91}\\\\ NGC1407& 18.0$\\times 10^{-14}$&0.9$\\times 10^{-14}$& \\cite{Goud94b}\\\\ &&$<$5.11$\\times 10^{-14}$& \\cite{Shi91}\\\\ &&4.82$\\times 10^{-14}$& \\cite{Macch96}\\\\ NGC2534& 1.13$\\times 10^{-14}$&6.4$\\times 10^{-14}$& \\cite{Fin10i} \\\\ NGC3377& 7.24$\\times 10^{-14}$&11.9$\\times 10^{-14}$& \\cite{Goud94b}\\\\ NGC3489& 15.2$\\times 10^{-14}$&51.89$\\times 10^{-14}$& \\cite{Macch96}\\\\ NGC3607& 35.0$\\times 10^{-14}$&7.83$\\times 10^{-14}$& \\cite{Macch96}\\\\ NGC4125& 14.1$\\times 10^{-14}$&39.5$\\times 10^{-14}$& \\cite{Goud94b}\\\\ &&16$\\times 10^{-14}$& \\cite{Kim89}\\\\ NGC4649& 9.79$\\times 10^{-14}$&11$\\times 10^{-14}$& \\cite{Trin91}\\\\ NGC4697& 31.8$\\times 10^{-14}$&29.5$\\times 10^{-14}$& \\cite{Goud94b}\\\\ &&4.7$\\times 10^{-14}$& \\cite{Trin91}\\\\ NGC5812& 25.3$\\times 10^{-14}$&5.83$\\times 10^{-14}$&\\cite{Macch96}\\\\ NGC5846& 18.9$\\times 10^{-14}$&16$\\times 10^{-14}$& \\cite{Trin91}\\\\ &&28.02$\\times 10^{-14}$& \\cite{Macch96} \\\\ NGC7052& 0.74$\\times 10^{-14}$&6.0$\\times 10^{-14}$& \\cite{Fin10i}\\\\ \\hline \\end{tabular} \\label{h_compare}\\end{center} } \\end{table} \\subsection{Association of dust and ionized gas} Association of dust and ionized gas and their possible origin in E/SO galaxies were studied recently by \\cite{Fin12}. The HST observations revealed presence of nuclear and filamentary/patchy dust in the central part of a significant fraction of early-type galaxies (\\citealt{Tran01, Mart04}). Most of the galaxies from our sample show dust in their central part. A study similar to that of \\cite{Fin12} was carried out, by including early-type galaxies with nuclear, filamentary/patchy dust to the sample of galaxies having dust lanes. Various relevant parameters like optical dust mass, IRAS dust mass, ionized gas mass, optical luminosity, IR luminosity, H$\\alpha$ luminosity etc. were collected from \\cite{Bus93}, \\cite{Goud94b}, \\cite{Goud95}, \\cite{Sahu96} \\cite{Sahu98}, \\cite{Dew99}, \\cite{Sing95}, \\cite{Tran01}, \\cite{Pat07}, \\cite{Fin10i}, \\cite{Fin12}, this we denote as combined sample. Wherever necessary the quantities have been normalized for H$_0$ = 73 km sec$^{-1}$ Mpc$^{-1}$. The lower optical depth and smaller sampling area in galaxies with patchy /filamentary dust makes it difficult to determine dust mass using extinction values. Dust mass using optical extinction could be estimated only for 12 galaxies with prominent dust features in our sample. It was demonstrated by \\cite{Tran01} that even in the case of patchy/filamentary dust, the dust mass determined using optical extinction and IRAS fluxes are well correlated. Further, the ionized gas mass and optical dust mass are found to be correlated in early-type galaxies (\\citealt{Fer99, Fin12}). The slope of log(M$_{HII}$) versus log(M$_{dust, opt}$) for the combined sample was found to be $\\sim$ 0.6, which is shallower than the slope ($\\sim$ 0.8) found by \\cite{Fin12} for dust lane galaxies. This indicates that in the prominent dust lane galaxies the source of ionization may be more efficient compared to the galaxies with patchy/filamentary dust. In Figure 5 dust mass derived using IRAS fluxes is plotted versus ionized gas mass. A good correlation between these two quantities is seen. The best fitting line to the observed data points with slope $\\sim$ 0.72 is also shown in the same figure. Our results about the co-existence of dust and ionized gas and their correlation are consistent with those of similar work by \\cite{Goud94b}, \\cite{Goud95}, \\cite{Macch96} and \\cite{Fin08, Fin10d}. \\begin{figure} \\begin{center} \\includegraphics[totalheight=3.4in,width=3.4in]{fig5.eps} \\caption{ Ionized gas mass plotted versus dust mass estimated from $IRAS$ fluxes} \\label{mass_ih} \\end{center} \\end{figure} \\subsection{Possible source of ionization in early-type galaxies} The common presence of ionized gas in early-type galaxies is well established (\\citealt{Kim89, Goud94b, Macch96, Sarzi06, Fin08}), however, the source of ionization is still debated. Possible sources of ionization in early-type galaxies were investigated (\\citealt{Mart04, Sarzi10}). \\cite{Sarzi10} explored various possible sources of ionization e.g. post-asymptotic giant branch (pAGB) stars, presence of AGN's, fast shocks, OB stars and interaction with hot ISM. On the basis of their ionizing balance arguments the pAGB stars were regarded as the best candidate for photoionization; it may be either associated to old stellar population or to recent star formation. However, on-going star formation is also responsible for ionization of gas in at least $\\sim$ 10\\% of their sample. The GALEX ultraviolet data revealed that the strong UV flux from $\\sim$ 30\\% of nearby bright early-type galaxies can not be explained without invoking $\\sim$ 1 - 3\\% (in total stellar mass) star formation rate in the last billion years. The fraction of galaxies showing recent star formation could be as high as $\\sim$ 50\\%, if corrected for extinction and proper care is taken for UV flux contribution of AGNs (\\citealt{Kav06}). Early-type galaxies are known to have cold molecular gas (\\citealt{Huch95, Mor06, Com07, Oos10}). More recently, in the ATLAS$^{3D}$ sample of early-type galaxies the detection rate of molecular gas is $\\sim$ 22\\% with H$_{2}$ mass $\\log$ M(H$_{2}$)/M$\\odot$ in the range 7.10 - 9.29. A strong correlation between presence of molecular gas and dust, blue features and young stellar ages seen in the H$\\beta$ absorption indicate that the detected molecular gas is often involved in the star formation (\\citealt{You11}). Based on the CO emission in a representative sample of early-type galaxies, \\cite{Com07} and \\cite{Croc11} showed that the CO-emitting early-type galaxies form a low star formation rate (SFR) extension to the empirical law in spirals. We searched for available data on CO emission in our sample galaxies. Twelve out of 26 galaxies with dust and $H{\\alpha}$ emission in our sample were mapped to check for CO emission while for remaining 14 galaxies we don't have any information. CO was detected in only four galaxies namely NGC 2768, NGC 3489 (\\citealt{Croc11}) NGC 3607, NGC 5866 (\\citealt{Davis11}) and upper limits are available for NGC 4125 (\\citealt{wiklind95}), NGC 4649 (\\citealt{Young02}). Using [OIII]/H$\\beta$ ratio \\cite{Croc11} have investigated possible source of ionization in his sample galaxies. It is concluded that in NGC 2768 old pAGB stars/AGN and in NGC 3489 young pAGB stars are the main source of the ionizing photons. However, to get a better estimate of fraction of CO emitting galaxies, where star formation is the dominant source of ionization, a detailed investigation similar to that done by \\cite{Croc11} is required. \\subsection{Origin of dust and gas in early-type galaxies} \\begin{figure}\\begin{center} \\includegraphics[totalheight=3.4in,width=3.4in]{fig6.eps} \\caption{Comparison IRAS dust mass, normalized by the luminosity} \\label{mass_lum} \\end{center} \\end{figure} In early-type galaxies presence of ISM in various forms is well established, but its origin is still an open issue. Many evidences support their co-existence and physical association, pointing towards a common origin (\\citealt{Goud94b, Goud95, Macch96, Sarzi06, Fin10i, Fin12}). As discussed earlier, $H{\\alpha}$ emission is detected in a significant fraction of our sample galaxies. The matching morphology of line emitting regions with that of optical extinction indicates co-existence of dust and ionized gas in them. The possible sources for origin of dust and gas in these systems are internal and external. In case of internal origin dust is deposited through mass loss from evolved stars. The dust particles also get destroyed due to sputtering in various environments. It has been demonstrated that the mass of dust observed in early-type galaxies is generally higher than that expected from stellar mass loss, indicating an external origin of dust such as galaxy interaction or mergers (\\citealt{Goud95, Pat07, Fin12} and references therein). To investigate possible origin of dust in early-type galaxies relationship between optical luminosity and dust mass normalized by luminosity is also explored (\\citealt{Goud95, Fer99, Pat07}). In Figure 6, % IRAS dust mass normalized with respect to blue luminosity is plotted against blue luminosity for the combined sample. The absence of any apparent correlation between these quantities suggests the possibility of external such as merger related origin of dust and gas in these galaxies. The observed misalignment between the angular momentum vectors of interstellar gas and stellar system in several early-type galaxies suggests that they are kinematically decoupled (\\citealt{Ber88, Kim89, van95, Caon00, Kraj08, Sarzi06}). Further, \\cite{Sarzi06} showed that the kinematical misalignment between the gaseous and stellar component is strongly dependent on apparent flattening and the level of rotational support in these galaxies. The flatter and faster rotating early-type galaxies are known to have preferentially co-rotating gaseous and stellar systems, indicating that an internal origin of gas may play an important role in fast rotating galaxies. Hence, it is necessary to invoke a balance between the internal and external origin of gas and dust to explain the recent observational results (\\citealt{Fin12}). Moreover, the observed blue colours (near-UV - $r$ $<$ 5.5) in a large sample of early-type galaxies can only be explained by introducing some amount of recent star formation in these galaxies (\\citealt{Kav06}). The recycled gas from stellar mass loss within the galaxy is not enough to account for their observed blue colour. Hence, an additional source of fuel for star formation is unavoidable, may be from external origin." }, "1402/1402.2686_arXiv.txt": { "abstract": "We present IRAM Plateau de Bure Interferometer observations of the CO(3--2) and CO(5--4) line transitions from a \\lya blob at $z$ $\\sim$ 2.7 in order to investigate the gas kinematics, determine the location of the dominant energy source, and study the physical conditions of the molecular gas. CO line and dust continuum emission are detected at the location of a strong MIPS source that is offset by $\\sim$1.5\\arcsec\\ from the \\lya peak. Neither of these emission components is resolved with the 1.7\\arcsec\\ beam, showing that the gas and dust are confined to within $\\sim$7\\,kpc from this galaxy. No millimeter source is found at the location of the \\lya peak, ruling out a central compact source of star formation as the power source for the \\lya emission. Combined with a spatially-resolved spectrum of \\lya and \\heii, we constrain the kinematics of the extended gas using the CO emission as a tracer of the systemic redshift. Near the MIPS source, the \\lya profile is symmetric and its line center agrees with that of CO line, implying that there are no significant bulk flows and that the photo-ionization from the MIPS source might be the dominant source of the \\lya emission. In the region near the \\lya peak, the gas is slowly receding ($\\sim$100\\kms) with respect to the MIPS source, thus making the hyper-/superwind hypothesis unlikely. We find a sub-thermal line ratio between two CO transitions, \\ICO{5}{4}/\\ICO{3}{2} = 0.97 $\\pm$ 0.21. This line ratio is lower than the average values found in high-$z$ SMGs and QSOs, but consistent with the value found in the Galactic center, suggesting that there is a large reservoir of low-density molecular gas that is spread over the MIPS source and its vicinity. ", "introduction": "\\setcounter{footnote}{0} Giant Ly$\\alpha$ nebulae (also known as ``\\lya blobs'') are large (50--100~kpc) spatially extended regions emitting copious amounts of Ly$\\alpha$ emission [$L$(\\lya)$\\sim$ 10$^{43-44}$ \\unitcgslum] \\cite[e.g.,][]{Keel99, Steidel00, Francis01, Matsuda04, Matsuda11, Dey05, Saito06, Smith&Jarvis07, Ouchi09, Prescott09, Prescott12a, Yang09, Yang10, Erb11}. They may represent sites of massive galaxy formation and their early interaction with the intergalactic medium. However, the questions of what powers these gigantic gas halos and whether the surrounding gas is outflowing from or infalling into the embedded galaxies are still debated. The proposed scenarios include photo-ionization by AGNs \\citep{Geach09}, shock-heated gas by galactic superwinds \\citep{Taniguchi&Shioya00}, cooling radiation from cold-mode accretion \\citep{Fardal01, Haiman00, Dijkstra&Loeb09, Goerdt10}, and resonant scattering of \\lya from star-forming galaxies \\citep{Steidel11, Hayes11}. However, observations of \\lya blobs currently allow no firm conclusions about their nature. For example, less than 20\\% of blobs are found to contain X-ray luminous AGN \\citep{Geach09} implying that a powerful AGN is not always required for producing the extended halo \\cite[see also][]{Yang09}. Yet \\citet{Overzier13} argue strongly that most are powered by AGN. It has long been believed that \\lya blobs undergo intense dusty starbursts like submillimeter galaxies \\citep{Ivison98, Chapman04, Geach05}, but recent studies show that most \\lya blobs are not as luminous at rest-frame far-infrared (FIR) wavelengths as submillimeter galaxies \\citep{Yang12, Tamura13}. The molecular gas content in \\lya blobs has to date remained unconstrained, despite numerous attempts to detect CO emission \\citep{Chapman04, Yang12, Wagg12a}. Observations of the molecular gas component in these systems are key to determining the physical conditions in the star-forming gas, the star formation efficiency, and constraining various relevant timescales and the role of feedback. CO emission lines are also a useful tool as a tracer of the systemic redshift, without which it is difficult to constrain the gas kinematics within a given system. For example, previous studies of the same \\lya blob using only \\lya emission line have led to radically different conclusions about the gas kinematics, suggesting that gas is outflowing in stellar or AGN winds \\citep{Wilman05}, inflowing due to gas accretion \\citep{Dijkstra06b}, or even static \\citep{Verhamme08}. These discrepancies mainly arise because the systemic velocities of the galaxies within the \\lya blob are not directly measured, but rather need to be assumed \\cite[see also][]{Bower04,Weijmans09}. While other nebular lines in the UV/optical such as \\oiii and \\ha have been used to constrain the systemic velocity within a small sample of \\lya blobs \\cite[e.g.,][]{Yang11,McLinden13}, they are strongly susceptible to extinction and are only detectable from the ground over a restricted redshift range due to atmosphere. In such cases, the far infrared emission lines such as CO and \\cii 158\\um are ideal probes to determine the systemic redshift of the system, should they harbor a cold molecular component. In this paper, we discuss new Plateau de Bure Interferometer (PdBI) observations of the CO $J$ = 3\\,$\\rightarrow$\\,2 and $J$ = 5\\,$\\rightarrow$\\,4 line transitions from one of the most luminous \\lya blobs \\cite[LABd05;][]{Dey05}. Originally discovered due to its strong {\\it Spitzer} MIPS 24\\um flux, LABd05 has been the subject of many multi-wavelength studies. {\\it Spitzer} and high resolution {\\sl HST} optical/NIR observations show that this giant \\lya nebula is composed of a strong MIPS source and $\\sim$17 small compact galaxies \\citep{Prescott12b}. The peak of the \\lya emission is substantially offset from all the galaxies (e.g., 1.5\\arcsec\\ from the MIPS source), but is coincident with a detection in \\heii emission. \\citet{Prescott08} found that LABd05 inhabits an overdense environment, suggesting that it may be the progenitor of a rich galaxy group or low-mass galaxy cluster. The SED of the bright MIPS source within LABd05 can be explained by an AGN-dominated template with a FIR luminosity of \\lfir(40--1000\\um) = 4\\E{12}\\lsun \\citep{Dey05, Bussmann09, Yang12}, although it is not yet clear whether the far-IR luminosity is driven primarily by star-formation or AGN activity \\citep{Colbert11}. The MIPS source contributes the bulk of the bolometric luminosity from the region and has a very extreme rest-frame UV-to-mid-IR color that characterizes it as a heavily dust-obscured galaxy \\cite[DOG;][]{Dey08, Prescott12b}. On the other hand, the kinematics of \\lya-emitting gas have not been fully constrained. While the Keck Low Resolution Imaging Spectrometer (LRIS) longslit spectroscopy shows that there is a monotonic (and approximately linear) velocity gradient across the \\lya blob (maybe due to rotation, outflow or infall), the interpretation of this signature was ambiguous because of the lack of the systemic redshift of the embedded galaxies, possibly the center of the gravitational potential of the blob \\citep{Dey05}. Previous attempts to detect the CO emission in LABd05 using the IRAM-30m yielded only upper limits \\citet{Yang12}. Here, we present new PdBI interferometric observations of LABd05 that reveal detections of the dust continuum emission and CO emission lines from the system. This paper is organized as follows. In \\S\\ref{sec:observation}, we describe our PdBI observations. In \\S\\ref{sec:detection}, we present the detection of molecular gas from LABd05. In \\S\\ref{sec:voffset}, we constrain the gas kinematics of the \\lya-emitting gas using the systemic redshift derived from CO and \\heii lines. In \\S\\ref{sec:COSED}, we use the CO line SED to study physical conditions in star-forming regions. In \\S\\ref{sec:SED_LABd05}, we present the FIR SED and derive constraints on the dust properties of LABd05. In \\S\\ref{sec:model}, we discuss some plausible physical models for LABd05. Section \\ref{sec:conclusion} summarizes our conclusions. Throughout this paper, we adopt the following cosmological parameters: $H_0$ = 70\\,${\\rm km\\,s^{-1}\\ Mpc^{-1}}$, $\\Omega_{\\rm M}=0.3$, and $\\Omega_{\\Lambda}=0.7$. ", "conclusions": "\\label{sec:conclusion} We have obtained IRAM PdBI observations of the CO $J$ = 3\\,$\\rightarrow$\\,2 and $J$ = 5\\,$\\rightarrow$\\,4 line transitions from a $z$=2.7 \\lya blob (LABd05) in order to investigate the molecular gas content and kinematics, to determine the location of the dominant energy source, and to study the physical conditions of star-forming regions within the \\lya blob. We detect CO line emission from the molecular gas associated with this \\lya blob. % The CO line emission and the dust continuum are detected at the location of a strong MIPS source, which is offset by $\\sim$1.5\\arcsec\\ (12\\,kpc in projection) from the peak of \\lya surface brightness distribution. Neither the CO line nor the dust continuum emission is resolved with our 1.7\\arcsec\\ beam, showing that the molecular gas and dust are confined to within a $\\sim$7kpc region around the MIPS source. In addition, no millimeter continuum source is found at the location of the \\lya peak, excluding the presence of a compact source of hidden star formation at SFR $>$ 260\\msun{yr$^{-1}$} which might be directly responsible for the \\lya emission. The CO line spectra show that the MIPS source is indeed located at the same redshift as the extended \\lya emission and that it is a massive galaxy ($M_{\\rm dyn}$ = 3--7\\E{11}\\msun) based on the broad CO line-width (FWHM = 700\\,\\kms). Combined with Keck/LRIS longslit spectroscopy of \\lya and \\heii, we constrain the kinematics of the extended gas using the CO emission as the best tracer of the systemic redshift. At the position of the MIPS source, the \\lya profile is symmetric and its line center agrees with those of the CO lines. This implies that there are no significant bulk flows and the photo-ionization from the MIPS source might be the dominant source of the \\lya emission. Near the peak of the \\lya nebula, the gas is slowly receding ($\\sim$100\\kms) with respect to the MIPS source, thus disfavoring the hyper-/superwind hypothesis where extreme galactic winds are responsible for the extended \\lya emission. However, we note that we cannot rule out the possibility that gas flow within LABd05 is tangential to our LOS. We find that a significantly sub-thermal line ratio between the two CO transitions, \\ICO{5}{4}/\\ICO{3}{2} = 0.97 $\\pm$ 0.21 [\\LpCO{5}{4}/\\LpCO{3}{2} = 0.35 $\\pm$ 0.08]. This line ratio is lower than the average values found in high-$z$ SMGs and QSOs, but is consistent with the value found in the center of the Milky Way. This line ratio indicates that there is a large reservoir of low-density molecular gas that could be spread over the vicinity of the MIPS source. Observations of CO(7--6) and CO(1--0) lines with higher spatial resolution are required to further constrain the properties of the star-forming regions within this \\lya blob. \\medskip" }, "1402/1402.7295_arXiv.txt": { "abstract": "To illustrate the complementarity of the linear collider and astrophysics bounds on the light (MeV-scale mass) dark matter (DM), we study the constraints on the magnetic dipole DM from the DM-electron interactions at the proposed International Linear Collider (ILC) and in supernova (SN) 1987A. We in particular focus on the $e^+ e^-$ annihilation which is the common process for producing DM pairs both at the ILC and in the SN. We estimate the bounds on the DM magnetic dipole moment from the mono-photon signals at the ILC and also from the energy loss rate due to the freely streaming DM produced in the SN. The SN bounds can be more stringent than those from the ILC by as much as a factor ${\\cal O}(10^5)$ for a DM mass below $10^2$ MeV. For larger DM masses, on the other hand, SN rapidly loses its sensitivity and the collider constraints can complement the SN constraints. ", "introduction": "The nature of the dark matter (DM) remains an outstanding question which can provide crucial clues for the physics beyond the Standard Model (SM). In particular, besides the commonly studied weakly interacting massive particles (WIMPs) with weak scale mass, there has been growing interest in light DM whose parameter region has not yet been experimentally fully explored. For instance, current direct DM search experiments have recoil energy sensitivity down to of order a keV which limits the DM mass to be larger than about a GeV, and it would be of great interest to investigate the lighter mass range below a GeV for the potential window to new physics beyond the SM. Facing the wide open possibilities for the properties of DM, we study the interaction of MeV scale DM particles which possess a magnetic dipole moment and therefore interact with the photon. The DM magnetic dipole moment can be easily generated in many extensions of the SM such as asymmetric DM models and there have been many studies of the dipole DM, in particular for the light DM whose interactions with the SM particles can enjoy infrared enhancement due to the small momentum transfer in the photon exchange \\cite{pos,kam,mas,kopp,dede,bank,zure2,ess,semi,light,barg2,heo3,paolo3,essi,ilidio,gary,nob}. We aim to illustrate the complementarity of linear collider and astrophysical probes on light dark matter in the MeV-scale mass range. We focus on the interactions of the DM and electron/positron pairs and estimate how they can can affect the ILC and SN signals. For the ILC, we study the impact of the magnetic dipole DM on the mono-photon events where pairs of DM particles arise from $e^+ e^-$ annihilations, and, for the SN, we calculate the DM emission rate potentially affecting SN cooling which is also due to pairs of DM particles produced by $e^+ e^-$ annihilations. We relate the collider and SN phenomenology through this common DM production channel of $e^+ e^-$ annihilations and clarify how collider and SN signals can complement each other in constraining the DM magnetic dipole moment. ", "conclusions": "" }, "1402/1402.0010_arXiv.txt": { "abstract": "Hybrid-kinetic numerical simulations of firehose and mirror instabilities in a collisionless plasma are performed in which pressure anisotropy is driven as the magnetic field is changed by a persistent linear shear $S$. For a decreasing field, it is found that mostly oblique firehose fluctuations grow at ion Larmor scales and saturate with energies $\\propto$$S^{1/2}$; the pressure anisotropy is pinned at the stability threshold by particle scattering off microscale fluctuations. In contrast, nonlinear mirror fluctuations are large compared to the ion Larmor scale and grow secularly in time; marginality is maintained by an increasing population of resonant particles trapped in magnetic mirrors. After one shear time, saturated order-unity magnetic mirrors are formed and particles scatter off their sharp edges. Both instabilities drive sub-ion-Larmor--scale fluctuations, which appear to be kinetic-Alfv\\'{e}n-wave turbulence. Our results impact theories of momentum and heat transport in astrophysical and space plasmas, in which the stretching of a magnetic field by shear is a generic process. ", "introduction": " ", "conclusions": "" }, "1402/1402.0592.txt": { "abstract": "{}{}{}{}{} % 5 {} token are mandatory \\abstract % context heading (optional) % {} leave it empty if necessary {The atmospheric composition of transiting exoplanets can be characterized during transit by spectroscopy. Detections of several chemical species have previously been reported in the atmosphere of gaseous giant exoplanets. For the transit of an Earth twin, models predict that biogenic oxygen (O$_2$) and ozone (O$_3$) atmospheric gases should be detectable, as well as water vapour (H$_2$O), a molecule linked to habitability as we know it on Earth.} % aims heading (mandatory) {The aim is to measure the Earth radius versus wavelength $\\lambda$ - or the atmosphere thickness $h(\\lambda)$ - at the highest spectral resolution available to fully characterize the signature of Earth seen as a transiting exoplanet.} % methods heading (mandatory) {We present observations of the Moon eclipse of December 21, 2010. Seen from the Moon, the Earth eclipses the Sun and opens access to the Earth atmosphere transmission spectrum. We used two different ESO spectrographs (HARPS and UVES) to take penumbra and umbra high-resolution spectra from $ \\approx 3100$ to 10400\\AA. A change of the quantity of water vapour above the telescope compromised the quality of the UVES data. We corrected for this effect in the data processing. We analyzed the data by three different methods. The first method is based on the analysis of pairs of penumbra spectra. The second makes use of a single penumbra spectrum, and the third of all penumbra and umbra spectra. } % results heading (mandatory) {Profiles $h(\\lambda)$ are obtained with the three methods for both instruments. The first method gives the best result, in agreement with a model. The second method seems to be more sensitive to the Doppler shift of solar spectral lines with respect to the telluric lines. The third method makes use of umbra spectra which bias the result by increasing the overall negative slope of $h(\\lambda)$. It can be corrected for this \\textit{a posteriori} from results with the first method. The three methods clearly show the spectral signature of the Rayleigh scattering in the Earth atmosphere and the bands of H$_2$O, O$_2$, and O$_3$. Sodium is detected. Assuming no atmospheric perturbations, we show that the E-ELT is theoretically able to detect the $O_2$ A-band in 8~h of integration for an Earth twin at 10~pc.} % conclusions heading (optional), leave it empty if necessary {Biogenic $O_2$, $O_3$, and water vapour are detected in Earth observed as a transiting planet, and, in principle, would be within reach of the E-ELT for an Earth twin at 10~pc.} % ", "introduction": "Observing planetary transits gives access to a set of properties of the star+planet system (e.g. \\cite{winn2009}). The relative size of the planet with respect to the star, given by the depth of the transit light-curve, is one of the most straightforward methods. Spectrally resolved observations have shown for gaseous giant exoplanets that transit the brightest stars that the transit depth varies with wavelength (the first to report this were \\cite{charbonneau2002}, \\cite{vidal-madjar2003}, \\cite{vidal-madjar2004}) which reveal the signature of the planetary atmosphere. At certain wavelengths, the atmosphere indeed absorbs the stellar light and becomes opaque, which makes the planet look larger in front of the star and produces a deeper transit. The atmosphere of smaller terrestrial exoplanets close to or in the habitable zone (HZ) is in principle also accessible by this method (early works by \\cite{schneider1992}, \\cite{schneider1994}, and more recent models by \\cite{ehrenreich2006}, \\cite{kaltenegger2009}, \\cite{ehrenreich2012}, \\cite{garciamunoz2012}, \\cite{snellen2013} and \\cite{betremieux2013}). Recent simulations show that the next generation of instruments in the infrared or visible domains, that is the James Webb Space Telescope (JWST) and the European Extremely Large Telescope (E-ELT), should be able to offer a first access to characterizing terrestrial planets in the HZ and detecting biosignatures (here atmospheric O$_2$ and O$_3$) and molecules linked to habitability (\\cite{rauer2011}; \\cite{hedelt2013}; \\cite{snellen2013}). In this context, it is of prime interest to observe Earth as a transiting exoplanet. Vidal-Madjar et al. (2010) observed the August 2008 partial lunar eclipse with the high-resolution SOPHIE \\'echelle spectrograph at the 193~cm Haute-Provence telescope (\\cite{perruchot2008}; \\cite{bouchy2009}) and showed that the Earth atmosphere thickness versus wavelength can be derived from such observations. Earth observed from the Moon during a lunar eclipse indeed transits in front of the Sun and gives the possibility to study the Earth atmosphere as during a transit. These first observations also allowed one to identify molecular species in the atmosphere, that is water vapour (H$_2$O), a molecule linked to Earth habitabilty, biogenic oxygen (O$_2$), and ozone (O$_3$). Ozone and oxygen are considered as reliable biosignatures for Earth-like planets (\\cite{owen1980}; \\cite{leger1993}; \\cite{seager2013}) except for terrestrial planets outside the liquid-water habitable zone (\\cite{schindler2000}). These species were also detected by Pall\\'e et al. (2009) from umbra spectra taken during the same eclipse. However, Vidal-Madjar et al. (2010) pointed out that the Pall\\'e et al. spectrum is not the transmission spectrum of Earth observed as a distant transiting planet. The umbra signal indeed results from photons that are highly scattered and refracted by the deep atmosphere. These photons strongly deviate from the star-planet direction and cannot be collected during the observation of a transit. From penumbra observations, Vidal-Madjar et al. (2010) additionally detected the thin high-altitude sodium Na I layer. The larger diameter of Earth in the blue due to increasing Rayleigh scattering at shorter wavelengths is also well visible, although it is partially blended with residuals from spectrograph orders correction. In this paper, we analyze new observations obtained during the lunar eclipse on Dec. 21, 2010 with the ESO spectrographs HARPS (from $ \\approx 3800$ to 6900\\AA, \\cite{mayor2003}) and UVES (from $\\approx 3200$ to 10400\\AA, \\cite{dekker2000}). We present first the principle of the observation and different reduction methods: the original Vidal-Madjar et al. (2010) data reduction method (method~1), which makes use of a pair of penumbra spectra, and a second method (method~2), which requires only one single penumbra spectrum to recover the atmosphere thickness profile (Sect.~\\ref{principle}). We also present a third alternative method to derive the thickness profile, which makes use of penumbra and umbra spectra. This third approach is biased by the refraction in the Earth atmosphere (Sect.~\\ref{method3}, method~3), but can be corrected \\textit{a posteriori} with the result of method~1. An observation log and more details about the spectrographs are given in Sect.~\\ref{log}. We give details on the data reduction in Sect.~\\ref{HARPS-data-processing} and Sect.~\\ref{UVES-data-processing} for HARPS and UVES, respectively. The altitude profiles versus wavelength obtained with the three different methods are detailed in Sect.~\\ref{Results}, as well as simulations of the E-ELT observing a transiting Earth twin at 10~pc. %__________________________________________________________________ ", "conclusions": "We have described the analysis of the observations of a lunar eclipse used as a proxy to observe Earth in transit. The results fully confirm those obtained with SOPHIE and the observation of the August 2008 eclipse (\\cite{vidal-madjar2010}) and extend the observed $h(\\lambda)$ profile towards the near infrared thanks to the UVES data. They agree well with the models (\\cite{ehrenreich2006}; \\cite{kaltenegger2009}; \\cite{snellen2013}; \\cite{betremieux2013}). The Rayleigh increase in the blue is visible. Biogenic oxygen and ozone are also well visible in the atmosphere altitude profiles, and are therefore relevant detectable biosignatures to look for during transits of Earth-like exoplanets in the [4000-10500]\\AA\\ range. The oxygen A and B bands show up clearly with HARPS or UVES. Water vapour is well detected at $\\lambda>$~8000\\AA\\ with UVES because water-vapour absorption is stronger at these wavelengths. Therefore even low densities of water vapour at these altitudes are detected. Water vapour is seen at an altitude of up to 20~km around 9500\\AA. Moreover, Rayleigh scattering and ozone absorption are essentially absent above 8000\\AA\\, therefore we are also able to probe deeper layers in the Earth atmosphere where low-altitude water vapour is also present at higher densities. In HARPS data at $\\lambda<$~7000\\AA\\, water vapour appears above 25~km with method~1 but remains below 25~km with method~2 and below 12~km with method~3, because water-vapour absorption is weaker in the visible range. The results from UVES are nevertheless less reliable than those of HARPS, because the PWV was not stable during the beginning of the night when the calibration spectra were recorded on the full-Moon before the eclipse, which compromised the quality of the eclipse observation. Although we were able to recover the water-vapour content in the calibration spectra \\textit{a posteriori}, the results for the detection of water vapour in the altitude profiles with UVES are not coherent between method~1, 2 and 3. This underlines the difficulties of these observations which were made through the local atmosphere above the telescope, while our goal was to measure atmospheric features at a planetary scale. Lunar eclipse observations done from space, either at low or high spectral resolution, would not suffer from the telluric pollution, with the additional benefit of an access to UV and the prominent Hartley band of ozone. Another bias in the recovering of the Earth atmosphere profile from a lunar eclipse is the effect of refraction in method~3 that makes use of refracted light from the umbra. We have shown that this effect can be partially corrected with the results from method~1, but the profiles - either from HARPS or UVES - still show a residual negative slope with respect to the profile from method~1 or 2. Finally these results provide input for the preparation of future observations with the next generation of giant telescopes. We showed that the E-ELT observing in the visible range at low spectral resolution and in the absence of atmospheric perturbations is in principle able to detect the O$_2$ A-band in the atmosphere of a transiting Earth twin. It will be extremely challenging in practice through the atmosphere. If we overcome this difficulty, especially with high spectral resolution spectroscopy, the O$_2$ signature may then be increasing from year to year (from transit to transit), as will the enthusiasm of the observers." }, "1402/1402.5635_arXiv.txt": { "abstract": "We describe the data flow in the operation of the VEGA/CHARA instrument. After a brief summary of the main characteristics and scientific objectives of the VEGA instrument, we explain the standard procedure from the scientific idea up to the execution of the observation. Then, we describe the different steps done after the observation, from the raw data to the archives and the final products. Many tools are used and we show how the Virtual Observatory principles have been implemented for the interoperability of these softwares and data bases. ", "introduction": "\\noindent Optical long baseline interferometry suffers from an image of a complex observing technique reserved to specialists of instrumentation. With the advent of widely open large facilities such as VLTI and KECK, this has change a lot these recent years. Interferometric instruments are now not only used by small teams but are widely shared in open time competition. This is true also on the Center for High Angular Resolution Astronomy (CHARA) Array \\citep{chara} where our instrument, VEGA (Visible spEctroGraph and polArimeter) \\citep{mourard09,Mourard2011}, has been in operation for many years. This effort of opening the access to a wider community has driven some of the developments we have done these last years. An interferometric program relies on an interferometric infrastructure and on one or many instruments. VEGA operates either with $2$, $3$ or $4$ telescopes depending on the science objectives. In all these modes, the choice of the telescopes among the $6$ that CHARA is operating opens a large number of possibilities. It's clear that, starting with a scientific idea, the problem of defining the best observing strategy is not obvious and needs tools. This is usually done by community softwares that simulate the interferometer. Moreover, doing the observations and optimizing the night with the different programs, is also a complicated task. In order to facilitate these two main aspects of the VEGA operation, we have developed a global vision of the information and data flow, starting from the scientific idea up to the final data. This paper aims to present the different steps of the VEGA use. We present the VEGA instrument and its main science programs in Section~\\ref{vega}. Then Section~\\ref{prepa} presents the preparation phase, Section~\\ref{observation} details the way the observations are conducted and finally Section~\\ref{processing} presents the necessary steps from the raw data to the final products. These steps are schematically summarized in Fig.~\\ref{fig:PIVOT} where the relations between all the software tools and the databases are presented. \\begin{figure} \\centering \\includegraphics[scale=0.4]{fig1.jpg} \\caption{Diagram of the relations between all the tools used from the preparation of observations to the exploitation of data. The arrows represent the data flow between the different applications.}\\label{fig:PIVOT} \\end{figure} ", "conclusions": "In this paper we have described the sequence of actions one has to do to proceed from an interesting scientific idea using high angular resolution up to the resolution of the question with data and comparison with models. Due to the intrinsic complexity of the interferometric measurements, mainly due to the possible large number of observing modes, it has been of utmost importance, for non specialists, to highly simplify the procedure. This was exactly the aim of the work we have developed in the last years and the fact that, now, non-interferometric astronomers are already using VEGA for their scientific programs is certainly a great step forward. Thanks to the development of interoperability by the JMMC and OCA teams, in the framework of the development of the Virtual Observatory principles, it has been possible to easily link the different applications and to use the powerful capabilities of databases to smooth all the interferometric process." }, "1402/1402.4891_arXiv.txt": { "abstract": "Different pieces of observational evidence suggest the existence of disks around isolated neutron stars. Such disks could be formed from supernova fallback when neutron stars are born in core-collapse supernova explosions. Efforts have been made to search for disks around different classes of pulsars, which include millisecond pulsars, young neutron star classes (magnetars, central compact objects, and X-ray dim isolated neutron stars), and regular radio pulsars. We review the main results from observations at wavelengths of from optical to sub-millimeter/millimeter. ", "introduction": "\\label{sec:intr} It has long been suggested that isolated pulsars (PSRs) might have disks and the existence of disks would affect evolution of pulsars and help explain some observed features in pulsars \\citep{md81}. The discovery of a planetary system around PSR B1257+12 \\citep{wf92}, which actually was the first discovered extrasolar planetary system, motivated numerous studies about how such a planetary system could be formed. The proposed formation mechanisms more or less involved the existence of an accretion/debris disk around the neutron star (see \\citealt{mh01} and references therein). To date, two additional pulsars, B1620$-$26 and J1719$-$1438, have been found with planetary systems, although the first is a triple system in the globular cluster M4 likely having been formed from a dynamical exchange interaction (\\citealt{sig+03} and references therein), and in the second the planet-mass companion is probably the leftover of a carbon white dwarf having lost most its mass during the phase of low-mass X-ray binary evolution \\citep{bai+11}. In addition to these three pulsars, \\citet{sha+13} recently have shown that the observed, long-term timing variations for PSR B1937+21 are possibly explained by considering an asteroid belt around the pulsar. The above neutron stars are old (ages $> 10^8$ yrs), so-called recycled pulsars with fast, millisecond spin periods. For regular pulsars formed from core-collapse supernovae for 10$^3$--10$^6$ yrs, rotational spin noise (i.e., timing noise) in them is several orders of magnitude larger than that in millisecond pulsars (MSPs) (e.g., \\citealt{sc10}). It is thus difficult to detect planetary bodies around them from pulsar timing. However certain phenomena have been seen hinting the existence of accretion/debris disks. \\citet{cs08} summarize four types of pulsar pulse emission variations: nulling, transient pulse emitting from so-called rotating radio transients (RRATs; \\citealt{mcl+06}), subpulse drifting, and emission mode changing, and consider that these phenomena are caused by migration of circumpulsar debris material into the magnetospheres of pulsars. In individual pulsars, for example, the recent measurement of the second period derivative of the spin of PSR J1734$-$3333 suggests that the magnetic field of this pulsar is increasing \\citep{esp+11}, or alternatively its spin properties might be affected by having an accretion disk \\citep{cal+13}. It has also been suggested that because jets are generally associated with accretion disks, young pulsars seen with jets might harbor accretion disks \\citep{bp04}. While it is not very clear how disks around MSPs would be formed (e.g., \\citealt{sha+13}), particularly for the planetary system case around B1257+12 \\citep{mh01}, young pulsars are believed to possibly have disks due to supernova fallback \\citep{che89}. During a core-collapse supernova explosion, part of ejected material may fallback and if the material has sufficient angular momentum, a disk might be formed around the newly born neutron star \\citep*{lwb91}. The existence of fallback disks has been suggested to be the cause of the diversity of young neutron stars (\\citealt{alp01}; \\citealt*{ace13} and references therein). Unlike what was once thought---young radio pulsars like those in the Crab and Vela supernova remnants were prototypical of newborn neutron stars, it has been realized that there are classes of magnetars \\citep{wt06}, central compact objects (CCOs) in young supernova remnants \\citep*{pst04,del08}, and X-ray dim isolated neutron stars (XDINSs; \\citealt{tur09, mer11}). The latter three classes of young neutron stars are rather `quiet' at multiple wavelength regions from optical to radio: they generally do not have strong non-thermal emission and are not surrounded by any bright, pulsar-wind powered nebulae. Neutron stars generally have non-thermal emission radiated from their magnetospheres, and sometimes a thermal component arising from their hot surfaces may be seen (e.g., \\citealt{bec09}; note that the thermal emission can be dominant in some cases). The non-thermal emission can be described by a power law with flux decreasing from X-ray to optical/IR wavelengths, while because of their $\\sim$10$^6$~K surface temperature, the thermal component's Rayleigh-Jeans tail may be detectable at ultraviolet/optical wavelengths (e.g., \\citealt{kap+11}). As a result, among $\\sim$2000 known neutron stars \\citep{man+05}, only over 20 neutron stars have been detected at optical/IR wavelengths. For a comparison, a disk would have thermal-like emission, and depending on temperature, which is often assumed to be a function of disk radius, the disk would be generally bright at optical/IR wavelengths \\citep{phn00}. Emission from putative disks would thus be distinguishable from that of neutron stars. In addition, if a debris disk consists of cold, $\\leq$100~K dust, it would also possibly be detectable at submillimeter (submm) and millimeter (mm) wavelengths \\citep{pc94}. Given all these reasons, searches for disks around neutron stars have been carried out with different telescopes. The goal of the searches is to find thermal-like emission from neutron stars at wavelengths of from optical to submm/mm, which is not expected to be radiated from neutron stars themselves. In this paper, we review the current status of searches and provide a summary of the main results. It should be note that since young neutron stars (for example the magnetars that are covered in this paper) may exhibit strong variability (e.g., \\citealt{kas07}), and hence (nearly) simultaneous. observations at multiple wavelengths are often required in order to identify the source of emission. ", "conclusions": "\\label{dis} Sufficient observational evidence has shown that debris disks are ubiquitous around from different main-sequence stars (e.g., see \\citealt{wya08} and references therein) to white dwarfs (e.g., \\citealt{fjz09}; \\citealt{xj12} and references therein). In main-sequence stars, debris disks may likely be the remnants of protoplanetary disks at early ages of $\\sim$10~Myr and then are replenished by planetesimal collisions. In white dwarfs, which like the neutron stars are also post main-sequence, compact stars, debris disks are thought to be formed from material produced by tidal disruption of planetary bodies \\citep{gra+90,jur03} that have survived through late phases of stellar evolution \\citep{ds02}. Comparing neutron stars to them, the major uncertainties would be the initial conditions and heating mechanisms if the current scenario of fallback disk formation is believed. For young neutron stars (magnetars, CCOs, and XDINSs) and regular radio pulsars, a debris disk would be the remnant of a fallback disk. The initial mass of the fallback disk and its interaction with a newly-born neutron star are quite uncertain, which would affect subsequent evolution and detectability. While dust grains around normal stars or white dwarfs are nicely heated to 100--1000~K temperature by ultraviolet/optical photons from host stars, revealing their existence, it is not clear how much a pulsar's particle wind can heat up a disk. Hopefully in the near future, with great capabilities of the next generation telescopes such as the thirty meter telescope and \\textit{James-Webb Space Telescope}, these uncertainties would be cleared out by detections or even sufficiently deep upper limits of a number of different classes of neutron stars. In summary, deep searches for disks around different types of neutron stars have been carried out. For magnetars, 6 of them were found with NIR emission, which was seen (except for 4U~0142+61) to be related to magnetars' outburst activities. Further observations to identify the origin of the NIR emission is needed. Among the 6 magnetars with NIR counterparts, 4U~0142+61 has been found to have a MIR component in its optical and IR broad-band spectrum and it is likely indicative of a dust disk around this X-ray emitting neutron star. Deep MIR imaging of the magnetars 1E~2259+586 and 1E 1048.1$-$5937 detected the first source at MIR 4.5~$\\mu$m but did not detect the second one, and the results suggest a similar dust disk to that of 4U~0142+61 around the first source (although only on the basis of the MIR and NIR $K_s$ band detections) and exclude the existence of a similar disk around the second one. For CCOs and XDINSs, the deepest optical and NIR observations were carried out. The non-detections of them rule out accretion (from a disk) as the power source to produce the observed X-ray emission from the CCOs, while the upper limits on the XDINSs are not very constraining due to possibly large inner radii considered for their putative accretion disks. Deep MIR imaging of CCOs was also conducted, but the obtained upper limits can not conclusively determine whether or not dust disks exist. A few millisecond and middle-aged radio pulsars were searched. The discovery of a candidate MIR counterpart to the Vela pulsar, if confirmed, would provide a certain case for our understanding of disk heating and detectability of disks around regular pulsars. Extensive searches using WISE imaging data are underway, and hopefully the results would bring to light the general existence of debris disks and their evolution under pulsars' extreme environments. We gratefully thank anonymous referees for constructive suggestions. This research was supported by the National Natural Science Foundation of China (11073042, 11373055) and the Strategic Priority Research Program ``The Emergence of Cosmological Structures\" of the Chinese Academy of Sciences (Grant No. XDB09000000). Z.W. is a Research Fellow of the One-Hundred-Talents project of Chinese Academy of Sciences." }, "1402/1402.0082_arXiv.txt": { "abstract": "The electromagnetic energy equation is analyzed term by term in a 3D simulation of kinetic reconnection previously reported by \\citet{vapirev2013formation}. The evolution presents the usual 2D-like topological structures caused by an initial perturbation independent of the third dimension. However, downstream of the reconnection site, where the jetting plasma encounters the yet unperturbed pre-existing plasma, a downstream front (DF) is formed and made unstable by the strong density gradient and the unfavorable local acceleration field. The energy exchange between plasma and fields is most intense at the instability, reaching several $\\unit{pW/m^3}$, alternating between load (energy going from fields to particles) and generator (energy going from particles to fields) regions. Energy exchange is instead purely that of a load at the reconnection site itself in a region focused around the x-line and elongated along the separatrix surfaces. Poynting fluxes are generated at all energy exchange regions and travel away from the reconnection site transporting an energy signal of the order of about $\\bfS \\approx 10^{-3} \\unit{W/m^2}$. ", "introduction": "Reconnection is one of the most studied processes capable of releasing magnetic energy into kinetic energy, in the form of flows and particle heating. The typical conditions for reconnection have a sheared magnetic field where across an interface at least one component of the magnetic field reverses sign. In that situation the field lines of opposite polarity break and reconnect in a new configuration. In the process the magnetic energy content is decreased in favor of energization of the particles, in the form of ordered flows or random particle heating. The exact point where the magnetic field breaks needs not bear any direct link with the region where energy is in fact released. Traditionally, two competing scenarios have been guiding the discussion. In the Sweet-Parker mechanism \\citep{sweet,parker} a diffusion region is present around the point of topological line breakage and the energy dissipation in the form of Ohmic heating and plasma acceleration is assumed to take place in the diffusion layer. In competition, the Petschek \\citep{petschek} model sees standing slow shocks form an interface across which energy is released from the magnetic field to the particles. More recently, the kinetic description has identified that aspects of both models are present in the real plasma (see \\citet{birn-priest} for a recent review). And real collisionless plasmas are kinetic of course. The electrons and ions become decoupled~\\cite{sonnerup1979solar,terasawa1983hall} and two layers reminiscent of the Sweet-Parker layer are present \\cite{laval1966instabilites,hesse1999diffusion,kuznetsova2000toward,gem1836}. An inner one has the electrons being accelerated and the outer one the ions. In the typical cartoon picture the two layers are nested boxes. Of course the cartoon is not real and the reality is more complex with the boxes extending along the separatrices, especially in presence of finite guide fields~\\cite{kleva,biskamp,ricciguide}. But in kinetic reconnection, the energy release is not limited to the electron and ion diffusion regions proper. Other energy releases are possible in conduction with reconnection. A main process receiving substantial attention both from the theoretical simulation side and from direct in situ observation is that of the regions downstream of a reconnection site, where the plasma flow caused by reconnection slams into its surrounding plasma. In these regions the outflowing plasma and magnetic field act as a snowplow releasing its mass and energy against the surrounding pre-existing plasma and field. Such fronts emanating from a reconnection site have been studied in recent kinetic simulations \\citep{sitnov2009dipolarization} and confirmed by direct observational evidence has been obtained with data from the THEMIS mission~\\citep{runov2009themis}. Many effects relative to these fronts can be studied in 2D, but one key process requires a full 3D study: the presence of an instability that perturbs the downstream front (DF)~\\citep{nakamura2002interchange,guzdar2010simple}. Fluid studies have been used to capture the effect~\\cite{nakamura2002interchange,guzdar2010simple}. In the fluid case, the instability squires the nature of an interchange mode similar to the Rayleigh-Taylor instability. Across the downstream front the density increases substantially. The vertical magnetic field and the density pile up at the DF. The curvature of the field lines and the braking of the front by the momentum exchange with the yet unperturbed plasma leads to an effective acceleration pointing contrary to the DF speed. Such configuration is unstable to interchange modes: the higher density region is ahead of the front and the density gradient is therefore opposite to the direction of the acceleration~\\citep{nakamura2002interchange}. A well know limitation of MHD in this circumstance is the inability to predict the fastest growing mode: the interchange instability has the same growth speed at any wavenumber $k$. This is circumvented in MHD by seeding the instability ad hoc~\\citep{guzdar2010simple} or self-contently as a consequence of other processes that produce the required seed~\\citep{lapenta2011self}. Full kinetic studies are more suitable for modeling the process, but of course 3D kinetic simulations are much more computationally demanding, limiting the accessible domain size. In kinetic theory, extra physics is present to determine the scale of the process, predicting a fastest growing mode~\\citep{pritchett2010kinetic}. In particular the presence of a density (and pressure) gradient at the DF induces also drift waves and instabilities that modify the nature of the DF instability~\\citep{divin2013dipolarization}. Recently, \\citet{vapirev2013formation} reported a fully kinetic simulation of the development of the secondary instability in DF emerging from a reconnection region in 3D simulations of sufficient domain size to track the evolution for several ion skin depths. Recent Cluster multispacecraft observations provide direct evidence for the presence of an interchange instability at DFs\\citep{runov2012multipoint} at scales comparable with that observed in simulations~\\citep{guzdar2010simple,lapenta2011self,vapirev2013formation}. In the present paper, we consider the issue of the energetic consequences of the DF instability. We track the electromagnetic energy through the system and observe its balance, as described by the electromagnetic energy equation, and its flow described by the Poynting vector. The issue has been considered from an observational angle in the recent work by \\citet{hamrin2012role}. The energy exchange between plasma and fields is measured by computing directly $\\bfJ \\cdot \\bfE$. The measure is not easy, requiring the estimation of the current from the four Cluster spacecraft data \\citep{hamrin2011energy}. Nevertheless, it has been done and the published results are used here to compare with our simulation data. Previous 2D studies have already investigated the energy exchanges. The energetics at the DF has been inferred from Cluster data by \\citet{huang2012kinetic}, finding an energy transfer from the fields to the plasma in DF, in agreement with 2D simulation results \\citep{sitnov2009dipolarization}. The 2D simulations show a more intense electron contribution to the energy deposition near the x-point and a more intense ion deposition elsewhere~\\citep{sitnov2011onset}. The extension to 3D can, and as shown below indeed does, explain several other features observed in data. First, the observations \\citep{hamrin2012role} show that the energy exchange is exclusively from the fields to the plasma near the reconnection site, but as one considers Cluster crossings closer to the Earth the presence of energy exchanges in both directions is found. The condition where energy is going to the plasma from the fields is called a load, with circuit terminology. The opposite situation of energy being transferred by the particles to the field is called a generator. The presence of transfer of energy in both directions is shown also in MHD models~\\citep{birn2005energy}, with loads concentrated in the mid-nigh region and generator regions located in the flanks where the cross-tail current is diverted to the field-aligned currents of the substorm current wedge. The mechanism at play here is different and due to the development of the DF instability. Second, there is observational evidence for an important role of waves generated at the DF \\citep{marghitu2006experimental} where generator regions radiate electromagnetic energy. The suggestion is made that the energy goes into kinetic \\alf waves. The study for the energy fluxes from Cluster observations shows that the dominant component of the energy flux is ion enthalpy flux, with smaller contributions from the electron enthalpy and heat flux and the ion kinetic energy flux\\citep{angelopoulos2002plasma,eastwood2013energy}. The Poynting flux is a minority contribution but it is not negligible, and in certain parts of the ion diffusion region the Poynting flux in fact dominates\\cite{eastwood2013energy} The present description revisits these issues using 3D fully kinetic simulations, measuring directly the energy balance and reproducing many of the observational features outlined above. In particular, we observe that electromagnetic energy is converted to plasma energy at the reconnection site in a region very elongated along the separatrices. At the DF, we find the instability to convert about one order of magnitude more energy than at the reconnection site itself. But this energy is alternating between generator and load regions, as in the observations mentioned above. Significant regions of intense Poynting flux emerges from both regions of energy exchange. At the reconnection site, the energy flux is again of definite sign, but at the DF alternates in sign as the source causing it. The remainder of the paper is organized as follows: Section 2 reports the details of the simulation approach followed and gives an overall view of he processes developing in the simulation. Section 3 considers the energy balance equation for the electromagnetic energy and investigates each term in turn. Section 4 provides a summary and an interpretation of the results of the energy analysis making a direct link with observations. The final overview is provided in Fig.~\\ref{scenario} that summarizes pictorially the findings of the present investigation. ", "conclusions": "" }, "1402/1402.2614_arXiv.txt": { "abstract": "The classical picture of a star-forming filament is a near-equilibrium structure, with collapse dependent on its gravitational criticality. Recent observations have complicated this picture, revealing filaments as a mess of apparently interacting subfilaments, with transsonic internal velocity dispersions and mildly supersonic intra-subfilament dispersions. How structures like this form is unresolved. Here we study the velocity structure of filamentary regions in a simulation of a turbulent molecular cloud. We present two main findings: first, the observed complex velocity features in filaments arise naturally in self gravitating hydrodynamic simulations of turbulent clouds without the need for magnetic or other effects. Second, a region that is filamentary only in projection and is in fact made of spatially distinct features can displays these same velocity characteristics. The fact that these disjoint structures can masquerade as coherent filaments in both projection and velocity diagnostics highlights the need to continue developing sophisticated filamentary analysis techniques for star formation observations. ", "introduction": "Observations with the Herschel satellite have revealed that the backbone of molecular clouds is a compex network of connecting and interacting filaments \\citep[e.g.][]{2010A&A...518L.102A,2010A&A...518L.100M,2011A&A...529L...6A,2012A&A...540L..11S}. The dynamics of this web of dense molecular gas not only determines the evolution and stability of molecular clouds, but also regulates their condensation into stars. Molecular cloud cores and single low-mass stars are almost always found in filaments, often aligned like pearls on a string \\citep{2002ApJ...578..914H,2008ApJ...672..410L,2010A&A...518L.102A}. This can be interpreted as a result of the gravitational instability of a supercritical filamentary section \\citep{1997ApJ...480..681I,2011A&A...533A..34H}. Where filaments intersect, more massive hubs form \\citep{2009ApJ...700.1609M,2013ApJ...764..140M} that later on could become the progenitors of star clusters. \\citet{2010ApJ...724..687L} \\citep[see also e.g.][]{2007ApJ...666..982E} demonstrated that the fraction of gas in the dense molecular web is of order 10 percent of the total molecular mass. Interestingly, the mass fraction of protostars to dense ($n > 10^4$ cm$^{-3}$) molecular gas is also a constant of the same order. \\citet{2013ApJ...773...48B} showed that this requires new filamentary segments to continuously form from the diffuse intra-filament medium on a gravitational collapse timescale. Classically, gas filaments have been treated as cylinders of gas in hydrostatic equilibrium \\citep{1964ApJ...140.1529O,2012A&A...542A..77F,2013A&A...558A..27R}. It is not clear, however, how such quiescent structures could form in the turbulent environment of a molecular cloud \\citep[e.g.][]{2010A&A...520A..17K,2013ApJ...769..115H}. In addition, millimeter line studies indicate that filaments have intrinsic, super-thermal linewidths \\citep{2013A&A...553A.119A}. More recently, observations by \\citet{2013A&A...554A..55H} revealed that filaments are often compact bundles of thin spaghetti-like subfilaments. They presented observations of a prominent filamentary feature in Taurus, dominated by the L1495 cloud \\citep{1962ApJS....7....1L} and several dark patches \\citep{1927cdos.book.....B}. They observed the region in the moderate density tracer C$^{18}$O, obtaining spectra along a $\\sim 10$ pc length of the region. Analysing the richest $\\sim 3$ pc section of the filament in the resultant position--position--velocity space, they found the intriguing result that the gas along the ridge is organized in velocity-coherent filamentary structures, with typical lengths $\\sim0.5$ pc. Each filament is internally subsonic or transsonic, though the collection of filaments is characterised by a mildly supersonic interfilamentary dispersion of $\\sim 0.5$ km s$^{-1}$. They describe the collection of velocity detections as \"elongated groups that lie at different `heights' (velocities) and present smooth and often oscillatory patterns\". Multiple velocity components along a single line of sight have also been observed in Serpens South \\citep{2013ApJ...778...34T}; this feature may therefore be common for many young star forming sites. The complexity of Hacar et al.'s position--position--velocity data, exemplified by their figure 9, invites theoretical and numerical exploration. In particular the apparent organization of filaments into bundles tends to bring to mind magnetic fields or other relatively complex physics. In this paper we explore whether this velocity signal is present in simulated molecular clouds with a more minimal set of physics, including only gravity and hydrodynamic forces acting on the initial turbulence. \\begin{figure*} \\begin{center} \\includegraphics[width=2\\columnwidth]{figure1.pdf} \\caption{\\label{finderimage} Surface density projections through the simulated volume at two times. The dashed cyan lines mark the boundary of the periodic simulation domain. We calculated the surfaced density using only the gas between $10^3 < n/{\\rm cm}^{-3} < 10^{4.5}$, and converted this to an approximation of an optically thin C$^{18}$O observation. The magenta boxes delineate the filaments we focus on in this paper. Sink particles are shown as orange dots.} \\end{center} \\end{figure*} ", "conclusions": "We simulated a 10pc region of a turbulent molecular cloud, allowing it to evolve under self gravity for 1.25 Myr. While the global structure of the region is still dominated by its turbulent structure, the denser gas has had time to become gravitationally organized. We treated the simulation as an observer might observe the sky; converting the density to an approximate line intensity, selecting filamentary regions for further study, and analysing the line-of-sight velocity information in these regions. Our first main finding is that velocity characteristics very similar to those observed by \\citet{2013A&A...554A..55H} form naturally in such a turbulent setup. Individually bound subfilaments display approximately sonic or subsonic dispersions, while the agglomerations that make up the larger filaments have transsonic to mildly supersonic relative motions. While we will further study the detailed evolution of these structures in future work, we speculate that this velocity structure is a relic of the supersonic turbulence that is generally taken as the initial conditions of star formation. The substructured (both in space and velocity) filaments appear without the need for magnetic fields, which are the physical mechanism that immediately spring to mind when considering a filament composed of a bundle of subfilaments. Secondly, the same velocity structures that appear in spatially coherent filaments can show up when observing a faux-filament. Based on the line-of-sight velocity information, there is not a clear case to be made that our filament B (particularly the section with $L > 2$pc; see Figure \\ref{filvels1}) is different from filaments C1 or D1. Examining the third dimension reveals it to in fact be composed of widely separated dense regions (Figure \\ref{filaments3D}). Apparent velocity coherence similar to that seen in our most linear structures is evidently not enough to diagnose a true filament. The presence of this imposter filament in our data, not merely in projected density but also in the line-of-sight velocity, highlights the need for the continued development of sophisticated filament diagnostics, using both simulation and observation and both spatial and velocity data, in order to interpret not only existing observations of star forming filaments but also the imminent onslaught of data from ALMA." }, "1402/1402.6684.txt": { "abstract": "The Automated Planet Finder (APF) is a facility purpose-built for the discovery and characterization of extrasolar planets through high cadence Doppler velocimetry of the reflex barycentric accelerations of their host stars. Located atop Mt. Hamilton, the APF facility consists of a 2.4-m telescope and its Levy Spectrometer, an optical echelle spectrometer optimized for precision Doppler velocimetry. APF features a fixed format spectral range from 374 nm - 970 nm, and delivers a ``Throughput\" (resolution $*$ slit width product) of 114,000 arc-seconds, with spectral resolutions up to 150,000. Overall system efficiency (fraction of photons incident on the primary mirror that are detected by the science CCD) on blaze at 560 nm in planet-hunting mode is 15\\%. First-light tests on the RV standard stars HD 185144 and HD 9407 demonstrate sub-\\ms\\ precision (RMS per observation) held over a 3-month period. This paper reviews the basic features of the telescope, dome, and spectrometer, and gives a brief summary of first-light performance. ", "introduction": "Around the turn of the millennium, the California-Carnegie Exoplanet Team launched a project to construct a dedicated ground-based precision radial velocity facility with which to find planets in the liquid water habitable zone \\citep{kas93, kas13} around low mass stars. An early discussion of the prevalence and possibility of potentially habitable planets around M dwarfs can be found in the discovery paper for GJ 876b, the first known planet around an M dwarf \\citep{mar98}. Such planets have periods in the 20-60 day regime that are quite hard to capture from a conventionally shared telescope that is scheduled around lunation phases. One needs telescope access throughout the lunar month to overcome aliasing and phase coverage problems for periods near the lunar month or integral multiples thereof. Also, one needs much higher cadence than is typically obtainable with a shared large telescope such as Keck in order to acquire enough data points to realize substantial $\\sqrt N$ gains in signal-to-noise to tease out the extremely weak signals of Earth-sized planets in habitable zones from a limiting background of essentially random stellar noise. Successful observational and modeling strategies for optimizing exoplanet detection against a limiting background of stellar jitter noise have been presented by \\cite{dum11a, dum11b, tuo12} and others. And, as this facility was to be operated every night of the year, we also wished to automate the telescope, and hence named it the Automated Planet Finder or APF. The APF project got its official start with cornerstone funding obtained in the form of a \\$6.4 million ear-mark in the 2002 Defense Appropriations Bill through the U.S. Naval Observatory (USNO). For that amount, the original intent was to simply purchase an existing clone of the four 1.8-m telescopes from Electro-Optical Systems Technologies (EOST) of Tucson, AZ that were originally slated to be used as outrigger telescopes for the Keck twin-telescope interferometer. That clone was in storage in EOST's Tucson warehouse. The plan was to house this 1.8-m telescope in a commercial dome sited in the parking lot adjacent to the Shane 3-m telescope's dome and fiber-feed its output into the existing Hamilton spectrometer \\citep{vog87} of the Shane 3-m. However, an additional \\$1.81M NASA grant was obtained that allowed the project to be super-sized up to a 2.4-m EOST telescope and dedicated new spectrometer, in an IceStorm-2 dome manufactured by EOST's parent company Electro-Optical Systems (EOS) of Queanbeyan, Australia. The new dome was sited next to the Lick 20\" Astrograph telescope atop Mt. Hamilton. The total cost-to-completion of the APF project was \\$12.37 million. Work on the project began in 2004 under supervision of the UCO/Lick Labs in Santa Cruz with S. Vogt as Principal Investigator, G. Marcy and D. Fischer as Co-I's, and M. Radovan as Project Manager and Principal Engineer. A large team of technical support personnel from UCO/Lick provided the requisite site engineering and construction management, as well as providing major engineering support to EOS for the dome construction and commissioning, to EOST for telescope construction and commissioning, and for building the Levy spectrometer. In 2007 the California-Carnegie Exoplanet team split into two teams, the California Planet Survey (CPS) team led by G. Marcy at UC. Berkeley, and the Lick-Carnegie Exoplanet Survey (LCES) team co-led by R. Butler at DTM and S. Vogt at UCO/Lick. Although the APF project was initiated under the California-Carnegie Exoplanet team, oversight and management of the telescope and dome contracts, fabrication of the spectrometer, and commissioning of the overall facility were carried out solely by the LCES team and technical staff of UCO/Lick Observatory. The project was initially expected to require only 2-3 years to complete and the contracts called for the delivery of a turn-key telescope and integrated dome facility. Getting to the June 2009 acceptance sign-off involved an arduous six-year slog, with major support on virtually all fronts by UCO/Lick technical staff to the prime contractors for both the telescope and dome. Both the telescope and dome systems had serious issues that required extensive assistance from Lick technical staff to resolve. Among these were UCO/Lick having to make a new M2 to correct a serious back-focal-distance error, providing major assistance in the polishing and figuring of M1, recoating of M3, and providing also the coating for M1. Since then June 2009 acceptance sign-off, UCO/Lick staff have struggled to overcome numerous post-acceptance issues with both the dome and telescope, a task made substantially more difficult by the fact that, while some of the critical documentation (software and firmware) is locked away in an escrow account until January 29, 2018, other critical firmware and board schematics were never delivered and were not included in the escrow deposit. ", "conclusions": "This paper presents a brief overview of the newly-commissioned Automated Planet Finder or APF. At the time of this writing, the facility is still in the final commissioning and testing phases. The facility is currently capable of working through an entire night autonomously, and has been mostly operating robotically since Jan. 1, 2014. Unlike other highly successful precision RV facilities such as HARPS, HARPS-N, and SOPHIE-HR, APF does not attempt to achieve precision through absolute stability of the instrument and the use of an image scrambler. Rather, an Iodine cell is used to provide the precision velocity reference, thereby greatly easing demands on long-term instrumental stability. Nevertheless, the Levy spectrometer on APF incorporates extensive passive athermalization in its optomechanical design to help enhance stability of the PSF and spectral position on the CCD. This passive athermalization, together with a temperature-controlled insulated housing, keeps the spectrum position stable on the CCD to typically under 0.2 pixels over at least several-months time scales, at or a bit less than the position drifts due to barometric pressure changes. Unlike HARPS, where the entire spectrometer is enclosed in a vacuum chamber, or CHIRON \\citep{tok13} and PFS where the echelle is enclosed in a vacuum chamber, no attempt was made with the Levy to control or stabilize atmospheric pressure. Such sub-pixel spectral shifts due to atmospheric pressure changes have no discernible deleterious affect on the resultant precision, and are faithfully removed by the Iodine reference lines. Accurate guiding is also crucial to obtaining high RV precision. Guiding on APF is done by picking off 4\\% of the incoming starlight with a beamsplitter and presenting a symmetric unvignetted seeing disk to the CCDTV for more precise guiding. Tight closed loop tracking specs on the servo-driven telescope also help to further stabilize guiding. Since the APF has forgone the use of fiber scrambling and/or image slicing (both of which can produce substantial losses if not extremely well executed technically), it achieves rather high overall optical efficiency. Under typical exoplanet hunting conditions, the APF facility delivers a peak efficiency (fraction of photons hitting the primary mirror that are detected by the science CCD) of about 15\\%, with a typical spectral resolution of 110,000. At the same time, despite the lack of an image scrambler to stabilize the PSF, observations of the RV-null stars Sigma Dra (HD 185144) and HD 9407 over a period of three months in summer of 2013 demonstrated sub-\\ms\\ (RMS per point) instrumental precision. The APF's Levy produces a fixed spectral coverage from 374-900 nm, much wider than the 490-600 nm Iodine region. As such, it should be useful for a wide range of conventional stellar spectroscopy programs other than precision RV's. While its nominal dispersion and resolution are quite high by traditional standards for fainter object work, it may prove useful for target of opportunity (ToO) observations in support of bright supernovae and GRB follow-up. APF's guider currently can reach V=15, with extension to at least V=18.5 when using the charge multiplication mode on the auto guider CCDTV. But for it's primary mission, high precision RV work, APF's high efficiency coupled with sub-\\ms\\ precision and dedicated nightly cadence abilities, should make it a valuable contributor over the coming years to exoplanet discovery and characterization." }, "1402/1402.4803_arXiv.txt": { "abstract": "Mergers of binary neutron stars (NSs) usually result in the formation of a hypermassive neutron star (HMNS). Whether- and when this remnant collapses to a black hole (BH) depends primarily on the equation of state and on angular momentum transport processes, both of which are uncertain. Here we show that the lifetime of the merger remnant may be directly imprinted in the radioactively powered \\emph{kilonova} emission following the merger. We employ axisymmetric, time-dependent hydrodynamic simulations of remnant accretion disks orbiting a HMNS of variable lifetime, and characterize the effect of this delay to BH formation on the disk wind ejecta. When BH formation is relatively prompt ($\\lesssim 100$ ms), outflows from the disk are sufficiently neutron rich to form heavy $r$-process elements, resulting in $\\sim$week-long emission with a spectral peak in the near-infrared (NIR), similar to that produced by the dynamical ejecta. In contrast, delayed BH formation allows neutrinos from the HMNS to raise the electron fraction in the polar direction to values such that potentially \\emph{Lanthanide-free} outflows are generated. The lower opacity would produce a brighter, bluer, and shorter-lived $\\sim$ day-long emission (a `blue bump') prior to the late NIR peak from the dynamical ejecta and equatorial wind. This new diagnostic of BH formation should be useful for events with a signal to noise lower than that required for direct detection of gravitational waveform signatures. ", "introduction": "Mergers of binary neutron stars (NSs) (hereafter `neutron star mergers', or NSMs) are the primary source of gravitational waves (GW) for upcoming ground-based interferometric detectors such as Advanced LIGO and Virgo (\\citealt{Abadie+10}). They are also promising central engines for short-duration gamma-ray bursts (GRBs; \\citealt{Paczynski86,Eichler+89}; see \\citealt{Berger13} for a recent review). General relativistic simulations of NSMs show that the merger process can result in two qualitatively different outcomes, depending primarily on the total mass of the binary $M_{\\rm t}$. If $M_{\\rm t}$ exceeds a critical value $M_{\\rm c}$, then the massive object produced by the merger collapses to a black hole (BH) on the dynamical time ($\\sim$few ~ms, e.g., \\citealt{sekiguchi2011}). On the other hand, if $M_{\\rm t} < M_{\\rm c}$ then the merger product is at least temporarily supported against gravitational collapse by differential rotation and/or thermal pressure. This meta-stable compact object is usually called a \\emph{hypermassive} NS (HMNS; e.g. \\citealt{kaplan2013}). The value of $M_{\\rm c}$ depends on the uncertain equation of state of (EoS) of nuclear matter. The recent discovery of massive $\\sim 2M_{\\odot}$ NSs (\\citealt{Demorest+10,Antoniadis+13}) excludes a soft EoS, placing a lower limit of $M_{\\rm c} \\gtrsim 2.6-2.8M_{\\odot}$ (\\citealt{Hotokezaka+13}; \\citealt{Bauswein+13}). It thus appears likely that the `canonical' 1.4 + 1.4 $M_{\\odot}$ binary merger goes through a HMNS phase. \\begin{figure*} \\includegraphics*[width=1.5\\columnwidth]{f1.eps} \\caption{ Relation between the observed kilonova and the properties of the ejecta that powers it. Material ejected dynamically in the equatorial plane is highly neutron rich ($Y_e < 0.1$), producing heavy r-process elements that include Lanthanides. This results in emission that peaks in the near-infrared and lasts for $\\sim 1$~week (`late red bump') due to the high opacity. Outflows from the remnant disk are more isotropic and also contribute to the kilonova. If the HMNS is long-lived, then neutrino irradiation can increase $Y_e$ to a high enough value ($Y_e\\sim 0.4$) that no Lanthanides are formed, resulting in emission peaking at optical wavelengths (`early blue bump'). If BH formation is prompt, outflows from the disk remain neutron rich, and their contribution is qualitatively similar to that of the dynamical ejecta. } \\label{f:cartoon} \\end{figure*} When a HMNS does form, its lifetime before collapsing into a BH depends on the timescale for thermal energy loss via neutrino emission (e.g., \\citealt{ruffert1999,Paschalidis+12,Galeazzi+13}) and the efficacy of angular momentum transport via gravitational waves and magnetohydrodynamic stresses (e.g., \\citealt{Duez+06,Stephens+08,Siegel+13}). For particularly nearby mergers, oscillations excited in the HMNS may be detectable in the GW strain data (\\citealt{Shibata05}; \\citealt{Bauswein+12}; \\citealt{Hotokezaka+13c}). However, the subsequent ring down phase following BH formation is unlikely to be detected by the initial generation of advanced detectors. Fortunately, NSMs are also accompanied by coincident electromagnetic (EM) signals that inform physical processes at work during the merger (e.g.~\\citealt{Metzger&Berger12}; \\citealt{Kelley+13}; \\citealt{Piran+13}). One such counterpart is a thermal IR/optical transient powered by the radioactive decay of heavy elements synthesized in the merger ejecta (a `kilonova'; \\citealt{Li&Paczynski98}; \\citealt{Metzger+10}; \\citealt{Roberts+11}; \\citealt{Goriely+11}; \\citealt{Piran+13}; \\citealt{Grossman+13}; \\citealt{Tanaka+14}). Kilonovae are particularly promising EM counterparts because (1) their generation is relatively robust, requiring only a modest amount of unbound ejecta; (2) their signal is independent of the existence of a dense surrounding external medium; and (3) unlike a GRB, kilonovae are relatively isotropic. A candidate kilonova was recently detected following the GRB 130603B (\\citealt{Tanvir+13}; \\citealt{Berger+13}; ). If the merger ejecta is sufficiently neutron-rich for $r$-process nucleosynthesis to reach the Lanthanides ($A \\gtrsim 139$), the optical opacity becomes much higher than that of iron-group elements \\citep{Kasen+13}, resulting in emission that is redder, dimmer, and more slowly evolving (\\citealt{Barnes&Kasen13,tanaka2013}). Although such unusually red colors may be beneficial in distinguishing NSM transients from unrelated astrophysical sources, the current lack of sensitive wide field infrared telescopes could make EM follow-up across the large sky error regions provided by Advanced LIGO/Virgo even more challenging (e.g.~\\citealt{Nissanke+13}; \\citealt{Metzger+13}; \\citealt{Hanna+13}; \\citealt{Kasliwal&Nissanke13}). The matter ejected dynamically following a NSM is likely to be sufficiently neutron-rich (as quantified by the electron fraction $Y_e~\\lesssim 0.3$) to produce a red kilonova (e.g., \\citealt{Rosswog05,Duez+09,bauswein2013}). Dynamical expulsion is not the only source of ejecta, however. A robust consequence of the merger process is the formation of a remnant torus surrounding the central HMNS. Outflows from this accretion disk over longer, viscous timescales also contribute to the merger ejecta (e.g.,~\\citealt{Surman+08,Metzger+08,Metzger+09a,Lee+09,Dessart+09,Wanajo&Janka12}). The more isotropic geometry of disk winds suggests that they may contribute a distinct component to the kilonova light curve for most viewing angles \\citep{Barnes&Kasen13, Grossman+13}. \\citet[ hereafter FM13]{FM13} calculated the viscous evolution of remnant BH accretion disks formed in NSMs using two-dimensional, time-dependent hydrodynamical simulations. Over several viscous times, FM13 found that a fraction $\\sim$several percent of the initial disk mass is ejected as a moderately neutron-rich wind ($Y_e \\sim 0.2$) powered by viscous heating and nuclear recombination. Although the higher entropy of the outflow as compared to the dynamical ejecta results in subtle differences in composition (e.g. a small quantity of helium), the disk outflows likely produce Lathanide elements with sufficient abundance to result in a similarly red kilonova as with the dynamical ejecta. FM13 included the effects of self-irradiation by neutrinos on the dynamics and composition of the disk. Due to the relatively low accretion rate and radiative efficiency at the time of the peak outflow, neutrino absorption had a sub-dominant contribution to the disk evolution. This hierarchy is important because a large neutrino flux tends to drive $Y_e$ to a value higher than that in the disk midplane (e.g.~\\citealt{Surman+08}; \\citealt{Metzger+08}; \\citealt{Surman+13}). If neutrino irradiation is sufficient to drive $Y_e \\gtrsim 0.3-0.4$, the nuclear composition of the disk outflows would be significantly altered, resulting in a distinct additional component visible in the kilonova emission. By ignoring the influence of a central HMNS, FM13 implicitly assumed a scenario in which BH formation was prompt or the HMNS lifetime very short. Here we extend the study of FM13 to include the effects of neutrino irradiation from a long-lived HMNS. As we will show, the much larger neutrino luminosity of the HMNS has a profound effect on the quantity and composition of the disk outflows, allowing a direct imprint of the HMNS lifetime on the kilonova (Figure~\\ref{f:cartoon}). As in FM13, our study includes many approximations that enable us to follow the secular evolution of the system. We focus here on exploring the main differences introduced by the presence of a HMNS, and leave more extensive parameter space studies or realistic computations for future work. The paper is organized as follows. In $\\S\\ref{s:model}$ we describe the numerical model employed. Our results are presented in $\\S\\ref{s:results}$, separated into dynamics of the outflow (\\S\\ref{s:evolution}) and composition (\\S\\ref{s:composition}). A summary and discussion follows in $\\S\\ref{s:discussion}$. Appendix~\\ref{s:neutrino_details} describes in more detail the upgrades to the neutrino physics implementation relative to that of FM13. ", "conclusions": "\\label{s:conclusions} \\label{s:discussion} We have explored the effects of a hypermassive neutron star (HMNS) on the long-term evolution of remnant accretion disks formed in neutron star binary mergers. Our main results can be summarized as follows: \\newline \\noindent 1. -- A long-lived HMNS results in the ejection of a significant fraction of the disk over a timescale of $\\sim 1$~s. The amount of mass increases monotonically with HMNS lifetime. This enhanced mass loss, up to a factor $\\gtrsim 10$ relative to prompt BH formation, results from enhanced neutrino heating and a reflecting inner boundary. \\newline \\noindent 2. -- The composition of the ejecta is latitude-dependent. Material within an angle $\\sim 30^\\circ$ of the polar axis is strongly irradiated by the HMNS. If the neutrino and antineutrino luminosities are similar (as we have assumed), the electron fraction of the outflow can be raised to values where Lanthanides are not produced. \\newline \\noindent 3. -- Material ejected equatorially is still expected to produce a strong r-process, similar to that of the dynamical ejecta, although the detailed composition is dependent on the HMNS lifetime. \\newline The criterion on the neutron star lifetime to appreciably change the electron fraction of the polar outflow ($t_{\\rm ns} \\gtrsim 100$ ms) is obviously satisfied if the remnant mass is below the maximum mass of a cold neutron star, or if the remnant is stabilized by solid body rotation (a {\\it supramassive} NS), the latter of which is only removed on much longer timescales via e.g. magnetic dipole spin-down. The $t_{\\rm ns} \\gtrsim 100$ ms condition is also likely satisfied if the HMNS is supported by thermal pressure (e.g.~\\citealt{Bauswein+10}; \\citealt{Paschalidis+12}; \\citealt{kaplan2013}), as the latter is removed on the HMNS cooling timescale, which is typically on the order of $\\sim$ seconds (e.g.~\\citealt{pons1999}). On the other hand, support via differential rotation may not last 100 ms, as the latter can be efficiently removed by magnetic stresses: the growth rate of the MRI is as short as milliseconds. Whether a blue bump indeed develops out of the polar outflow in the case of a long-lived HMNS will depend on a number of factors. First, enough mass needs to be ejected so that the contribution to the lightcurve from radioactive decay becomes detectable. Second, the level of irradiation and the ratio between neutrino and antineutrino luminosities must be sufficient to raise $Y_e$ to values close to $0.5$. The results of \\citet{Dessart+09}, who employ a much more realistic treatment of neutrino transport but did not include the viscous evolution, seem to align with our findings. The ejecta mass required for a detectable signal depends on how much radioactive heating is supplied by synthesized elements, which are lighter than the $r$-process nuclei but potentially more neutron-rich than $^{56}$Ni (e.g.~\\citealt{Grossman+13}). A final requirement for producing a blue bump is that the high $Y_e$ (Lanthanide-free) material must be ejected first, such that it resides {\\it exterior} to any lower $Y_e$ (Lanthanide-rich) material that would otherwise block its emission. Our calculations support this requirement as well: after rising quickly, $Y_e$ of the polar ejecta decreases monotonically with time (Fig.~\\ref{f:ye_average}). This is a direct consequence of the decrease in the neutrino luminosities with time, either smoothly as in the case of a long-lived HMNS (Appendix~\\ref{s:neutrino_details}), or via a sudden drop when a BH forms. Future work will explore in more detail the observational consequences of this bimodal outflow. Figure~\\ref{f:summary_plot} provides estimates of various quantities relevant to the kilonova emission derived from our calculations, as a function of the HMNS collapse time. The peak luminosities and time to peak are estimated using the `Arnett rule' for a kilonova (e.g., \\citealt{Li&Paczynski98,Metzger+10}), \\begin{eqnarray} \\label{eq:L_peak} L_{\\rm peak} & \\simeq & 4.3\\times 10^{41}\\textrm{ erg s}^{-1}\\left(\\frac{f}{3\\times 10^{-6}} \\right) \\left(\\frac{v_r}{0.1c} \\right)^{1/2}\\nonumber\\\\ &&\\qquad\\qquad\\qquad\\qquad \\times \\left(\\frac{M_{\\rm ej}}{0.01M_\\sun} \\right)^{1/2}\\kappa^{-1/2}\\\\ \\label{eq:t_peak} t_{\\rm peak} & \\simeq & 1.4\\textrm{ d } \\left(\\frac{v_r}{0.1c} \\right)^{-1/2} \\left(\\frac{M_{\\rm ej}}{0.01M_\\sun} \\right)^{1/2}\\kappa^{1/2}, \\end{eqnarray} where $f$ is a factor quantifying radioactive energy deposition, and $\\kappa$ is the opacity of the material in units of cm$^2$~g$^{-1}$. In making Figure~\\ref{f:summary_plot}, we have used $\\kappa = 10$~cm$^2$~g$^{-1}$ for the equatorial material (Lanthanide-dominated), and $\\kappa = 1$~cm$^2$~g$^{-1}$ for the polar material (Lanthanide-free), see \\citet{Kasen+13}. The velocity is the mass-flux weighted value at $r = 10^9$~cm considering only unbound material. \\begin{figure} \\begin{overpic}[width=\\columnwidth]{f6.eps} \\put(2.5,93.3){\\tiny $\\odot$} \\end{overpic} \\caption{Kilonova properties as a function of HMNS lifetime. Shown are the ejecta mass in unbound material (a), the mass-flux weighted velocity at $10^9$~cm (b), the mass-flux weighted electron fraction (c), the peak luminosity (d) (eq.~[\\ref{eq:L_peak}]), and the time to peak (e) (eq.~[\\ref{eq:t_peak}]). Red squares and blue dots indicate equatorial and polar material. The opacity is assumed to be Lanthanide-dominated for the former and Lanthanide-free for the latter. Note that these numbers underestimate the `red' component since we have not included contributions from the dynamical ejecta. } \\label{f:summary_plot} \\end{figure} Most of the resulting kilonova properties are a monotonic function of the HMNS lifetime. Peak luminosities increase from $\\sim 10^{40}$ to $\\sim 10^{41}$~erg~s$^{-1}$ for the red component, while the blue component, when present, is brighter by a factor $\\sim 2$. Similarly, the peak times for the red component range from about a week to a month, increasing with longer HMNS lifetime due to the larger ejected mass. The blue component can last from a few- to several days. While the blue component is faster with longer HMNS lifetime, the average velocity of the red component saturates at $\\sim 0.05c$. The large amount of ejecta mass found by our calculations suggests that outflows from the disk could easily overwhelm that from the dynamical ejecta. One implication of this result relates to the Galactic production of $r$-process elements. FM13 estimated that ejection of $\\sim 10\\%$ of the disk mass would contribute with $\\sim 20\\%$ of the production rate of elements with $A\\gtrsim 130$ assuming reasonable values for the disk mass and neutron star merger rate. The increase in the ejected fraction of the disk by a factor of several relative to the prompt BH case implies that disks with long-lived HMNS could become a dominant contribution to the galactic $r$-process element production (\\citealt{Freiburghaus+99}; \\citealt{Rosswog+13}; \\citealt{Piran+14}). This is the case even when the polar outflow is Lanthanide free (Fig~\\ref{f:histogram}). The large ejecta masses we infer for a moderately long-lived HMNS ($t_{\\rm ns} = 100$ ms) may also help alleviate the tension between large ejecta mass $\\sim 3\\times 10^{-2}M_{\\odot}$ required to fit the NIR excess observed following GRB 130603B with current models (\\citealt{tanaka2013}; \\citealt{Piran+14}), without the need to invoke less likely scenarios such as the merger of a NS with a low mass BH. If the long-lived HMNS is magnetized, its rotational spin-down could also power the excess X-rays observed following this event (\\citealt{Fong+14}; \\citealt{Metzger&Piro13}; \\citealt{Fan+13}). Given that early blue emission appears to require the presence of a long-lived HMNS, detection of such a component in future events provides a relatively clean way to rule out a NS-BH merger (although the {\\it absence} of early blue emission would not rule out a NS-NS merger). Our models include many approximations in order to make the evolution to a time $\\sim 10$ s computationally feasible. In addition, we have focused here on the key differences introduced by the HMNS applied to models with a particular choice of parameters. Much more work remains in order to make reliable predictions for kilonovae emission. In addition to a more extensive exploration of parameter space, models with realistic angular momentum - and neutrino transport will be needed." }, "1402/1402.5752_arXiv.txt": { "abstract": "We analyze the recent released $HST$/WFC3 IR images in the GOODS-N region to study the formation and evolution of Quiescent galaxies (QGs). After examining the reliability with artificial galaxies, we obtain the morphological parameters with S\\'ersic profile of 299 QGs and 1,083 star-forming galaxies (SFGs) at $z\\sim0.5$--3.0, finding the evolution of $r_\\mathrm{e}$ and $n$ of massive ($M_*\\geq10^{10.5}$~M$_\\sun$) QGs while weaker evolution of SFGs and less massive ($M_*<10^{10.5}$~M$_\\sun$) QGs. The regression of the size evolution of massive QGs follows $r_\\mathrm{e}\\propto (1+z)^{-\\alpha_{r_\\mathrm{e}}}$ with $\\alpha_{r_\\mathrm{e}}=1.06\\pm0.19$ (a factor of $\\sim2.2$ increase from $z\\sim2.5$ to $\\sim0.5$), which is consistent with the general picture of the significant size growth. For the further understanding of the evolution scenario, we study the evolution of S\\'ersic index, $n$, and find that of massive QGs to significantly evolve as $n\\propto(1+z)^{-\\alpha_n}$ with $\\alpha_n=0.74\\pm0.23$ ($n\\sim1$ at $z\\sim2.5$ to $n\\sim4$ at $z\\sim0.5$), while those of the other populations are unchanged ($n\\sim1$) over the redshift range. The results in the present study are consistent with both of observation and numerical simulations, where gas-poor minor merger is believed to be the main evolution scenario. By taking account of the connection with less massive QGs and SFGs, we discuss the formation and evolution of the massive QGs over{\\it ``Cosmic High Noon\"}, or the peak of star-formation in the universe. ", "introduction": "The study of the high-redshift (high-$z$) early-type galaxies (ETGs) provides us clues to understanding the formation and evolution of massive galaxies in the local universe. Their star-formation activity peaked during the cosmological epoch at $1 < z < 3$ (e.g., Dickinson et al.~\\citeyear{dickinson03}; Heavens et al.~\\citeyear{heavens04}; Papovich et al.~\\citeyear{papovich06}; Hopkins \\& Beacom~\\citeyear{hopkins06}) and galaxy morphologies have changed dramatically (Kajisawa \\& Yamada~\\citeyear{kajisawa06}). For galaxy sizes, many studies have corroborated that massive galaxies at high-$z$ were much smaller than local galaxies with comparable mass (Daddi et al.~\\citeyear{daddi05}; Trujillo et al.~\\citeyear{trujillo06},~\\citeyear{trujillo07}; Cimatti et al.~\\citeyear{cimatti08}; van~Dokkum et al.~\\citeyear{vandokkum08}; Akiyama et al.~\\citeyear{akiyama08}; Franx et al.~\\citeyear{franx08}; Szomoru et al.~\\citeyear{szomoru10},~\\citeyear{szomoru12}; van~der~Wel et al.~\\citeyear{vanderwel11}; Barro et al.~\\citeyear{barro13}). At a fixed stellar mass, ETGs are claimed to have been significantly compact at high-$z$ and have evolved with rapid increase of their effective radius by a factor of $\\sim$ 4 or even larger from $z\\sim 2$ (Buitrago et al.~\\citeyear{buitrago08}; Carrasco et al.~\\citeyear{carrasco10}) and by a factor $\\sim$ 2 from $z \\sim 1$ (van~der~Wel et al.~\\citeyear{vanderwel08}; Trujillo et al.~\\citeyear{trujillo11}). To reach the size of local ETGs, rapid and violent evolutions by major merger (Hopkins et al.~\\citeyear{hopkins09a}) or minor merger (Bezanson et al.~\\citeyear{bezanson09}; Naab et al.~\\citeyear{naab09}) have been demanded. Recent very deep infrared observations of high spatial resolution with the $Hubble\\ Space\\ Telescope\\ (HST)$ have shed light on morphological details and shapes of galaxies at high-$z$. Bruce et al.~(\\citeyear{bruce12}) studied over 200 massive galaxies at $1 < z < 3$ in the CANDELS-UDS field and found that these galaxies had much smaller size at a given mass than that of local ETGs. On the other hand, it has also been argued that the compact galaxies have apparent smaller effective radii because of low signal to noise ratio (S/N) (e.g., Ryan et al.~\\citeyear{ryan12}). The lack of the consideration for AGN component would also make the radius smaller (Yoshino \\& Ichikawa~\\citeyear{yoshino08}; Pierce et al.~\\citeyear{pierce10}). In addition, the best-fit morphological outputs with, for example, {\\ttfamily GALFIT} (Peng et al.~\\citeyear{peng02}), which is one of the most frequently used fitting codes for galaxy morphology, could be significantly changed with small differences of fitting inputs (e.g., initial guess, point spread function (PSF), weight image) and the image properties (e.g., size of postage stamp, sky background noise). Although {\\ttfamily GALFIT} are frequently used, it sometimes gives inappropriate results, mostly when used without careful considerations to image quality of galaxies and to the contamination by neighboring objects (H\\\"{a}u{\\ss}ler et al.~\\citeyear{haussler07}, hereafter H07; Barden et al.~\\citeyear{barden12}). Some previous studies (e.g., Trujillo et al.~\\citeyear{trujillo06}; H07; Carollo et al.~\\citeyear{carollo13}; Mosleh et al.~\\citeyear{mosleh13}) estimated the errors in effective radius, $r_\\mathrm{e}$, and S\\'ersic index, $n$, using artificial galaxies (AGs), and derived simple relations between the original and output values. Szomoru et al.~(\\citeyear{szomoru10}) contrived to compensate the faint extended wings of galaxies. They estimated the limit of surface brightness and fitted the S\\'ersic profile to the galaxy images above the surface brightness limit with {\\ttfamily GALFIT}. Then, they corrected the result $r_\\mathrm{e}$ by calculating the residual counts between the original and model images. Based on a careful study of the bias of image quality and PSF profiles, van~der~Wel et al.~(\\citeyear{vanderwel12}) presented global structural parameters of more than 100,000 galaxies in the CANDELS survey. Bruce et al.~(\\citeyear{bruce12}) applied {\\ttfamily GALFIT} to three-component fitting (bulge, disk, and central components) for high-$z$ galaxies with a careful attention to the background noise and PSF convolution. In addition, we should take account of the different analysis for local galaxies when comparing the morphological properties at high-$z$. The half-light radii of the SDSS local galaxies used in Shen et al.~(\\citeyear{shen03}, hereafter S03) were claimed to be underestimated (Guo et al.~\\citeyear{guo09}, hereafter G09; Simard et al.~\\citeyear{simard11}). The comparison of the size-stellar mass relations between the different definitions of stellar mass would also lead to inappropriate results (Mosleh et al.~\\citeyear{mosleh13}). The comparison of the structural parameters for high-$z$ galaxies with those in the local universe should be based on the consistent definition and analysis of galaxy data. In this paper, we investigate the reliability and limit of {\\ttfamily GALFIT} to obtain the morphological properties of high-$z$ galaxies. Then, using deep near infrared (NIR) observations with Wide Field Camera 3 (WFC3) instrument installed on $HST$, we apply {\\ttfamily GALFIT} to galaxies in the Great Observatories Origins Deep Surveys-North (GOODS-N) region, in which Ichikawa et al.\\ (\\citeyear{ichikawa12}) (hereafter Ic12) studied the size evolution of galaxies in a non-parametric way with $K_\\mathrm{s}$-band ground-based images of MOIRCS Deep Survey (MODS). The ground-based images were not reliable enough for the morphological study of galaxies at $z>1$ with {\\ttfamily GALFIT} (Konishi et al.~\\citeyear{konishi11}). As such, Ic12 obtained the size-stellar mass relations based on half- and 90 percent light radii. On the other hand, deep images by WFC3 with much higher spatial resolution will allow us to apply {\\ttfamily GALFIT} to high-$z$ galaxies, including compact galaxies, for the morphological study. An outline of the paper is as follows. In Section~\\ref{sec:sec2}, we describe the samples of massive galaxies in the GOODS-N. Using the background noise and PSF of WFC3 images, we make AGs with various shape parameters. We analyze them with {\\ttfamily GALFIT} and compare the results with the original parameters in Section~\\ref{sec:sec3}. We examine the reliability and the systematic errors of $r_\\mathrm{e}$ and $n$ of the AGs obtained with {\\ttfamily GALFIT} under some conditions. After examining the validity of the results, we apply the fitting method to massive galaxies in the GOODS-N. The results are described in Section~\\ref{sec:sec4} and Section~\\ref{sec:sec5}. Finally we discuss our results in comparison with those of previous parametric and non-parametric studies in Section~\\ref{sec:sec6}. Throughout this paper, we assume $\\Omega_m$ = 0.3, $\\Omega_\\mathrm{\\Lambda}$ = 0.7 and $H_0$ = 70 kms$^{-1}$Mpc$^{-1}$. We use the AB magnitude system (Oke \\& Gunn~\\citeyear{oke83}; Fukugita et al.~\\citeyear{fukugita96}). ", "conclusions": "\\label{sec:sec6} We have obtained the size and shape of the massive galaxies, using MODS and $HST$/WFC3 CANDELS data in the GOODS-N region. Thanks to the high image quality and depth, we were allowed to analyze as faint objects as with $H_\\mathrm{AUTO}\\leq25$, which reaches less massive ($\\sim10^{10}$~M$_{\\sun}$) SFGs at $z\\sim2.5$ (QGs at $z\\sim1.5$) and massive ($\\geq10^{10.5}$~M$_{\\sun}$) SFGs at $z\\sim3.0$ (QGs at $z\\sim2.5$). With a careful test of {\\ttfamily GALFIT} analysis, applying magnitude criteria to AGs, we obtained unbiased morphological results for the samples. The tests for different PSFs and color effect also improved the reliability. First, we discuss the evolution of $r_\\mathrm{e}$. As shown in Fig.~\\ref{fig:fig.6} and Fig.~\\ref{fig:fig.8}, \\textcolor{black}{ we found a number of compact QGs ($r_\\mathrm{e}\\sim1$~kpc or less) at $1.5 < z \\leq 2.5$, which is consistent with the previous studies (van~Dokkum et al.~\\citeyear{vandokkum09}; Szomoru et al.~\\citeyear{szomoru12}). } The size for massive QGs from $z\\sim2.5$ to $\\sim0.5$ is represented as $r_\\mathrm{e}\\propto (1+z)^{-\\alpha_{r_\\mathrm{e}}}$ with $\\alpha_{r_\\mathrm{e}} \\sim1.06$ (or a factor of $\\sim2.5$ increase from $z\\sim2.5$ to $\\sim0.5$ at a given stellar mass), which is consistent with the previous results of $\\alpha_{r_\\mathrm{e}} \\sim$ 0.7--1.5 (a factor of $\\sim$ 1.8--3.6 size increase). It is noted that SFGs have weaker size evolution ($\\alpha_{r_\\mathrm{e}} \\sim0.5$), irrespective of their stellar mass bins, which is inconsistent with previous studies (e.g., Mosleh et al.~\\citeyear{mosleh11}). However, the sample definition of SFGs in the studies are different ($UV$-bright galaxies in Mosleh et al.), and their results could not be compared with the present results. The scenario of the size evolution for QGs is in dispute for a decade both in theoretical (e.g., Fan et al.~\\citeyear{fan08}; Hopkins et al.~\\citeyear{hopkins09a}; Naab et al.~\\citeyear{naab09}; Oser et al.~\\citeyear{oser12}) and observational studies (Newman et al.~\\citeyear{newman10}; van~Dokkum et al.~\\citeyear{vandokkum10}). According to Eq.~4 in Naab et al.~(\\citeyear{naab09}), it takes $\\sim4~(\\sim2)$ minor merger events with 1:10 (3:10) mass ratio to explain the size evolution of the massive QGs (see also Bezanson et al.~\\citeyear{bezanson09}). In a merger scenario, however, about 10\\% of the massive galaxies at $z\\sim2$ are expected to have survived without \\textcolor{black}{ equal-mass merging (Hopkins et al.~\\citeyear{hopkins09a}), which leaves the superdense QGs morphologically unchanged to the local universe. } In the SDSS and other surveys for the local galaxies, however, there is not enough number of such a superdense relic (S03; Taylor et al.~\\citeyear{taylor10}; M\\'armol-Queralt\\'o et al.~\\citeyear{marmol12}). These facts suggest that the observed size evolution is hard to be explained only by the merger scenario (see also van~der~Wel et al.~\\citeyear{vanderwel09}; Barro et al.~\\citeyear{barro13}). Secondly, we discuss the shapes of the galaxies by using $n$. As the shape of galaxy is believed to exhibit the evolution trace, irrespective of the galaxy masses, we discuss the evolution of $n$. In Fig.~\\ref{fig:fig.10}, we see that the typical $n$ of QGs is larger than that of SFGs. There is also difference between the evolution of $n$ of massive and less massive QGs, though the latter are limited to $z\\leq1.5$ due to the completeness limit. We found that $n$ of massive QGs significantly evolved with $\\alpha_n\\sim0.74$ of $n\\propto (1+z)^{-\\alpha_n}$, while those of less massive QGs and SFGs were unchanged on average ($\\alpha_n\\sim0.17$ and $\\sim 0$, respectively). The findings of the evolution in $n$ would give us a clue to understanding the morphological evolution of high-$z$ galaxies. \\textcolor{black}{ One of the most favorable scenarios is the inside-out mass growth, where the mass accretion in the outer parts of compact galaxies increase $r_\\mathrm{e}$ and $n$ toward low-$z$ (van~Dokkum et al.~\\citeyear{vandokkum10}). In van~Dokkum et al., they showed the evolution of $n$ ($n\\sim4$ to $\\sim2$ from $z\\sim2$ to $\\sim0.5$), as well as the size evolution ($r_\\mathrm{e}\\sim3$~kpc to 8~kpc over the same redshift range). } In addition to the observational results, numerical simulation by Naab et al.~(\\citeyear{naab09}) explained the evolution of $n$ by minor merging of satellite galaxies. As done in Chevance et al.~(\\citeyear{chevance12}), the investigation for the distribution and evolution in of $b/a$ over the redshift range may help our understandings for the low-$n$ QGs, which could be the progenitor of local bulges (see also Trujillo, Carrasco \\& Ferr\\'e-Mateu~\\citeyear{trujillo12}), though the small sample in the present study would not give the robust conclusion. Based on the results of $r_\\mathrm{e}$ and $n$ in the present study, we then discuss the formation and evolution of massive QGs, taking account of the evolution of less massive QGs and SFGs as well. First, focusing on $n$, less massive SFGs at the whole redshift range ($0.5\\leq z\\leq2.5$) are thought to be the progenitor of the less massive QGs after exhausting their gas and passively evolve with little change in size and shape. In the context, the progenitors of comparatively less massive QGs at high-$z$ could be small or amorphous SFGs located at higher redshift ($z\\gtrsim3.0$) (Kajisawa \\& Yamada~\\citeyear{kajisawa01}), most of which are not included in the present study due to our detection limit. \\textcolor{black}{ While the sizes of QGs are smaller than those of SFGs at the whole redshift, it is believed that there are shrinkage of extended SFGs when they transform into QGs (Barro et al.~\\citeyear{barro13b}). The shrinkage of SFGs is also investigated by Dekel et al.~(\\citeyear{dekel13}), in which migration of star-forming clumps forms the blue compact galaxies (or blue nuggets), then quench into red nuggets (see also Noguchi~\\citeyear{noguchi99}). We need more detailed investigation whether the compact galaxies keep $n$ lower after the migration of star-forming clumps, while Williams et al.~(\\citeyear{williams14}) showed that blue nuggets had $\\langle n\\rangle \\sim$ 2-3 and red nuggets $\\langle n\\rangle \\sim$ 3-4. Whitaker et al.~(\\citeyear{whitaker12}) found very little difference in the sizes of young and old quenched galaxies. } After the birth of low-$n$ red QGs, in the later epoch ($z<1.5$) major or minor mergers of those dry (gas-poor) QGs dominates the evolution of massive QGs, enlarging their size rapidly over the cosmic time (Gao et al.~\\citeyear{gao04}). The dry merger is also consistent with the evolution of $n$ because it is efficiently change the light profile of galaxy (Barnes~\\citeyear{barnes92}; Hernquist~\\citeyear{hernquist92}; Naab et al.~\\citeyear{naab09}), while wet (gas-rich) merger would reproduce disk-like galaxies (Steinmetz \\& Navarro~\\citeyear{steinmetz02}; Springel \\& Hernquist~\\citeyear{springel05}; Robertson et al.~\\citeyear{robertson06}; Cox et al.~\\citeyear{cox06}). On the other hand, we found no significant evolution in $r_\\mathrm{e}$ and $n$ of less massive QGs. \\textcolor{black}{ To explain this, we might need to consider the environmental effect on their formation (or halo mass size; Dekel et al.~\\citeyear{dekel13}). However, we avoid discussing this because of the small sample and limited redshift range of the less massive QGs in the present study. } \\textcolor{black}{ It should be noted that the argument above (see also Carollo et al.~\\citeyear{carollo13}) is complicated by the fact that samples of massive galaxies, not separated into star-forming and quiescent, have been demonstrated to grow smoothly with time (e.g., van~Dokkum et al.~\\citeyear{vandokkum10}; Patel et al.~\\citeyear{patel13}). To corroborate the evolution scenario avobe, the discussion based on comoving number density for each group of galaxies would be helpful. Patel et al.~(\\citeyear{patel13}) studied the structural evolution of massive galaxies by linking progenitors and descendants at a constant cumulative number density. However, small number of samples in the present study would not be enough for the robust discussion in statistical sense. The field variance in a small field would also hamper the reliable conclusion. In addition, the present study investigated the evolution of $r_\\mathrm{e}$ at given two mass bins while those previous studies investigated the evolution taking account of the stellar mass evolution ({\\it differential size evolution}) by using the constant number density method or mass-normalized size. The mass-normalized size is useful not only to see the differential size evolution, but also to compensate the weak completeness, which could give rise to the biased (or spurious) size evolution. In Fig.~\\ref{fig:norm} we show the mass-normalized size evolution for QGs and SFGs. $r_\\mathrm{e,norm}$ is derived as $r_\\mathrm{e,norm}=r_\\mathrm{e}/(M_*^\\mathrm{cor}/M_\\mathrm{c})^{\\alpha_\\mathrm{M}}$, where ${\\alpha_\\mathrm{M}}$ is the linear slope of the size-stellar mass relation for each redshift bin (see Section~\\ref{sec:sec5.1}). It is noted that the relations are derived only using the sample within the completeness limits. Then, we see the size evolution with $r_\\mathrm{e}\\propto(1+z)^{-\\alpha_{r_\\mathrm{e}}}$ finding that $\\alpha_{r_\\mathrm{e}}=1.07\\pm0.15$ for QGs and $\\alpha_{r_\\mathrm{e}}=0.59\\pm0.05$ for SFGs, which is consistent with the results of massive QGs and SFGs at fixed mass bins. The results ensure our results of the size evolution over the redshift ranges, while we need more sample to investigate the weak evolution of $r_\\mathrm{e}$ (and $n$) of less massive QGs. } \\begin{figure} \\figurenum{12} \\plotone{norm.eps} \\caption{ Mass-normalized size evolution for QGs (top) and SFGs (bottom). Black dotted lines represent the regressions of $r_\\mathrm{e,norm}\\propto(1+z)^{-\\alpha_{r\\mathrm{e}}}$. } \\label{fig:norm} \\end{figure} Finally, we should refer to the scatters in size at whole redshift range comparable to that of the local populations (Fig.~\\ref{fig:fig.9}). This suggests that there had already been massive QGs at $z\\sim$ 1.5--3, whose morphological properties were similar to that of the local ones. This is encouraged by the fact that the velocity dispersion of QGs at $z\\sim1.6$ were found to be comparable to those of the local galaxies (Onodera et al.~\\citeyear{onodera12}). It should be noted, however, that a single S\\'ersic profile may not be appropriate to fit galaxies at high-$z$, even if they seem to be well-virialized. Bruce et al.~(\\citeyear{bruce12}) applied {\\ttfamily GALFIT} with two components (bulge, disk) massive galaxies ($M_*>10^{11}$~M$_{\\sun}$) at $11$, making it hard to find the \"best-fit\" results, while we adopted a single component S\\'ersic fit in the present study. \\textcolor{black}{ To discuss the origin of the scatters in size, we need further investigation with, for example, mass-normalized size, which is beyond the present study. }" }, "1402/1402.1470.txt": { "abstract": "{We present a {\\it Herschel} far-IR and sub-mm study of a sample of 120 galaxies in 28 Hickson compact groups. Fitting their UV to sub-mm spectral energy distributions with the model of da Cunha et al. (2008), we accurately estimate the dust masses, luminosities, and temperatures of the individual galaxies. We find that nearly half of the late-type galaxies in dynamically ``old'' groups, those with more than 25\\% of early-type members and redder UV-optical colours, also have significantly lower dust-to-stellar mass ratios compared to those of actively star-forming galaxies of the same mass found both in HCGs and in the field. Examining their dust-to-gas mass ratios, we conclude that dust was stripped out of these systems as a result of the gravitational and hydrodynamic interactions, experienced owing to previous encounters with other group members. About 40\\% of the early-type galaxies (mostly lenticulars), in dynamically ``old'' groups, display dust properties similar to those of the UV-optical red late-type galaxies. Given their stellar masses, star formation rates, and UV-optical colours, we suggest that red late-type and dusty lenticular galaxies represent transition populations between blue star-forming disk galaxies and quiescent early-type ellipticals. On the other hand, both the complete absence of any correlation between the dust and stellar masses of the dusty ellipticals and their enhanced star formation activity, suggest the increase in their gas and dust content due to accretion and merging. Our deep {\\it Herschel} observations also allow us to detect the presence of diffuse cold intragroup dust in 4 HCGs. We also find that the fraction of 250$\\mu$m emission that is located outside of the main bodies of both the red late-type galaxies and the dusty lenticulars, is 15-20\\% of their integrated emission at this band. All these findings are consistent with an evolutionary scenario in which gas dissipation, shocks, and turbulence, in addition to tidal interactions, shape the evolution of galaxies in compact groups. } ", "introduction": "%%%%%%%%%%%%%%%%%%%%%%% Since most galaxies are found within large structures, such as groups and clusters, studying how the environment can affect their properties (i.e. stellar populations and morphology) is crucial to understanding their evolution. It is now known that dynamical interactions and the merging of galaxies can affect both their morphologies (induced bars, bridges, tails, and other tidal distortions) and the star formation and nuclear activity \\citep[i.e.][]{Struck99}. Nevertheless, galaxy environments can differ depending on the number and the kinematic behaviour of their members. Compact groups of galaxies are systems of several galaxies that display high galaxy densities, similar to those found in the central regions of rich clusters, unlike clusters, they have much lower average velocity dispersions (\\citealt{Hickson97} quotes a $\\sigma \\sim$250 km/s). There are exceptions, such as Stephan's Quintet, where an intruder is colliding with the group with a velocity difference of over 800 km/s \\citep{Hickson92}. As a result, galaxies in compact groups can experience a series of strong tidal encounters with other group members during their lifetimes. For these reasons, compact groups are considered ideal systems for studying the effects of dense environments in galaxy evolution. To examine the properties of galaxies in compact groups we used the catalogue defined by \\citet{Hickson82}; it consists of 100 groups, containing 451 galaxies, in compact configurations (less than 5 arcmin), within relatively isolated regions where no excess of other surrounding galaxies can be seen. Even though all Hickson compact groups (HCGs) were selected as having four or more galaxies, the original sample was later reduced to 92 groups, since spectroscopic observations revealed the inclusion of interlopers among their members \\citep[see][]{Hickson92}. \\begin{figure*} \\begin{center} \\includegraphics[scale=0.50]{figure_example.png} \\caption{Spitzer/IRAC ``true'' colour image of HCG 07 (panel a). The blue channel traces the 3.6$\\mu$m emission, the green the 4.5$\\mu$m, and the red the 8.0$\\mu$m. The {\\it Herschel}/PACS 160$\\mu$m (panel b), and the {\\it Herschel}/SPIRE 250$\\mu$m (panel c) images of the same group. In the bottom left corners of panels b and c we include the beam sizes of the {\\it Herschel} instruments (11$''$ and 22$''$, respectively). In panel c) we also display for comparison the beams of the {\\it AKARI} 160$\\mu$m band (70$''$) and {\\it IRAS} 100$\\mu$m (1.2$'\\times$5$'$). } \\label{fig:fig_example} \\end{center} \\end{figure*} Detailed studies in the past three decades have revealed that HCGs occupy a unique position in the framework of galaxy evolution. They display an excess of elliptical galaxies (relative to the field), with their spiral galaxy fraction nearly half of what it is observed in the field (43\\%; \\citealt{Hickson82}). \\citet{Mendes94} showed that almost half of their galaxies display morphological features of interactions, and \\citet{Zepf93} revealed indications of mergers in the irregular isophotes of the elliptical galaxies. In addition, \\citet{Ponman96} found hot X-ray gas in $\\sim$75\\% of the groups, implying that their members reside in a common dark matter halo, while \\citet{Desjardins13} show that the X-ray luminosity increases as group HI-to-dynamical mass ratio decreases. However, the hot gas is not in hydrostatic equilibrium and, thus, these systems are not the low-mass analogues of rich groups or clusters. Single-dish radio measurements reveal that galaxies in compact groups are generally deficient in HI with a median mass two times less than what is observed in loose groups \\citep{Verdes01}. Furthermore, \\citet{MartinezBadenes12} show an increase in the conversion of HI-to-H$_{2}$ caused by the on-going tidal interactions in the spiral galaxies of groups found in early stages of evolution, followed by HI stripping and the decrease in molecular gas because of the lack of replenishment. Nearly 40\\% of HCG members for which nuclear spectroscopy has been obtained display evidence of active galactic nuclei \\citep{Shimada00, Martinez10}. Although tidal stripping is clearly a strong contender for explaining HI deficiencies, there are cases where the gas seems to have been almost completely stripped from the galaxies, most likely involving both hydrodynamic, as well as pure gravitational, processes. The best studied example is Stephan's Quintet (SQ = HCG92), where HI and molecular studies show that almost all of the cold ISM now resides in the intragroup medium \\citep{Williams02,Gao00,Guillard12}. Strong dissipative processes, in addition to gravity, are clearly at work on a large scale in this system, leading to extensive shock heating in the interfaces between galaxies and the intragroup medium, as is evident by the presence of hot X-ray gas, warm molecular hydrogen, and diffuse ionized carbon \\citep{Trinchieri05,O'Sullivan09,Appleton06,Cluver10,Appleton13}. While systems like SQ seem rare, they are not unique and may represent a rapid hydrodynamic phase in which head-on collisions between the group members can cause rapid sweeping of the gas (e. g. \\citealt{Appleton96} for the ring galaxy VII Zw 466) in addition to disruptive tides, conventional mergers, and gas depletion. For example, a recent study of HCG57 by Alatalo et al. (in preparation) also shows a highly disruptive collision between two galaxies in which the molecular and atomic gas is dramatically changed by direct collision. This may be similar to the Taffy system \\citep{Condon94}, where a head-on collision has pulled gas into a highly turbulent bridge between the galaxies, heating the molecular gas \\citep{Peterson12}. \\citet{Cluver13} studied the mid-IR spectra of 74 galaxies in 23 HCGs and discovered that more than 10\\% of them contain enhanced warm molecular hydrogen emission that, like SQ, may be indicative of shocks and turbulent heating of their ISM. Furthermore, these H$_{2}$ enhanced galaxies (called molecular hydrogen emission galaxies -- MOHEGs by \\citealt{Ogle10}) seem to primarily occupy a special place in mid-IR colour space referred to as the ``gap'' by \\citet{Johnson07}. Although not strictly a ``gap'' \\citep[see][]{Bitsakis11}, Johnson et al. as well as \\citet{Walker12} argue that there is a deficiency of galaxies in the {\\it Spitzer} IRAC colour-colour space defined primarily as lying in the range $-0.4$1) and strong polarisation ($\\sim$ 20-30 \\% at 1.4 GHz). Their origin is subject to debate and not yet understood. There is a general consensus that they are related to shock waves, occurring in the ICM during mergers. Shock waves should be able to amplify magnetic fields and accelerate electrons up to relativistic energies, hence producing synchrotron radio emission \\citep{Bruggen11,2012MNRAS.423.2781I,Vazza12,sk13}. Although this picture is roughly consistent with observations, some important issues remain unexplained. First, a radio relic is not always detected when a shock wave is present in the ICM, as revealed by X-ray observations \\citep{Russell11} and recently \\citet{2013MNRAS.433..812O} have found a displacement between the X-ray shock wave and the radio relic in 1RXS J0603.3+4214. Secondly, most of the relics do not appear to be co-located with a shock wave in X-ray observations (see review by \\citealt{Bruggen11} and ref. therein). Thirdly, shock waves in the ICM are characterised by Mach numbers of the order of 2 - 4 \\citep{sk08,va09shocks} so that their efficiency in accelerating particles is expected to be too low to account for the radio emission.\\\\ Some authors have proposed that shock waves by themselves are not sufficient to accelerate the particles from the thermal pool to relativistic energies. This would indicate that a seed population of relativistic electrons must be already present before the shock passage \\citep{KangRyu11,KangRyu12,Pinzke13}. This seed population of old radio plasma could be re-energised by a shock wave, explaining why the connection between a relic and a shock wave is not one to one. Since the old radio plasma is much more energetic than the thermal gas in the ICM, a low acceleration efficiency could be sufficient to re-accelerate the particles and power the radio emission. The problem is now to find the source for this seed population of relativistic electrons. Some authors have suggested that they come from a previous episode of shock acceleration (\\citealt{Macario11}). Another possibility is that the old plasma comes from the activity of radio galaxies.\\\\ Recent results by \\citet{VazzaBruggen13} have shown that the gamma-ray upper limits to the cluster emission from the {\\it Fermi} satellite are at odds with the hypothesis that radio relics are generated by Diffusive Shock Acceleration (DSA) mechanism, at least in the way it appears to work for supernova remnants. In fact, if shock waves accelerate electrons in the ICM, also protons should be accelerated, with a higher efficiency. In inelastic collisions with the thermal protons, relativistic protons produce pions and gamma-ray photons, which {\\it Fermi} should have detected.\\\\ \\noindent {\\bf Radio Halos}\\\\ Radio halos are diffuse Mpc-sized objects found in merging galaxy clusters \\citep[e.g.][]{Buote01,Cassano10} and characterised by steep spectra (see e.g. reviews by \\citealt{Ferrari08,Feretti12}). Two different classes of models have been proposed so far: the hadronic models \\citep{2010ApJ...722..737K,2011A&A...527A..99E} and the turbulent re-acceleration models \\citep{Fujita03,CassanoBrunetti05}. In this paper, we study the peculiar and complex emission of the cluster PLCKG287.0 +32.9 to test DSA models. PLCKG287.0 +32.9 is a massive cluster ($M_{500} \\sim 1.4 \\times 10^{15}$ solar masses, Planck Collaboration 2013) located at redshift $z=0.39$, showing a disturbed X-ray morphology (Bagchi et al. 2011) which indicates that the cluster is undergoing a merger event. The X-ray centre of the cluster has a Right Ascension equal to $11h50m51.02s$ and a declination of $-28d04'09.37\"$ PLCKG287.0 +32.9 is known to host two radio relics \\citep{Bagchi11}. Our new radio observations give a different picture of the system. The radio emission is more complex and more extended than previously observed, and offers a unique opportunity to unravel the origins of radio relics. The paper is organised as follows: In Sec. \\ref{radio} we present our new observations. The X-ray and optical data are discussed in Sec. \\ref{sec:xray} and \\ref{sec:optical}, respectively. In Sec. \\ref{radioanalysis} the radio properties are studied, and the spectral properties are presented in Sec. \\ref{sec:spix}. Results are discussed in \\ref{discussion} and we present our conclusions in Sec. \\ref{conclusions}.\\\\ Throughout this paper, we assume a concordance $\\rm{\\Lambda CDM}$ cosmological model, with $H_0=$ 71 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_M=$ 0.27, and $\\Omega_{\\Lambda}=$ 0.73. One arcmin corresponds to 316 kpc at $z=$0.39. \\begin{figure} \\vspace{90pt} \\begin{picture}(90,90) \\put(-20,0){\\includegraphics[width=9cm]{otticoradiox.eps}} % \\end{picture} \\caption{X-ray emission in red (XMM-Newton), radio emission at 323 MHz in blue (low resolution, beam FWHM $\\sim 22'' \\times 18''$) and in green (high resolution: beam FWHM$\\sim 13'' \\times 8''$).} \\label{fig:xray} \\end{figure} \\begin{table*} \\centering \\caption{Radio observations.} \\begin{tabular}{c c c c c c c c} \\hline Frequency & Radio telescope & Observing date & time & bandwidth & Int. time & rms noise & Beam \\\\ MHz & & & h & MHz & s & mJy/beam &\\\\ \\hline 325 & GMRT & Jan 2013 & \t 7 & 33 & 8 \t\t & 0.1 & 13$'' \\times 8''$ \\\\ 610 & GMRT & May 2013& 10 & 33 & 8\t\t & 0.065 & 7$'' \\times 5''$ \\\\ 150 & GMRT &\tJul 2011 & 8 & 16 & 8\t\t & 1.3 &25$'' \\times 18''$ \\\\ 3000 & JVLA &\tJun 2013 & 6\t & 2000 & 3 & 0.05 & 19$'' \\times 11''$ \\\\ \\hline \\multicolumn{7}{l}{\\scriptsize Col. 1: Observing Frequency, Col. 2: Radio telescope. Col. 3: Date of the observation. Col. 4: total observing time. }\\\\ \\multicolumn{7}{l}{\\scriptsize Col. 5: Observing bandwidth. Col 6: Integration time per visibility.}\\\\ \\multicolumn{7}{l}{\\scriptsize Col 7: 1$\\sigma$ rms noise reached in the high resolution images.}\\\\ \\multicolumn{7}{l}{\\scriptsize Col 8: FWHM of major and minor axes of the restoring beam in the high resolution images. }\\\\ \\end{tabular} \\label{tab:obs} \\end{table*} ", "conclusions": "\\label{conclusions} We have presented a multi-wavelength analysis of the galaxy cluster PLCKG287.0 +32.9, that might shed new light on the origin of radio relics. Our results can be summarised as follows: \\begin{itemize} \\item{Optical data suggests that PLCKG287.0 +32.9 is located within a 6-Mpc long intergalactic filament. The cluster has undergone a major merger (NWc -SEc), slightly misaligned with respect to the main direction of accretion. Along the filament two sub-clumps are detected. The galaxies along the intergalactic filament, as well as the two sub-clumps (SEext and NW), are likely on the process of accretion onto the main cluster.} \\item{Two radio relics and a radio halo are present in the cluster. Additional emission is detected NW of the relic. Another relic is located SE of the cluster, at a projected distance of 2.8 Mpc. A radio halo is located in the cluster center, and filamentary emission is detected around both relics.} \\item{The large projected distance of the SE relic and the small projected distance on the NW relic are interpreted as shock waves produces at different times during a minor merger event. A sub-cluster (with a mass $\\sim$ 0.1 of the main one) could have caused a first shock wave at his first infall onto the main cluster, and a second shock wave during the second core-passage (NW relic). The radio halo could be due to the major merger between the SEc and NEc sub-clusters.} \\item{The NW relic emission fades into the lobes of a radio galaxy, indicating that radio relics originate from electrons previously injected by AGN and reaccelerated, likely by a shock wave. The spectral index analysis supports this interpretation. However, it is necessary to assume that two or more radio sources have injected electrons into the pre-shock region to account for the relic size.} \\end{itemize}" }, "1402/1402.1171_arXiv.txt": { "abstract": "We present a generalized Newtonian description of particle dynamics valid for any spherically symmetric, static black hole spacetime. This approach is derived from the geodesic motion of test particles in the low-energy limit. It reproduces exactly the location of the marginally stable, marginally bound, and photon circular orbits; the radial dependence of the energy and angular momentum of circular orbits; parabolic motion; pericentre shift; and the spatial projection of general trajectories. As explicit examples of the new prescription, we apply it to the Schwarzschild, Schwarzschild--de Sitter, Reissner--Nordstr{\\\"o}m, Ay{\\'o}n-Beato--Garc{\\'i}a, and Kehagias--Sfetsos spacetimes. In all of these examples, the orbital and epicyclic frequencies are reproduced to better than $10\\%$. The resulting equations of motion can be implemented easily and efficiently within existing Newtonian frameworks. ", "introduction": "Since its birth almost 100 years ago, the theory of general relativity (GR) has successfully withstood continuous experimental testing: from the anomalous perihelion precession of Mercury and the bending of light rays from distant stars by the Sun to the indirect confirmation of the existence of gravitational waves from the orbital decay of the Hulse--Taylor pulsar \\citep{will06}. All of these examples, however, only probe the weak field limit of the theory while the strong field regime remains largely uncharted \\citep{psaltis}. One of the most dramatic consequences of the strong field limit of GR is the prediction of black holes and, although current astrophysical observations are consistent with their existence, the available evidence does not allow to discriminate between the objects predicted by GR and those coming from alternative gravitational theories \\citep[see, e.g.][]{barausse13,bambi13}. It is generally believed that astrophysical black holes possess a certain degree of intrinsic angular momentum, either since birth or as a result of subsequent accretion processes \\citep{narayan}. For this reason, it is expected that the spacetime around astrophysical black holes should be well approximated by the solution of a rotating black hole found by \\cite{kerr}. Nevertheless, many key relativistic features can already be investigated in the non-rotating, spherically symmetric black hole solution of Schwarzschild. Moreover, one of the first steps in the development of an alternative theory of gravity is to look for the kind of black hole solutions that it admits, and, given their simplicity, a prominent role in this search is played by static, spherically symmetric black hole spacetimes. The Schwarzschild black hole solution of GR is the best known example of a static and spherically symmetric spacetime, but there exist other metrics that might be relevant for different astrophysical scenarios, especially in the context of alternative theories of gravity and/or extra degrees of freedom for the central object. Examples include \\begin{itemize} \\item[{\\bf-}] The Reissner--Nordstr\\\"om solution describing a charged black hole within Einstein--Maxwell equations. For a discussion of particle motion in this spacetime see, e.g~\\cite{bicak89,pugliese,grunau}. \\item[{\\bf-}] The Schwarzschild--de Sitter spacetime describing a Schwarzschild black hole in an expanding universe with cosmological constant. See \\cite{stuchlik08,hackmann08b,hackmann08a} for a general discussion of the motion of test particles in this spacetime. \\item[{\\bf-}] The Kehagias--Sfetsos spacetime \\citep{kehagias}, a black hole solution of the Lorentz-violating gravitational theory of \\cite{horava}. The motion of test particles in this spacetime has been discussed in, e.g.~\\cite{abdu,enolskii,vieira}. \\item[{\\bf-}] The Boulware--Deser black hole solution in (4+1)-dimensional Gauss--Bonnet gravity \\citep{boulware}. \\item[{\\bf-}] Regular black holes. These are curvature singularity-free, exact solutions of Einstein equations coupled to some nonlinear electrodynamics \\citep[see, e.g.][]{bardeen68,ABG}. See \\cite{zhou,garcia} for a discussion of test particle motion in these spacetimes. \\item[{\\bf-}] A Schwarzschild black hole pierced by a cosmic string \\citep{aryal}. For a discussion of geodesic motion in this spacetime see, e.g.~\\cite{hackmann}. \\item[{\\bf-}] Schwarzschild and Reissner--Nordstr\\\"om black holes in spacetimes with higher dimensions \\citep{tangherlini,emparan}. For a general discussion of geodesic motion in these spacetimes see, e.g.~\\cite{hackmann08}. \\end{itemize} In this article we present a generalized Newtonian description of the motion of test particles in any given static, spherically symmetric spacetime. This work extends the scheme of \\cite{TR} (referred to as Paper~I in the following) where, by considering the low-energy limit of geodesic motion, we introduced an accurate Newtonian description of particle motion around a Schwarzschild black hole. The approach presented in Paper~I reproduces exactly certain key relativistic features of this spacetime (e.g.~radial location, angular momentum and energy of distinct circular orbits), while several other properties (e.g.~Keplerian and epicyclic frequencies) are described with a better accuracy than with commonly used pseudo-Newtonian potentials. Moreover, it showed a very good agreement with the thin disc accretion model of \\cite{novikov73} and the analytic model for relativistic accretion of \\cite{tejeda2,tejeda3}. The paper is organized as follows. In Section~\\ref{S2} we present the model and discuss the motion of general test particles within the new approach. In Section~\\ref{S3} we derive explicit expression for the equations of motion that can be easily implemented within existing Newtonian frameworks. Circular orbits are discussed in detail in Section~\\ref{S4} while perturbations away from them are considered in Section~\\ref{S5}. Finally, we summarize our results in Section~\\ref{S6}. ", "conclusions": "\\label{S6} We have presented a generalized Newtonian description of the motion of test particles valid for any static, spherically symmetric spacetime based on the low-energy limit of geodesic motion. This work constitutes a generalization of the approach developed in Paper~I for a Schwarzschild black hole. Among the various examples of spherically symmetric spacetimes that might be of relevance for different astrophysical contexts, we have considered the use of the new approach for a Schwarzschild black hole with cosmological constant (Schwarzschild--de Sitter spacetime), a charged black hole in Einstein--Maxwell theory (Reissner--Nordstr\\\"om spacetime), a charged regular black hole in Einstein theory coupled with a non-linear electrodynamics (Ay\\'on-Beato--Garc\\'ia spacetime), and a black hole solution of the Lorentz-violating gravitational theory of Ho{\\v r}ava-Lifshitz (Kehagias-Sfetsos spacetime). We have shown that the new approach reproduces exactly the radial location of the photon, marginally bound, and marginally stable circular orbits. Moreover, the radial dependence of the binding energy and angular momentum of circular orbits is also reproduced exactly. In addition, the spatial projection and the pericentre shift of general particle trajectories coincide exactly with the corresponding relativistic expressions. On the other hand, we have shown that, for the spacetimes considered in this paper, the Keplerian and epicyclic frequencies are reproduced to better than $10\\%$ for all of the radii that allow stable circular motion. We believe that the ability of the present approach to reproduce all of these key relativistic features simultaneously and, in addition, its ease of implementation within existing Newtonian frameworks, make it a simple, yet powerful tool for studying astrophysical accretion flows onto a broad family of black hole spacetimes. \\ack We would like to thank John C.~Miller and Zdenek Stuchl\\'ik for insightful discussions and critical comments on the manuscript. Part of this work was started during the conference Prague Synergy 2013. ET would like to thank the organizers for their hospitality and the participants for stimulating discussions. This work has been supported by the Swedish Research Council (VR) under the grant 621-2012-4870. \\begin{landscape} \\begin{table} \\caption{Application to examples of static and spherically symmetric black hole spacetimes. The function $\\alpha$ is connected with the spacetime metric as in \\eq{e1}. The ratio $\\Omega_*/\\Omega = \\Omega_*^\\parallel/\\Omega^\\parallel$ corresponds to the function $\\xi$ defined in \\eq{e22}. The second to last column gives the maximum percentage error obtained for $\\Omega_*$, $\\Omega_*^\\parallel$ and $\\Omega_*^\\perp$ (we have considered only values of $r$ for which circular motion is stable, see Figure~\\ref{f1}).} \\label{t1} \\begin{tabular}{llllp{5.7cm}} \\br\\\\[-7pt] Spacetime & $\\alpha$ & $\\frac{\\Omega_*}{\\Omega}$ & $\\frac{\\Omega-\\Omega_*}{\\Omega}$ & Notes \\\\[6pt]\\mr\\\\[-7pt] Schwarzschild & $1 - \\frac{ 2\\,\\rg }{ r }$ & $\\frac{r-2\\,\\rg}{\\sqrt{r(r-3\\,\\rg)}}$ & $\\leqslant 5.7\\%$ & $\\rg = \\G M/\\cc^2$ is the gravitational radius and $M$ is the mass of the black hole.\\\\ % Schwarzschild--de Sitter & $1 - \\frac{ 2\\,\\rg }{ r } - \\frac{\\Lambda\\,r^2}{3}$ & $\\frac{r-2\\,\\rg-\\Lambda\\,r^3/3}{\\sqrt{r(r-3\\,\\rg)}}$ & $\\leqslant 7.1\\%$ & $\\Lambda$ is the cosmological constant. The condition $\\Lambda\\,\\rg^2 \\leqslant 7.11\\times10^{-4}$ should be satisfied in order to have stable circular orbits. \\\\ % Reissner--Nordstr\\\"om & $ 1 - \\frac{ 2\\,\\rg }{ r } +\\frac{\\rQ^2}{r^2}$ & $\\frac{r^2-2\\,\\rg\\,r+\\rQ^2}{r\\sqrt{r^2-3\\,\\rg\\,r+2\\,\\rQ^2}}$ & $\\leqslant 8.2\\%$ & $\\rQ = \\sqrt{\\G}\\,Q/\\cc^2$ and $Q$ is the electrical charge of the black hole. To prevent the formation of a naked singularity, the condition $|\\rQ| < \\rg$ should be satisfied.\\\\ % Ay\\'on-Beato--Garc\\'ia & $ 1 -\\frac{ 2\\,\\rg \\,r^2}{\\left(r^2+\\rQ^2\\right)^{3/2} }+\\frac{ \\rQ^2r^2}{\\left(r^2+\\rQ^2\\right)^{2} }$ & $\\frac{1 -\\frac{ 2\\,\\rg \\,r^2}{\\left(r^2+\\rQ^2\\right)^{3/2} }+\\frac{ \\rQ^2r^2}{\\left(r^2+\\rQ^2\\right)^{2} }}{\\left[1 -\\frac{ 3\\,\\rg \\,r^4}{\\left(r^2+\\rQ^2\\right)^{5/2} }+\\frac{ 2\\,\\rQ^2r^4}{\\left(r^2+\\rQ^2\\right)^{3} }\\right]^{\\nicefrac{1}{2}}}$ & $\\leqslant 7.5\\%$ & We have considered values of $q\\leqslant q_\\s{\\text{crit}} \\simeq 0.634\\,\\rg$ in order to ensure a monotonic dependence of the ISCO on $q$. \\\\ % Kehagias-Sfetsos & $ 1 +\\omega\\,r^2\\left[ 1-\\left(1+\\frac{4\\,\\rg}{\\omega\\,r^3}\\right)^{\\nicefrac{1}{2}}\\right]$ & $\\frac{1 +\\omega\\,r^2\\left[ 1-\\left(1+\\frac{4\\,\\rg}{\\omega\\,r^3}\\right)^{\\nicefrac{1}{2}}\\right]}{\\left[1-\\frac{3\\,\\rg}{r}\\left(1+\\frac{4\\,\\rg}{\\omega\\,r^3}\\right)^{-\\nicefrac{1}{2}}\\right]^{\\nicefrac{1}{2}}}$ & $\\leqslant 6.3\\%$ & $\\omega$ gives the deviation away from GR (which is recovered in the limit \\mbox{$\\omega\\rightarrow \\infty$}). The condition $\\omega\\,\\rg^2\\geqslant 1/2$ should be satisfied to prevent the emergence of a naked singularity.\\\\ \\br \\end{tabular} \\end{table} \\end{landscape}" }, "1402/1402.6798_arXiv.txt": { "abstract": "We explore the inflationary phase of a scalar field with a kinetic term non-minimally coupled to gravity. We find that one of the slow-roll conditions is naturally consequence of the equation of motion of the scalar field. Thus, slow-roll conditions impose fewer constraints on potentials than other inflationary models. Moreover, it is demonstrated that the inflationary phase can be described by just one slow-roll parameter. By investigating the metric perturbations, it is shown that except for one potential, almost all potentials have the same pattern in the ($n_{s}$, $r$) plane. We provide an exact solution for the exceptional case. The exact solution represents the condensed scalar field and results in an accelerated expansion. \\pacs{98.80.Cq} ", "introduction": "INTRODUCTION} Observation of the cosmic microwave background and large scale structure are consistent with slow-roll inflation paradigm in which a scalar field rolls slowly down its potential \\cite{plank-data}.\\\\ To have slow-roll inflationary phase, certain slow-roll conditions are considered that lead to some constraints on parameters of a model \\cite{weinberg}. For example, consider a simple theory of inflation described by the following Lagrangian \\begin{equation}\\label{simple} \\mathcal{L_{s}}=\\sqrt{-g}\\left[\\frac{R}{2\\kappa^{2}}-\\frac{1}{2}g^{\\mu\\nu}\\partial_{\\mu}\\phi\\partial_{\\nu}\\phi-V(\\phi)\\right], \\end{equation} where $\\kappa^{2}=8\\pi G$, in the flat Friedmann-Robertson-Walker metric with signature $(-,+,+,+)$ and $a$ as the scale factor. There exist two \\emph{independent} slow-roll conditions. The first of these conditions is actually the definition of inflationary phase states that $\\dot{H}/H^{2}\\equiv\\epsilon\\ll1$. Since the evolution of the scalar field is given by a second-order differential equation, it is not clear that the first condition holds over an extended period. So, the second condition is required in which we \\emph{demand} that $\\ddot{\\phi}\\ll H\\dot{\\phi}$.\\\\ Also, to determine the details of models from observations, various slow-roll parameters are used such as $\\epsilon, \\eta\\equiv\\dot{\\epsilon}/H\\epsilon, \\chi\\equiv\\dot{\\eta}/H\\eta $. The parameters are used to determined the specific potential term in \\eqref{simple} for the inflation era \\cite{plank-data}. Since the shape of the potential term is important to study high energy physics, the values of these parameters connect cosmology to particle physics.\\\\ In this paper we introduce a model for which we need just one parameter, $\\epsilon$, to describe slow-roll inflation phase. The model is given by the following action \\begin{equation}\\label{0-1} S=\\int d^{4}x\\sqrt{-g}\\left[\\frac{R}{2\\kappa^{2}}-g^{\\mu\\nu}\\partial_{\\mu}\\varphi\\partial_{\\nu}\\varphi++\\frac{1}{2}\\alpha^{2}G^{\\mu\\nu}\\partial_{\\mu}\\varphi\\partial_{\\nu}\\varphi-V(\\varphi) \\right], \\end{equation} where $G^{\\mu\\nu}$ is the Einstein's tensor and $\\alpha$ is an inverse mass parameter. We have chosen $+$ sign for the second term because for the de Sitter space we have $G_{\\mu\\nu}\\propto-g_{\\mu\\nu}$, so the scalar filed in \\eqref{0-1} has the same dynamics as the scalar filed in \\eqref{simple} in the de Sitter space.\\\\ The third term is one of the operators of the Horndeski's scalar-tensor theory \\cite{Horndeski}. Thus, in this model we confront with second-order differential equations. C. Germani et al studied $V(\\varphi)=\\lambda\\varphi^4$ for \"the new higgs inflation\" \\cite{higgs,higgse} and extended their model for other cases in \\cite{higgs-extend}. It has been shown that a new type of inflationary phase arises when the third term dominates over the standard kinetic term \\cite{higgs,higgs-extend}. So in this paper we will focus on this type of inflation and we will neglect the second term in the action \\eqref{0-1}.\\\\ Compared to the other works \\cite{higgs,higgs-extend}, we have two new results. The first result is that for $V(\\varphi)=M^{6}/\\varphi^2$ background equations can be solve exactly and metric perturbations will be obtained. The second result is to show that up to the first order in $\\epsilon$, in contrast to the simple model of inflation \\eqref{simple}, the values of measurable quantities of the inflationary regime of this model are almost independent from the form of the potential term in \\eqref{0-1}, as is shown in Fig. 1. Also, the model gives values for the parameters which are agreement with the Planck data in \\cite{plank-data}.\\\\ The organization of this paper is as follows: in Sec. II an exact solution for a special potential is given and the metric perturbations of the solution is discussed. we show that the scalar field is condensed and results in accelerated expansion. Sec. III is devoted to study the background and the metric perturbations of general potentials. We show that many features of the exact solution are similar to other potentials. We summarize our findings and discuss about the results in Sec. IV. In Appendix the second order actions for the metric perturbations, in which $V(\\varphi)\\neq0$, is given. ", "conclusions": "" }, "1402/1402.6525.txt": { "abstract": "{In protoplanetary disks, the inner boundary between the turbulent and laminar regions could be a promising site for planet formation, thanks to the trapping of solids at the boundary itself or in vortices generated by the Rossby wave instability. %% Activity at this %% radius will also influence the thermodynamic structure of much of the %% dead-zone (including features such as the ice line). At the interface, the disk thermodynamics and the turbulent dynamics are entwined because of the importance of turbulent dissipation and thermal ionization. Numerical models of the boundary, however, have neglected the thermodynamics, and thus miss a part of the physics.} {The aim of this paper is to numerically investigate the interplay between thermodynamics and dynamics in the inner regions of protoplanetary disks by properly accounting for turbulent heating and the dependence of the resistivity on the local temperature.} {Using the Godunov code RAMSES, we performed a series of 3D global numerical simulations of protoplanetary disks in the cylindrical limit, including turbulent heating and a simple prescription for radiative cooling.} {We find that waves excited by the turbulence significantly heat the dead zone, and we subsequently provide a simple theoretical framework for estimating the wave heating and consequent temperature profile. In addition, our simulations reveal that the dead-zone inner edge can propagate outward into the dead zone, before staling at a critical radius that can be estimated from a mean-field model. The engine driving the propagation is in fact density wave heating close to the interface. A pressure maximum appears at the interface in all simulations, and we note the emergence of the Rossby wave instability in simulations with extended azimuth.} {Our simulations illustrate the complex interplay between thermodynamics and turbulent dynamics in the inner regions of protoplanetary disks. They also reveal how important activity at the dead-zone interface can be for the dead-zone thermodynamic structure.} ", "introduction": " ", "conclusions": "" }, "1402/1402.6090_arXiv.txt": { "abstract": "{We combined new near-infrared VLT/HAWK-I data of the globular clusters (GCs) in the isolated edge-on S0 galaxy \\object{NGC\\,3115} with optical and spectroscopic ones taken from the literature, with the aim of analyzing the multiband GC color distributions. A recent study from the SLUGGS survey has shown that the GCs in this galaxy follow a bimodal distribution of Ca II triplet indices. Thus, \\object{NGC\\,3115} presents a critical example of a GC system with multiple, distinct, metallicity subpopulations, and this may argue against the ``projection'' scenario, which posits that the ubiquitous color bimodality mainly results from nonlinearities in the color-metallicity relations. Using optical, NIR, and spectroscopic data, we found strong and consistent evidence of index bimodality, which independently confirms the metallicity bimodality in \\object{NGC\\,3115} GCs. At the same time, we also found evidence for some color--color nonlinearity. Taken in the broader context of previous studies, the multicolor consistency of the GC bimodality in \\object{NGC\\,3115} suggests that in cases where GC systems exhibit clear differences between their optical and optical--NIR color distributions (as in some giant ellipticals), the apparent inconsistencies most likely result from nonlinearities in the color--metallicity relations.} ", "introduction": "The study of globular cluster (GC) systems in galaxies is one of the keystones for understanding the processes at the base of the formation and evolution of galaxies \\citep{ashman92,brodie06}. Recently, the interpretation of one of the most intriguing properties of GC systems in early-type galaxies, the nearly universal presence of two distinct peaks in the optical color distribution, has inspired a vigorous and prolific debate \\citep[][]{yoon06,kundu07,yoon11b,chiessantos12,blake12a,usher12}. The importance of GC bimodality was recognized before it was a commonly observed property in early-type galaxies \\citep[ETGs,][]{schweizer87}. Historically, the bimodal GC color in optical bands has been equated to metallicity ([Fe/H]) bimodality, implying a fundamental constraint on GC and galaxy formation scenarios. Metallicity bimodality requires two distinct epochs or mechanisms of formation, or both, for the blue (metal-poor) and red (metal-rich) GC subpopulations. There are various proposed explanations for the GC color bimodality in ETGs: {\\it dissipational} merging of spirals, in which a merger-formed population of red, metal-rich GCs is assumed to appear distinct from the blue, metal-poor GCs of the progenitor spirals \\citep[][]{ashman92}; the {\\it dry} hierarchical assembly, which begins with a massive ``seed'' ETG that has a unimodal metal-rich GC distribution, and in which it is possible to produce a bimodal metallicity distribution through dissipationless accretion of many early-type dwarfs \\citep[][]{cote98}; and the {\\it insitu} formation scenario \\citep{forbes97}. Most of these proposed mechanisms, though, have assumed a simple linear conversion between [Fe/H] and color, which seemed justified from the small fractional age variations among the GCs \\citep[][]{cohen98,kuntschner02,puzia05}. However, this assumption became the subject of debate when three independent works, using observations and stellar population models, pointed out non-negligible nonlinearities in the color-metallicity relations of GCs \\citep{peng06,richtler06,yoon06}. In particular, Yoon and colleagues and \\citeauthor{richtler06} demonstrated that these nonlinearities naturally produce bimodal color histograms from nonbimodal [Fe/H] distributions. This interpretation, dubbed the projection effect, provided an alternative explanation based on stellar evolution for the ubiquity of bimodal GC color distributions. In this regard, \\citet{cantiello07d} suggested the use of multicolor GC histograms to verify the consistency of [Fe/H] distributions derived from different colors. These authors highlighted the role of optical to near-infrared (NIR) colors to distinguish between genuine bimodality in [Fe/H] and projected bimodality in color. If the nonlinear projection is at work, then the [Fe/H] distributions inferred from linear inversion of different color indices for the same GC sample will show some degree of inconsistency (discordant [Fe/H] peaks and/or fractions of GCs in each [Fe/H] component). The analysis of optical-to-NIR GC colors in various ETGs, as well as $u$ to $z$ photometry in some Virgo cluster members, indicates that the nonlinear projection effect is present at some level in these galaxies \\citep{blake12a,chiessantos12,yoon11b,yoon13}. At the same time, the results of the SLUGGS survey\\footnote{\\url{http://sluggs.ucolick.org.}}, which collected spectra of $\\sim1000$ GCs in 11 galaxies and derived [Fe/H] from the calcium II triplet index, CaT, support true [Fe/H] bimodality in at least some galaxies in addition to the \\object{Milky Way}. \\citet{usher12} found evidence for bimodal CaT distributions in six of eight galaxies with sufficient numbers of GC spectra. Nevertheless, the spectroscopically and photometrically derived [Fe/H] distributions show non-negligible differences in several galaxies of the SLUGGS sample, thus lacking the aforementioned multi-index coherence. However, the case of \\object{\\object{NGC\\,3115}}, an isolated lenticular galaxy at a distance of $\\sim10$~Mpc \\citep[][]{tonry01}, revealed highly consistent [Fe/H] and color distributions, leading the authors to present this galaxy as a critical test of [Fe/H] bimodality \\citep{brodie12}. Previous optical VLT FORS2 spectroscopy for 17 GCs in \\object{NGC\\,3115} showed hints of both a bimodal metallicity distribution \\textit{and} color or spectral index nonlinearity \\citep[][]{kuntschner02}, but it was limited by sample size. In this Letter, we combine new NIR photometry with literature data to investigate the consistency of the bimodality in the optical, NIR, optical-NIR colors, and CaT, of the GCs in \\object{NGC\\,3115}. ", "conclusions": "Our analysis of the GC system in \\object{NGC\\,3115} provides additional definitive proof of GC [Fe/H] bimodality in this S0 galaxy, the first to be firmly established beyond the \\object{Milky Way}. Various studies have noted that to rule out the possibility that nonlinearities project a nonbimodal [Fe/H] distribution into a bimodal color distribution, one must recover consistent [Fe/H] distributions from multiple different photometric indices. In particular, \\citet{cantiello07d} showed that optical-to-NIR colors are the most useful in constraining the underlying metallicities. This is in part because the broad color baselines imply a lower sensitivity to the detailed shape of the index-metallicity relation; for the converse reason, bimodal [Fe/H] distributions may not be evident in the purely NIR colors. The present study confirms these expectations: $i)$ the optical and optical-to-NIR colors are clearly bimodal and consistent with each other in terms of proportions of red and blue GCs, and $ii)$ the pure NIR distributions give ambiguous results, in the sense that different colors and/or statistical indicators indicate the presence or lack of bimodality. The comparison of GC colors with SSP models confirms earlier results \\citep{brodie12}, using CaT as a proxy for metallicity, which derived a bimodal [Fe/H] distribution with a dip at $-1.3\\lsim [Fe/H] \\lsim -0.3$ dex. The model comparison also suggests that the metal-rich GCs have slightly younger ages than the blue/metal-poor ones. Although this conclusion is subject to model uncertainties, the size of the age difference is consistent with that found between the metal-rich and metal-poor GCs in the \\object{Milky Way} \\citep{vandenberg13}. Some previous studies of optical-to-NIR GC colors in giant ellipticals have found significantly different optical versus optical-to-NIR color distributions. For instance, using HST ACS and WFC3/IR data, \\citet{blake12a} showed that while the optical color distributions of the Fornax giant elliptical \\object{NGC\\,1399} are clearly bimodal \\citep[supporting earlier results from][]{forte05}, the $V{-}H$ and $I{-}H$ distributions are not, or they imply significantly \\textit{different} bimodal breakdowns than found for the optical alone. Similarly, \\citet{chiessantos12} studied the $gzK$ color distributions of 14 elliptical galaxies, mainly in the Virgo cluster, and reported that double-peaked color distributions are more common in $g{-}z$ than in the optical-NIR colors. Both studies found significant nonlinearity between the purely optical and optical-to-NIR colors, and both concluded that bimodal optical color distributions are not necessarily indicative of underlying bimodality in metallicity. The lack of consistency between the purely optical and optical--NIR colors for \\object{NGC\\,1399} and some Virgo ellipticals contrasts strongly with the coherent color bimodality observed in \\object{NGC\\,3115}, for which the metallicities are also clearly bimodal. In light of this contrast, and with the evidence for nonlinear color--color relations, the inconsistent optical and optical-NIR color bimodalities in some galaxies imply that nonlinearities do indeed play an important role in shaping the GC color distributions in those galaxies. Thus, while our multicolor photometric analysis confirms the [Fe/H] bimodality of \\object{NGC\\,3115} GCs, we conclude that the nonlinear projection effect remains a viable explanation for the \\textit{ubiquity} of optical color bimodality and is the most likely cause in cases where the optical-NIR colors lack obvious bimodality. In summary: $i)$ optical and optical-to-NIR colors and CaT indices of GCs in \\object{NGC\\,3115} are bimodal; $ii)$ the bimodal distributions derived for different photometric and spectroscopic indices show good consistency; $iii)$ evidence for bimodality is weak or absent for purely NIR colors; $iv)$ our results agree with model predictions for GC systems with truly bimodal [Fe/H] distributions, which provides definitive proof of [Fe/H] bimodality in \\object{NGC\\,3115}, perhaps the first galaxy beyond the Local Group for which this is the case; $v)$ comparison with SSP models confirms earlier metallicity results based on CaT indices; $vi)$ despite the consistency of the color distributions, we also observe color-color nonlinearities, most clearly in the case of $i{-}H$ versus $g{-}z$. Thus, the metallicity distributions of extragalactic GC systems and, more specifically, the existence or lack of a universal bimodality of [Fe/H] in ETGs, remains a matter of debate. Indirectly, when taken in the broader context of previous work, our analysis indicates that optical GC color bimodalities have different causes in different galaxies, with nonlinear color-metallicity relations playing an important role for some previously studied giant ellipticals. Finally, our study shows the effectiveness of optical-to-NIR colors as an unambiguous test for underlying metallicity bimodality." }, "1402/1402.6329_arXiv.txt": { "abstract": "We present observations of the Auriga-California Molecular Cloud (AMC) at 3.6, 4.5, 5.8, 8.0, 24, 70 and 160 \\micron\\ observed with the IRAC and MIPS detectors as part of the {\\it Spitzer} Gould Belt Legacy Survey. The total mapped areas are 2.5 deg$^2$ with IRAC and 10.47 deg$^2$ with MIPS. This giant molecular cloud is one of two in the nearby Gould Belt of star-forming regions, the other being the Orion A Molecular Cloud (OMC). We compare source counts, colors and magnitudes in our observed region to a subset of the SWIRE data that was processed through our pipeline. Using color-magnitude and color-color diagrams, we find evidence for a substantial population of 166 young stellar objects (YSOs) in the cloud, many of which were previously unknown. Most of this population is concentrated around the \\lkha 101 cluster and the filament extending from it. We present a quantitative description of the degree of clustering and discuss the fraction of YSOs in the region with disks relative to an estimate of the diskless YSO population. Although the AMC is similar in mass, size and distance to the OMC, it is forming about 15 -- 20 times fewer stars. ", "introduction": "\\label{sec:intro} The cycle 4 {\\it Spitzer Space Telescope} Legacy project ``The Gould Belt: Star Formation in the Solar Neighborhood'' (PID: 30574; PI: L.E. Allen) completed the {\\it Spitzer} survey of the large, nearby star-forming regions begun by the c2d Legacy Project \\citep{Evans2003, Evans2009}. The cloud with the least prior study included in the survey is the cloud we have designated as ``Auriga'' which lies on the Perseus-Auriga border. This cloud has also been designated the California Molecular Cloud by \\cite{Ladaetal2009} since it extends from the California Nebula in the west to the \\lkha 101 region and associated NGC 1529 cloud in the east. We adopt the name Auriga-California Molecular Cloud (AMC) to encompass both nomenclatures. Despite the AMC's proximity to two of the most well-examined star-forming clouds, Taurus-Auriga and Perseus, it is a relatively unstudied region. Several dark nebulae were noted along its length by \\cite{Lynds1962}, and CO associated with many Lynds objects was measured by \\cite{Ungerechts1987}, who note the presence of a CO ``cloud extending from the California nebula (NGC 1499) in Perseus along NGC 1579 and \\lkha 101 well into Auriga'' (their cloud 12). Only very recently has a giant molecular cloud been unambiguously associated with the series of Lynds nebulae through high resolution extinction maps by \\cite{Ladaetal2009} who placed its distance firmly within the Gould Belt (GB) at $450 \\pm 23$ pc. At this distance, the cloud's extent of 80 pc and mass of $\\sim10^5 \\ M_{\\odot}$ rivals that of the Orion Molecular Cloud (L1641) for the most massive in the Gould Belt. For the remainder of this paper, we adopt this distance of 450 pc for the entire AMC. This is consistent with the distance of $510^{+100}_{-40}$ pc found by \\citep{Wolketal2010} on their study of \\lkha 101 with Chandra. We note that this distance differs from that adopted by \\cite{Gutermuthetal2009} for \\lkha 101 of 700 pc. We have mapped a significant fraction of the AMC with the Infrared Array Camera (IRAC; \\citealt{Fazio2004}) and the Mid-Infrared Photometer for \\Spitzer\\ (MIPS; \\citealt{Rieke2004}) on board the \\textit{Spitzer Space Telescope} \\citep{Werner2004}, with a total overlapping coverage of 2.5 deg$^2$ in the four IRAC bands (3.6, 4.5, 5.8 and 8.0 \\micron) and 10.47 deg$^2$ in the three MIPS bands (24, 70, and 160 \\micron). The mapped areas are not all contiguous and were chosen to include the areas with \\Av~$> 3$, as given by the \\cite{Dobashi2005} extinction maps. The goal of these observations is to identify and characterize the young stellar object (YSO) and substellar object populations. The data presented here are the first mid-IR census of the YSO population in this region. The area around \\lkha 101 and its associated cluster was observed as part of a survey of 36 clusters within 1 kpc of the Sun with \\Spitzer\\ by \\cite{Gutermuthetal2009} and those data have been incorporated into our dataset through the c2d pipeline. More recently, the AMC has been observed by the {\\it Herschel Space Observatory} at 70 -- 500 \\micron, and by the Caltech Submillimeter Observatory with the Bolocam 1.1 mm camera \\citep{Harveyetal2013}. These observations characterize the diffuse dust emission and the cooler Class 0 and Class I objects which can be bright in the far-IR. We do not analyze the large scale structure of the cloud in this paper as \\cite{Harveyetal2013} present such an analysis with the \\Herschel~observations, which are more contiguous and have a higher resolution than our MIPS observations. \\cite{Harveyetal2013} also include a comparison to these MIPS data and so further analysis is not required here. We describe the observations and data reduction (briefly as it is well-documented elsewhere) in $\\S$ \\ref{sec:obs}. In $\\S$ \\ref{sec:yso}, we describe the source statistics, the criteria for identifying and classifying YSO candidates and we compare the YSO population to other clouds. The SEDs and disk properties of YSOs are modeled in $\\S$ \\ref{sec:sed}. We characterize the spatial distribution of YSOs in $\\S$ \\ref{sec:spatial} and summarize our findings in $\\S$ \\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} We observed the AMC with IRAC and MIPS aboard the \\textit{Spitzer Space Telescope} and identify 138 YSOs in the cloud. As our IRAC coverage is segmented, we complemented our more contiguous MIPS coverage with WISE data to further eliminate galaxies from the sample, leaving 28 MIPS-only YSOs remaining, bringing the total number of YSOs in the AMC to 166. We classified the YSOs based on the spectral slope of their SEDs between 2 \\micron\\ and 24 \\micron\\ and find 37 Class I objects, 21 Class F objects (flat spectrum sources), 91 Class II objects, and 17 Class III objects. The high fraction of Class Is and Class Fs suggests that the AMC is relatively unevolved compared to other star-forming clouds. Despite the similarity in cloud properties between the AMC and the OMC, there is a distinct difference in the star formation properties. The star formation in the AMC is also concentrated along its filament, however, it is also forming a factor of about 20 fewer stars than the OMC. \\cite{Ladaetal2009} find that there is much less material at high density in the AMC than in the OMC and attribute the difference in star formation to this. Further studies of the star formation and YSO population in the AMC are needed to highlight the differences of the two clouds given their similar age. We modelled the SEDs of the Class II and Class III sources and their excesses by first fitting a K7 stellar spectrum to the optical and near-IR fluxes. The spectrum is normalized to the 2MASS flux (or the IRAC1 flux when 2MASS is unavailable) and we use an \\Av~value to match the spectrum of the stellar model to the de-reddened observed optical fluxes. An A0 stellar spectrum is used in the eight cases where a K7 spectrum is unable to provide a reasonable fit. Fitting a stellar spectrum allows us to measure the disk luminosities and characterize the excess. The excesses of the Class II and Class III sources were further parameterized by \\lambdaturnoff, the longest wavelength before an excess greater that 80\\% is measured, and \\alphaexcess, the slope of the SED at wavelengths longward of \\lambdaturnoff. \\lambdaturnoff\\ is a useful tracer for the proximity of dust to the star and consequently we identify fourteen classical transition disk candidates. The bulk of the star formation in the AMC is in the southern region of the cloud. We included a clustering analysis to quantify the densest areas of star formation and to identify groups within the cloud. We find four groups with 10 or more members all in the region around \\lkha 101 and its adjoining filament. We find three smaller groups with 5 -- 9 members scattered throughout the cloud. The largest group is that around \\lkha 101 and contains 49 members. We note that there are likely even more YSOs in this group since our YSO identification criteria of S/N $\\geq 3$ in IRAC1-4 and MIPS1 are difficult to attain in this bright region." }, "1402/1402.1215_arXiv.txt": { "abstract": "\\begin{changeb} The design of an interstellar microwave digital communication system at interstellar distances is considered, with relevance to starships and extraterrestrial civilizations (SETI and METI). Distances are large and thus require large transmitted power and/or large antennas, while the atmospheric/interstellar microwave is transparent over a wide bandwidth. Recognizing the known tradeoff between wider signal bandwidth and lower energy, a reduced requirement for energy delivered to the receiver at the expense of wide bandwidth is advantageous. In addition, wide bandwidth results in significant simplifications to the design and implementation. It is shown that appropriate transmit signal design can circumvent dispersion and scattering arising in the interstellar medium and motion, obviating any related processing in transmitter or receiver. The concept of interstellar coherence hole is introduced, capturing the time and frequency coherence constraint. Because other impairments can be circumvented, the fundamental limit on the energy delivered to the receiver for each bit of information is determined by cosmic microwave background alone. Designs that exchange larger bandwidth for extreme simplicity (recognizing the lack of coordination) and approaching the fundamental energy limit are identified. The Morse code invented for the telegraph in 1836 comes closer to this ideal than most approaches used in modern terrestrial radio by mapping a single bit onto a carrier burst. Rather than the terrestrial approach of adding phases and amplitudes to increase information capacity while minimizing bandwidth, adding multiple locations for carrier bursts in time and frequency increases information capacity while minimizing energy per information bit. The resulting location code can approach the fundamental energy limit as bandwidth is expanded, and is consistent with blind discovery with straightforward modifications to existing SETI pattern recognition algorithms. A discovery search can detect individual carrier bursts absent detailed knowledge of the signal structure, and observations of the interstellar medium and motion by transmitter and receiver desirably constrain the parameterization of these bursts. \\end{changeb} ", "introduction": "There has been considerable experience in communication with ''deep space'' probes in and around our solar system at radio wavelengths \\citeref{660}, and there is growing interest in optical \\citeref{661} as well. Communication of information at much greater distances, such as with starships and extraterrestrial civilizations, introduces new challenges, and has not been addressed either theoretically or empirically. Some of those challenges are addressed here from the perspective of communication engineering, emphasizing radio (rather than optical) wavelengths. \\begin{changea} The insights here are highly relevant to both the search for extraterrestrial intelligence (SETI) by informing the discovery of information-bearing signals, and messaging for extraterrestrial intelligence (METI), which transmits information-bearing signals. \\end{changea} \\begin{changec} They are also relevant to the design of two-way links with interstellar spacecraft (often called starships). \\end{changec} \\doctable {acronyms} {Acronyms} {l p{6cm}} { \\bfseries Acronym & \\bfseries Definition \\\\ \\hline AWGN & Additive white Gaussian noise \\\\ CMB & Cosmic background radiation \\\\ CSP & Coding and signal processing \\\\ ICH & Interstellar coherence hole \\\\ ISM & Interstellar medium \\\\ SETI & Search for interstellar intelligence \\\\ \\hline } \\begin{changea} This context is illustrated in Fig. \\ref{fig:blockDiagram}, showing the subsystems of an end-to-end system communicating information by radio. This paper concerns the coding and signal processing (CSP) subsystem. It is presumed that a message input is represented digitally (composed of discrete symbols), in which case it can always be represented as a stream of bits. The transmit-end CSP inputs information bits and outputs a baseband waveform \\begin{changec} (the spectrum of which is concentrated about d.c.). \\end{changec} The radio subsystem \\begin{changec} (including modulation to passband, centered about a carrier frequency $f_c$, \\end{changec} a radio-frequency transmitter, transmit/receive antennas, and demodulation \\begin{changec} back to baseband) \\end{changec} delivers a replica of this baseband waveform to the receive-end CSP with impairments (such as distortion and noise) introduced by the physical environment (the interstellar propagation and motion). The modeling of these impairments, which profoundly affect the CSP, is summarized here and addressed in greater depth elsewhere \\citeref{664}. \\incfig {blockDiagram} {1} {The major functional blocks in an interstellar communication system, with transmit on the top row and receive on the bottom row. These blocks are separated into (from left to right) information, baseband processing, passband radio, and physical propagation. } The CSP realizes the essential function of mapping information bits into a continuous-time baseband waveform suitable for transmission as an electromagnetic wave. The radio subsystem attempts to minimize impairments, for example by reducing noise introduced in the receiver. Through the choice of antennas, in the context of a given transmission distance it determines the radiated power necessary to deliver a needed level of power at receive baseband. The remaining performance characteristics of the system are determined by the CSP, including the fidelity of the message replica delivered by the receive-end CSP and the resources consumed in achieving this fidelity. The fidelity is usually measured by probability of error, and the primary resources of interest are the bandwidth of the radiated signal and the signal power that must be delivered to receive baseband. These resources can be manipulated over many orders of magnitude through the design of the CSP, and thus it is the subsystem with the greatest opportunity to manipulate the message fidelity and resources. A complete messaging system combines the expertise of several disciplines, including astronomy and astrophysics to model the physical, radio astronomy and electronics to design the radio, and applied statistics to design the CSP. The CSP has been the target of extensive research and practice in the context of commercial terrestrial wireless systems, and that experience is directly applicable to the interstellar case. The goal here is to communicate to a wider audience the opportunities afforded by the CSP. Since the knowledge and techniques applied in the CSP may be unfamiliar to many readers, the emphasis here is on reviewing results and developing intuition, with details documented elsewhere \\citeref{656}. One of the tradeoffs determined by the CSP is between bandwidth and received power; inevitably, increasing either allows the other to be reduced. \\end{changea} We argue that in communication at the large interstellar distances, large propagation losses place a premium on minimizing energy requirements. Other challenges applying specifically to communication with other civilizations are dispersive effects arising in the interstellar medium (ISM) at distances greater than hundreds of light years, relative motion, a lack of prior design coordination between the two ends, and the challenges of discovering a signal with unknown design and parameterization. It turns out that these considerations are closely linked, with both energy minimization and interstellar impairments contributing beneficially to both an implicit form of coordination between transmitter and receiver, as well as informing a discovery search. \\begin{changea} In the case of starships, relativistic effects also become critical due to large relative velocities \\citeref{665}. \\end{changea} ", "conclusions": "End-to-end design of a digital communication system at interstellar distances has been considered, with an emphasis on minimizing the energy per bit or energy per message delivered to the receiver. A solution with all the desired properties is identified. \\begin{changeb} A location code is conceptually simple, and a signal of this type is straightforward to discover at the receiver using a strategy based on detection of individual energy bursts followed by pattern recognition. \\end{changeb} This design can also approach the bandwidth-unconstrained limit on received energy, even in the presence of scintillation (when it is combined with time diversity). It does, however, require a significant bandwidth expansion. A similar approach can also achieve energy-efficient information-free beacons \\citeref{656}. Perhaps most significantly, the fundamental principle of minimizing received energy without bandwidth constraint and in consideration of jointly observable ISM and motion impairments leads to a simple and highly-constrained signal design. One can hope that the transmitter and receiver designers addressing this joint challenge without the benefit of coordination might arrive at compatible conclusions. Interstellar impairments are fortuitous for uncoordinated communication since they constrain the signal structure and parameterization. The tighter constraints at lower carrier frequencies and at radio as opposed to optical wavelengths are also helpful in this regard, because they more tightly constrain the signal without adversely impacting the energy requirements. It is reasonable to ask two skeptical questions. First, is it likely that another civilization is aware of the opportunity to reduce received energy, and aware of the fundamental limit on received energy? The history of earthbound communications leads to optimism. Energy-limited communication near the fundamental limit was understood much earlier (in the 1950`s) than bandwidth-limited communication (in the 1990`s) because of the simplicity of the solution. Optical communication is typically much simpler than radio precisely because bandwidth has never been considered a limiting resource at the shorter optical wavelengths. Second, is another civilization likely to be motivated by and act upon the opportunity to reduce the received energy? This is a more difficult question, since a more advanced civilization may well have tapped into cheaper sources of energy. Even if so, they may be motivated by the simplicity of the solutions and the benefits of that simplicity to implicit coordination. In addition, even if energy is more plentiful, there are many beneficial ways to consume more energy other than deliberate inefficiency. They could increase message length, reduce the message transmission time, transmit in more directions simultaneously, or transmit to greater distances. Overall it is unlikely that a civilization would use more energy than necessary unless for some reason they consider a reduction in bandwidth to be a higher priority. For communication with a starship, trading for greater bandwidth remains useful as a way to minimize received energy, which is a particular benefit on the downlink from the starship because of the reduction in transmit power and/or transmit antenna size. Because there is the luxury of joint design of transmitter and receiver, simplicity is not a goal in itself and if desired the bandwidth expansion can be reduced without a significant penalty in received energy using something considerably more complex than the location code." }, "1402/1402.0814_arXiv.txt": { "abstract": "{ The treatment of radiation transport in global circulation models (GCMs) is crucial for correctly describing Earth and exoplanet atmospheric dynamics processes. The two-stream approximation and correlated-$k$ method are currently state-of-the-art approximations applied in both Earth and hot Jupiter GCM radiation schemes to facilitate the rapid calculation of fluxes and heating rates. Their accuracy have been tested extensively for Earth-like conditions, but verification of the methods' applicability to hot Jupiter-like conditions is lacking in the literature. We are adapting the UK Met Office GCM, the Unified Model (UM), for the study of hot Jupiters, and present in this work the adaptation of the Edwards--Slingo radiation scheme based on the two-stream approximation and the correlated-$k$ method. We discuss the calculation of absorption coefficients from high-temperature line lists and highlight the large uncertainty in the pressure-broadened line widths. We compare fluxes and heating rates obtained with our adapted scheme to more accurate discrete ordinate (DO) line-by-line (LbL) calculations ignoring scattering effects. We find that, in most cases, errors stay below $\\SI{10}{\\percent}$ for both heating rates and fluxes using $\\sim 10$ $k$-coefficients in each band and a diffusivity factor $D = 1.66$. The two-stream approximation and the correlated-$k$ method both contribute non-negligibly to the total error. We also find that using band-averaged absorption coefficients, which have previously been used in radiative-hydrodynamical simulations of a hot Jupiter, may yield errors of $\\sim \\SI{100}{\\percent}$, and should thus be used with caution.} ", "introduction": "For Earth's atmosphere, irradiation from the Sun is the primary source of energy. Any model of the Earth's atmosphere therefore needs a robust and accurate treatment of radiation transport. Global circulation models (GCMs) of the Earth are used for both weather prediction and climate research, and they include a dynamical core that solves some variant of the Navier--Stokes equations and a radiation scheme that calculates the radiative heating rate. The dynamical cores are tested using benchmarks \\cite[see e.g.][]{Held1994,Reed2011,Ullrich2013}, and both dynamical cores\\footnote{See \\url{http://earthsystemcog.org/projects/dcmip-2012/}} and radiation schemes \\citep{Ellingson1991,Collins2006,Oreopoulos2012} are tested through intercomparison projects. GCMs have also been successfully applied to other solar system planets such as Jupiter, Saturn, Mars and Venus \\citep[see for example][respectively]{Yamazaki2004, Muller-Wodarg2006, Hollingsworth2010, Lebonnois2011}. In the past decade, GCMs have been used to study large-scale circulation on hot Jupiters \\citep{Showman2002,Showman2009,Rauscher2012,Cho2008,Thrastarson2010,Dobbs-Dixon2008}, a class of extrasolar planets which are approximately the size of Jupiter but orbit less than $\\SI{0.1}{\\astronomicalunit}$ from their parent star. These planets, thought to have tidally locked circular orbits due to strong tidal interactions between the planet and its parent star \\citep{Baraffe2010}, experience intense irradiation yielding a significant temperature contrast between the (permanent) day-side and night-side. Winds in the atmosphere of these planets are therefore expected to transport heat from the day-side to the night-side. Atmospheric properties of hot Jupiters are obtained by various observational techniques, mainly transmission spectroscopy and secondary eclipse measurements. Brightness maps \\citep{Knutson2007,Majeau2012} and wind velocities \\citep{Snellen2010} are now accessible, and constraints on the composition are becoming available, but large uncertainties remain. Observations indicate a hotspot shifted eastward of the substellar point \\citep{Knutson2007,Majeau2012} and temperature contrasts smaller than what is expected for these planets without winds \\citep{Knutson2007,Knutson2009}, indicating transport of heat from the day-side to the night-side \\citep{Watkins2010,Perez-Becker2013}. HD~209458b appears to have a temperature inversion in its upper atmosphere \\citep{Knutson2008,Burrows2007} while HD~189733b does not \\citep{Charbonneau2008,Barman2008,Knutson2009}, indicating that, despite similar orbital properties, hot Jupiters may have very different circulation patterns that still need to be understood. GCMs are therefore very valuable when trying to understand the increasing amount of observations of these systems. Benchmarking of the dynamical cores of GCMs applied to hot Jupiters has been used to investigate stability of the codes and discrepancies between them \\citep{Heng2011,Menou2009,Bending2013,Mayne2013a,Mayne2013b,Polichtchouk2014}. These benchmarks, and early GCMs applied to hot Jupiters, used simple, parametrised radiation schemes termed ``Newtonian cooling'' or ``temperature forcing'', where the temperature is relaxed towards assumed equilibrium pressure--temperature ($P$--$T$) profiles on a given timescale \\citep{Showman2002}. Pressure--temperature profiles and timescales can be estimated using one-dimensional time-dependent radiative transfer calculations \\citep{Iro2005}, though such an approach has flaws: (i) The equilibrium $P$--$T$ profiles used in the forcing may have a limited accuracy, (ii) radiative timescales may also have a limited accuracy and will vary in a non-trivial way as a function of latitude, longitude and depth, (iii) the forcing parametrisation itself may not be physically realistic, though the use of time-averaged equilibrium states when analysing model results may make this less of an issue, (iv) the model flexibility is poor since for each new planet modelled, the forcing must be changed. Later studies used more complicated schemes such as flux-limited diffusion \\citep{Dobbs-Dixon2008} and the two-stream approximation \\citep{Showman2009,Rauscher2012,Dobbs-Dixon2013}. For the opacity treatment, grey schemes \\citep{Rauscher2012}, binning and averaging of the absorption coefficients \\citep{Dobbs-Dixon2013} and the correlated-$k$ method \\citep{Showman2009} have been used. The correlated-$k$ method has also been used for retrieval analysis and characterisation of hot Jupiter atmospheres \\citep{Irwin2008} and to model brown dwarf atmospheres \\citep{Burrows1997}. Brown dwarfs atmospheres have many similarities with hot Jupiter atmospheres (e.g. temperature range and composition), but local conditions are very different due to the strong irradiation from the parent stars for hot Jupiters. There is a notable lack of analysis of the accuracy of these schemes when applied to hot Jupiter-like atmospheres and of details on how opacities have been calculated from line lists, preventing rigorous comparison with results previously published in the literature. These are serious shortcomings in a field of research which develops quickly and will deliver more and more accurate data requiring reliable tools for their interpretation. Both the two-stream approximation and the correlated-$k$ method \\citep[see e.g.][]{Thomas2002} are widely used in GCM simulations of the Earth, and the literature on the methods' applicability to the Earth atmosphere and their accuracy is extensive \\citep[see e.g.][]{Toon1989,Meador1980,Zdunkowski1980,Goody1989,Lacis1991,Mlawer1997}. They have both been found to yield results with satisfactory accuracy when comparing to more accurate solutions obtained from e.g. discrete ordinate (DO), line-by-line (LbL) calculations and when different schemes are compared through intercomparison projects \\citep{Ellingson1991,Collins2006,Oreopoulos2012}. They are, however, still under investigation \\citep{Goldblatt2009} and are still one of the limiting factors of the accuracy of both weather prediction and climate modelling. We are currently adapting the UK Met Office GCM, the Unified Model (UM), to hot Jupiter-like conditions. The main advantage of this model is its dynamical core, which solves the full 3D Euler equations and is coupled to a radiation scheme based on the correlated-$k$ method and two-stream approximation. This GCM is state-of-the-art for both Earth and hot Jupiter atmospheric dynamics modelling. In previous papers we have tested and confirmed the dynamical core's suitability to model hot Jupiter-like atmospheres \\citep{Mayne2013a,Mayne2013b}. This paper presents the adaptation of the radiation scheme, the Edwards--Slingo (ES) radiation scheme \\citep{Edwards1996a}, to conditions prevailing in hot Jupiter atmospheres. Opacities from high-temperature line lists have been calculated for the dominant absorbers in these atmospheres and the radiation scheme, using $k$-coefficients calculated from these opacities, has been tested against more accurate DO LbL calculations. Observations of absorbing and scattering species in hot Jupiter atmospheres have so far been limited to the detection of molecular absorbers \\citep{Huitson2013,Wakeford2013,Tinetti2007}, with some observations suggesting Rayleigh/Mie scattering clouds \\citep{Pont2008,Sing2013}. Due to the large uncertainties related to scatterers in hot Jupiter atmospheres and the complexity it adds to radiation transport, we limit the discussions in this paper to purely absorbing atmospheres and postpone the inclusion of scattering to a future work. The motivation for the present work is the lack of accurate tests and analysis of these radiation schemes now widely used by the community. This work will help in the implementation of similar schemes in the future and provide some guidelines for further progress. In \\cref{sec:opacities} we first discuss our opacity data including the high-temperature line lists we use and the calculation of cross-sections from these line lists by using estimates for the line widths. In \\cref{sec:ES} we briefly summarise the implementation of the correlated-$k$ method and two-stream approximation in the UM and move on to testing this scheme for gradually more complicated hot Jupiter-like atmospheres in \\cref{sec:testing}. These tests will be useful to other groups for comparison and benchmark purposes. Our conclusions follow in \\cref{sec:conclusions}. Combination of the adapted dynamical core and radiation schemes is in progress and will be presented in a future work, which will result in a state-of-the-art GCM that can be applied to a variety of exoplanet atmospheres. ", "conclusions": "\\label{sec:conclusions} The accuracy of radiation schemes used in GCMs has been studied extensively for Earth-like conditions, but detailed analysis for hot Jupiter-like conditions are lacking. In this paper we have analysed the accuracy and uncertainties in state-of-the-art radiation schemes used in several GCMs applied to hot Jupiters. Opacity sources and calculation of absorption coefficients from high-temperature line lists have been discussed. We present a line profile cut-off scheme that decreases the computation time required to calculate absorption coefficients by a factor of $\\sim 100$ compared to other methods used in the literature, while still giving accurate results. Both the two-stream approximation and correlated-$k$ method's applicability to hot Jupiter atmospheres have been analysed by comparing the Edwards--Slingo radiation scheme to discrete ordinate line-by-line calculations. The ES radiation scheme's performance in these tests shows that we have successfully adapted it to hot Jupiter-like atmospheres. Our main conclusions are: \\begin{itemize} \\item Pressure broadening parameters for high-temperature molecular lines are very uncertain and usually extrapolated from room temperature and pressure and small quantum numbers. Improvements in this area will become important as higher accuracy will be required to analyse results from future exoplanet characterisation projects (e.g. JWST, EChO, Sphere and ELT). \\item A diffusivity factor of $D = 1.66$, already widely used in both Earth and hot Jupiter GCMs, yields the smallest errors from the two-stream approximation, although $D = \\sqrt{3} \\approx 1.73$ is only slightly less accurate. \\item About $10$ $k$-coefficients in each band for molecular line absorption yield satisfactory accuracy. Using $\\sim 100$ $k$-coefficients per band does improve the overall accuracy, but errors decrease by less than $\\SI{50}{\\percent}$, while the radiative transfer computation time increases by a factor of $10$. We therefore choose to adopt the former as a balance between accuracy and computational cost. \\item Both the two-stream approximation and the correlated-$k$ method contribute non-negligibly to the total error, with overall heating rate errors of $\\lesssim \\SI{10}{\\percent}$ in regions with significant heating/cooling. Flux errors are similar or smaller. \\item Whether a black-body spectrum, solar spectrum or uniform (in wavenumber) weighting scheme is used has little effect on the overall accuracy given the band structure used here (\\cref{tbl:bands}). We therefore choose to adopt a uniform weighting scheme, enabling the use of the same $k$-coefficients in both the thermal and stellar spectral regions and for different irradiation spectra. \\item Using a mean absorption coefficient in each band, as in \\citet{Dobbs-Dixon2013}, yields inaccurate fluxes and heating rates for molecular absorption. Heating rate errors can reach $\\SI{100}{\\percent}$ or more, even in regions with significant heating. Band-averaged absorption coefficients should thus be used with caution. \\end{itemize} Any radiation scheme applied to hot Jupiters should be checked against the tests we have presented here. These tests and the detailed descriptions of our methods and approximations will be useful for future adaptation of radiation schemes in other GCMs. Current observational constraints on exoplanets do not require the level of accuracy we have applied in this work. The field develops at an amazing pace, however, and modellers should now develop the best theoretical and numerical tools to tackle the challenges posed by the increasing accuracy expected from future large observational projects." }, "1402/1402.2485_arXiv.txt": { "abstract": "In order to determine the composition of the dust in the circumstellar envelopes of oxygen-rich asymptotic giant branch (AGB) stars we have computed a grid of {\\sc modust} radiative-transfer models for a range of dust compositions, mass-loss rates, dust shell inner radii and stellar parameters. We compare the resulting colours with the observed oxygen-rich AGB stars from the SAGE-Spec Large Magellanic Cloud (LMC) sample, finding good overall agreement for stars with a mid-infrared excess. We use these models to fit a sample of 37 O-rich AGB stars in the LMC with optically thin circumstellar envelopes, for which 5--35-\\mum {\\em Spitzer} infrared spectrograph (IRS) spectra and broadband photometry from the optical to the mid-infrared are available. From the modelling, we find mass-loss rates in the range $\\sim 8\\times10^{-8}$ to $5\\times10^{-6}$ M$_{\\odot}\\ \\mathrm{yr}^{-1}$, and we show that a grain mixture consisting primarily of amorphous silicates, with contributions from amorphous alumina and metallic iron provides a good fit to the observed spectra. Furthermore, we show from dust models that the {\\em AKARI} [11]--[15] versus [3.2]--[7] colour-colour diagram, is able to determine the fractional abundance of alumina in O-rich AGB stars. ", "introduction": "Stars on the asymptotic giant branch (AGB) lose a significant fraction of their matter through slow dense winds at rates of $10^{-10}$ to $10^{-4}$ M$_\\odot$ yr$^{-1}$ \\citep{Bowen1991}. During this phase of evolution the effective stellar temperature is low enough for molecules to form and dust grains to condense in the circumstellar outflow. In general, dust formation occurs in stages known as the `dust condensation sequence', and depends on the physical conditions in the envelope \\citep[][]{Tielens1990, GailSed1999, Gail2010}. The mineralogy of the dust is dominated by the chemistry of the envelope (which is set by a star's photospheric C/O atomic ratio and initial elemental abundance), and by the envelope's temperature and density (which determines which condensation products are stable). Although not understood in detail, the thermodynamic condensation sequence for oxygen-rich outflows (C/O $<$ 1) predicts that refractory oxides such as alumina (Al$_{2}$O$_{3}$) would be the first astrophysically significant species to form. These grains can exist relatively close to the star ($T_{\\rm cond} \\sim 1400$ K at pressures relevant to stellar outflows) and may act as seed nuclei for other grains \\citep[e.g.][]{Onaka1989, Stencel1990, Sogawa1999}. At larger radii, where the temperature is lower ($T \\sim 1000$ K) silicate grains form. It is well established that the dust in oxygen-rich outflows consists mainly of silicates \\citep{WoolfNey1969, Hackwell1972, TreffersCohen1974}. However, alumina dust has been detected in the spectra of some low mass-loss rate ($\\dot M$) O-rich AGB stars in the Milky Way \\citep{Onaka1989, Speck2000, Dijkstra2005}. These stars show a broad low-contrast emission feature in the 7 -- 14 $\\mu$m region which is best fit with a blend of amorphous alumina and amorphous silicates. As the star evolves along the AGB the shape and peak position of this feature changes; silicate dust becomes more important and dominates the 10-$\\mu$m emission \\citep{Little-Marenin1990,Sloan1995}. At high mass-loss rates the 10-$\\mu$m feature is only due to silicates. This sequence has been quantified by \\cite{Sloan1995, Sloan1998} using flux ratios in the 9--12 $\\mu$m region. The \\emph{Spitzer} Space Telescope has taken spectra of a large number of individual evolved stars in the Large and Small Magellanic Clouds, covering a wide range of colours and magnitudes \\citep{Kemper2010}. Although the mineralogy of the oxygen-rich AGB stars in the Magellanic Clouds is dominated by amorphous silicates both at low and high mass loss rates \\citep{Sargent2010, Riebel2012} some sources also contain a small fraction of crystalline silicates \\citep{Jones2012}. There is an apparent absence of Al$_{2}$O$_{3}$ in the \\emph{Spitzer} spectra of O-AGB stars in the Magellanic Clouds \\citep{Sloan2008}, with the 10-$\\mu$m feature of low density (low $\\dot M$) winds better reproduced by amorphous silicate grains than the expected mixture of Al$_{2}$O$_{3}$ and silicates. This implies that either alumina dust is depleted in the LMC or that observational biases hinder its detection. However, an extensive systematic study into the Al$_{2}$O$_{3}$ content of evolved stars in the LMC has not yet been carried out. In this study we evaluate the alumina content of oxygen-rich AGB stars in the Magellanic Clouds, by creating a grid of radiative transfer models that explores a range of physical parameters relevant to evolved stars. In Section~\\ref{model}, we describe how the radiative transfer models for the circumstellar dust shell were computed. Colour-colour diagrams of the grid of models are presented in Section~\\ref{results} and will be used to study the alumina dust in a sample of O-rich AGB stars in the Magellanic Clouds in Section~\\ref{AluminaLMC}. Finally, we discuss our results in Section~\\ref{Discussion} and summarise this study in Section~\\ref{conclusions}. ", "conclusions": "We have presented a grid of dust radiative transfer models which explores a range of alumina and silicate dust compositions and stellar and dust shell parameters. The models have been used to simultaneously fit the spectra and broadband SED of 37 oxygen-rich AGB stars in the LMC with optically-thin circumstellar envelopes. The mass-loss rates of our sample range from $\\sim 8\\times10^{-8}$ to $5\\times10^{-6}$ M$_{\\odot}\\ \\mathrm{yr}^{-1}$. We find that a combination of amorphous silicates, amorphous alumina and metallic iron provides a good fit to the spectra. All our sources we found to be silicate-rich, though alumina and iron is often present in significant amounts. This dust composition is consistent with the thermodynamic dust condensation sequence for oxygen-rich AGB stars. Furthermore, we show from dust models that the {\\em AKARI} [11]--[15] versus [3.2]--[7] colour, is able to determine the fractional abundance of alumina in oxygen-rich AGB stars." }, "1402/1402.5093_arXiv.txt": { "abstract": "% A quantum electrodynamics (QED) correction surface for the simplest polyatomic and polyelectronic system \\hp{} is computed using an approximate procedure. This surface is used to calculate the shifts % to vibration-rotation energy levels due to QED; such shifts have a magnitude of up to 0.25~\\cm{} for vibrational levels up to 15~000~\\cm{} and are expected to have an accuracy of about 0.02~\\cm. Combining the new \\hp\\ QED correction surface with existing highly accurate Born-Oppenheimer (BO), relativistic and adiabatic components % suggests that deviations of the resulting {\\it ab initio} energy levels from observed ones % are largely due to non-adiabatic effects. \\pacs{Valid PACS appear here}% ", "introduction": "\\emph{Ab initio} studies of diatomic and triatomic systems containing less than ten electrons are nowadays able to produce rotation-vibrational energy levels with better than spectroscopic accuracy, i.e. with errors of less than 1~\\cm. To improve on this accuracy one needs to account for several small effects which are routinely neglected, including electronic relativistic and adiabatic corrections, as well as --- most notably for this work --- non-adiabatic effects and corrections due to quantum electrodynamics (QED). General discussions of relativistic and QED effects in molecular physics and quantum chemistry can be found in several recent reviews \\cite{Pyykko2012,Liu2013,Autschbach2012,Liu2012,Kutzelnigg2012,Ilias2010,anatomy} and textbooks \\cite{Reiher.book,Dyall.book}. In this study we follow the convention of calling `relativistic effects' corrections to the non-relativistic Schr{\\\"o}dinger equation of second order in the fine-structure constant $\\alpha$ (i.e., all effects correctly described by the many-electron no-pair Dirac-Coulomb-Breit equation), while so-called radiative corrections due to the quantization of the electromagnetic field and appearing in higher powers of $\\alpha$ are referred to as QED effects. The hydrogen molecular ion H$_2^+$ is the simplest physical system with a rotational-vibrational spectrum and serves as an important benchmark. Rotational-vibrational energy levels for H$_2^+$ were notably presented by Moss \\cite{Moss1999} with an estimated accuracy of $10^{-4}$~\\cm{} and included non-adiabatic, relativistic as well as leading QED corrections. More recent studies have considerably improved the achievable accuracy and, for selected rotation-vibrational transitions, QED corrections up to $\\alpha^5$ have been computed \\cite{Korobov2006,Korobov2008,Zhong2012} leading to uncertainties of about $2\\times 10^{-6}$~\\cm. Next in terms of size and complexity is the hydrogen molecule \\htwo, for which an accuracy of $10^{-4}$ \\cm\\ has recently been achieved {\\it ab initio} \\cite{KomasaJCTC2011, Ubachs,Ubachs1} by careful inclusion of non-adiabatic corrections and of QED corrections to order $\\alpha^4$. Studies of H$_2^+$ and H$_2$ represent the current state-of-the-art for calculations of molecular rotational-vibrational energy levels; for larger systems the achievable accuracy is considerably lower. In particular, for \\hp\\ the highest accuracy achieved so far is 0.10~\\cm\\ for all known energies up to 17~000~\\cm\\ \\cite{jt512}, which is therefore several orders of magnitude worse than for H$_2^+$ and H$_2$. Higher accuracy energy levels are necessary for proper analysis of \\hp\\ experimental spectra. More specifically, about 30 years ago Carrington and co-workers \\cite{CB82,CK84,cm89} measured very dense near-dissociation spectra of \\hp\\ and its isotopologues with an average line spacing of less than 0.01 \\cm; these spectra, which remain unassigned and substantially uninterpreted \\cite{jt157}, clearly require very high accuracy to be analysed from theoretical calculation. Another source of motivation is provided by the recent studies by Wu {\\it et al} \\cite{13WuLiLi.H3+} and Hodges {\\it et al} \\cite{13HoPeJe.H3+}, who have concentrated on high-precision and high-accuracy frequency measurements on the \\hp\\ $\\nu_2$ fundamental band. Measurements were made by both groups at the sub-MHz ($3 \\times 10^{-5}$ \\cm) level but currently do not agree with each other within the claimed uncertainties. The assigned \\hp\\ experimental data has recently been the subject of an analysis using the MARVEL procedure \\cite{jt412}, producing a comprehensive set of rotation-vibration energy levels \\cite{H3pMARVEL,H3pMARVEL1} which we use for comparison throughout this study. Given the present experimental situation it is therefore very desirable to improve the accuracy of theoretical \\hp\\ energy levels beyond the 0.1~\\cm\\ level. The main non-relativistic, clamped nuclei Born-Oppenheimer (BO) potential energy surface (PES) from Pavanello {\\it et al} \\cite{jt512,jt526} and the associated relativistic and adiabatic surfaces, all of which we use in this work, are probably sufficiently well-determined to predict energy levels with an accuracy of about $10^{-2}$ \\cm\\ for low-lying levels up to about 15,000~\\cm. % There are currently two factors limiting the accuracy in \\hp\\ to the 0.1~\\cm{} level, namely a proper treatment of, \\emph{i)}, non-adiabatic and, \\emph{ii)}, QED effects. Non-adiabatic effects in \\hp\\ and its isotopologues are known to affect line positions by up to 1.0~\\cm{} \\cite{jt236} and therefore must be accounted for accurately. Polyansky and Tennyson (PT) \\cite{jt236} introduced a simple model based on the use of fixed, effective vibrational and rotational masses taken from Moss's \\cite{Moss1993} studies on H$_2^+$; PT were able to improve the accuracy of calculations from 1~\\cm\\ to 0.1~\\cm. Further improvements require more sophisticated treatments of non-adiabatic effects; a step in this direction has been made by Diniz {\\it et al} \\cite{jt566}, who obtained non-adiabatic rotational-vibrational energies for the $\\nu_2$ band with an accuracy of 0.01 ~\\cm\\ but did not consider higher vibrational states. The second factor limiting the final accuracy of \\hp\\ energy levels are QED effects. As discussed above, QED effects have been computed accurately for H$_2^+$ \\cite{Moss1993} and H$_2$ \\cite{KomasaJCTC2011,Ubachs,Ubachs1} and have an effect in the region 0.1---0.2~\\cm\\ on the corresponding rotation-vibration energy levels. In the case of \\hp, QED effects have so far been entirely neglected but must clearly be taken into account to achieve accuracies better than 0.1~\\cm. Pyykk{\\\"o} \\emph{et al.} \\cite{jt236} suggested a simple scheme for describing leading QED effects in molecules (see section \\ref{section.QED} for details). This scheme has been already applied to the water molecule \\cite{jt236,jt309} --- for which QED corrections % are of the order of 1~\\cm --- and was instrumental in recent studies achieving an accuracy of 0.1~\\cm\\ for levels up to 15~000~\\cm\\ \\cite{jt550} and of 1~\\cm\\ for the dissociation energy \\cite{jt549}. In this study we use the model of Pyykk{\\\"o} \\emph{et al.} \\cite{jt236} to provide a QED correction surface for \\hp. This correction energy surface, when combined with the existing non-relativistic, relativistic and adiabatic surfaces from previous studies \\cite{jt512,jt526} and with a future, accurate treatment of non-adiabatic effects is expected to provide rotation-vibration energy levels with a typical accuracy of 0.01 \\cm. The paper is organised as follows. Section \\ref{sec.basis-set} presents a comparison of the Born-Oppenheimer PES computed using explicitly correlated Gaussians \\cite{jt512,jt526} and surfaces computed using standard quantum chemistry methods based on full configuration interaction (FCI) and Gaussian basis sets. We show that available basis sets provide an accuracy between 0.1 \\cm\\ and 1 \\cm\\ for rotation-vibration energy levels. Section III compares results of accurate QED calculations for \\htwo\\ \\cite{KomasaJCTC2011,Ubachs,Ubachs1} with our calculations using the approximate method of Pyykk{\\\"o} et al. \\cite{jt236}. QED corrections for \\hp\\ using the same methodology are presented. Section IV presents results of nuclear motion calculations using a BO PES, relativistic and adiabatic corrections \\cite{jt512,jt526} and our QED correction surface. Nuclear motion calculations are given both without non-adiabatic corrections and with a simple non-adiabatic treatments based either on the Polyansky-Tennyson (PT) model \\cite{jt236} or on the model by Diniz {\\it et al} \\cite{jt566}. Analysis of the residual deviations between theory and experiment is given. Section V presents a final discussion and conclusions. ", "conclusions": "We calculated a QED energy correction surface for \\hp\\ using the approximate method of Pyykk{\\\"o} \\etal \\cite{jt265}. This method is benchmarked against accurate QED calculations for H$_2^+$ and \\htwo; the comparisons suggest that our QED surface for \\hp\\ should provide QED corrections to rotational-vibration energy levels with an accuracy better than $0.02$~\\cm. The effect of QED on low-lying energy levels is of the order of 0.2~\\cm{} and hence is much larger than the accuracy of $10^{-2}$ \\cm\\ which has already been achieved for all components of \\ai\\ calculations on \\hp\\ with the notable exception of non-adiabatic effects. Inclusion of QED effects leads to \\hp\\ energy levels being reproduced with a RMS % deviation which is reduced from 0.99 \\cm\\ to 0.84 \\cm{} when no allowance is made for non-adiabatic effects (nuclear masses used for energy levels calculation). These calculations, which include highly accurate BO, adiabatic, relativistic and QED effects but no provision for non-adiabatic effects, therefore represent an accurate characterisation of the value of non-adiabatic effects for each \\hp\\ level. Allowance for non-adiabatic effects using the simple model of PT \\cite{jt236} results in a further reduction of this deviation to 0.22~\\cm. Use of the non-adiabatic model of Diniz \\etal shows that in this model the use of QED corrections reduces the errors in the results by almost a factor of two from 0.33~\\cm\\ to 0.19~\\cm. This demonstrates the necessity of including QED corrections in accurate {\\it ab initio} treatments of \\hp\\ rotation-vibration energy levels; it opens the way for the development of an accurate non-adiabatic model which could potentially reach the $10^{-2}$ \\cm\\ accuracy necessary for the assignment of Carrington -- Kennedy \\cite{CB82} near-dissociation spectrum of \\hp\\ and its isotopologues." }, "1402/1402.2350_arXiv.txt": { "abstract": "We used six simultaneous {\\em XMM-Newton} and {\\em Rossi X-ray Timing Explorer} plus five {\\em Suzaku} observations to study the continuum spectrum and the iron emission line in the neutron-star low-mass X-ray binary 4U 1636$-$53. We modelled the spectra with two thermal components (representing the accretion disc and boundary layer), a Comptonised component (representing a hot corona), and either a Gaussian or a relativistic line component to model an iron emission line at $\\sim 6.5$ keV. For the relativistic line component we used either the {\\sc diskline}, {\\sc laor} or {\\sc kyrline} model, the latter for three different values of the spin parameter. The fitting results for the continuum are consistent with the standard truncated disc scenario. We also find that the flux and equivalent width of the iron line first increase and then decrease as the flux of the Comptonised component increases. This could be explained either by changes in the ionisation state of the accretion disc where the line is produced by reflection, or by light bending of the emission from the Comptonised component if the height at which this component is produced changes with mass accretion rate. ", "introduction": "Low-mass X-ray binaries (LMXBs) consist of a compact object (a neutron star or a black hole) and a late-type companion star with a mass of less than $\\sim $1 M$_{\\odot}$. Material from the outer layers of the companion is stripped off and accretes onto the compact object via the inner Lagrangian point and an accretion disc \\citep{accretion_book}. The inner parts of the accretion disc, and in the case of neutron stars the neutron-star surface and boundary layer, emit mostly in the X-ray band. These systems show also high-energy (up to a few 100 keV) emission, likely produced by Comptonisation of the soft X-ray photons in a hot electron corona \\citep{mcc2000,sanna13}. An extra hard tail has been observed in some LMXBs, e.g., GX 17$+$2, 4U 1636$-$53, GX 349$+$2, Sco X-1, 4U 1608$-$522, GX 13$+$1, 4U 1705$-$44 \\citep{salvo00,fiocchi06,paizis06,piraino07}. Based on X-ray spectral and rapid variability properties, \\citet{hasinger89} classified the neutron-star low-mass X-ray binaries into two categories: the Z sources and the Atoll sources, owing their names to the shapes that the source traces in an X-ray colour-colour diagram. The Z sources have higher luminosities ($\\sim 0.5-1~L_{\\rm Edd}$) than the Atoll sources \\citep[$0.01-0.2 ~L_{\\rm Edd}$; e.g.,][]{done07,homan07,ford2000}. In Atoll sources the X-ray spectrum softens and the time-scale of the majority of the variability components decreases as the luminosity of the source generally increases \\citep{hasinger89,van03}. More specifically, at low luminosities these sources are in the so-called hard state, in which the Comptonised component dominates the energy spectrum; this component can be reasonably described by a power law with a photon index of $\\sim 1.6 - 2.5$ \\citep{yoshida93,mendez97}. The temperature at the inner edge of the accretion disc is relatively low, $0.3 - 0.5$ keV \\citep{sanna13}, and the disc usually contributes less than 20\\% of the emission in the $1-20$ keV energy range. In the truncated disc scenario \\citep[see, e.g.,][and references therein]{done07}, the accretion disc is truncated at a large radius in the hard state, that being the reason for the relatively low inner-disc temperature and thermal component flux \\citep{gierlinski03}. At high luminosities, in the soft state, the disc emission in the $1-20$ keV range becomes more significant. The standard accretion-disc model \\citep{ss73} predicts that in this case the disc extends down to the innermost stable circular orbit radius, leading to a high disc temperature and a strong thermal component. Compared to the hard state, in the soft state the inner-disc temperature increases to $0.7-1.0$ keV \\citep{sanna13}; since the number of soft photons increases as $T^4$, the electrons in the corona are efficiently cooled down via the inverse Compton scattering process, and the Comptonised spectrum steepens \\citep[photon index $\\sim 2 - 2.5$, e.g.,][]{miya93,mendez97}. In the soft state, the thermal components dominate the X-ray spectrum below $\\sim 20$ keV, and little hard emission is detected (Gierli\\'nski \\& Done. 2003). Both in the hard and the soft state the neutron star surface or boundary layer, usually fitted with a blackbody component \\citep{white88,sanna13}, contributes significantly to the emission at energies below $\\sim 10$ keV. Going from the hard to the soft state, or vice-versa, the spectra of these sources display some intermediate properties, and the source is said to be in the intermediate, or transitional, state. (These basic states receive several names depending on the class of the source; see, e.g., \\citealt{hasinger89}). The mechanism driving the transition between the hard to the soft state is still unclear. However, it is generally assumed that the change of mass accretion rate and the disc geometry are connected to the state evolution \\citep{hasinger89,mendez99,lin07}. Besides the emission components described above, the accretion disc is likely illuminated by the Comptonised photons and the thermal spectrum from the neutron star and its boundary layer, and as a consequence it produces a reflection spectrum \\citep[e.g.][]{fabian10}. Due to the high abundance and fluorescence yield, an iron emission line at $\\sim$6$-$7 keV may appear in the spectrum, with an intrinsic line width of the order of 1eV \\citep{basko78}. The iron line profile is asymmetrically broadened by the fast disc rotation (Doppler-broadening) and special and general relativistic effects (e.g., Doppler boosting and gravitational redshift) near the central compact object \\citep{fabian89}. The final line profile is determined by parameters of the system, like the inclination angle of the disc with respect to the line of sight, the inner radius of the disc, and the spin parameter of the central object. Thus measurements of the iron line profile provide an excellent way to study the physics and geometry of the accretion process \\citep{bhatta07,cackett08,miller13}. 4U 1636$-$53 is an Atoll LMXBs, consisting of a neutron star and a 0.4 M$_{\\odot}$ companion in a 3.8 hr orbit \\citep{pedersen82}, at 6 kpc distance \\citep{galloway06}. The source shows the full range of spectral states \\citep{belloni07,alta08}. Highly coherent burst oscillations indicate that the system harbors a millisecond pulsar with a spin frequency of 581 Hz \\citep{zhang97,stro02}. A pair of quasi-periodic oscillations (QPOs) at kHz frequencies were discovered by \\citet{zhang96} and \\citet{wijnands97}. \\citet{kaaret99} and \\citet{marcio13} found that the soft X-ray emission in the lower kHz QPO (from the pair of kHz QPOs the one at lower frequency) lags the hard X-ray emission, suggesting that the emission is due to reprocessing of hard X-rays in a cooler Comptonising corona or the accretion disc, in a region with a size of at most a few kilometres. In the last ten years 4U 1636--53 showed a regular state transition cycle of $\\sim$40 days \\citep{shih05,belloni07}, making it an excellent source to study the variations of the broadband spectrum and iron line as a function of spectral state. In this work, we study the spectrum of 4U 1636$-$53 in different states. We use eleven observations from Suzaku, XMM-Newton, and the Rossi X-Ray Timing Explorer (RXTE) satellites covering a wide range in luminosity, allowing us to investigate the evolution of the different spectral components as a function of spectral state, and study possible correlations between the continuum spectrum and the iron line. Since the disc is illuminated and photon-ionised by the continuum emission to produce the iron line, correlations between the iron line flux and the flux of different continuum components may provide an important clue to understand the origin of the iron emission line and the evolution of the accretion flow geometry. ", "conclusions": "We used all available Suzaku and XMM-Newton observations, the latter complemented with simultaneous RXTE observations, to study the evolution of the continuum spectrum and the iron line emission in the neutron star low-mass X-ray binary 4U 1636$-$53 across different spectral states. We found that the temperature of the neutron star and that at the inner edge of the accretion disc increase, whereas the electron temperature of the corona decreases, as the source moves across the colour-colour diagram from the hard to the soft state. Simultaneously, the inner radius of the accretion disc deduced from the {\\sc diskbb} component decreases rapidly by a factor of $\\sim 10$, and then remains more or less constant as the source reaches the soft state, suggesting that the inner edge of the disc may have reached the ISCO. The power-law photon index of the component used to fit the corona emission increases from $\\sim 1.7$ in the hard state up to $\\sim 2.8$ (depending on the model used to fit the line) in the soft state. Simultaneously, the total unabsorbed flux and the flux of the Comptonised component remain more or less constant, whereas the flux of the disc and that of the neutron star plus boundary layer increase, as the source moves from across the colour-colour diagram from the hard to the soft state (all fluxes calculated in the $0.5 -130$ keV range). Interestingly, the flux and equivalent width of the iron line first increase and then decrease as the flux of the Comptonised component increases. The maximum of this relation takes place when the source is close to the vertex in the colour-colour diagram. The relation between the flux or the equivalent width of the line on one side and the flux of the hard spectral component on the other appears to contradict the expected behaviour if the line is due to reflection off the accretion disc. In the reflection scenario, the corona (and in neutron-star systems possibly also the neutron-star or boundary-layer) illuminates and photon-ionises the disc, where the reflection spectrum (continuum + emission line) is then produced. In this scenario the flux or equivalent width of the iron line should be positively correlated with that of the illuminating component. We find that, while this is the case when at low {\\sc nthcomp} flux values, the opposite is true when the {\\sc nthcomp} flux is high (see Figure \\ref{corrflx}). However, the disc becomes more ionised when the illuminating flux increases, provided that there are enough photons with energies above the ionisation potential of H- and He-like iron. As the flux of the illuminating component increases further, the material in the disc would eventually become fully ionised, and the line will disappear. This scenario agrees with recent calculations of the reflection spectrum of an ionised slab at different ionisation levels \\citep{garcia13}. \\cite{garcia13} found (see their Figure 5) that the flux of the iron emission line in the reflected spectrum decreases gradually as the material in the slab is ionised further. Similarly, \\cite{matt93} showed that the equivalent width of the iron line in the reflected spectrum of an ionised slab initially increases and then decreases as the flux of the ionising source increases. Our results are generally consistent with this idea \\citep[compare our Figure \\ref{corrflx} with Figures 1 and 2 in][]{matt93}. However, in this scenario, as the ionisation state of the disc increases, the fitted energy of the emission line should gradually increase, as the fraction of H-like iron ions relative to the He-like iron ions in the disc increases. Our results do not show any correlation between the fitted energy of the line and the flux of the ionising source (see Tables \\ref{sline} and \\ref{xline}), regardless of the model we used to fit the line. While this lack of correlation could be due to limitations of the models that we used to fit the iron line, or to the fact that we only fitted the iron line instead of the full reflection spectrum, using the same XMM-Newton data presented here \\citet{sanna13} found no correlation between the ionisation parameter of the full reflection model they fitted to the spectra and the position of the source in the colour-colour diagram, reinforcing this conclusion. Alternatively, the relation between iron line flux or equivalent width and the flux of the hard spectral component in 4U 1636--53 shown in Figure \\ref{corrflx} could be interpreted in terms of light bending \\citep{miniutti04}. In this model, the observed flux of the hard illuminating source (in our fits the {\\sc nthcomp}) depends strongly upon the height of the source of photons in the corona above the disc; the observed flux of the direct emission can change by up to a factor $\\sim20$ as the height of the illuminating source changes, even if the luminosity of the illuminating source remains constant. On the other hand, the reflected spectrum of the disc, and hence the flux or equivalent width of the iron line, is much less sensitive to the height of the illuminating source. \\cite{miniutti04} find that there are three regimes in which, as the flux of the direct emission increases, the reflection component (and hence the iron line flux or equivalent width) is first correlated, then insensitive, and finally anti-correlated with the flux of the direct emission. \\cite{rossi05} found that, in the black-hole candidate XTE J1650--500, as the total flux of the source above 7 keV increases, the flux of the iron line initially remains constant and eventually decreases. They found that the source flux level at which the line flux starts to decrease coincides with the transition from the hard to the soft state in this source. \\cite{park04} also found a complex relation between the iron-line flux and the flux of the hard and the soft components in the black-hole candidate 4U 1543--475. In this case, however, the relation does not follow any clear trend, and an interpretation in terms of the light bending model is not apparent. Rossi et al. (2005) suggested the direct and reflected components in XTE J1650--500 could be related to the existence of a radio jet in this source \\citep{corbel04}, with possible changes of the height of the base of the jet as the mechanism that drives the changes of the direct emission in XTE J1650--500. It is unclear whether this could also be the case in 4U 1636--53, since this source has not been detected in radio, and hence the existence of a strong jet in this system is at best doubtful \\citep{thomas79,russell12}. The lack of a strong radio jet in 4U 1636--53 could also be the reason that the flux of the hard (direct) component in this source changes only by a factor $\\sim 5$, whereas in XTE J1650--500 the flux of the direct ionising source changes by $\\simmore 10$. We finally note that if the relation between the flux or equivalent width of the iron line and the {\\sc nthcomp} flux in 4U 1636--53 is due to light bending, this would be the first case of a neutron-star system in which this effect is observed. The light bending model was developed for the case of a rapidly spinning black hole, and therefore it is unclear whether it would also apply for a moderately spinning neutron star. In the case of 4U 1636--53 the point at which the flux of the iron line switches from correlated to anti-correlated with the flux of the {\\sc nthcomp} component coincides with the vertex in the colour-colour-diagram, at $S_{a} \\sim 2.1$. The existence of the vertex in this diagram directly indicates a sudden change of the spectral properties of the source which, going from the hard to the soft state, quickly softens there. We note also that the vertex of the colour-colour diagram is the place where the quality factor of the kHz QPOs \\citep{barret06,mendez06} and the intrinsic coherence between the variability in the hard and soft energy bands \\citep{marcio13} are the highest. Furthermore, \\citet{zhang11} studied 298 type-I X-ray bursts in the 4U 1636$-$53 using RXTE observations, and they found that in this area of the colour-colour diagram most photospheric radius expansion bursts and a super-burst are observed. All these properties suggest that, in this source several properties of the accretion flow change significantly in this area of the colour-colour diagram. The changes of the properties of the continuum spectrum of 4U 1636--53 as the source moves in the colour-colour diagram are generally consistent with those of other neutron-star LMXBs. For instance, \\citet{farinelli11} measured the power-law index and the electron temperature $kT_{\\rm e}$ of the Comptonising component in the neutron star LMXBs Sco X-1, GX 349$+$2, X 1658$-$298, 1E 1724$-$3045, GX 17$+$2, Cyg X$-$2, GX 340$+$0, GX 3$+$1, and GS 1826$-$238, and found that the power-law energy index ($\\Gamma -1$ in the {\\sc nthcomp} component) remains more or less constant at around 1$\\pm$0.2. Using RXTE and BeppoSAX observations, \\citet{seifina11,seifina12} showed that in 4U 1728$-$34 and GX 3$+$1, the power-law index also remains almost constant as the temperature of the corona changes dramatically. Recently, \\citet{titar13} found that in another neutron star binary, 4U 1820$-$30, the power-law index remains almost constant for different source states. The power-law index in 4U 1636$-$53 changes from $\\sim 1.7$ in the hard state to $\\sim 2.8$ in the soft state (depending on the model used to fit the iron line), which is a somewhat larger than the variations in the sources studied by \\citet[][note, however, that the model that they used to fit the Comptonised component is not the same as the one we used here]{farinelli11}. \\citet{titar13} concluded that the power-law index quasi-stability is an intrinsic property of neutron star binaries, which is fundamentally different from that of black-hole binary systems. \\citet{seifina11} suggested that this stability of the power-law index happens when the energy released in the corona itself is much higher than the one from the disc intercepted by the corona. The changes of the continuum spectrum are also broadly consistent with the truncated disc scenario in LMXBs \\citep[e.g.][and references therein]{done07}: The temperature of the disc increases and the inner radius of the disc decreases as the inferred mass accretion rate increases. The electron temperature of the {\\sc nthcomp} component and the temperature of the neutron-star surface are also consistent with this scenario. It is intriguing, however, that the inner radius of the accretion disc deduced from the relativistically broaden iron line does not follow the same trend. The model of the disc that we used to fit the data does not include spectral hardening, and the iron line profile may be further affected by mechanisms other than relativistic broadening. It remains to be seen whether these effect could explain this discrepancy." }, "1402/1402.4483_arXiv.txt": { "abstract": "We report on a comprehensive X-ray spectral analysis of the nearby radio-quiet quasar MR~2251$-$178, based on the long-look ($\\sim 400$ ks) \\textit{XMM--Newton} observation carried out in November 2011. As the properties of the multiphase warm absorber (thoroughly discussed in a recent, complementary work) hint at a steep photoionizing continuum, here we investigate into the nature of the intrinsic X-ray emission of MR~2251$-$178 by testing several physical models. The apparent 2--10 keV flatness as well as the subtle broadband curvature can be ascribed to partial covering of the X-ray source by a cold, clumpy absorption system with column densities ranging from a fraction to several $\\times 10^{23}$ cm$^{-2}$. As opposed to more complex configurations, only one cloud is required along the line of sight in the presence of a soft X-ray excess, possibly arising as Comptonized disc emission in the accretion disc atmosphere. On statistical grounds, even reflection with standard efficiency off the surface of the inner disc cannot be ruled out, although this tentatively overpredicts the observed $\\sim 14$--150 keV emission. It is thus possible that each of the examined physical processes is relevant to a certain degree, and hence only a combination of high-quality, simultaneous broadband spectral coverage and multi-epoch monitoring of X-ray spectral variability could help disentangling the different contributions. Yet, regardless of the model adopted, we infer for MR~2251$-$178 a bolometric luminosity of $\\sim 5$--$7 \\times 10^{45}$ erg s$^{-1}$, implying that the central black hole is accreting at $\\sim15$--25 per cent of the Eddington limit. ", "introduction": "Among the X-ray brightest Active Galactic Nuclei (AGN) in the local Universe due to its 2--10 keV luminosity largely exceeding 10$^{44}$ erg s$^{-1}$, MR~2251$-$178 ($z \\simeq 0.064$; Canizares, McClintock \\& Ricker 1978) is a spectacular object in every respect. It was the first quasar to be detected and identified through X-ray observations (Cooke et al. 1978; Ricker et al. 1978), as well as the first one where the presence was established of \\textit{warm} absorption by photoionized gas, variable in both ionization state and possibly column density over timescales of less than one year (Halpern 1984). The source was later found to experience appreciable changes in the X-ray flux over periods of $\\sim 10$ days (Pan, Stewart \\& Pounds 1990), with a tight correlation between the ionization parameter of the absorbing gas\\footnote{The ionization parameter is defined as $\\xi = L_\\rmn{ion}/nr^2$, where $n$ is the electron density of the gas and $r$ is its distance from an ionizing source with 1--1000 Ry luminosity $L_\\rmn{ion}$ (Tarter, Tucker \\& Salpeter 1969).} and the continuum luminosity (Mineo \\& Stewart 1993). Narrow absorption lines with a systematic blueshift of $\\sim 300$ km s$^{-1}$ have been detected in the ultraviolet (UV) due to Ly$\\alpha$, N~\\textsc{v} and C~\\textsc{iv} (Monier et al. 2001). In particular, the C~\\textsc{iv} doublet apparently vanished in less than four years (Ganguly, Charlton \\& Eracleous 2001), implying that the UV absorber is truly local to the AGN, and that it is possibly one and the same with the soft X-ray warm absorber. Much deeper insights came with the advent of high-resolution X-ray spectroscopy. Early \\textit{XMM--Newton} Reflection Grating Spectrometer (RGS) observations hinted at a multiphase configuration for the warm absorber, likely consisting of two or three distinct components (Kaspi et al. 2004), while the \\textit{Chandra} High Energy Transmission Grating (HETG) spectrum revealed a highly ionized absorption feature in the iron K band, interpreted as the Fe~\\textsc{xxvi} Ly$\\alpha$ line at the sizable outflow velocity of $v_\\rmn{out} \\sim 0.04c$ (Gibson et al. 2005). The corresponding mass-loss rate was calculated to be at least an order of magnitude larger than the accretion rate, unless the covering fraction of the outflow is very small. The environment of MR~2251$-$178 is likewise exceptional. The quasar lies in the outskirts of a loose, irregular cluster with several tens of galaxies (Phillips 1980), and is surrounded by a huge emission-line nebula detected in H$\\alpha$ and [O~\\textsc{iii}] at optical wavelengths out to a distance of $\\sim 100$ kpc from the central source (Bergeron et al. 1983; Shopbell, Veilleux \\& Bland-Hawthorn 1999). The knotty and filamentary appearance of this gaseous envelope is more typical of powerful radio galaxies, yet MR~2251$-$178 only shows weak radio emission, whose elongated morphology resembles a double-lobed jet-like structure (Macchetto et al. 1990). A similar spatial extent and orientation marks the ionization cones recently found through deep [O~\\textsc{iii}]$\\lambda$5007/H$\\beta$ and [N~\\textsc{ii}]$\\lambda$6583/H$\\alpha$ flux ratio maps (Kreimeyer \\& Veilleux 2013). Overall, the quasar radiation field can easily account for the ionization state of the nebula, which is arguably the most extended around a radio-quiet source. Its origin, however, is still unclear. Diffuse X-ray emission along some directions was preliminary reported by Gibson et al. (2005) based on the smoothed HETG zeroth-order image taken in 2002. Unfortunately, we cannot safely corroborate these findings through the much deeper 2011 HETG data set, since the artificial broadening of the instrumental Point Spread Function due to photon pile-up overrides any faint contribution from the halo.\\footnote{The AGN flux in the 2011 \\textit{Chandra} observation was larger by $\\sim 50$ per cent with respect to 2002. While the effects of pile-up on the Point Spread Function can be qualitatively simulated, they prevent any accurate image reconstruction.} Due to its key role as the underlying source of ionizing photons for the ambient gas, from nuclear to intergalactic scales, a proper knowledge of the shape and behaviour of the intrinsic X-ray continuum is highly desirable, yet most of the effort in the wealth of X-ray analyses performed so far has concentrated on improving the characterization of the multi-component warm absorber. Incidentally, all the X-ray spectral studies actually agree with describing the absorbed continuum by means of a power law of photon index $\\Gamma \\sim 1.6$, with a high-energy exponential cutoff at $\\sim 100$ keV (Orr et al. 2001) and a soft excess below $\\sim 1$ keV (Kaspi et al. 2004). In a recent paper, we have taken advantage of coordinated $\\sim 400+400$ ks long \\textit{Chandra} HETG and \\textit{XMM--Newton} RGS observations to bring our grasp on the properties of the intervening ionized gas to unprecedented detail and energy resolution (Reeves et al. 2013; hereafter R13). Both campaigns were conducted as large programs in late 2011, with just a few weeks of separation from one another, and yielded intriguing information on the illuminating continuum itself, which below 2 keV is required to be much steeper ($\\Gamma > 2$) than typically estimated at hard X-rays. Indeed, the analysis of the \\textit{Suzaku} spectrum carried out by Gofford et al. (2011) had already suggested that the broadband X-ray emission of MR~2251$-$178 can be reproduced equally well through a softer power law with $\\Gamma \\simeq 2$, provided that an additional cold-gas column of $N_\\rmn {H} \\sim 10^{23}$ cm$^{-2}$ covering a moderate fraction of the source is introduced. In the wake of these indications, here we present a thorough investigation of the 0.3--10 keV EPIC/pn spectrum obtained in the 2011 \\textit{XMM--Newton} observation, with the aim of understanding the nature of the primary photoionizing continuum and unveiling the high-energy physical processes at work in the very central regions of this powerful quasar. This work is organized as follows. In Section~2 we provide the basic details about the \\textit{XMM--Newton} large observing program on MR~2251$-$178, and describe the principal steps of our data reduction. Section~3 is dedicated to the spectral analysis, whose results and main implications are discussed in Section~4. Our conclusions are drawn in Section~5. Throughout this paper we have assumed $H_0=70$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_m=0.27$ and $\\Omega_\\Lambda=0.73$, in agreement with the latest values of the concordance cosmological parameters (Hinshaw et al. 2013). ", "conclusions": "Based on a recent \\textit{XMM--Newton} long-look observation with $\\sim 270$ ks of net exposure, we have reported on the 0.3--10 keV EPIC/pn spectral analysis of MR~2251$-$178, one of the brightest radio-quiet quasars in the local Universe. Following a companion paper dedicated to the study of the high-resolution grating spectra and of the physical properties of the complex warm absorber, here we have focused on the nature of the intrinsic X-ray emission. The broadband X-ray continuum of MR~2251$-$178 is known to exhibit substantial curvature up to $\\sim 100$ keV, where a possible cutoff is also present. Together with the apparent soft excess below 0.7 keV and the warm absorption trough at $\\sim 1$--2 keV, this gives rise to a peculiar steep--flat--steep spectral shape, which has been inspected within the frameworks of partial covering absorption, ionized reflection, and Comptonized disc emission. As their application turns out to be nearly statistically equivalent over the 0.3--10 keV band, all these models have also been compared with the coeval OM photometric data and the $\\sim 14$--150 keV \\textit{Swift}/BAT spectrum, averaged over 70 months. Assuming no cross normalization, the entire SED shows that the optical and hard X-ray fluxes lie at almost the same level in $\\nu f_\\nu$ units. Indeed, the optical to X-ray spectral index is extremely small ($\\alpha_\\rmn{ox} < 1.2$), irrespective of the exact absorption correction. The corresponding bolometric output of MR~2251$-$178 is estimated to be $L_\\rmn{bol} \\sim 5$--$7 \\times 10^{45}$ erg s$^{-1}$, i.e. $\\sim 15$--25 per cent of the source Eddington luminosity. None of the alternative interpretations taken into account in this work are conclusive by themselves, displaying both strengths and limitations at the same time, which can be summarized as follows: 1) The pure partial covering scenario reveals two distinct low-ionization components with columns of 0.6 and $6.6 \\times 10^{23}$ cm$^{-2}$, eclipsing a fraction of $\\sim 0.23$ and 0.49 of the X-ray source, respectively. We argue that these absorbers are located at BLR scales, but are unlikely to belong to different layers. Density gradients might be entailed, pointing to a single clump of gas with a dense core and a lighter halo. The intrinsic photon index $\\Gamma \\sim 2.2$ is close to what is required by the photoionization models of the warm absorber, but still somewhat lower; even so, excess curvature is left at hard X-rays around 50 keV. 2) At EPIC/pn resolution, X-ray reflection provides a reasonable fit without implying any extreme gravity regime. The blurred disc component, however, is unusually cold, and $\\sim 50$ times less ionized than its distant counterpart. This is possibly an artefact to compensate for the inadequate results with the soft X-ray emission lines (resolved by the gratings), and also leads to a misrepresentation of the warm absorption properties (for which the illumination is anyway flat). The degree of smoothing appears too large for the iron K band, and a clear pattern is found in the residuals above 9 keV. Moreover, the BAT spectrum is largely overpredicted. 3) The optically-thick/thin Comptonization of thermal photons from the inner disc into both the soft excess and the hard power law, with the outer regions emitting as a colour temperature corrected blackbody, is established as the most obvious connection between the optical and X-ray bands. Nevertheless, in spite of the remarkably good extrapolation to the OM data and of the ample leverage on the physical parameters, a basic self-consistent model either fails to reproduce the 0.3--10 keV features or invariably plummets too quickly at higher energies. Any agreement is lost on the three warm absorption components. A composite model allows us to successfully overcome most of the above shortcomings. While reflection from the accretion flow and/or the ambient material might still be involved to a certain extent, in the preferred picture the hard power-law continuum ($\\Gamma \\sim 1.75$) steepens below $\\sim 1$ keV into a soft excess with $kT \\sim 0.14$ keV, whose origin remains unclear although it can be tentatively associated with Comptonization in the disc atmosphere. A single cloud with $N_\\rmn{H} \\sim 3 \\times 10^{23}$ cm$^{-2}$ is present along the line of sight, covering about one sixth of the X-ray source. In conclusion, the X-ray observations of MR~2251$-$178 brought to light a complex and stratified environment close to the central source, which deeply transforms the shape of the intrinsic X-ray spectrum. Our study suggests that only a combination of time variability analysis, simultaneous broadband coverage and high spectral resolution can help disentangling the effects of the different physical processes responsible for the observed X-ray emission of AGN." }, "1402/1402.3677.txt": { "abstract": "The emission of radio waves from air showers has been attributed to the so-called geomagnetic emission process. At frequencies around 50 MHz this process leads to coherent radiation which can be observed with rather simple setups. The direction of the electric field induced by this emission process depends only on the local magnetic field vector and on the incoming direction of the air shower. We report on measurements of the electric field vector where, in addition to this geomagnetic component, another component has been observed which cannot be described by the geomagnetic emission process. The data provide strong evidence that the other electric field component is polarized radially with respect to the shower axis, in agreement with predictions made by Askaryan who described radio emission from particle showers due to a negative charge-excess in the front of the shower. Our results are compared to calculations which include the radiation mechanism induced by this charge-excess process. ", "introduction": "\\label{sec:introduction} When high-energy cosmic rays penetrate the atmosphere of the Earth they induce an air shower. The detailed registration of this avalanche of secondary particles is an essential tool to infer properties of the primary cosmic ray, such as its energy, its incoming direction, and its composition. Radio detection of air showers started in the 1960's and the achievements in these days have been presented in reviews by Allan \\cite{Refworks:188} and Fegan \\cite{Refworks:189}. In the last decade, there has been renewed interest through the publications of the LOPES \\cite{Refworks:101} and CODALEMA \\cite{Refworks:91} collaborations. We have deployed and are still extending the Auger Engineering Radio Array (AERA) \\cite{Refworks:178,Refworks:180,Refworks:183,vandenBergECR2012,SchroederICRC2013} as an additional tool at the Pierre Auger Observatory to study air showers with an energy larger than $10^{17}$ eV. In combination with the data retrieved from the surface-based particle detectors \\cite{Refworks:176} and the fluorescence detectors \\cite{Refworks:177} of this observatory, the data from radio detectors can provide additional information on the development of air showers. %An important step in the interpretation of the data obtained with radio-detection methods is the understanding of the emission mechanisms. In the early studies of radio emission from air showers it was conjectured that two emission mechanisms play an important role: the geomagnetic emission mechanism as proposed, amongst others, by Kahn and Lerche \\cite{Refworks:190} and the charge-excess mechanism as proposed by Askaryan \\cite{Refworks:191}. %BEGIN NEW TEXT An important step in the interpretation of the data obtained with radio-detection methods is the understanding of the emission mechanisms. In the early studies of radio emission from air showers it was conjectured that two emission mechanisms play an important role: the geomagnetic emission mechanism as proposed, amongst others, by Kahn and Lerche \\cite{Refworks:190} and the charge-excess mechanism as proposed by Askaryan \\cite{Refworks:191}. Essential for the geomagnetic effect is the induction of a transverse electric current in the shower front by the geomagnetic field of the Earth while the charge excess in the shower front is to a large extent due to the knock-out of fast electrons from the ambient air molecules by high-energy photons in the shower. The magnitude of the induced electric current as well as the induced charge excess is roughly proportional to the number of particles in the shower and thus changing in time. The latter results in the emission of coherent radio waves at sufficiently large wavelengths \\cite{Refworks:190, REF12a}. The shower front, where both the induced transverse current and the charge excess reside, moves through the air with nearly the velocity of light. Because air has a refractive index which differs from unity, Cherenkov-like time compression occurs \\cite{REF12b, REF12c}, which affects both the radiation induced by the transverse current as well as by the charge excess. The polarization of the emitted radiation differs for current-induced and charge-induced radiation, but its direction for each of these individual components does not depend on the Cherenkov-like time compression caused by the refractive index of air. For this reason we will distinguish in this paper only geomagnetic (current induced) and charge-excess (charge induced) radiation. The contribution of the geomagnetic emission mechanism, described as a time-changing transverse current by Kahn and Lerche \\cite{Refworks:190}, has been observed and described in several papers; see e.g. Refs. \\cite{Refworks:101,Refworks:74,Refworks:35,Refworks:186,Refworks:221}. Studies on possible contributions of other emission mechanisms from air showers have also been reported \\cite{Refworks:208,Refworks:209,Refworks:210,Refworks:211,Prescott1971}. An observation of the charge-excess effect in air showers has been reported by the CODALEMA collaboration \\cite{Refworks:198}. %END NEW TEXT We present the analysis of two data sets obtained with two different setups consisting of radio-detection stations (RDSs) deployed at the Pierre Auger Observatory. The first data set was obtained with a prototype setup \\cite{Refworks:67} for AERA; the other one with AERA itself \\cite{Refworks:178,Refworks:180,Refworks:183} during its commissioning phase while it consisted of only 24 stations. In addition, we will compare these data sets with results from different types of calculations outlined in Refs. \\cite{Refworks:3, Refworks:11, Refworks:196, Refworks:220, Refworks:212, Refworks:213, Refworks:223}. This paper is organized as follows. We discuss in Section \\ref{sec:detectionsystem} the experimental equipment used to collect our data. In Section \\ref{sec:dataanalysis} we present the data analysis techniques and the cuts that we applied on the data, whilst in Section \\ref{sec:comparison} we compare our data with calculations. In Section \\ref{sec:conclusion} we discuss the results and we present our conclusions. For clarity, Sections \\ref{sec:dataanalysis} and \\ref{sec:comparison} contain only those figures which are based on the analysis for the data obtained with AERA during the commissioning of its first 24 stations; the results of the prototype are shown in Appendix C. We mention that analyses of parts of the data have been presented elsewhere \\cite{Refworks:186,Refworks:199,Refworks:214,TimHuegeICRC2013}. %\\clearpage ", "conclusions": "\\label{sec:conclusion} We have studied with two different radio-detection setups deployed at the Pierre Auger Observatory the emission around 50 MHz of radio waves from air showers. For a sample of 37 air showers the electric field strength has been analyzed as a tool to disentangle the emission mechanism caused by the geomagnetic and the charge-excess processes. For the present data sets, the emission is dominated by the geomagnetic emission process while, in addition, a significant fraction of on average $(14 \\pm 2)\\%$ is attributed to a radial component which is consistent with the charge-excess emission mechanism. Detailed simulations have been performed where both emission processes were included. The comparison of these simulations with the data underlines the importance of including the charge-excess mechanism in the description of the measured data. However, further refinements of the models might be required to fully describe the present data set. A possible reason for the incomplete description of the data by the models might be an underestimate of (systematic) errors in the data sets or the effect of strong electric fields in the atmosphere. The successful installation, commissioning, and operation of the Pierre Auger Observatory would not have been possible without the strong commitment and effort from the technical and administrative staff in Malarg\\\"ue. We are very grateful to the following agencies and organizations for financial support: Comisi\\'on Nacional de Energ\\'ia At\\'omica, Fundaci\\'on Antorchas, Gobierno De La Provincia de Mendoza, Municipalidad de Malarg\\\"ue, NDM Holdings and Valle Las Le\\~nas, in gratitude for their continuing cooperation over land access, Argentina; the Australian Research Council; Conselho Nacional de Desenvolvimento Cient\\'ifico e Tecnol\\'ogico (CNPq), Financiadora de Estudos e Projetos (FINEP), Funda\\c{c}\\~ao de Amparo \\`a Pesquisa do Estado de Rio de Janeiro (FAPERJ), S\\~ao Paulo Research Foundation (FAPESP) Grants \\#2010/07359-6, \\#1999/05404-3, Minist\\'erio de Ci\\^{e}ncia e Tecnologia (MCT), Brazil; AVCR, MSMT-CR LG13007, 7AMB12AR013, MSM0021620859, and TACR TA01010517 , Czech Republic; Centre de Calcul IN2P3/CNRS, Centre National de la Recherche Scientifique (CNRS), Conseil R\\'egional Ile-de-France, D\\'epartement Physique Nucl\\'eaire et Corpusculaire (PNC-IN2P3/CNRS), D\\'epartement Sciences de l'Univers (SDU-INSU/CNRS), France; Bundesministerium f\\\"ur Bildung und Forschung (BMBF), Deutsche Forschungsgemeinschaft (DFG), Finanzministerium Baden-W\\\"urttemberg, Helmholtz-Gemeinschaft Deutscher Forschungszentren (HGF), Ministerium f\\\"ur Wissenschaft und Forschung, Nordrhein-Westfalen, Ministerium f\\\"ur Wissenschaft, Forschung und Kunst, Baden-W\\\"urttemberg, Germany; Istituto Nazionale di Fisica Nucleare (INFN), Ministero dell'Istruzione, dell'Universit\\`a e della Ricerca (MIUR), Gran Sasso Center for Astroparticle Physics (CFA), CETEMPS Center of Excellence, Italy; Consejo Nacional de Ciencia y Tecnolog\\'ia (CONACYT), Mexico; Ministerie van Onderwijs, Cultuur en Wetenschap, Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWO), Stichting voor Fundamenteel Onderzoek der Materie (FOM), Netherlands; Ministry of Science and Higher Education, Grant Nos. N N202 200239 and N N202 207238, The National Centre for Research and Development Grant No ERA-NET-ASPERA/02/11, Poland; Portuguese national funds and FEDER funds within COMPETE - Programa Operacional Factores de Competitividade through Funda\\c{c}\\~ao para a Ci\\^{e}ncia e a Tecnologia, Portugal; Romanian Authority for Scientific Research ANCS, CNDI-UEFISCDI partnership projects nr.20/2012 and nr.194/2012, project nr.1/ASPERA2/2012 ERA-NET, PN-II-RU-PD-2011-3-0145-17, and PN-II-RU-PD-2011-3-0062, Romania; Ministry for Higher Education, Science, and Technology, Slovenian Research Agency, Slovenia; Comunidad de Madrid, FEDER funds, Ministerio de Ciencia e Innovaci\\'on and Consolider-Ingenio 2010 (CPAN), Xunta de Galicia, Spain; The Leverhulme Foundation, Science and Technology Facilities Council, United Kingdom; Department of Energy, Contract Nos. DE-AC02-07CH11359, DE-FR02-04ER41300, DE-FG02-99ER41107, National Science Foundation, Grant No. 0450696, The Grainger Foundation USA; NAFOSTED, Vietnam; Marie Curie-IRSES/EPLANET, European Particle Physics Latin American Network, European Union 7th Framework Program, Grant No. PIRSES-2009-GA-246806; and UNESCO. \\clearpage \\appendix" }, "1402/1402.3111.txt": { "abstract": "{} %aims {Period study of 321 fundamental mode RR Lyrae type stars (RRab), which had appropriate data in ASAS and SuperWASP surveys, was performed to complement and extend the list of known Blazhko stars in galactic field with bright stars up to 12.5 mag in maximum light.} %methods {An individual approach was applied to each studied star. Permanent visual supervision was maintained to each procedure in data analysis (data cleaning, frequency spectra examination) to avoid missing any possible sign of the Blazhko effect. Period analysis was performed using \\textsc{Period04} software.} %results {We found 100 stars to be definitely modulated. In 25 cases, previously unknown modulation was revealed and 8 new candidates for Blazhko stars were identified. Their modulation needs to be confirmed. In 18 previously found Blazhko stars, no modulation was detectable. Multiple modulation was revealed for eight stars that were previously proposed to show simple modulation. In total, there were twelve stars with some peculiarity in their modulation in the sample. This brings the incidence rate of multiple/irregularly-modulated stars to 12~\\%. The ratio of the modulation periods of five of the double-modulated stars was within the ratios of small integers. One of stars studied, IK Hya, showed a very interesting frequency spectrum, which we interpret as changing Blazhko period between 71.81 and 75.57~days and an additional 1403-day-long cycle analogous to a four-year cycle of the prototype RR Lyr. The limits of the shorter period produce a beating period that is approximately twice as big as a 1403-day period. The newly revealed Blazhko star RZ~CVn seems to undergo changes in the amplitude of the modulation, as well as in the basic pulsation and Blazhko periods. We found that the incidence rate of the Blazhko RR Lyraes is at least 31~\\%, more likely even higher. It was also found that the majority of the Blazhko variables show triplet structures in their frequency spectra and that in 89~\\% of these cases, the peak with larger amplitude is on the right-hand side of the main pulsation component.} {} ", "introduction": "Data from automatic sky surveys provide an invaluable opportunity for finding new variables and for long-term monitoring of the sky. The most extensive survey of Galactic field variable star research are without doubt Polish ASAS \\citep[\\textit{All Sky Automated Survey}; see e.g.][]{pojmanski1997,pojmanski2002} and SuperWASP \\citep[\\textit{Super Wide Angle Search for Planets}, hereafter SW; ][]{pollaco2006,butters2010} managed by the University of Leicester. Both projects have been run for several years. Field RR Lyraes have been studied many times on the basis of various sky surveys. Among others, \\citet{kinemuchi2006} and \\citet{wils2006} utilized data from \\textit{Northern sky variability survey} \\citep[NSVS, ][]{wozniak2004}, e.g. \\citet{kovacs2005}, \\citet{szczygiel2009} studied RR~Lyraes based on ASAS data, and \\citet{drake2013} dealt with RR Lyraes measured in \\textit{Catalina sky survey}\\footnote {http://www.lpl.arizona.edu/css}. Probably the most precise ground-based survey data were provided by \\textit {the Konkoly Blazhko survey}, which have been operated since 2004 \\citep[e.g.][]{jurcsik2009a,sodor2007,sodor2012}. Crucial advances in Blazhko effect research have been made recently by space missions \\textit{CoRoT} and \\textit{Kepler}. Ultra-precise data provided by these space telescopes allowed findings of, say, cycle-to-cycle variations now known as period doubling \\citep{szabo2010,kolenberg2010,benko2010} or long-term changes in the Blazhko effect and excitation of additional modes \\citep{poretti2010,chadid2010,guggenberger2012}. Large samples of Blazhko RR Lyrae stars in galactic bulge were also studied in the framework of MACHO and OGLE surveys \\citep{soszynski2011, moskalik2003}. Since the stars in the bulge are faint and are not included in ASAS or SW surveys, we did not deal with them. Special papers based on ASAS data, which are at least partially devoted to Blazhko stars in our Galaxy, were published by \\citet{kovacs2005}, \\citet{wils2005} (hereafter WS05), and \\citet{szczygiel2007} (hereafter SF07), who all revealed many new Blazhko variables and gave a list of them. An extensive study of the Blazhko effect is presented in the paper of \\citet{leborgne2012}, which is based on observations made with \\textit{TAROT} telescopes \\citep{klotz2008,klotz2009}. \\citet{leborgne2012} also made use of $O-C$ information stored in the \\textit{GEOS RR Lyr database} \\citep{leborgne2007} and gave a list of galactic field Blazhko stars. A regularly updated database of known Blazhko RR Lyraes situated in the galactic field \\citep[\\textit{BlaSGalF}, ][]{skarka2013b}\\footnote{http://physics.muni.cz/$\\sim$blasgalf/} recently contained about 270 objects. There are a few studies of particular RR Lyraes proceeded from SW mesurements, e.g. the TV Boo study made by \\citet{skarka2013a}, or \\citet{srdoc2012}, who revealed the Blazhko effect in GSC02626-00896, but a comprehensive study of RR Lyraes based on SW data has not been performed yet. RR Lyraes with appropriate ASAS and SW data were in the spotlight. The goal was to search for the presence of the Blazhko modulation and to describe its characteristics (pulsation and modulation periods and multiplicity/irregularity of the modulation). We applied commonly used techniques of light curve and frequency spectra examination to complement and extend the list of known Blazhko stars with bright RR Lyraes (up to 12.5 mag in maximum light). To be as precise as possible, we carefully cleaned and analysed each light curve and frequency spectra individually. ", "conclusions": "A period study of 321 fundamental mode galactic field RR Lyraes with brightness at maximum light higher than 12.5 mag was performed based on available data from the ASAS and/or SW database. These stars were chosen from a more extended sample consisting of 557 stars after careful pre-selection based on conditions described in sec. 2. Each light curve was cleaned and analysed individually. In addition, each frequency spectrum was scanned through careful visual inspection to avoid omitting any sign of the Blahzko effect. One hundred stars were identified and confirmed to be Blazhko variables (25 previously unknown). Eight stars, marked as Blazhko candidates, showed some ambiguous indications of modulation like characteristic scattering around maximum of phased light curve. In contrast no indication of the Blazhko behaviour was found in 18 stars, which were supposed to be modulated. However, five of these eighteen stars are certain Blazhko variables, but with too small amplitude of modulation to be detectable in ASAS or SW data. The incidence rate of RRab Blazhko stars from our analysis is at least 31~\\%, which is a much higher percentage than given by previous studies based on data from surveys with similar data quality. This increase would be logically expected, since more time-extended datasets were examined (in many cases simultaneously in ASAS and SW surveys). However, it is still less than the incidence rates determined in the framework of precise ground-base and space surveys. With certainty there remained undetected low-amplitude Blazhko variables, as well as stars with the shortest and the longest modulation periods, simply due to nature of the data from ASAS and SW surveys. A surprising result was a very high incidence of stars with some peculiarity in their Blazhko effect -- each eighth star of our Blazhko sample showed signs of the changes in Blazhko effect or compound modulation. An interesting examples are IK Hya and RZ CVn. Both stars probably undergo changes in their pulsation and modulation characteristics. IK Hya could be another star with a few-year-long cycle analogical to four-year cycle in RR~Lyrae itself. Five of ten identified multiple-modulated stars showed the ratio between their modulation periods in ratio of small integers. Our study confirmed that classification and statistics of modulated stars according to their side peaks, which are based on low-quality data, are not reliable. It was also shown that the sensitivity of automatic procedures used to identify the Blazhko effect is insufficient -- one fourth of the Blazhko variables were identified for the first time. It is likely that the incidences and characteristics of Blazhko variables identified in other surveys \\citep[e.g. in LMC based on OGLE, ][]{alcock2003} could be slightly different than proposed. Our conclusions should serve as a challenge for developing better tools for automatic data analysis, because a manual examination of\tmany thousand stars is impossible to perform." }, "1402/1402.1329_arXiv.txt": { "abstract": "We study an isothermal system of semi-degenerate self-gravitating fermions in general relativity. Such systems present mass density solutions with a central degenerate core, a plateau and a tail which follows a power law behaviour $r^{-2}$. The different solutions are governed by the free parameters of the model: the degeneracy and temperature parameters at the center, and the particle mass $m$. We then analyze in detail the free parameter space for a fixed $m$ in the keV regime, by studying the one-parameter sequences of equilibrium configurations up to the critical point, which is represented by the maximum in a central density ($\\rho_0$) Vs. core mass ($M_c$) diagram. We show that for fully degenerate cores, the known expression for the critical core mass $M_c^{cr}\\propto m_{pl}^3/m^2$ is obtained, while instead for low degenerate cores, the critical core mass increases showing the temperature effects in a non linear way. The main result of this work is that when applying this theory to model the distribution of dark matter in galaxies from the very center up to the outer halos, we do not find any critical core-halo configuration of self-gravitating fermions, which be able to explain the super massive dark object in their centers together with an outer halo simultaneously. ", "introduction": "Systems of self-gravitating semi-degenerate fermions in general relativity were studied in \\cite{rr} and more recently with applications to dark matter in galaxies in \\cite{charly}. It was shown that, for a given central temperature parameter ($\\beta_0$) in agreement with the corresponding observed halo circular velocity, there are lower bounds for the central degeneracy parameter ($\\theta_0$) and particle mass ($m\\gtrsim0.4$ keV) above which the \\textit{observed} halo mass and radius are fulfilled. The density profiles solutions in this approach present a novel core-halo morphology composed by a quantum degenerate core followed by a low degenerate plateau until they reach the $r^{-2}$ Boltzmannian regime. This interesting overall morphology provides the flat rotation curves in the outermost part of the galaxies as well as a possible alternative to massive black holes in their centers (see \\cite{rrchina} and \\cite{firstwork}). The system of Einstein equations are written in a spherically symmetric space-time metric $g_{\\mu \\nu}={\\rm diag}(e^{\\nu},-e^{\\lambda},-r^2,-r^2\\sin^2\\theta)$, where $\\nu$ and $\\lambda$ depend only on the radial coordinate $r$, together with the thermodynamic equilibrium conditions of Tolman \\cite{tolman}, and Klein \\cite{klein}, \\begin{equation} e^{\\nu/2} T=const.\\, , \\quad e^{\\nu/2}(\\mu+m c^2)=const, \\nonumber \\end{equation} where $T$ is the temperature, $\\mu$ the chemical potential, $m$ the particle mass and $c$ the speed of light. We then write the system of Einstein equations in the following dimensionless way, \\begin{align} &\\frac{d\\hat M}{d\\hat r}=4\\pi\\hat r^2\\hat\\rho \\label{eq:1}\\\\ &\\frac{d\\theta}{d\\hat r}=\\frac{\\beta_0(\\theta-\\theta_0)-1}{\\beta_0} \\frac{\\hat M+4\\pi\\hat P\\hat r^3}{\\hat r^2(1-2\\hat M/\\hat r)}\\\\ &\\frac{d\\nu}{d\\hat r}=\\frac{\\hat M+4\\pi\\hat P\\hat r^3}{\\hat r^2(1-2\\hat M/\\hat r)} \\\\ &\\beta_0=\\beta(r) e^{\\frac{\\nu(r)-\\nu_0}{2}}\\, . \\label{eq:2} \\end{align} The variables of the system are the mass $M$, the metric factor $\\nu$ , the temperature parameter $\\beta=k T/(m c^2)$ and the degeneracy parameter $\\theta=\\mu/(k T)$. The dimensionless quantities are: $\\hat r=r/\\chi$, $\\hat M=G M/(c^2\\chi)$, $\\hat\\rho=G \\chi^2\\rho/c^2$ and $\\hat P=G \\chi^2 P/c^4$, with $\\chi=2\\pi^{3/2}(\\hbar/mc)(m_p/m)$ and $m_p=\\sqrt{\\hbar c/G}$ the Planck mass. The mass density $\\rho$ and pressure $P$ are given by Fermi-Dirac statistics (see also \\cite{charly}). This system is solved for a fixed particle mass $m$ in the keV range, with initial conditions $M(0)=\\nu(0)=0$, and given parameters $\\theta_0>0$ (depending on the chosen central degeneracy), and $\\beta_0$. We thus construct a sequence of different thermodynamic equilibrium configurations where each point in the sequence has different central temperatures $T_0$ and central chemical potential $\\mu_0$, so that satisfy the $\\theta_0$ fixed condition. Defining the core radius $r_c$ of each equilibrium system at the first maximum of its rotation curve, or equivalently at the degeneracy transition point in which $\\theta(r_c)=0$, we represent the results obtained for each sequence in a central density ($\\rho_0$) vs. core mass ($M_c$) diagram (see Fig.\\ref{fig:1}). It is shown that the critical core mass $M_c^{cr}$ is reached at the maximum of each $M_c(\\rho_0)$ curve. \\begin{figure}[!hbtp] \\centering \\includegraphics[width=\\linewidth]{criticalMassT.eps} \\caption{Different sequences of equilibrium configurations plotted in a central density ($\\rho_0$) Vs. core mass ($M_c$) diagram. The critical core mass is reached at the maximal value of $M_c$. Each sequence is built for selected values of $\\theta_0=\\mu_0/kT_0$ and different values of $T_0, \\mu_0$ varying accordingly.} \\label{fig:1} \\end{figure} It is important to emphasize that we are not interested in follow the history of equilibrium states of one specific system. Thus, the standard stability analysis as done for compact stars or in dense stellar cluster (see e.g. \\cite{rrcluster}), which is based on the constancy of the entropy per nucleon (S/N) along the equilibrium sequence of a given configuration, does not apply here. Nonetheless, in computing the $M_c(\\rho_0)$ curves in Fig.~\\ref{fig:1} we have explored the full range of $\\theta_0>1$ and $\\beta_0>10^{-10}$ parameters (including the critical ones). Then the equilibrium sequences with constant specific entropy (S/N), which differ from the ones with constant $\\theta_0$ considered here, necessarily must be contained within the full ($T_0,\\mu_0$) parameter space covered in Fig.~\\ref{fig:1}. In Table~\\ref{table:1} we show a set of central critical parameters of the model together with the correspondent critical core masses, for a very wide range of fixed central degeneracy parameters $\\theta_0$ and $m=8.5$ keV$/c^2$. \\begin{table} \\begin{ruledtabular} \\begin{tabular}{cccc} $\\theta_0$ & $\\beta_0^{cr}$ & $\\mu_0^{cr}/mc^2$ & $M_c^{cr} (M_\\odot)$ \\\\ \\colrule 1 & $6.45\\times10^{-2}$ & $6.45\\times10^{-2}$ & $1.59\\times10^{10}$ \\\\ 5 & $2.23\\times10^{-2}$ & $1.11\\times10^{-1}$ & $7.91\\times10^9$ \\\\ 40 & $8.33\\times10^{-3}$ & $3.33\\times10^{-1}$ & $7.44\\times10^9$ \\\\ 55 & $6.06\\times10^{-3}$ & $3.33\\times10^{-1}$ & $7.44\\times10^9$ \\\\ 100 & $3.33\\times10^{-3}$ & $3.33\\times10^{-1}$ & $7.44\\times10^9$ \\\\ \\end{tabular} \\caption{Critical temperature parameter and normalized chemical potential at the center of each different critical configuration, for different fixed central degeneracies.} \\end{ruledtabular} \\label{table:1} \\end{table} Defining the halo radius of each configuration at the onset of the flat rotation curve, we show in Table II the critical halo magnitudes $r_h^{cr}$, $M_h^{cr}$ and $v_h^{cr}$ corresponding to the same set of critical parameters as given in Table I. \\begin{table} \\begin{ruledtabular} \\begin{tabular}{cccc} $\\theta_0$ & $r_h^{cr} (pc)$ & $M_h^{cr}/mc^2 (M_\\odot)$ & $v_h^{cr} (km/s)$ \\\\ \\colrule 1 & $4.4\\times10^{-1}$ & $5.7\\times10^{11}$ & $7.5\\times10^4$ \\\\ 5 & $4.0\\times10^{-1}$ & $4.3\\times10^{11}$ & $6.2\\times10^4$ \\\\ 40 & $4.3\\times10^{3}$ & $1.1\\times10^{15}$ & $3.3\\times10^4$ \\\\ 55 & $2.9\\times10^{5}$ & $6.0\\times10^{16}$ & $2.9\\times10^4$ \\\\ 100 & $2.0\\times10^{11}$ & $2.3\\times10^{22}$ & $2.2\\times10^4$ \\\\ \\end{tabular} \\caption{Critical halo magnitudes of different critical configurations, for different fixed central degeneracies as given in Table I.} \\end{ruledtabular} \\label{table:2} \\end{table} The results obtained in Tables I and II imply a marked division in two different families depending on the value of $M_c^{cr}$. \\textit{i}) The first family: the critical mass has roughly a constant value $M_c^{cr}=7.44\\times10^9 M_\\odot$. This family corresponds to large values of the central degeneracy ($\\theta_0\\geq40$), where the critical temperature parameter is always lower than $\\beta_0^{cr}\\lesssim8\\times10^{-3}$ and the critical chemical potential $\\mu_0^{cr}\\approx$ const. Physically, these highly degenerate cores are entirely supported against gravitational collapse by the degeneracy pressure. In this case the critical core mass is uniquely determined by the particle mass according the relation $M_c^{cr}\\propto m_{pl}^3/m^2$ (see also section III). \\textit{ii}) The second family: the critical core mass increases from $M_c^{cr}=7.44\\times10^9 M_\\odot$ up to $M_c^{cr}\\sim10^{10} M_\\odot$. This case corresponds to critical cores with a lower central degeneracy compared with the former family ($1<\\theta_0<40$). Here the critical temperature parameter ($\\beta_0\\sim10^{-2}$), is closer to the relativistic regime with respect to the first family. This result physically indicates that the thermal pressure term has now an appreciable contribution to the total pressure, which supports the critical core against gravitational collapse. In this case $M_c^{cr}$ is completely determined by the particle mass $m$, the central temperature $T_0^{cr}$ and the central chemical potential $\\mu_0^{cr}$ (see section III). In Figs.~(\\ref{fig:2}) and (\\ref{fig:3}) we show a critical metric factor $e^{\\nu/2}$ and a critical temperature $kT$ as function of the radius for the two different families mentioned above. \\begin{figure} \\centering \\includegraphics[width=.9\\linewidth]{nutemphot.eps} \\caption{The critical temperature profile of the system (in keV) and the critical metric, for $\\theta_0=5$ and $\\beta_0^{cr}=2.23\\times 10^{-2}$. The dashed line corresponds to the isothermality condition, $Te^{\\nu/2}=const$.} \\label{fig:2} \\end{figure} \\begin{figure} \\centering \\includegraphics[width=.9\\linewidth]{nutemp.eps} \\caption{The critical temperature of the system (in keV) and the critical metric, for $\\theta_0=55$ and $\\beta_0^{cr}=6.06\\times10^{-3}$. The red line corresponds to the isothermality condition, $Te^{\\nu/2}=const$.} \\label{fig:3} \\end{figure} ", "conclusions": "" }, "1402/1402.5166_arXiv.txt": { "abstract": "Inferences of sub-surface flow velocities using local domain ring-diagram helioseismology depend on measuring the frequency splittings of oscillation modes seen in acoustic power spectra. Current methods for making these measurements utilize maximum-likelihood fitting techniques to match a model of modal power to the spectra. The model typically describes a single oscillation mode, and each mode in a given power spectrum is fit independently. We present a new method that produces measurements with greater reliability and accuracy by fitting multiple modes simultaneously. We demonstrate how this method permits measurements of sub-surface flows deeper into the Sun while providing higher uniformity in data coverage and velocity response closer to the limb of the solar disk. While the previous fitting method performs better for some measurements of low-phase-speed modes, we find this new method to be particularly useful for high phase-speed modes and small spatial areas. ", "introduction": "\\label{sec:intro} Helioseismology determines the structure and dynamics of the solar interior through analysis of seismic waves observed at the surface. Ring-diagram helioseismology (\\opencite{hill_1988}; \\opencite{basu_1999}; \\opencite{haber_2002}) investigates sub-surface horizontal flows by measuring the direction-dependent frequency shift of oscillation modes. The Doppler shift of the frequency of an oscillation mode due to a sub-surface flow is expressed as \\begin{equation} \\delta \\omega_{n}(\\bvec{k}) = \\bvec{k} \\cdot \\bvec{u}_{n}(k), \\label{eqn:freqshift} \\end{equation} where $\\bvec{k}$ and $n$ are the horizontal wavenumber and radial order of the oscillation mode, and $\\bvec{u}_{n}(k)$ is a spatial average of the horizontal velocity within the Sun, \\begin{equation} \\bvec{u}_{n}(k) = \\langle \\bvec{v}(\\bvec{r}) \\rangle = \\int \\! K_{n}(\\bvec{r};k) \\ \\bvec{v}(\\bvec{r}) \\ \\mathrm{d}^{3}r. \\label{eqn:kerneldef} \\end{equation} Here, $\\bvec{v}(\\bvec{r})$ is the true horizontal sub-surface flow velocity at any point \\noparen{$\\bvec{r}$} in the Sun, and $K_{n}(\\bvec{r};k)$ is the weighting function\\emdash or sensitivity kernel\\emdash associated with each mode ($k$,$n$), which describes the spatial extent over which the true velocity is averaged to create a single frequency shift \\cite{birch_2007}. The variation of the frequency shift as a function of direction (the frequency splitting) is measured to provide an estimate of $\\bvec{u}_{n}(k)$. While the interpretation of the frequency splitting is straightforward, the method of extracting it from the data is not. Traditionally, a model of the spectral power is fit to oscillation modes visible in power spectra of line-of-sight velocity observed in the photosphere. The model accounts for a frequency shift of the modal power as a function of horizontal direction, and this shift is directly translated into a velocity as shown in Equation \\eqnref{eqn:freqshift}. The specifics of the fitting procedure determine how well the frequency splittings are measured, as well as what other qualities of the power spectra are taken into account. There are currently two commonly used fitting procedures in the \\textit{Helioseismic and Magnetic Imager} (HMI) Ring-Diagram Pipeline \\cite{bogart_2011a}. The first method considered in this article is one introduced by \\inlinecite{haber_2002}, which fits a frequency-shifted Lorentzian model to individual modes. Since this method analyzes single ridges of modal power sequentially, we refer to it as the Single-Ridge Fitting method (SRF). The second method used in the HMI Pipeline, which will not be considered here, also fits modes independently \\cite{basu_1999}, but uses a significantly different power model that includes asymmetries in modal power, two background terms, and a host of other wavenumber- and direction-dependent modifications. In this article, we present a new fitting method that utilizes a model similar to that used in the SRF method, but modified to permit multiple radial orders to be fit simultaneously. The development of this Multi-Ridge Fitting (MRF) method is an attempt to improve upon the performance of the SRF method in terms of reliability and accuracy of measured frequency splittings. We present a comparison of the frequency splittings (interpreted as an average velocity) from a common data set processed with both fitting methods. We focus on the performance for measurements of modes that reach deepest into the Sun and for measurements made near the solar limb. As metrics of reliability and accuracy we consider the fit success rate, the typical random errors in the measurements, and how well each method recovers a known velocity. In Section \\ref{sec:data} we outline the data-processing steps taken to prepare solar observations for helioseismic mode fitting. In Section \\ref{sec:fitting} we describe the SRF and MRF fitting methods and how they differ in procedure. In Section \\ref{sec:comparison} we compare the performance of the two fitting methods using a common data set. In Section \\ref{sec:discussion} we discuss the implications of the results in the context of improving accuracy and data coverage in ring-diagram helioseismology. ", "conclusions": "\\label{sec:discussion} Extracting frequency shifts that are interpreted as sub-surface velocities is done through fitting a model of acoustic power to a three-dimensional power spectrum. By fitting a model that includes various effects seen in the data to different segments of the three-dimensional spectrum, many unique fitting methods can be created. We have presented two such fitting procedures that differ both in the model used and the selection of data considered in a single optimization. These methods use identical expressions for the velocity-induced frequency splitting for each oscillation mode, yet produce significantly different results for the velocity measurements themselves. \\subsection{Improvements in Depth for Inversions} The MRF method is able to obtain a greater number of velocity measurements for higher phase-speed modes in a given tile size than the SRF method. The increase in the success rate is also prominent for small tile sizes, suggesting that the treatment of mode blending plays an important role. Smaller tiles exhibit more leakage of modal power across wavenumbers than larger tiles due to spatial apodization. The spreading of power increases the effective width of modes at a single wavenumber. High-phase-speed modes are already spaced closer in wavenumber than other modes, enhancing the mode blending for their region of each spectrum. The SRF model does not accommodate any overlap of power from neighboring modes when fitting a single mode, causing a disparity between model and data that worsens at higher phase-speeds and smaller tile sizes. The MRF method, in accounting for mode overlap, returns a higher number of successful measurements for these cases that can be used for studying deep flows and high horizontal resolution flows. Referring to Equation \\eqnref{eqn:kerneldef}, the set of measurements obtained from a single tile \\noparen{$\\bvec{u}_n(k)$} are measures of the true sub-surface velocity \\noparen{$\\bvec{v}(\\bvec{r})$} within a single region \\noparen{$d^3r$} weighted with different sensitivity kernels \\noparen{$K(\\bvec{r})$}. The sensitivity kernels generally do not have sufficient isolation of sensitivity at any given depth, so while having a larger number of high-phase-speed measurements results in reaching deeper into the Sun, these measurements alone do not provide a localized determination of sub-surface flows. The velocity measurements from many modes can be merged through an inversion to produce an estimate of the true sub-surface flow velocity in an isolated region. After finding a linear combination of sensitivity kernels that produces a single isolated peak of sensitivity at a target location, the estimated velocity for that region is constructed using the same linear combination of the associated velocity measurements: \\begin{equation} \\bvec{w} = \\sum_{i} a_{i} \\bvec{u}_{i}. \\label{eqn:invsoln} \\end{equation} \\noindent Here $\\bvec{w}$ is the estimated sub-surface flow velocity at a selected target point, $\\bvec{u_{i}}$ are the velocity measurements obtained from each mode $(n,k)$, and $a_{i}$ are the coefficients of the linear combination determined by the inversion. To estimate the localization of each solution point, we can construct an averaging kernel that represents the spatial distribution of sensitivity for a solution point, just as a sensitivity kernel does for measured velocities: \\begin{equation} H(\\bvec{r}) = \\sum_{i} a_{i} K_{i}(\\bvec{r}). \\label{eqn:avgkercombination} \\end{equation} The set of modes that we are able to fit reliably using either method determine the basis set of kernels that can be used to construct a locally isolated averaging kernel $H(\\bvec{r})$. It is not only how high in phase-speed the fitting method can reach, but also how reliably it can obtain successful measurements at all phase-speeds, that determines how well we can constrain sub-surface flows. To determine how well averaging kernels can be constructed using the mode set available from each fitting method operating on $16^{\\circ}$ tiles, we use a simple 1D optimally localized averaging (OLA) inversion (\\opencite{dalsgaard_1990}; \\opencite{schou_1998}). The target functions are Gaussian profiles in depth and we restrict the set of sensitivity kernels available for inversion to those with a $\\mathrm{80\\,\\%}$ or higher success rate. Figure \\ref{fig:kernels} compares averaging kernels constructed from each kernel set to the target functions used. \\begin{figure}[] \\centering \\includegraphics{kernels2_16.ps} \\caption{(a) Target functions for the 1D OLA inversion. (b) Averaging kernels constructed from sensitivity kernels from the SRF mode set. (c) Averaging kernels constructed from sensitivity kernels from the MRF mode set. The colors indicate different target depths. The SRF method is only able to produce averaging kernels down to a depth of about $30 \\ \\mathrm{Mm}$, while the MRF method reaches down to $45 \\ \\mathrm{Mm}$.} \\label{fig:kernels} \\end{figure} For shallow targets, both sets are able to accurately recreate the Gaussian target function with minimal sidelobes in depth. As the target depth increases, the number of kernels available for constructing averaging kernels decreases. Past a target depth of 30 Mm, the SRF set does not contain enough kernels to create a sufficiently isolated peak of sensitivity. The MRF set still contains enough unique kernels that reach past this depth to construct reasonable averaging kernels down to 45 Mm. While the ability to construct reasonable averaging kernels through inversion is crucial for interpreting sub-surface flow results, this ignores the influence of measured errors. For OLA inversions, an optimization is done to balance the shape of the averaging kernel against the propagated error: \\begin{equation} \\sigma^2 = \\sum_{i} ( a_{i} \\sigma_{i} )^2. \\label{eqn:inverr} \\end{equation} In this way, it is both the available mode set and the associated errors that determine the usefulness of the inversion results. Figure \\ref{fig:inverr} shows the propagated inversion error as a function of the averaging kernel center-of-gravity depth for each fitting method. To mimic the result of inverting an average of many days of data over all longitudes to look at large-scale mean flows, the error for each mode has been scaled: \\begin{equation} \\sigma_i \\rightarrow \\frac{\\sigma_{i}}{\\sqrt{N_d N_l r_i}}, \\label{eqn:inverr2} \\end{equation} where $r_i$ is the success rate for a given mode, $N_d$ is the number of days to average over, and $N_l$ is the number of longitudes to average over. We set $N_d$ to 180 to mimic inversions of six months of data and $N_l$ to 17 for the typical number of tiles spanning the solar disk in longitude. The various curves for different regions of the solar disk demonstrate how much the dependence of the success rate and average error on disk position affect inversion results. The sharp downturn seen at the deep end of each curve indicates where the averaging kernels no longer resemble the target functions and are therefore useless. Due to the higher success rate for each mode, the MRF method is able to produce useful inversion results to a greater depth than the SRF method. As the distance from disk center increases, the MRF results remain mostly constant while the SRF results start to deteriorate both in error magnitude and kernel depth. Despite larger average error in the MRF results for shallow modes seen in Figure \\ref{fig:avgerr}, the higher success rate for these modes causes the inversion errors to not show a significant increase relative to the SRF results. \\begin{figure}[] \\centering \\includegraphics{inverr.ps} \\caption{Typical error magnitudes propagated through 1D OLA inversion as a function of depth for multiple distances from disk center. The sharp turnoff at the deep end of each curve demonstrates where proper averaging kernels can no longer be produced. The inversions done with SRF results tend to turn off between 33 and 35 Mm depth while those done with MRF results turn off between 38 and 40 Mm.} \\label{fig:inverr} \\end{figure} \\subsection{Spatial Variability and Implications} Anisotropy along the azimuthal direction of the power spectrum appears to be the primary cause of fitting problems near the solar limb. The SRF method tackles this issue by operating on spectra that have been ``flattened\" along azimuth. This step not only amplifies the modal power in certain directions in order to eliminate anisotropy, but also amplifies the noise. By accounting for the natural variation of power along each ridge as well as in the background power, the MRF method is able to push closer to the solar limb. As seen in Figure \\ref{fig:successdisk}, the mode set measureable by the SRF method is position-dependent. The higher-order modes see a drop in success rate closer to disk center than lower-order modes. The depth to which one can measure is then shallower near the limb. The MRF method, in contrast, allows for consistent determinations of sub-surface flows within $75^{\\circ}$ of disk center. Large tile sizes tend to provide the most accurate velocity measurements for both fitting methods, with a slight tendency for both to overestimate at high latitudes (Figure \\ref{fig:track4}). Both methods exhibit spatial variability in the velocity response for smaller tile sizes, but only for low-phase-speed modes. The MRF response is radially symmetric with the worst velocity response occurring near disk center, while the worst regions of the SRF response appear far from disk center and have opposite signs between the north--south and east--west direction. While the magnitude of this variability is relatively small, there are significant implications when attempting to analyze small residual flows left over from the subtraction of large-scale flows. Tiles at high latitudes tracked at the Carrington rate see an overall flow speed of nearly $200 \\ \\mathrm{m \\ s^{-1}}$ due to differential rotation. The systematic underestimation seen in the SRF method introduces an anomalous retrograde flow of around $20 \\ \\mathrm{m \\ s^{-1}}$, similar to the magnitude of the center-to-limb velocity systematic for low-phase-speed modes \\cite{greer_2013}. It is unclear what causes the systematic inaccuracy in either fitting method, although the strong dependence on disk position suggests a possible coupling between velocity measurements and power anisotropy. As mentioned in Section \\ref{sec:mrf}, there is a significant increase in the computational cost when switching from the SRF method to the MRF method. While processing a day's worth of tiles with the SRF method is typically one of the fastest steps in the HMI Pipeline, the MRF method brings the process of fitting more in line with the time needed for tracking. It is important to consider when this additional computational burden is justified. The SRF method produces high-quality results for low-phase-speed modes and for large tiles near disk center. This leaves three distinct cases where the MRF method provides improvements: measuring higher-phase-speed modes, pushing closer to the limb, and using small tiles. The implications of this new procedure on determining sub-surface flows follow these three technical improvements. For a given tile size, we are able to extend our analysis deeper into the Sun while maintaining a constant horizontal and temporal resolution. Analysis to these extended depths can be performed consistently across most of the solar disk, providing uniform coverage over a larger fraction of the Sun. The increased reliability of small tile sizes permits higher horizontal resolution analysis of sub-surface flows. \\textbf{Acknowledgements} This work was supported by NASA through NASA grants NNX08AJ08G, NNX08AQ28G, and NNX09AB04G. The data used here are courtesy of NASA/SDO and the HMI science team. SDO is a NASA mission, and the HMI project is supported by NASA contract NAS5-02139." }, "1402/1402.5399_arXiv.txt": { "abstract": "Very light WIMPs ($\\chi$), thermal relics that annihilate late in the early Universe, change the energy and entropy densities at BBN and at recombination. BBN, in combination with the CMB, can remove some of the degeneracies among light WIMPs and equivalent neutrinos, constraining the existence and properties of each. Depending on the nature of the light WIMP (Majorana or Dirac fermion, real or complex scalar) the joint BBN + CMB analyses set {\\bf lower} bounds to \\mchi~in the range $0.5 - 5\\,{\\rm MeV}$ ($m_{\\chi}/m_{e} \\gsim 1 - 10$), and they identify {\\bf best fit} values for \\mchi~in the range $5 - 10\\,{\\rm MeV}$. The {\\bf joint} BBN + CMB analysis finds a {\\bf best fit} value for the number of equivalent neutrinos, \\Deln~$\\approx 0.65$, nearly independent of the nature of the WIMP. In the absence of a light WIMP (\\mchi~$\\gsim 20\\,{\\rm MeV}$), \\neff~$= 3.05(1 + \\Deln/3)$. In this case, there is excellent agreement between BBN and the CMB, but the joint fit reveals \\Deln~$= 0.40\\pm0.17$, disfavoring standard big bang nucleosynthesis (SBBN) (\\Deln~= 0) at $\\sim 2.4\\,\\sigma$, as well as a sterile neutrino (\\Deln~= 1) at $\\sim 3.5\\,\\sigma$. The best BBN + CMB joint fit disfavors the absence of dark radiation (\\Deln~= 0 at $\\sim 95\\%$ confidence), while allowing for the presence of a sterile neutrino (\\Deln~= 1 at $\\lsim 1\\,\\sigma$). For all cases considered here, the lithium problem persists. These results, presented at the TAUP 2013 Conference, are based on \\citet{kngs}. ", "introduction": "\\label{intro} In the absence of ``extra\", equivalent neutrinos (dark radiation) or light ($\\lsim 20\\,{\\rm MeV}$), weakly interacting massive particles (WIMPs), the particle content relatively late in the early Universe is quite simple. After the \\epm pairs (and all the other more massive standard model (SM) particles) have annihilated ($T \\lsim m_{e}$), the only remaining SM particles are the CMB photons and the three SM neutrinos ($\\nu_{e}$, $\\nu_{\\mu}$, $\\nu_{\\tau}$). At these early epochs the Universe is ``radiation dominated\" and the energy density may be written as $\\rho_{\\rm R} = \\rho_{\\gamma} + 3\\,\\rho_{\\nu}$, where $3\\,\\rho_{\\nu}$ accounts for the contributions from the three SM neutrinos. During these early epochs the contributions to the total mass/energy density from the baryons (B) and the dark matter (DM), as well as any dark energy (DE), are very subdominant compared to $\\rho_{\\rm R}$. More generally, in addition to the SM neutrinos, there may be extra, ``beyond the standard model\" particles that, like the SM neutrinos, are extremely light ($\\lsim 10\\,{\\rm eV}$) and very weakly interacting. During the early (or, even, relatively late) evolution of the Universe these neutrino-like particles, so called ``equivalent neutrinos\", will contribute to the energy density. The energy density controls the early Universe expansion rate. If \\Deln~counts the contribution of equivalent neutrinos, often referred to as ``dark radiation\", $\\rho_{\\rm R} = \\rho_{\\gamma} + (3 + \\Deln)\\,\\rho_{\\nu}$. The contribution to \\Deln~of an equivalent neutrino that decouples along with the SM neutrinos (at $T = T_{\\nu d}$) will be \\Deln~= 1 for a Majorana fermion (\\eg, a sterile neutrino), \\Deln~= 2 for a Dirac fermion or, \\Deln~= 4/7 for a real scalar. In general, \\Deln~is an integer (fermions) or an integer multiple of 4/7 (bosons). However, an equivalent neutrino that is more weakly interacting than the SM neutrinos, will have decoupled earlier in the evolution of the Universe and its contribution to \\Deln~will be suppressed by the heating of the SM neutrinos (and photons) when the heavier SM particles decay and/or annihilate. Therefore, in principle, there is no reason that \\Deln~should be an integer or an integer multiple of 4/7 (for further discussion see \\citet{steig13}; for a specific example of three, very weakly coupled, right-handed neutrinos, see \\citet{haim} and for the example of a weakly coupled scalar particle see \\citet{weinberg}). After the SM neutrinos decouple, when $T = T_{\\nu d} \\approx 2 - 3\\,{\\rm MeV}$, the \\epm pairs annihilate, heating the photons but not the neutrinos. Prior to neutrino decoupling (and \\epm annihilation), the neutrinos, \\epm pairs, and the photons are in equilibrium at the same temperature, $T_{\\nu} = T_{e} = T_{\\gamma}$ but, after \\epm annihilation, the photons are hotter than the relic neutrinos (or, equivalently, the neutrinos are cooler than the photons). In most simplified analyses it is assumed that the neutrinos decoupled instantaneously and that the electrons were effectively massless at neutrino decoupling. With these approximations the late time (after \\epm annihilation is complete) ratio of neutrino and photon temperatures, $(T_{\\nu}/T_{\\gamma})_{0} = (4/11)^{1/3}$, follows from entropy conservation. The late time ratio of the energy density in one species of neutrino ($\\rho^{0}_{\\nu}$) to that in the photons is $(\\rho^{0}_{\\nu}/\\rho_{\\gamma})_{0} = 7/8\\,(T_{\\nu}/T_{\\gamma})^{4}_{0} = 7/8\\,(4/11)^{4/3}$. However, at neutrino decoupling $m_{e}/T_{\\nu d} \\approx 0.2 \\neq 0$ and, as a result, $\\rho_{\\nu}$ differs (by a small amount) from $\\rho^{0}_{\\nu}$ \\citep{steig13}. Furthermore, the neutrinos don't really decouple instantaneously. While the neutrinos are partially coupled to the annihilating \\epm pairs they share a small amount of the energy released by the annihilation \\citep{dolgov,dolgov-osc,enqvist,hannestad,mangano}. These effects can be accounted for by introducing \\neff, the ``effective number of neutrinos\" where, at late times ($T_{0} \\ll m_{e}$), $\\rho_{\\rm R\\,0} \\equiv \\rho_{\\gamma\\,0} + {\\rm N}_{\\rm eff}\\,\\rho^{0}_{\\nu\\,0}\\,,$ \\beq {\\rm N}_{\\rm eff} = 3\\,\\bigg[{11 \\over 4}\\bigg({T_{\\nu} \\over T_{\\gamma}}\\bigg)^{3}_{0}\\bigg]^{4/3}\\bigg(1 + {\\Deln \\over3}\\bigg)\\,. \\eeq It should be kept in mind that while the relative contributions of neutrinos and photons to the total radiation density may be evolving before and during BBN, \\neff~is a ``late time\" quantity, evaluated long after BBN has ended, when the only relativistic particles remaining are the photons and the neutrinos. Under the assumptions of instantaneous neutrino decoupling and $m_{e} \\ll T_{\\nu d}$, \\neff~= 3 + \\Deln. Keeping the instantaneous decoupling approximation but correcting for the finite electron mass, \\neff~$\\approx 3.02(1 + \\Deln/3)$ \\citep{steig13}. Following the non-instantaneous neutrino decoupling and allowing for the finite electron mass, \\neff~$\\approx 3.05(1 + \\Deln/3)$ \\citep{mangano}. It should be noted that in this latter case there is, in addition, a very small, but not entirely negligible correction to the BBN predicted primordial helium abundance \\citep{mangano}. Since the expansion rate (Hubble parameter) of the radiation dominated early Universe is $H \\propto \\rho_{\\rm R}^{1/2}$, the presence of dark radiation (\\Deln~$\\geq 0$) results in a speed up to the early Universe expansion rate. \\begin{figure}[!t] \\begin{center} \\includegraphics[width=0.45\\columnwidth]{neff0_vs_mchi.pdf} \\hskip .4in \\includegraphics[width=0.45\\columnwidth]{neff_vs_mchi.pdf} \\caption{The left panel shows N$^{0}_{\\rm eff}$ (\\neff~when \\Deln~= 0) as a function of the WIMP mass for electromagnetically coupled light WIMPs in the absence of equivalent neutrinos. From bottom to top, the solid red curve is for a Dirac WIMP, the dashed green curve is for a complex scalar, the solid black curve is for a Majorana fermion and, the dashed blue curve is for a real scalar. The horizontal, red/pink bands are the Planck CMB 68\\% and 95\\% allowed ranges for N$_{\\rm eff}$. The right panel specializes to the case of a Majorana fermion WIMP, showing N$_{\\rm eff}$ as a function of the WIMP mass for \\Deln~equivalent neutrinos. The solid curve is for $\\Delta {\\rm N}_{\\nu} = 0$, the short dashed curve is for $\\Delta {\\rm N}_{\\nu} = 1$ and, the long dashed curve is for $\\Delta {\\rm N}_{\\nu} = 2$. The horizontal red bands are the Planck CMB 68\\% and 95\\% allowed ranges for N$_{\\rm eff}$, including baryon acoustic oscillations in the CMB constraint.} \\label{fig:neffvsmchi} \\end{center} \\end{figure} So far, the possibility of a very light, weakly interacting, massive particle, a WIMP $\\chi$, has been ignored in the discussion here. The difference between a WIMP and an equivalent neutrino is that a WIMP remains thermally coupled to the SM particles until after it has become non-relativistic and begins annihilating. As a result, the light WIMP annihilation heats the remaining SM particles (either the photons and, possibly, the \\epm pairs if the WIMP couples electromagnetically or, the SM neutrinos if the WIMP only couples to them). Note that in the analysis and discussion here, the WIMP {\\bf need not} be a dark matter candidate; the WIMP could be a sub-dominant component of the dark matter ($\\Omega_{\\chi} < \\Omega_{\\rm CDM}$). For example, the WIMP could be a light, millicharged particle such as that proposed by \\citep{millicharged} and discussed by Dolgov at this conference. Here, as in \\citet{kngs}, we specialize to the case of a light WIMP coupled only to the photons and \\epm pairs. The relevant role for BBN and the CMB played by such a light WIMP is that its annihilation heats the photons relative to the decoupled SM neutrinos, changing (reducing) $(T_{\\nu}/T_{\\gamma})_{0}$. After \\epm and WIMP annihilation, at fixed photon temperature, the neutrinos, SM and equivalent, are cooler than the photons. In this case, \\neff~is a function of \\mchi~(see \\citet{steig13} and references therein). Since the expansion rate of the early Universe is controlled by the energy density, any modification of \\neff~will be reflected in a non-standard expansion rate (\\eg, during BBN and at recombination). In addition, extremely light WIMPs (\\mchi~$\\lsim m_{e}$) will annihilate so late that, if their annihilation produces photons, they will modify the baryon-to-photon ratio ($\\eta_{10} = 10^{10}(n_{\\rm B}/n_{\\gamma})_{0} = 273.9\\,\\omb$) during or after BBN. BBN can probe \\neff~(through the effects of the neutrinos on the expansion rate and on the weak rates that regulate the neutron to proton ratio) as well as the universal ratio of baryons-to-photons. At late times in the early Universe, \\eg, at recombination, the CMB can also probe \\omb~and \\neff. As independent probes of the effective number of neutrinos (\\neff) or the number of equivalent neutrinos (\\Deln) and the universal baryon density (\\omb~or $\\eta_{10}$), BBN and the CMB can help to break the degeneracies among these parameters and the WIMP mass (and spin/statistics) and to constrain their allowed ranges (see, \\citet{steig13} \\& \\citet{kngs} and Fig.\\,\\ref{fig:neffvsmchi}). \\subsection{Planck CMB Constraints} In their analysis, \\citet{kngs}, whose results are described and summarized here, adopted the CMB constraints on \\omb~and \\neff~from the Planck $\\Lambda\\mathrm{CDM}+N_\\mathrm{eff}$ fit including supplementary baryon acoustic oscillation (BAO) data \\citep{planck}. The correlations between these quantities were included in \\citep{kngs} and in the analysis here. From the CMB we adopted \\omb~$= 0.0223 \\pm 0.0003$ ($\\eta_{10} = 6.11 \\pm 0.08$) and \\neff~$= 3.30 \\pm 0.27$. In Fig.\\,\\ref{fig:neffvsmchi}, the Planck 68\\% and 95\\% constraints on \\neff~are shown as a function of the WIMP mass (the CMB constraints are independent of the WIMP mass). Also shown are the curves corresponding to \\neff~as a function of \\mchi~for a Majorana fermion WIMP and for three choices of the number of equivalent neutrinos. The behavior seen here is qualitatively similar for a Dirac or scalar WIMP (see, \\eg, \\citep{steig13} \\& \\citep{kngs}). This figure illustrates the degeneracies between \\neff~and \\mchi. For example, for \\Deln~= 0 the CMB can set a {\\bf lower} bound to \\mchi. In contrast, for \\Deln~= 1 or 2, it is high values of \\mchi~that are excluded. \\subsection{BBN Constraints} Of the light nuclides produced during BBN, D and \\4he are the relic nuclei of choice. To account for, or minimize, the post-BBN modifications of the primordial abundances, observations at high redshift (z) and/or low metallicity (Z) are preferred. Deuterium (and hydrogen) is observed in high-z, low-Z, QSO absorption line systems and helium is observed in relatively low-Z, extragalactic \\hii~regions. Even so, it may still be necessary to correct for any post-BBN nucleosynthesis that may have modified their primordial abundances. The post-BBN evolution of D and \\4he is simple and monotonic. As gas is cycled through stars, D is destroyed and \\4he produced. Finally, D and \\4he provide complementary probes of the parameters (\\Deln~and \\omb) of interest. $y_{\\rm DP} \\equiv 10^{5}{\\rm (D/H)}_{\\rm P}$ is mainly sensitive to the baryon density at BBN (\\omb) and is less sensitive to \\Deln. In contrast, the \\4he mass fraction, \\Yp, is very insensitive to \\omb, but is quite sensitive to \\Deln. This complementary, nearly orthogonal, dependence of D and \\Yp~on $\\eta_{10}$ and \\Deln~is illustrated in Fig.\\,\\ref{fig:YvsD}. For the analysis here (and in \\citet{kngs}), we have adopted, $y_{\\rm DP} = 2.60 \\pm 0.12$ \\citep{pettini} and \\Yp~$= 0.254 \\pm 0.003$ \\citep{izotov}. \\begin{figure}[!t] \\begin{center} \\includegraphics[width=0.55\\columnwidth]{hevsd13c.pdf} \\caption{BBN predicted curves of constant baryon-to-photon ratio and equivalent number of neutrinos in the \\Yp~-- $y_{\\rm DP}$ plane. From left to right (blue), $\\eta_{10} = 7.0, 6.5, 6.0, 5.5$. From bottom to top (red) \\Deln~= 0, 1, 2. Also shown by the filled circle and error bars are the observationally inferred values of \\Yp~and $y_{\\rm DP}$ adopted here (see the text).} \\label{fig:YvsD} \\end{center} \\end{figure} In contrast, the post-BBN evolution of \\3he and \\7li, the other two nuclides produced in significant abundances during BBN, is complex and model dependent. \\3he has only been observed in the relatively metal-rich interstellar medium of the Galaxy and its BBN-predicted abundance is less sensitive to \\omb~and \\Deln~than is the BBN D abundance. \\3he is not used in our BBN analysis, but we have confirmed that its observationally inferred primordial abundance \\citep{rood} is in good agreement with our BBN-predicted results. \\7li suffers from some of the same issues as \\3he. Although \\7li is observed in very metal poor stars, its post-BBN evolution is complicated and model dependent, especially the connection between the surface lithium abundances observed and those in the gas out of which the stars formed. While in principle \\7li could be as useful as D in constraining \\omb~(and, to a lesser extent, \\Deln), there is the well known ``lithium problem\" (see, \\eg, \\citet{fields} and \\citet{spite} for recent reviews) that, as will be seen below, persists. In the BBN analyses, with and without a light WIMP, only D and \\4he are used to constrain \\omb~and \\Deln~(or, \\neff) and these BBN constraints are compared to the independent constraints from the CMB. ", "conclusions": "In the absence of a light WIMP and equivalent neutrinos (no dark radiation), BBN (SBBN) depends on only one parameter, the baryon abundance. For the adopted primordial D and \\4he abundances, SBBN predicts a best fit baryon density of $\\eta_{10} = 6.0 \\pm 0.2$ \\citep{kngs}, in excellent agreement with the corresponding value of the baryon abundance ($\\eta_{10} = 6.06 \\pm 0.07$) inferred from the Planck analysis with \\neff~fixed \\citep{planck}. However, as shown in \\citet{kngs}, for either of these baryon abundances, BBN predicts \\Yp~$= 0.247\\,(\\pm\\, 0.0005)$, which is a poor fit to the observationally inferred primordial helium abundance of \\Yp~$= 0.254\\pm0.003$ \\citep{izotov}. In the absence of a light WIMP, but now allowing for dark radiation (\\Deln~and \\neff~free to vary), the effective number of neutrinos and the number of equivalent neutrinos are related by \\neff~$= 3.05(1 + \\Deln/3)$. From the Planck CMB analysis alone, \\neff~$= 3.30 \\pm 0.27$ \\citep{planck}, constraining the number of equivalent neutrinos to \\Deln~$= 0.25 \\pm 0.27$, consistent with the absence of dark radiation (\\Deln~= 0) at $\\lsim 1\\,\\sigma$ and, inconsistent with a sterile neutrino (\\Deln~= 1) at $\\sim$\\,$2.8\\,\\sigma$. In addition to the constraint on \\Deln, the CMB alone also provides a constraint on the universal baryon density, \\omb~$= 0.0223 \\pm 0.0003$ ($\\eta_{10} = 6.11 \\pm 0.08$). For the Planck determined combination of the baryon abundance and the number of equivalent neutrinos, the BBN predicted D and \\4he abundances ($y_{\\rm DP} = 2.56 \\pm 0.08$ and \\Yp~$= 0.2505 \\pm 0.0005$) are in very good agreement with the observationally inferred primordial abundances, while lithium (A(Li)~$= 2.73 \\pm 0.03$) is too high. For the same case (no light WIMP, \\Deln~free to vary), the BBN fit to the observed D and \\4he abundances is in excellent agreement with the CMB inferred parameter values (BBN: $\\eta_{10} = 6.19 \\pm 0.21$, \\Deln~$= 0.51 \\pm 0.23$, \\neff~$= 3.56 \\pm 0.23$). A joint BBN + CMB analysis predicts $\\eta_{10} = 6.13 \\pm 0.07$ (\\omb~$= 0.0224 \\pm 0.0003$) and \\Deln~$= 0.40 \\pm 0.17$ (\\neff~$= 3.46 \\pm 0.17$). BBN with these joint fit parameter values predicts $y_{\\rm DP} = 2.60 \\pm 0.08$ and \\Yp~$= 0.2525 \\pm 0.0005$. The corresponding lithium abundance, A(Li)~$= 2.72 \\pm 0.03$, is a factor of $\\sim$ 3 higher than the observationally inferred, primordial value. In the absence of a light WIMP, BBN and the CMB are in excellent agreement, but neither \\Deln~$= 0$ (SBBN) nor \\Deln~$= 1$ (a sterile neutrino) is favored by BBN or by the combined BBN + CMB fit. In the presence of a sufficiently light WIMP (\\mchi~$\\lsim 20\\,{\\rm MeV}$) the CMB results are unchanged, although the connection between \\neff~and \\Deln~is modified depending on the WIMP mass, \\neff~$= {\\rm N}^{0}_{\\rm eff}(\\mchi)(1 + \\Deln/3)$. Now there is a degeneracy between the CMB constraints on \\neff~and \\Deln~(and \\mchi). As may be seen from Fig.\\,\\ref{fig:neffvsmchi}, for some choices of \\Deln, the CMB constraint on \\neff~sets a lower limit to \\mchi, while for other choices the CMB sets an upper limit to the WIMP mass. The independent constraints from BBN help to break these degeneracies. In the presence of a light WIMP BBN now depends on three parameters: \\Deln, \\neff, \\omb~(or, \\Deln, \\mchi, $\\eta_{10}$), but there are only two BBN constraints (from D and \\4he). For each choice of \\mchi, corresponding to a fixed value of ${\\rm N}_{\\rm eff}^{0}$, there is always a pair of \\Deln~and $\\eta_{10}$ values for which BBN predicts -- exactly -- the observationally inferred primordial D and \\4he abundances adopted here. However, the corresponding BBN inferred values (and ranges) of \\neff~and \\omb~need not necessarily agree with the values (and ranges) set by the CMB. By comparing the BBN and CMB constraints, the degeneracies may be broken, leading to a lower bound, as well as a best fit value, of the WIMP mass (depending on the nature of the WIMP). For the case of a Majorana fermion WIMP shown in the figures here, \\mchi~$\\gsim 1.7\\,{\\rm MeV}$ and the best fit is for \\mchi~$ = 7.9\\,{\\rm MeV}$. Depending on the nature of the WIMP, the lower bound to \\mchi~ranges from $\\sim$\\,$m_{e}$ to $\\sim$\\,$10\\,m_{e}$, while the best fit WIMP masses lie in the range $\\sim$\\,$5 - 10\\,{\\rm MeV}$ (see \\citet{kngs}). In all cases, very nearly independent of the nature of the WIMP, N$_{\\rm eff}^{0} \\approx 2.71$ and \\Deln~$\\approx 0.65$. While the joint BBN + CMB analysis is dominated by the CMB values for \\neff~and \\omb, the presence of an additional free parameter (\\mchi) relaxes the constraints (increases the error) on \\Deln~compared to the no light WIMP case. In the presence of a sufficiently light WIMP a sterile neutrino is now permitted at $\\lsim 68\\%$ confidence (see the right hand panel of Fig.\\,\\ref{fig:neffvsomb}). However, the absence of dark radiation (\\Deln~= 0) remains disfavored at $\\sim 95\\%$ confidence. For the joint BBN + CMB analysis the BBN predicted primordial lithium abundance is A(Li)~$= 2.73 \\pm 0.03$, essentially identical to that for the no light WIMP case. The persistence of the lithium problem is largely a result of the strong coupling between the BBN predicted abundances of D and \\7li, and cannot be resolved by an extension of SBBN to include equivalent neutrinos (\\Deln~$\\neq 0$) or light WIMPs. It should be noted that since TAUP 2013, \\citet{cooke} published new results on the primordial abundance of deuterium, $y_{\\rm DP} = 2.53 \\pm 0.04$. Although their new central value agrees very well with the earlier, \\citet{pettini} result adopted here, the new uncertainty is smaller by a factor of three. In the analysis described here (and, in more detail in \\citet{kngs}), this small change in the primordial deuterium abundance has the effect of increasing $\\eta_{10}$ by $\\sim$\\,$0.1$ and decreasing \\Deln~by $\\sim$\\,$0.01$. These small changes, well within the current errors, leave the results and conclusions presented here unaffected. \\begin{center} {\\bf Acknowledgements}\\\\ \\end{center} We are grateful to the Ohio State University Center for Cosmology and Astro-Particle Physics for hosting K. M. N's visit during which most of the work described here was done. K. M. N is pleased to acknowledge support from the Institute for Nuclear and Particle Physics at Ohio University. G. S. is grateful for the hospitality provided by the Departamento de Astronomia of the Instituto Astron$\\hat{\\rm o}$mico e Geof\\' \\i sico of the Universidade de S\\~ao~Paulo, where these proceedings were written. The research of G. S. was supported at OSU by the U.S.~DOE grant DE-FG02-91ER40690." }, "1402/1402.7073_arXiv.txt": { "abstract": "We present the evolution of the structure of relaxed cold dark matter haloes in the cosmology from the Planck satellite. Our simulations cover 5 decades in halo mass, from dwarf galaxies to galaxy clusters. Due to the increased matter density and power spectrum normalization the concentration mass relation in the Planck cosmology has a $\\sim 20\\%$ higher normalization at redshift $z=0$ compared to WMAP cosmology. We confirm that CDM haloes are better described by the Einasto profile; for example, at scales near galaxy half-light radii CDM haloes have significantly steeper density profiles than implied by NFW fits. There is a scatter of $\\sim 0.2$ dex in the Einasto shape parameter at fixed halo mass, adding further to the diversity of CDM halo profiles. The evolution of the concentration mass relation in our simulations is not reproduced by any of the analytic models in the literature. We thus provide a simple fitting formula that accurately describes the evolution between redshifts $z=5$ to $z=0$ for both NFW and Einasto fits. Finally, the observed concentrations and halo masses of spiral galaxies, groups and clusters of galaxies at low redshifts are in good agreement with our simulations, suggesting only mild halo response to galaxy formation on these scales. ", "introduction": "\\label{sec:intro} In the standard theoretical framework for structure formation in the Universe, the mass-energy budget is dominated by a cosmological constant and cold dark matter (CDM). In this paradigm, initially small density perturbations grow via gravitational instability, forming bound structures known as dark matter haloes. The structure of dark matter haloes are of particular interest as they provide a non-linear scale test of the cold dark matter paradigm, cosmological parameters, and more generally for the nature of dark matter itself (e.g., Moore 1994; Flores \\& Primack 1994; de Blok \\etal 2001; Zentner \\& Bullock 2002; McGaugh 2004). They also provide the backbone for the structural properties of galaxies and galaxy scaling relations (e.g., Mo \\etal 1998; Dutton \\etal 2007, 2013). From the theoretical side there are two hurdles that need to be overcome before an accurate prediction for the structure of CDM haloes is possible: 1) halo structure is sensitive to cosmological parameters (e.g., Macci\\`o \\etal 2008) and 2) the galaxy formation process can cause haloes to both contract or expand (e.g., Di Cintio \\etal 2014). The subject of this paper is to constrain the effects of the cosmological parameters on the ``baryon free'' dark halo structure, and to quantify the evolution of the structure of CDM haloes as a population across cosmic time. This work continues on from our earlier studies (Macci\\`o \\etal 2007, 2008; Mu{\\~n}oz-Cuartas \\etal 2011), as well as numerous studies in the literature (e.g., Navarro, Frenk \\& White 1996, 1997; Bullock \\etal 2001; Eke \\etal 2001; Zhao \\etal 2003, 2009; Duffy \\etal 2008; Gao \\etal 2008; Klypin \\etal 2011; Prada \\etal 2012; Ludlow \\etal 2013a,b). As shown by Macci\\`o \\etal (2008) relatively small changes in cosmological parameters have a non-negligible effect on the structure of CDM haloes. For example, the mean concentrations of CDM haloes varied by a factor of $1.5$ between the various WMAP cosmologies (Spergel \\etal 2003, 2007). The cosmology advocated by the Planck satellite (the Planck Collaboration 2013) has a significantly higher matter density, $\\Omega_{\\rm m}$, than adopted in all previous high-resolution simulations (See Fig.~\\ref{fig:omega_sigma} and Table~\\ref{tab:cosmo_param}). Compared to the WMAP 5th year cosmology (Komatsu \\etal 2009), the Planck cosmology also has higher $\\sigma_8$. These differences are expected to result in increased dark halo concentrations (e.g., using the model of Bullock \\etal 2001). \\begin{table*} \\centering \\caption{Cosmological Parameters. All cosmologies are flat, i.e., $\\Omega_{\\Lambda}+\\Omega_{\\rm m}=1$. } \\begin{tabular}{lcccccl} \\hline Name & $\\Omega_{\\rm m}$ & $h$ & $\\sigma_8$ & $n$ & $\\Omega_{\\rm b}$ & Simulation Reference\\\\ \\hline Planck & 0.3175 & 0.671 & 0.8344 & 0.9624 & 0.0490 & This paper \\\\ WMAP5 & 0.258\\phantom{1} & 0.72\\phantom{1} & 0.796\\phantom{1}& 0.963\\phantom{1} & 0.0438 & Macci\\`o \\etal 2008\\\\ WMAP3 & 0.238\\phantom{1} & 0.73\\phantom{1} & 0.75\\phantom{11}& 0.95\\phantom{11} & 0.042\\phantom{1} & Macci\\`o \\etal 2008\\\\ WMAP1 & 0.268\\phantom{1} & 0.71\\phantom{1} & 0.90\\phantom{11}& 1.0\\phantom{111} & 0.044\\phantom{1} & Macci\\`o \\etal 2008\\\\ Millennium & 0.25\\phantom{11} & 0.73\\phantom{1} & 0.90\\phantom{11}& 1.0\\phantom{111} & 0.045\\phantom{1} & Springel \\etal 2005\\\\ Bolshoi & 0.27\\phantom{11} & 0.70\\phantom{1} & 0.82\\phantom{11}& 0.95\\phantom{11} & 0.0469 & Klypin \\etal 2011\\\\ \\hline \\end{tabular} \\label{tab:cosmo_param} \\end{table*} \\begin{figure} \\psfig{figure=fig1.eps,width=0.47\\textwidth} \\caption{Constraints on the cosmic matter density ($\\Omega_{\\rm m}$) and the power spectrum normalization ($\\sigma_8$) from the Planck Collaboration (2013) (red point with 68\\% confidence intervals). For comparison, the various sets of cosmological parameters from the WMAP satellite are shown with blue hexagons (1st, 3rd and 5th year results), and the parameters used by the Millennium and Bolshoi N-body simulations are given by the black pentagon and square, respectively.} \\label{fig:omega_sigma} \\end{figure} Traditionally the structure of CDM haloes has been described by the two parameter NFW profile (Navarro, Frenk, \\& White 1996; 1997). This has a divergent inner density profile of $\\rho(r)\\propto r^{-1}$, and an outer profile of $\\rho(r)\\propto r^{-3}$. A common parametrization uses the halo mass, $M_{200}$, and the concentration parameter, $c\\equiv r_{200}/r_{-2}$ (where $r_{200}$ is the virial radius, and $r_{-2}$ is the scale radius). The mass and concentration are correlated, with a shallow slope ($c\\propto M_{200}^{-0.1}$), and small scatter ($\\sigma_{\\log c}\\sim 0.1$), so the structure of CDM haloes is almost scale free and described by a universal profile (NFW 97; Bullock \\etal 2001). Recent work has shown that three parameters provide a more accurate description of spherically average CDM density profiles - especially at small radii ($\\sim 1\\%$ of the virial radius). The most common generalizations are to allow the inner logarithmic density slope, $\\gamma$, to be a free parameter (recall the NFW profile has $\\gamma=-1$), sometimes known as a generalized NFW profile (or gNFW); or the Einasto profile (Einasto 1965), which is $d\\ln\\rho / d\\ln r \\propto r^{\\alpha}$. A number of studies have shown that the Einasto profile provides, in general, a better description of CDM haloes than the NFW or gNFW profiles (e.g., Navarro \\etal 2004, 2010; Merritt \\etal 2005, 2006; Stadel \\etal 2009; Reed \\etal 2011). Going one step further, using stacks of CDM haloes, Gao \\etal (2008) showed that the Einasto shape parameter, $\\alpha$, depends on halo mass. As in our earlier studies we use a large suite of cosmological N-body simulations with different box sizes to cover the entire halo mass range from $\\sim 10^{10} \\hMsun$ (haloes that host dwarf galaxies) to $\\sim 10^{15} \\hMsun $ (massive clusters). We use these simulations to investigate the structure of CDM haloes across cosmic time. In estimating halo concentrations we consider fits to the density profiles using both the Einasto and NFW functions, as well as a non-parametric approximation utilizing $V_{\\rm max}/V_{200}$ following Klypin \\etal (2011). This paper is organized as follows: in \\S\\ref{sec:nbody} the simulations and the determination of the halo parameters are presented. In \\S\\ref{sec:einasto} we discuss the quality of Einasto vs NFW fits as well as different methods for measuring halo concentrations. In \\S\\ref{sec:concentration} we compare the concentration mass relation of Planck cosmology to that from WMAP, and discuss the effects of different cosmological parameters. In \\S\\ref{sec:evolution} we present the evolution of the NFW concentration mass relation, while in \\S\\ref{sec:evolution_einasto}, we present the evolution of the Einasto concentration and shape parameters. In \\S\\ref{sec:observations} we compare the concentration mass relation from our simulations with observations of spiral galaxies and clusters of galaxies. Finally, we summarize our results in \\S\\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} In this paper we have used a large set of cosmological N-body simulations to study the evolution of the structure of cold dark matter haloes over 90\\% of the age of the Universe. At redshift zero our simulations span five orders of magnitude in halo mass ($10^{10}-10^{15} \\hMsun$), covering haloes those that host individual dwarf galaxies to those associated with massive clusters. We adopt the cosmological parameters derived from the first data release of the Planck Satellite (the Planck Collaboration 2013). We summarize our results as follows: \\begin{itemize} \\item The concentration mass relation in the Planck cosmology has a 20\\% higher normalization (at redshift $z=0$) than in the WMAP 5th year cosmology. Despite significant differences in cosmological parameters the Planck concentration mass relation is very similar to that from the WMAP1 cosmology. By coincidence, the increased $\\Omega_{\\rm m}$ (in the Planck vs WMAP1 cosmology) is almost perfectly balanced by the decrease in $\\sigma_8, n$, and $h$. \\item Propagating the uncertainties in cosmological parameters given by the Planck Collaboration results in just a 3.5\\% uncertainty in the concentrations of Milky Way mass haloes at redshift $z=0$, which is smaller than typical systematic uncertainties in measuring halo concentrations. \\item In agreement with previous studies we find that the spherically averaged density profiles of CDM haloes are better described by the Einasto (1965) profile than the NFW (1997) profile. For example, between 2\\% and 4\\% of the virial radius, simulated haloes of mass $M_{200}\\sim10^{13}\\hMsun$ (which host massive elliptical galaxies) have average logarithmic density slopes $d\\log\\rho/d\\log r \\simeq -1.6$, compared to $\\simeq -1.4$ for NFW fits. \\item At fixed redshift the average Einasto shape parameter, $\\alpha$, increases with halo mass, from $\\alpha\\sim 0.16$ for dwarf haloes to $\\alpha\\sim 0.25$ for cluster haloes. At fixed halo mass the average, $\\alpha$, increases systematically with redshift. This evolution is well described by the relation between $\\alpha$ and dimensionless peak height $\\nu(M,z)=\\delta_{\\rm crit}(z)/\\sigma(M,z)$ proposed by Gao \\etal (2008) --- who used simulations in the WMAP1 cosmology. \\item The distribution in $\\alpha$ about the $\\alpha-\\nu$ relation is well described by a log-normal function, with a standard deviation in $\\log_{10}\\alpha$ of $\\sim 0.2$ and a slight dependence on redshift. \\item We find systematic differences of order $\\sim 10\\%$ in halo concentrations due to different fitting methods. Relative to Einasto fits to the density profile, NFW fits give concentrations that differ by up to 15\\%, while the $V_{\\rm max}/V_{200}$ method (e.g., Klypin \\etal 2011; Prada \\etal 2012) gives differences of up to 25\\%. A consequence of these systematic differences is the upturn in the concentration mass relation at high masses and redshifts (e.g., Klypin \\etal 2011; Prada \\etal 2012) is not present for our Einasto fits. \\item None of the analytic models in the literature (e.g., NFW 1997; Bullock \\etal 2001; Gao \\etal 2008; Zhao \\etal 2009; Prada \\etal 2012) accurately reproduce the evolution of the concentration mass relation. We provide simple fitting formulae for NFW (Eqs.~\\ref{eq:c200_slope}--\\ref{eq:cvir_zero}) and Einasto fits (Eqs.~\\ref{eq:cm_slope_einasto} \\& \\ref{eq:cm_zero_einasto}), which are valid between redshifts $z=5$ and $z=0$. \\item The observed concentrations and halo masses from NFW fits to data of spiral galaxies from the DiskMass project (Bershady \\etal 2010; Martinsson \\etal 2013), groups and clusters of galaxies (Ettori \\etal 2010; Auger \\etal 2013) are in good agreement with our simulations suggesting only mild halo response to galaxy formation on these scales. \\end{itemize}" }, "1402/1402.2656_arXiv.txt": { "abstract": " ", "introduction": "There is evidence that long duration gamma-ray bursts (GRBs) are produced when a massive star dies at the end of its nuclear burning life and its core collapses to a neutron star or a blackhole (e.g. Galama et al. 1998, Hjorth et al. 2003, Stanek et al. 2003, Modjaz et al. 2006, Campana et al. 2006, Starling et al. 2011, Sparre et al. 2011, Melandri et al. 2012). The newly formed compact object at the center of the progenitor star produces a pair of relativistic jets that make their way out of the star along polar regions. Punching their way to the stellar surface these jets shock heat the material they encounter and push it both sideways and along the jet's direction. Therefore, the jet is surrounded by this shock heated plasma, or a hot cocoon, which provides collimation for it (see fig. 1). The total amount of thermal energy in the cocoon is equal to the work done by the jet on stellar material it encounters while inside the star and that is estimated to be of order $L_j t_*\\sim 10^{51}$erg --- where $L_j$ is the jet luminosity and $t_*$ is the travel time for the jet inside the star at sub-relativistic speed (e.g. Meszaros \\& Rees, 2001; Ramirez-Ruiz et al. 2002; Matzner, 2003). We are here making the rough assumption that the bulk of the work goes to thermal energy though a significant amount goes into the cocoon forward (in jet direction) momentum. \\begin{figure}[ht!] \\centering \\includegraphics[scale=0.17]{cocoon-IC-sketch.eps} \\caption{Schematic sketch of a hot cocoon surrounding the jet while the jet is still inside the star is shown on the left side of this figure. The sketch to the right shows half of the system (the hemisphere with right to the center jet) at a later time when the jet and the cocoon have punched through the surface of GRB progenitor star. Thermal radiation from the cocoon is inverse-Compton scattered by the same jet, and by any relativistic jet produced by the central engine at a later time when the {\\it late jet} emerges above the cocoon surface. The IC scattered photons form a halo peaked near the edge of the jet and roughly as wide as the jet. \\label{fig1} } \\end{figure} The jet and the cocoon emerge from the stellar surface more or less at the same time. The central engine of long-GRBs remains active for $\\sim$10s to 10$^5$s after the jet emerges from the stellar surface (e.g. Burrows et al. 2005, Chincarini et al. 2007, 2010). The isotropic equivalent of luminosity from the cocoon at its peak, shortly after it emerges from the stellar surface, is of order 10$^{49}$ erg s$^{-1}$ at a few keV (Ramirez-Ruiz et al. 2002). The cocoon luminosity decreases slowly with time for about 10$^3$s and then drops to zero rapidly when the cocoon becomes transparent to Thomson scattering of photons. Any relativistic jet that is launched after the cocoon breaks through the stellar surface is expected to encounter this intense radiation field when it reaches the cocoon photosphere, and will produce a short-lived bright pulse of inverse-Compton scattered photons which form a halo peaked near the edge of the jet and roughly as wide as the jet (fig. 1)\\footnote{The jet while inside the cocoon is shielded from the IC drag of its intense radiation by an optically thick layer of electron-positron plasma that is created at the interface of the jet and cocoon. These $e^\\pm$ pairs are produced by the collision of thermal photons from the cocoon with IC scattered photons at this interface (Ceccobello \\& Kumar, 2014).}. Furthermore, radiation from the cocoon is Thomson scattered by electrons in the wind of GRB progenitor star, and some of these wind-scattered photons collide with the relativistic jet at large angles to produce an intense pulse of high energy inverse-Compton radiation. We provide in this work an estimate of IC flux from these different interactions of cocoon radiation with the GRB relativistic jet, and suggest how the detection of this high energy radiation (or an upper limit) can be used to constrain GRB jet and progenitor star properties. In section 2 we describe cocoon dynamics and radiation which follows closely the work of Ramirez-Ruiz et al. (2002) and Matzner (2003), and is included here for the sake of completeness and ease to follow the rest of the paper. Section 3 describes the IC scattering of cocoon radiation by a relativistic jet directly, and via an intermediate process of scattering first by electrons in the circum-burst medium (CBM). Application to GRBs is provided in \\S4. ", "conclusions": "Relativistic jets of long duration GRBs push aside stellar material, and evacuate a cavity, through the progenitor star on their way out to the surface. This process creates a hot cocoon of plasma surrounding the jet with energy of order $10^{52}$ erg (isotropic equivalent). A fraction of this energy is radiated away on time scale of a few hundred seconds (in observer frame) when the cocoon punches through the stellar surface. The interaction of this cocoon radiation with the relativistic jet has been investigated in this work and shown to be useful for exploring GRB jet and progenitor star properties. The basic idea is easy to explain. The radiative luminosity of the cocoon is of order 10$^{48}$ erg/s (isotropic equivalent), and photons from the cocoon collide with the jet at an angle of order $\\Gamma_c^{-1}$ wrt jet axis; $\\Gamma_c$ is the Lorentz factor of the cocoon. The cocoon luminosity as viewed in the jet comoving frame is a factor $\\Gamma_j^2/(3\\Gamma_c^4)$ larger. The rate at which radiation is produced by the jet --- which we see as prompt $\\gamma$-ray or X-ray flare radiation --- is of order $10^{51}\\Gamma_j^{-2}$ erg s$^{-1}$ (isotropic equivalent) in its comoving frame. Therefore, the ratio of cocoon thermal-radiation and jet radiation energy densities is $\\sim 3\\times10^{-4} (\\Gamma_j/\\Gamma_c)^4$ in the jet comoving frame. This ratio is larger than 1 for $\\Gamma_j/\\Gamma_c>10$, and in that case cocoon radiation is more important than the radiation produced within the relativistic jet for radiative cooling of electrons. Figures \\ref{ic-lum1} and \\ref{ic-lum2} show that the IC scattered cocoon radiation --- which forms a halo peaked near the edge of the jet and roughly as wide as the jet --- is of order the luminosity carried by the relativistic jet if electrons in the jet are heated to a thermal Lorentz factor larger than about 10$^2$ within a distance from the central engine of $\\sim 10^{15}$cm. The interaction of cocoon radiation with jet and predictions for high energy emission have been investigated in detail in this work. A lack of detection of IC scattered cocoon thermal-radiation suggests either that the jet energy is not dissipated and imparted to electrons out to a radius of at least 10$^{15}$ cm --- which would rule out a certain class of models for GRB prompt emission --- or that the cocoon moves outward with a high Lorentz factor such that $\\Gamma_j/\\Gamma_c \\lae 5$ (this possibility can be constrained by afterglow observations). Photons from the cocoon scattered by electrons in the circum-burst medium (CBM) can collide with the jet at a larger angle than photons traveling from the cocoon to the jet directly. These collisions result is very high energy photons ($\\sim$GeV) of considerable luminosity even for a modest thermal Lorentz factor for electrons (fig. \\ref{ic-lum2}). There is considerable uncertainty, however, in this estimate because the CBM could have been partially evacuated by an earlier passage of a relativistic jet through this region. Detection of this signal, or an upper limit, would provide a handle on the stellar mass loss rate during the last few years of the life of the GRB progenitor star. \\medskip \\noindent{\\bf Acknowledgment:} The work of GS is funded in part by PCCP. \"Paris Center for Cosmological Physics acknowledges the financial support of the UnivEarthS Labex program at Sorbonne Paris Cit\u00e9 (ANR-10-LABX-0023 and ANR-11-IDEX-0005-02)\"." }, "1402/1402.5240_arXiv.txt": { "abstract": "Recent observational and theoretical studies of classical Be stars have established the utility of polarization color diagrams (PCD) in helping to constrain the time-dependent mass decretion rates of these systems. We expand on our pilot observational study of this phenomenon, and report the detailed analysis of a long-term (1989-2004) spectropolarimetric survey of 9 additional classical Be stars, including systems exhibiting evidence of partial disk-loss/disk-growth episodes as well as systems exhibiting long-term stable disks. After carefully characterizing and removing the interstellar polarization along the line of sight to each of these targets, we analyze their intrinsic polarization behavior. We find that many steady-state Be disks pause at the top of the PCD, as predicted by theory. We also observe sharp declines in the Balmer jump polarization for later spectral type, near edge-on steady-state disks, again as recently predicted by theory, likely caused when the base density of the disk is very high, and the outer region of the edge-on disk starts to self absorb a significant number of Balmer jump photons. The intrinsic $V$-band polarization and polarization position angle of $\\gamma$ Cas exhibits variations that seem to phase with the orbital period of a known one-armed density structure in this disk, similar to the theoretical predictions of Halonen \\& Jones. We also observe stochastic jumps in the intrinsic polarization across the Balmer jump of several known Be+sdO systems, and speculate that the thermal inflation of part of the outer region of these disks could be responsible for producing this observational phenomenon. Finally, we estimate the base densities of this sample of stars to be between $\\approx 8\\times 10^{-11}$ to $\\approx 4 \\times 10^{-12}\\,\\rm g cm^{-3}$ during quasi steady state periods given there maximum observed polarization. ", "introduction": "\\label{intro} Classical Be stars are a subset of B-type main sequence stars which are characterized by their rapid rotational velocities ranging from 60\\% to 100\\% of their critical rate \\citep{riv13}. They have a geometrically flattened decretion disk that is fed from material from the stellar photosphere as diagnosed from studies of their optical/IR emission lines, polarization, and interferometric signatures (see e.g. \\citealt{por03,stee11}). A large volume of observations suggest the kinematic properties of these gas disks is best represented by near Keplerian rotation \\citep{hum00, me07a, pot10, whe12, kra12}. For the most up to date review of Classical Be stars see \\cite{riv13}. As summarized in \\citet{car11}, the viscous decretion disk model developed by \\citet{lee91} can explain many of the observational signatures of Be disks, although other models have been explored to explain the structure of these disks \\citep{bjo93,cas02,bro08}. One key unanswered question in the study of Be disks is what mechanism(s) are responsible for injecting material into these disks. Non-radial pulsations have been suggested to be one contributing factor to supplying material to some of these disks \\citep{cra09,riv98,nei02}, while periastron passage of binary companions may contribute in other systems such as $\\delta$ Scorpii \\citep{mir01,mir03}. This scenario for $\\delta$ Scorpii is now questionable given the disk's growth prior to the periastron passage of 2011 \\citep{miro13}. Nonetheless, binarity may play a role in the phenomenon and disk variability. For example, the source of material and angular momentum in non-classical Be stars can be the result of a red giant phase binary transferring material to create a Be+sdO system \\cite{gies98}. Characterizing the evolution of Be stars' mass-loss rates is another promising approach to constrain the disk-feeding mechanism. On short time-scales, \\cite{car07} noted polarimetric variability in Archernar likely arising from injections of discrete blobs of mass into the inner disk, that subsequently circularize into rings. Studying longer duration disk-loss and disk-regeneration events \\citep{und82,doa83,cla03,vin06,hau12}, including the time-scales \\citep{wis10} and statistical frequency \\citep{mcs08,mcs09} of these episodes, is another way to diagnose the mechanism feeding Be disks. Polarimetry has been used to study the Be phenomenon for both individual Be stars \\citep{qui97,woo97,cla98} and larger statistical surveys \\citep{coy69,mcl78,poe79,wi07b}. It is widely believed that Thompson scattering (free-electron scattering) is the source of polarization in Be stars \\citep{wo96a,wo96b,hal13}. Pre- or post-scattering absorption of photons within the disk can imprint wavelength dependent signature on top of the wavelength independent Thompson scattering \\citep{woo95}. Because the polarization across the Balmer jump traces material in the innermost regions of disks ($\\sim$6 R$_{\\star}$; \\citealt{car11}, \\citealt{ha13a}) and the $V$-band polarization is a tracer of the total scattering mass of the disk, studying the time evolution of the wavelength-dependence of polarization in Be stars can be used to constrain the time dependence of the mass decretion rate and the $\\alpha$ parameter in these systems. Specifically, the slope, shape, and temporal evolution of polarization color diagrams (PCD) observed in Be systems \\citep{dra11} have been theoretically reproduced, for the first time, by time-dependent radiative transfer models in which the mass decretion rate, $\\alpha$ parameter, and inclination angle are the primary variables \\citep{hau13}. In addition to density changes caused by changes in the mass decretion rate, it has been suggested that PCD diagram loops can also be produced by one-armed density perturbations \\citep{ha13b}. In this paper, we implement the PCD diagram diagnostic developed in \\citet{dra11} on a broader sample of Be stars, including systems showing evidence of experiencing disk-loss events and those whose disks appear roughly stable over time. We describe the data sample in Section 2, and discuss the techniques we used to remove the interstellar polarization component from each dataset in Section 3. We discuss the temporal evolution of each Be disk in PCD diagram parameter space in Section 4, and also detail evidence of variability in the disk position angle in select systems. Finally, we summary the major results of this manuscript in Section 5. ", "conclusions": "\\label{discussion} \\subsection{Additional Disk-Growth and Loss Phases} In paper I of this series, we analyzed well sampled, long-term spectropolarimetric data for two classical Be stars that clearly underwent a complete disk-loss phase. In this paper, we explore the behavior of a larger sample of classical Be stars, including systems that exhibit disks that are stable over time periods of more than a decade, as well as systems that exhibit at least partial disk-loss/disk-renewal phases. While detailed modeling of each system that exhibits a non-stable disk is clearly necessary and warranted, albeit outside the scope of this paper, we discuss some of the basic observational properties of these systems below. \\subsubsection{$\\psi$ Per} The intrinsic polarization and H$\\alpha$ EW of $\\psi$ Per both exhibit clear evidence of significant variability over the $\\sim$13 year time coverage of our data (Fig. \\ref{fig:ip_all_b}). Throughout the early 1990s, the steady increase in the system's intrinsic polarization, lasting for $\\sim$1400 days, accompanied by a small strengthening in its H$\\alpha$ EW is indicative of a growth in density and size of the disk. Although the data sampling is sparse, there is evidence that the disk generally stabilized in strength between 1995-2001. Figure \\ref{fig:ip_all_b} also exhibits a dramatic, monotonic drop in intrinsic polarization over a period of $\\sim$190 days starting in August 2002, along with a more gradual, time-delayed, significant decrease in H$\\alpha$ EW. These trends are consistent with a major inside-out clearing of significant mass from the disk \\citep{wis10}. Since a small level of both intrinsic polarization and H$\\alpha$ emission remained after the event, we characterize this as a incomplete disk-loss event. Characterizing the time-scale of disk-loss events is important in order to constrain the viscosity parameter, $\\alpha$, assuming the disk dissipates on a viscous time-scale. Equation (19) of \\citet{bjo05} describes such an assumption and can be rewritten to give $\\alpha$ as function of the diffusion timescale, $t_{\\rm diff}$. \\begin{equation} \\alpha = (0.2 \\rm yr / t_{\\rm diff}) * (r/R_{\\star})^{0.5} \\end{equation} Assuming the bulk of the scattering events producing the observed polarization occurs at a radial distance of $\\sim$5 R$_{\\star}$ and the (partial) disk-loss time-scale of $\\sim$ 190 days yields $\\alpha$ $\\sim$0.86 for $\\psi$ Per. This value is intermediate compared to estimates for 60 Cyg ($\\sim$0.1; \\citealt{wis10}) and robustly measured for 28 CMa (1 $\\pm$ 0.2; \\citealt{car12}). Because the $\\psi$ Per disk did not completely clear out, our quoted disk-loss time-scale is a lower limit and correspondingly the quoted $\\alpha$ parameter is an upper limit. Clearly, characterizing the viscosity parameter for a larger number of systems is required to assess whether there is a preferred parameter for most Be disks. \\subsubsection{$\\omega$ Ori} The overall steady decline in H$\\alpha$ EW strength throughout most of the 13 years of coverage in our dataset (Fig. \\ref{fig:ip_all_b}) is indicative of a gradual, albeit incomplete, loss of the system's disk. The $\\sim$1500 day duration decline in intrinsic $V$-band polarization starting at the beginning of our dataset is consistent with a gradual loss of the system's disk. Mirroring the scenario observed in $\\pi$ Aqr reported by \\citet{wis10}, the subsequent $\\sim$1100 day long increase in intrinsic polarization is likely responsible for the temporary halt in the decline of H$\\alpha$ EW strength between 1995-1998 and indicative of temporary replenishment of material to the inner regions of the disk. The cessation of this mass injection to the inner disk is marked by a fast decrease in the intrinsic polarization level, as well as a resumption of the gradual decline in H$\\alpha$ EW. Near the end of our dataset, the sharp rise in both intrinsic polarization and H$\\alpha$ EW indicate significant mass injection into the disk. Overall, these disk-loss and growth epochs occurred in the span of 2-4 years. \\subsubsection{66 Oph} The H$\\alpha$ EW of 66 Oph also exhibited a steady decline throughout the entire 13 year baseline of our dataset (Fig. \\ref{fig:ip_all_b}). Supplementary H$\\alpha$ EWs from spectra obtained at Ritter Observatory improve the time sampling during the end of our spectropolarimetric data, and confirm that the emission strength decline continues during these epochs. The Ritter EWs end their decline after the last epoch of our spectropolarimetric data (see e.g. Table \\ref{66Oph_rit_data}), suggesting the system's mass-loss rate changed and a complete disk-loss phase was avoided. The intrinsic polarization data (Fig \\ref{fig:ip_all_b}) also exhibit a slow decline throughout the 13 year baseline of the HPOL data, supporting the interpretation of a gradual decline in the disk. \\subsection{Intrinsic Polarization Color Diagrams} \\label{intpol} \\citet{dra11} noted the evolution of the ratio of the polarization across the Balmer jump versus the $V$-band polarization occasionally traced out distinctive loops in PCD diagrams. \\citet{hau13} explored these diagrams in detail with theoretical models. Since the polarization change across the Balmer jump is roughly proportional to the density squared whereas the $V$-band polarization is proportional to the density, PCD diagrams offer a useful method to investigate the evolution of Be disks when spectral type and inclination can be constrained \\citep{hau13}. PCD loops observed in 60 Cyg and $\\pi$ Aqr were qualitatively modeled using the radiative transfer code HDUST and 1D hydrodynamical code SINGLEBE by turning the mass decretion rate on and off. Subsequent detailed modeling of PCDs demonstrated that the shape, slope, and time-scale of loops depends on spectral type, inclination angle, the base density of the disk, and the temporal behavior of the mass decretion rate \\citep{hau13}. One important conclusion from these modeling efforts is that steady state disks pause at the top of a PCD loop, while the disk is fed by either constant mass-loss or semi-regular mass injections from the star \\citep{hau13}. Our steady disk sample broadly reproduces this behavior. For example, both $\\phi$ Per and $\\gamma$ Cas (Fig \\ref{fig:ip_all}) cluster at high $V$-band polarization and high Balmer jump ratios in their PCDs during their steady-state disk stage.\\\\ However, the following systems have deviations from this steady state disk behavior: \\\\ \\subsubsection{48 Lib} 48 Lib exhibits clear evidence of a partial PCD loop (Fig \\ref{fig:ip_all_a}) in our dataset. The rise in 48 Lib's intrinsic $V$-band polarization (blue points; Fig \\ref{fig:ip_all_a}) indicates a disk growth phase, which eventually plateaus into a steady state disk phase. Unlike other steady state disks in our sample that sit at the top of the PCD loop during the steady state, 48 Lib's Balmer jump ratio rapidly drops during this phase. Since 48 Lib is a near edge-on inclination \\citep{riv06,ste12}, its PCD behavior seems consistent with the modeling scenario outlined in Fig. 9 of \\citet{hau13}. Specifically, these results suggest that 48 Lib was characterized by a high base density ($\\approx$ 10$^{-11}$ g cm$^{-3}$) disk and disk orientation which gives a high opacity. This manifests during the onset of the disk-growth event with a large Balmer jump ratio which later decreases at a nearly constant $V$-band polarization as it approaches a steady state phase. Furthermore, it has a steeper decline which is consistent with a later spectral type of B2 to B5 given 48 Lib is B3. \\\\ \\subsubsection{$\\psi$ Per} $\\psi$ Per experienced a notable rise in intrinsic $V$-band polarization during the early 1990's (blue points; Fig \\ref{fig:ip_all_b}) indicating a disk growth phase, that seemingly plateaued during the later part of the decade into a stable phase. $\\psi$ Per exhibited a dramatic drop in its Balmer jump polarization as it transitioned to a steady state system. Given the mid spectral type (B5) and near edge-on inclination ($\\sim$75$^{\\circ}$; Table 1) of the system, like 48 Lib, we suggest that the $\\psi$ Per PCD behavior could be caused by the outer regions of the high inclination, high density disk absorbing significant Balmer jump photons at the onset of the disk growth event \\citep{hau13}. This is further supported by the fact the polarization blueward of the Balmer jump was nearly zero, or essentially optically thick, to polarization of the disk. \\\\ \\subsubsection{28 Cyg} During 28 Cyg's overall long-term decline in disk strength, as measured by its declining intrinsic $V$-band polarization and H$\\alpha$ EW (Fig. \\ref{fig:ip_all_c}), the system experienced three short, large jumps in Balmer jump polarization. During a 90 day period when the star was monitored regularly, in some cases nightly, the PCD diagram illustrates 3 instances where the Balmer jump ratio spiked to $>$2.5 (compared to the median value of 1.4). The separation between these three peaks was 24 and 18 days respectively. Given its spectral type (B2.5), the Balmer jump polarization is most sensitive to the inner regions of the disk when compared to the $V$-band polarization \\citep{hau13}. These Balmer jump ratio spikes (that occur at a constant $V$-band polarization) are likely caused by discrete events in the inner disk. Specifically, we speculate that the spikes stem from stochastic mass injections into the disk. \\\\ Although outside the scope of this paper, it is clear that detailed modeling of 48 Lib, $\\psi$ Per, and 28 Cyg using codes such as \\citet{ha13a} and \\citet{hau13} should be improved and pursued to reproduce the PCD variability observed in these systems. Moreover, given the diagnostic potential for systems experiencing full disk growth/disk-loss episodes ($\\pi$ Aqr, 60 Cyg; \\citealt{dra11}) and episodic disk growth (48 Lib, $\\psi$ Per, 28 Cyg), it is certainly clear that enhanced observational monitoring of these types of systems should be aggressively pursued. Without consistent temporal monitoring, it can become challenging to constrain the exact time-scale for disk growth and loss events. Our results for 28 Cyg provide quantitative evidence that even high cadence (i.e. nightly) observations yield interesting PCD phenomenon that could be better exploited to diagnose episodic mass injection events. \\\\ \\subsection{One-armed Density Waves} \\label{PAvar} \\citet{ha13b} presented ad hoc model predictions of the effects of global one-armed oscillations \\citep{oka91,oka97} on the time-dependent linear polarization of Be disks, including models that used a perturbation pattern characterized by a pattern of opposite overdense and underdense regions \\citep{oka97} and models that used a spiral shaped perturbation pattern as adopted by \\citet{car09}. These models (and those of \\citealt{car09}) predict that the $V$-band polarization and polarization across the Balmer jump should exhibit a clear inclination-dependent modulation with phase (Figs 6-8 of \\citealt{ha13b}), although they will be out of phase with one another due to the different radial locations in the disks over which the scattering events occur in each bandpass. The $V$-band polarization position angle is also predicted to exhibit complex changes with phase (Fig 10; \\citealt{ha13b}). In-spite of these predictions, no conclusive observational evidence of this phenomenon has been reported \\citep{car09}. $\\gamma$ Cas is known to exhibit V/R variations indicating the presence of a one-armed density feature in its disk, as seen in its phase-folded He I V/R ratios (Fig \\ref{wacky}) compiled from data presented in \\citet{mir02}. Observations were made at the Ritter observatory from 1993 to 2002 of the He I at 5876 \\AA, amongst other lines, with a resolving power of $\\sim$26000. The line profile was consistently double peaked and had clear V/R variations $\\gamma$ Cas exhibits a generally stable disk, as diagnosed by its intrinsic polarization (Fig. \\ref{fig:ip_all_a}), which indicates its inner disk is being supplied by a generally stable decretion rate. It therefore makes an ideal system to search for long-term polarimetric effects related to its one-armed density feature. The intrinsic $V$-band polarization, polarization across the Balmer jump, and $V$-band polarization position angle phased to the V/R period of the He I data is shown in Fig \\ref{wacky}. The intrinsic $V$-band polarization generally exhibits a double-oscillation pattern over one period that is predicted by \\citet{ha13b}. The intrinsic polarization across the Balmer jump exhibits variability, albeit no clear indication of the phase-lagged behavior predicted by \\citet{ha13b}. The intrinsic $V$-band polarization position angle suggestively exhibits evidence of cohesive variations as a function of phase, but such variations are an order of magnitude greater than that predicted in the ad hoc models of \\citet{ha13b}. We speculate that one reason these data exhibit a stronger indication of the predicted behavior in $V$-band, rather in the Balmer jump polarization, is that the former is less sensitive to small changes in the inner disk caused by changes in the mass decretion rate. We encourage future modeling efforts of this phenomenon to explore the ramifications of abrupt and gradual changes in the mass decretion rate have on the polarization in disks having one-armed density waves. \\subsection{PA Deviations} \\citet{wis10} reported numerous instances of the intrinsic polarization of 60 Cyg and $\\pi$ Aqr deviating from their linear trends on a Stokes QU diagram, and found that these deviations in intrinsic polarization position angle were more prominent during large outburst events. These authors interpreted this behavior as either evidence of the injection and subsequent circularization of new blobs of mass into the inner disk region, similar to that noted in \\citet{car07}, or as evidence of the injection and subsequent circularization of new blobs at an inclined orbit to the plane of pre-existing disk material. Many of the Be systems explored in our current study exhibit generally stable, strong disks over most of the duration of our dataset, which suggests we should see analogous evidence of PA deviations in our dataset. As seen in Fig. \\ref{fig:gamCasdev}, $\\gamma$ Cas exhibits PA deviance as a function of the intrinsic $V$-band polarization in the system. Analogous figures for our other targets are available in the online-version of this paper. Overall, we do see clear evidence of strong PA deviations in most of our Be stars even though they cannot be correlated to specific outburst events like 60 Cyg and $\\pi$ Aqr were. The system that exhibits the smallest level of PA deviations, $\\omega$ Ori, is noteworthy as it exhibited evidence of a gradual, albeit incomplete, loss of its disk throughout the time-frame covered by our data. While errors in the ISP estimate could be inducing errors in the intrinsic PA, we note that the deviations are symmetric about the mean. If there were errors in our PA determinations, then one would expect the deviations to be systematically offset to one side of the mean. If there were a PA estimate error then one would expect the deviations systematically offset to one side of the mean. In general, our results are consistent with systems which have had more precise ISP determinations and time resolved outbursts which suggest the PA deviations are also ``clumpy'' injection events (e.g. 60 Cyg and $\\pi$ Aqr). \\subsection{Be+sdO systems}\\label{sdO} Three of our targets are known to have a sub dwarf companion, $\\phi$ Per \\citep{gies98}, FY CMa \\citep{peters08}, and 59 Cyg \\citep{peters13}, and all three systems exhibit steady state disks in our data. This apparent stability could be influenced by the sdO truncating these disks to the maximum allowed radius. Interestingly, we do observe significant variability in the polarization across the Balmer jump in both FY CMa and 59 Cyg (Fig \\ref{fig:ip_all_c}). We remind the reader that the ISP for 59 Cyg was poorly constrained; however, we explored different assumed total ISP PA values (about PA = 0) and still observed the jumps in the polarization across the Balmer jump, suggesting that they are likely real. As the sdO is likely heating the outer region of the disk nearest its orbital position, we speculate that this outside heating source may inflate the scale height of this region of the disk. Given the short periods of order 29-37 days \\citep{peters13}, we speculate that when the inflated scale heights of these moderately high inclination disks (Table 1) pass in front of our line of sight with the star, they cause the observed changes in polarization across the Balmer jump. We further speculate that the amount of disk material heated and inflated by $\\phi$ Per's more distant (127 day period) sdO companion may contribute to why this system exhibits no analogous Balmer jump polarization jumps. Higher cadence observations that sufficiently sample the suggested puffed-up outer disk over the orbital period of these systems' sdO companion would help to further establish our interpretation of this observational phenomenon. \\subsection{Maximum Polarization} The maximum polarization for a viscous disk in equilibrium is found to be correlated with inclination, spectral type, and base density \\citep{hau13,ha13b}. In most cases the disks in this sample reach some form of steady state. We then assume that the maximum observed polarization at some point in the 15 year survey is the maximum polarization of a disk around these stars. Since the effective temperature and inclination can be determined by other means (e.g. spectroscopy and interferometry), we use literature values to then derive the range of base densities for this sample of Be stars (See Table \\ref{star_sum}). Given the model tracks of \\cite{hau13}, we find that the base density for these stars lie within $ 8\\times 10^{-11}$ to $ 4 \\times 10^{-12}\\,\\rm g cm^{-3}$ (See Fig.\\ref{pmax}). Several items are key to interpreting the base density estimates extracted from Figure 5. $\\pi$ Aqr, for example, exhibits a lower maximum V-band polarization than the 4.2e-11 g cm${-3}$ base density track compared to other stars having a similar stellar effective temperature (Figure 5; right panel) due in part because the system's inclination angle (33.6$^{\\circ}$ is lower than that used for the model track (70$^{\\circ}$. Similarly, $\\pi$ Aqr exhibits a higher maximum than expected for the 4.2e-11 g cm${-3}$ base density track (Figure 5; left panel) compared to other systems having a similar inclination due in part to $\\pi$ Aqr having a warmer stellar effective temperature than adopted for these models. Due to this degeneracy, it likely has a similarly large base density as $\\phi$ Per around $4.2 \\times 10^{-11} \\rm g cm^{-3}$. $\\psi$ Per and 59 Cyg have similar maximum polarization and inclination yet have very different effective temperatures. This then requires a low base density of $8.4 \\times 10^{-12} \\rm g cm^{-3}$ which exhibits maximum polarization independent of effective temperature. An object like $\\gamma$ Cas, shows behavior similar to that of $\\pi$ Aqr in relation to the model tracks. It is likely limited in maximum polarization due to its low inclination rather then its spectral type, so it is consistent with a low base density of $4.2 \\times 10^{-11} \\rm g cm^{-3}$. Due to the degeneracy and variable nature of some of the stars observed by HPOL, these limits are not meant to be absolute but rather a first look into the disk properties for the class as a whole. Each star will require more detailed modeling, but potentially broader statistics could be applied to a wider sample if polarimetry, spectroscopy, and interferometry can be obtained simultaneously to derive the applicable parameters." }, "1402/1402.0379_arXiv.txt": { "abstract": "{Low-mass X-ray binaries (LMXBs) are a natural workbench to study accretion disk phenomena and optimal background sources to measure elemental abundances in the Interstellar medium (ISM). In high-resolution XMM-\\textit{Newton} spectra, the LMXB SAX~J1808.4$-$3658 showed in the past a neon column density significantly higher than expected given its small distance, presumably due to additional absorption from a neon-rich circumstellar medium (CSM).} {It is possible to detect intrinsic absorption from the CSM by evidence of Keplerian motions or outflows. For this purpose, we use a recent, deep (100\\,ks long), high-resolution \\textit{Chandra}/LETGS spectrum of SAX~J1808.4$-$3658 in combination with archival data.} {We estimated the column densities of the different absorbers through the study of their absorption lines. We used both empirical and physical models involving photo- and collisional-ionization in order to determine the nature of the absorbers.} {The abundances of the cold interstellar gas match the solar values as expected given the proximity of the X-ray source. For the first time in this source, we detected neon and oxygen blueshifted absorption lines that can be well modeled with outflowing photoionized gas. The wind is neon rich (Ne/O\\,$\\gtrsim$\\,3) and may originate from processed, ionized gas near the accretion disk or its corona. The kinematics ($v=500-1000$\\,km\\,s$^{-1}$) are indeed similar to those seen in other accretion disks. We also discovered a system of emission lines with very high Doppler velocities ($v\\sim24\\,000$\\,km\\,s$^{-1}$) originating presumably closer to the compact object. Additional observations and UV coverage are needed to accurately determine the wind abundances and its ionization structure.} {} ", "introduction": "\\label{sec:introduction} Accretion disks provide an important workbench to probe magnetized plasma dynamics, photoionization, thermal and ionization equilibria. The response of the ionized plasma to changes in the continuum and outflow kinematics may constrain the source that drives winds, the connection between jets, disk, and corona, and possibly the disk geometry. Low-mass X-ray binaries (LMXBs) consist of a neutron star (NS) or black hole (BH) in orbit with a $<$1 M$_{\\odot}$ companion. These sources are characterized by X-ray emission due to hot gas accreted by the compact objects in a form of disks. Some of the NS binaries exhibit X-ray bursts when enough of this material accumulates on the surface of the NS, eventually giving rise to a runaway thermonuclear explosion. Bursts can last up to an hour, but they provide only a few\\,\\% of the total fluence (see e.g. \\citealt{Lewin1993}). LMXBs display high flux and simple spectra, whose continuum is typically well described with blackbody (BB) and powerlaw (PL) emission. The BB component arises from the thermal emission of the accretion disk around the neutron star. The powerlaw describes the energy gain of the disk soft photons by scattering in the accretion disk corona, a process known as Comptonization. A few sharp features are seen in the high-energy part of the spectra, e.g. above 1\\,keV. Prominent Fe\\,K emission lines are commonly seen in LMXB high-quality spectra and a clear asymmetry is seen in the line profile, as would be expected if the lines originate from the innermost region of the accretion disk and therefore are subject to strong relativistic effects (see e.g. \\citealt{Cackett2012} and references therein). Most LMXB spectra do not exhibit intrinsic sharp absorption features in the soft X-ray energy band, but they are rich with absorption lines originating in the interstellar medium (ISM) that lies along their line-of-sight (LOS). The ISM influences the Galactic evolution through the exchange of matter with the stars and shows a complex structure consisting of phases at different equilibrium temperatures \\citep[for a review, see][]{Draine2011}. The K-shell transitions of carbon, nitrogen, oxygen, neon, and magnesium, and the L-shell transitions of iron fall in the soft X-ray energy band. The launch of the {XMM-\\textit{Newton}} and \\textit{Chandra} satellites, provided with high spectral resolution gratings, permitted to determine interstellar ionization states and amounts of dust in the LOS towards several X-ray sources (see e.g. \\citealt{paerels, devries, JuettI, costantini2005, Costantini2012, costantini, kaastra09, Lee2009, Pinto2010}). Recently, \\citet{Pinto2013}, hereinafter P13, have performed an extended analysis of the ISM towards nine LMXBs with high-quality {XMM-\\textit{Newton}} spectra, extracted total abundances, dust depletion factors, and attempted to constrain dust chemistry. They also measured the abundance trends in the Galactic disk and confirmed the well known metallicity gradient (see e.g. \\citealt{GradPedicelli}). However, their interstellar abundances significantly deviate from a monotonic distribution. The extreme case is given by the LMXB \\object{SAX~J1808.4$-$3658} (hereinafter SAX~J1808) where the neon abundance exceeds by 5$\\sigma$ the value predicted by the source location. P13 argue that this large scatter may be due to absorption by metal-rich material surrounding the X-ray sources. Absorption occurring near the source may be distinguished from Galactic absorption when velocity shifts or Doppler broadening are observed. However, the XMM-\\textit{Newton} spectral resolution may not be high enough to resolve them. Recent work by \\citet{Schulz2010} on the high-resolution \\textit{Chandra} spectra of LMXB 4U\\,0614$+$091 showed a variable Ne\\,K edge with an average velocity smear of $\\sim3500$\\,km\\,s$^{-1}$ implying a characteristic radius $<10^7$\\,m, consistent with an ultra-compact binary (UCB) nature. The variability proves that the excess is intrinsic to the source, but it is not yet clear whether this is due to either neon~/~oxygen absorption or some other process. \\citet{Costantini2012} found an absorption-line system in the RGS spectrum of the LMXB 4U1820--30, which may also be described as a $\\sim1200$\\,km\\,s$^{-1}$ outflow. \\citet{Ioannou2003} found evidence of a $\\sim1500$\\,km\\,s$^{-1}$ \\ion{C}{iv} outflow in the \\textit{Hubble Space Telescope} (HST) UV spectrum of the LMXB X2127+119. The X-ray transient SAX~J1808 is the first discovered accretion-powered millisecond X-ray pulsar in a LMXB (\\citealt{zand1998, Wijnands1998}). It is relatively nearby ($d=3.5\\pm0.1$\\,kpc, see \\citealt{Galloway2006}) and has a column density of interstellar medium $N_{\\rm H} = (1.40\\pm0.03)\\,\\times 10^{25}$\\,m$^{-2}$ (see P13). This $N_{\\rm H}$ value combined with the high source flux produces several strong interstellar absorption features in soft X-ray spectra. The source exhibits X-ray bursts showing photospheric expansion (see e.g. \\citealt{Galloway2008}). These most likely originate in a flash of a pure helium layer that is produced by stable hydrogen burning. This source is thus particularly suitable to study X-ray bursts in great detail and in a large bandpass. Therefore, \\citet{zand2013}, hereinafter ZA13, observed SAX~J1808 with Chandra and RXTE during the November 2011 outburst. They detected a single thermonuclear (type-I) burst, the brightest yet observed by Chandra from any source, and the second-brightest observed by RXTE. However, they found no evidence for discrete spectral features during the burst $-$ though absorption edges have been predicted to be present in such bursts $-$ and argued that a greater degree of photospheric expansion may be required. The persistent spectrum is instead optimal to study the absorption edges produced by diffuse (ISM) and circumstellar (CSM) media. In this work we used the deep observation of SAX~J1808 presented in ZA13, which was taken with the Advanced CCD Imaging Spectrometer (ACIS) in combination with the Low-Energy Transmission Grating (LETG) on board of {\\textit{Chandra}} during a high-flux source state as well as the archival observations taken with the {\\textit{Chandra}} Medium-Energy Grating (MEG) and the {XMM-\\textit{Newton}} Reflection Grating Spectrometer (RGS). The paper is organized as follows. In Sect.~\\ref{sec:data} we present the data reduction. In Sect.~\\ref{sec:spectra} we describe the spectral features and the fitting procedure. Alternative methods of analysis are treated in detail in Sect.~\\ref{sec:spectral_models1} and \\ref{sec:spectral_models2}. In Sec.~\\ref{sec:variability} we analyze the spectral variability. All results and their comparison with previous work are discussed in Sect.~\\ref{sec:discussion}. Conclusions are given in Sect.~\\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} In this paper we have presented a detailed analysis of the soft X-ray spectra at different epochs of the LMXB SAX\\,J1808.4-3658. We have carefully measured the abundances of the ISM which lies in the LOS towards this source. They all agree with the abundances of the Solar System as expected due to the nearness of the source. In this paper we took into account not only the cold ISM phase, but we also looked for absorption by gas in the source system. A better modeling of the absorption edges region allowed us a precise estimates of elemental abundances. We found an ISM Ne abundance consistent with the Solar value, which is less than previously estimated for this source. We detected gas outflowing at $500-1000$\\,km\\,s$^{-1}$ as shown by strong, blueshifted absorption lines at different ionization states that are well described with photoionized absorbers. We ascribed the outflow to the photoionized winds that have been observed in several sources (see e.g. \\citealt{DiazTrigo2012}). These winds are thought to originate from thermal and radiation pressure near the accretion disk. Our velocity estimates are in agreement with the dynamics of the accretion disks, but our ionization parameters are slightly lower than usual. However, we notice that outflows of intermediate-ionization gas have also been observed. The Ne/O abundance ratio in the wind is supersolar which would suggest high Ne/O abundances for the donor star if the wind originates from the accretion disk. This would disagree with the hypothesis of the companion to be a brown dwarf unless the matter is expelled from the stellar interior. {We also detected a system of emission lines in the spectrum of the source after a very powerful burst was exhibited. These lines can be well modeled with collisionally-ionized gas, originating presumably closer to the compact object, but they require very high Doppler velocities ($v\\sim0.1c$). Additional, higher-resolution X-ray (MEG) and far-UV (HST) observations covering the $1-20$\\,{\\AA} and $1000-2000$\\,{\\AA} spectral ranges can improve the abundance estimates of the winds and the interpretation of the emitting plasma.} Our method can be easily tested in several X-ray binaries in order to search for wind signatures and to probe the local circumstellar and the diffuse interstellar media." }, "1402/1402.0523_arXiv.txt": { "abstract": "Each of the two pulsars in the double pulsar PSR J0737-3039A/B system exhibits not only the pulses emanating from itself, but also displays modulations near the pulse period of the other. Freire et al. (2009, MNRAS, 396, 1764) have put forward a technique using the modulation of B by A to determine the sense of rotation of pulsar A relative to its orbital motion, among other quantities. In this paper, we present another technique with the same purpose. While the Freire et al. approach analyzes pulse arrival times, ours instead uses periods or frequencies (their inverses), which can be experimentally determined via power spectral analysis similar to that used in pulsar searches. Our technique is based on the apparent change in spin period of a body when it is measured from an orbiting platform (the other pulsar), and is shown to be entirely analogous to the difference between the sidereal and solar spin period of the Earth (i.e., the sidereal and solar day). Two benefits of this approach are its conceptual and computational simplicity. The direct detection of spin with this technique will observationally validate the rotating lighthouse model of pulsar emission, while the detection of the relative directions of spin and orbital angular momenta has important evolutionary implications. Our technique can be used on other binary systems exhibiting mutually induced phenomena. ", "introduction": "The double pulsar system PSR J0737-3039A/B was discovered by \\citet{Burgay03} and \\citet{Lyne04}. This system consists of a 22-ms pulsar (hereafter A) and a 2.8-s pulsar (hereafter B) with an orbital period of 2.4 hours. This discovery has provided a laboratory for the study of relativistic gravity and gravitational radiation \\citep{KramerWex2009}. The system has several strange features that challenge the current understanding of pulsars and provide an uncommon opportunity to improve pulsar theories. One of the most interesting properties is the observed modulation of each pulsar's signal by the energy flux from the other, as evidenced by each pulsar's modulation period being approximately equal to the other pulsar's pulse period \\citep{McLaughlin04a,McLaughlin04b}. \\citet{Freire09a} proposed a technique for analysing the arrival times of the pulsars' pulses and their mutual modulations which could yield the sense of rotation of each pulsar with respect to its orbital motion, among other quantities. In this paper, a complementary technique is presented with the same objective, but using measured periods rather than arrival times. The principal benefit of our approach is that it is simpler and more intuitive. If the validity of either of these techniques is confirmed, not only will new insights be gained but also the correctness of the lighthouse model will be empirically assessed beyond dispute. \\begin{table*} \\centering \\caption{Earth-Sun and double pulsar analogies. All quantities are defined as magnitudes.} \\begin{minipage}{184mm} \\begin{tabular}{lcc} \\hline Rotating object:& Earth $(\\earth)$&Pulsar A \\\\ \\hline Orbital period, Mean orbital frequency & $P_{\\rm orb} (\\equiv$ year), $f_{\\rm orb}$ & $P_{\\rm orb}, f_{\\rm orb}$\\\\ \\hline \\hspace{0.4in} Instantaneous orbital frequency & \\hspace{0.7in} $f_{\\rm orb}(t)$ & \\hspace{0.43in} $ f_{\\rm orb}(t) $ \\\\ \\hline Rotating object's sidereal spin period, frequency & $P_{\\earth,0} (\\equiv$ sidereal day), $f_{\\earth,0}$ & $P_{\\rm A,0}, f_{\\rm A,0} $ \\\\ \\hline Rotating object's spin period, frequency (measured at other body) & $P_{\\earth\\ \\rm at\\ \\sun} (\\equiv$solar day), $f_{\\earth\\ \\rm at\\ \\sun}$ & $P_{\\rm A\\ at\\ B}, f_{\\rm A\\ at\\ B}$ \\\\ $\\equiv\\ $Modulation period, frequency at other body due to rotating object & $\\equiv P_{\\rm m\\ at\\ \\sun\\ due\\ to\\ \\earth},$ & $\\equiv P_{\\rm m\\ at\\ B\\ due\\ to\\ A}, $ \\\\ &\\ \\ \\ $ f_{\\rm m\\ at\\ \\sun\\ due\\ to\\ \\earth} $& \\ \\ \\ $f_{\\rm m\\ at\\ B\\ due\\ to\\ A}$ \\\\ \\hline \\end{tabular} \\end{minipage} \\label{tab:one} \\end{table*} ", "conclusions": "\\cite{McLaughlin04a} presented a modulation pattern similar to drifting subpulses in the signal of PSR B, but with a frequency of 44 Hz, close to the pulse frequency of PSR A. The presence and sense of rotation of A is encoded in the observed modulation pattern, and can be revealed through an arrival-time-based analysis \\citep{Freire09a} or the frequency-based analysis presented above. Our procedure offers the benefits of relative conceptual simplicity and close analogy with familiar phenomena in the Earth-Sun system. We present a frequency-based procedure, building upon that used in binary pulsar search software, to distinguish among direct, retrograde, or no rotation of PSR A by searching synchronously for one of the three possible modulation signals over the full span of available data. Although the lighthouse model has been widely accepted, there has nevertheless been no direct observational evidence in its support up to this time. A strength of our technique is its ability to provide such a test, empirically supporting or refuting the model. \\citet{ferdman13} have shown that A's spin and orbital axes are aligned to within $3 \\degr$, but they have no direct means of distinguishing parallel from antiparallel alignment. By assuming parallel alignment, they are able to conclude that the second supernova, which created B, was relatively symmetric. Therefore the presence and sense of the rotation, as revealed by this analysis, will further test their and others' (e.g., \\citet{kramerstairs08,farret11}) evolutionary scenarios. While \\citet{McLaughlin04b} also found that A's signal is modulated at B's frequency near A's eclipse, the mechanism is different and not dependent on B's rotation. If, however, the modulation is observed in the future away from eclipse, it will provide an easier test for B's rotation than does the approach delineated above for A. Our analysis can also be used on other binary systems discovered to possess phenomena caused by mutual interactions." }, "1402/1402.4300_arXiv.txt": { "abstract": "We present a photometrical and morphological study of the properties of low redshift (z $<$ 0.5) quasars based on a large and homogeneous dataset of objects derived from the Sloan Digital Sky Survey (DR7). This study over number by a factor $\\sim$ 5 any other previous study of QSO host galaxies at low redshift undertaken either on ground or on space surveys. We used $\\sim$ 400 quasars that were imaged in the SDSS Stripe82 that is up to 2 mag deeper than standard Sloan images. For these quasars we undertake a study of the host galaxies and of their environments. In this paper we report the results for the quasar hosts. We are able to detect the host galaxy for more than 3/4 of the whole dataset and characterise the properties of their hosts. We found that QSO hosts are dominated by luminous galaxies of absolute magnitude M*-3 $<$ M(R) $<$ M*. For the unresolved objects we computed a upper limit to the host luminosity. For each well resolved quasar we are also able to characterise the morphology of the host galaxy that turn out to be more complex than what found in previous studies. QSO are hosted in a variety of galaxies from pure ellipticals to complex/composite morphologies that combine spheroids, disk, lens and halo. The black hole mass of the quasar, estimated from the spectral properties of the nuclei, are poorly correlated with the total luminosity of the host galaxy. However, taking into account only the bulge component we found a significant correlation between the BH mass and the bulge luminosity of the host. ", "introduction": "Accretion onto a supermassive black hole (SMBH) is the main mechanism that sustains the powerful activity of active galactic nuclei but may also represent a common phase in the evolution of normal galaxies. A number of fundamental question about the formation of the QSO phenomenon like the fuelling and triggering mechanisms are strictly related to the immediate environments of the active nucleus and in particular to its host galaxy \\citep{merloni10}. SMBHs may well have a period of maximum growth (maximum nuclear luminosity) contemporaneous with the bulk of the initial star formation in the bulge of galaxies. Studies of the co-evolution of SMBH and their host spheroids are therefore obviously critical to understanding how and when galaxies in the local Universe formed and evolved. The last ten years have yielded considerable progress in characterising AGN host galaxies. At variance with inactive galaxies their study is often hampered by the presence of the luminous central source that outshines the light of the host galaxy. A problem that becomes more serious for high luminosity AGNs and for sources at high redshift. In spite of these limitations the characterization of the properties of the host galaxies offers the unique opportunity to investigate the link between the central black hole mass and its host galaxy at moderate to high redshift and to trace the possible co-evolution at different cosmic epochs. This is because for broad line AGN like quasars it is possible to estimate the mass of the central BH using kinematic arguments that are not directly dependent on the host galaxies properties. Both ground-based and HST studies have shown that virtually all luminous low redshift (z$<$0.5) quasars reside in massive, spheroid-dominated host galaxies, whereas at lower luminosities quasars can also be found in early- type spiral hosts (e.g. Bahcall et al. 1997; Dunlop et al. 2003; Pagani et al. 2003; Floyd et al. 2004; Jahnke et al. 2004). This is in good agreement with the BH -- bulge relationship in inactive galaxies (e.g. Gultekin et al. 2009), since very massive BHs power luminous quasars. Only a small fraction of the host galaxies ($\\sim$15\\% ) are found in merger systems but it is difficult to determine clear merger signatures from morphology alone. At low redshifts a major contribution to the properties of quasar host galaxies has been provided by images from the Hubble Space Telescope (HST). The improved spatial resolution has allowed the characterization of the structure and the detailed morphology of the host galaxies \\citep{bahcall97, kukula01, ridgway01,dunlop03,peng06,zakamska06}. It turned out that QSO are hosted in luminous galaxies that are often dominated by the spheroidal component. At high redshift (z $>$1) HST observations of quasar host galaxies (e.g. \\cite{peng06,floyd13} and references therein) have been complemented by significant contributions from 8-m class ground-based telescopes under superb seeing conditions \\citep{kotilainen07,kotilainen09} and/or with adaptive optics \\citep{falomo08}. Comparison of host galaxies of AGN at high and low redshift constrain host galaxy evolution, as compared with the evolution of normal (inactive) galaxies. Most of the old studies of quasar host considered few tens of objects therefore in order to derive a picture of the host properties at various redshift one should combine many different samples often obtained with different telescopes and filters. Observations carried out by HST are certainly more homogeneous (although different filters were used) and allow to investigate a somewhat large sample based on high quality data. Nevertheless the size of these samples remain relatively small For instance in the range 0.25 $ < z < $ 0.5 about 50 QSO were imaged by HST (see references above). In order to explore a significantly larger dataset of QSO one should refer to large surveys that include both imaging and spectroscopic data. In this respect one of the most productive recent surveys is the Sloan Digital Sky Survey that allowed to find 105783 quasars \\citep{schneider2010} from (DR-7). Standard SDSS images are, however, too shallow and the faint nebulosity around the nucleus of quasars is not detected. This problem has been overcome in the case of the special sky region mapped by SDSS for the SDSS Legacy Survey. The central stripe in the South Galactic Gap, namely the Stripe82 \\citep{annis2011} is a stripe along the Celestial Equator in the Southern Galactic Cap. It is 2.5$^{\\deg}$ wide and covers -50$^{\\deg}\\leq$RA$\\leq$+ 60$^{\\deg}$, so its total area is 275deg$^2$. Stripe 82 was imaged by the SDSS multiple times, these data were taken in 2004 only under optimal seeing, sky brightness, and photometric conditions (i.e., the conditions required for imaging in the main Legacy Survey; York et al. (2000)). In 2005-2007, 219 additional imaging runs were taken on Stripe 82 as part of the SDSS supernova survey \\citep{frieman08}, designed to discover Type Ia supernovae at 0.1$< z <$0.4. The total number of images reaches $\\sim$100 for the S strip and $\\sim$ 80 for the N strip. The final frames were obtained by co-adding selected fields in r-band, with seeing (as derived from 2D gaussian fit of stars and provided by SDSS pipeline) better than 2\", sky brightness $\\leq$19.5 $mag/arcsec^2$ and less than 0.2 mag of extinction. In this area there are 12434 quasars. Recently \\cite{matsuoka14} analyzed the stellar properties of about 800 galaxies hosting optically luminous, unobscured quasars at z$<$ 0.6 using Stripe82 images. They focused on the color of the host galaxies and found that the quasar hosts are very blue and almost absent on the red sequence with a marked different distribution from that of normal (inactive) galaxies. For our study we selected QSO with redshift less than 0.5 for which the stripe 82 images allow us also to study the QSO galaxy environments. We adopt the concordance cosmology with H$_0$ = 70 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_m$ = 0.3 and $\\Omega_\\Lambda$ = 0.7. In this first paper of a series we focus on the properties of quasar hosts and their relationship with the central BH mass in the explored redshift range. In forthcoming papers we investigate the galaxy environments \\citep{karhunen13} and galaxy peculiarities \\citep{bettoni14}. A preliminary account of these results was presented in \\cite{kotilainen13}. ", "conclusions": "We have investigated the properties of the host galaxies from a large ($\\sim$ 400 objects) and homogeneous dataset of low redshift (z $<$ 0.5) quasars using th SDSS images in the Stripe82 region that are significantly deeper that standard SDSS data. The 2D analysis of the images allowed us to well resolve the quasar host for 3/4 of the objects in the sample, marginally resolve other 40 quasars and derive limits for the galaxy luminosity for the unresolved targets (60 objects). The following properties of quasar hosts are derived: \\begin{enumerate} \\item the luminosity of the host galaxies of low z quasars span a range from M(R) $\\sim$ -21.5 to M(R) $\\sim$ -24.0; the bulk of the host galaxies are located in the region corresponding to M*-1 and M*-2; there is a mild increase of the host luminosity with the redshift that is consistent with the passive evolution of the underlying stellar population \\item the morphology of the host galaxies turned out to be rather complex with both bulge and disc dominated galaxies; about one third of the objects in our sample show features characteristics of bulge and disc components \\item irrespective of the host morphology the size of the galaxies (as derived from the half light radius) ranges from compact (few kpc) objects up to extended galaxies (10-15 kpc); in the observed redshift range we do not find any significant trend of change of the galaxy size with z \\item the nuclear and host galaxy luminosities are not correlated suggesting that accretion rate, BH mass, and galaxy masses and morphology combine together to smear significantly the correlation between BH and host masses \\item the BH mass of quasars estimated from the QSO continuum luminosity and the width of the broad emission lines is poorly correlated with the total luminosity/mass of the whole host galaxy; on the contrary when the fraction of bulge to disc component is considered we find a significant correlation between the BH mass and the bulge luminosity of the host. \\end{enumerate} Another important source of information to characterise the properties of low redshift QSO come from the analysis of their galaxy environments as compared with those of similar galaxies with no active nuclei. These aspects will be pursued in forthcoming papers of this series \\citep{bettoni14,karhunen13}." }, "1402/1402.5210.txt": { "abstract": "We present a new approach for estimating the 3.6 $\\mu m$ stellar mass-to-light ratio $\\Upsilon_{3.6}$ in terms of the [3.6]-[4.5] colors of old stellar populations. Our approach avoids several of the largest sources of uncertainty in existing techniques using population synthesis models. By focusing on mid-IR wavelengths, we gain a virtually dust extinction-free tracer of the old stars, avoiding the need to adopt a dust model to correctly interpret optical or optical/NIR colors normally leveraged to assign the mass-to-light ratio $\\Upsilon$. By calibrating a new relation between NIR and mid-IR colors of giant stars observed in GLIMPSE we also avoid the discrepancies in model predictions for the [3.6]-[4.5] colors of old stellar populations due to uncertainties in the molecular line opacities assumed in template spectra. We find that the [3.6]-[4.5] color, which is driven primarily by metallicity, provides a tight constraint on $\\Upsilon_{3.6}$, which varies intrinsically less than at optical wavelengths. The uncertainty on $\\Upsilon_{3.6}$ of $\\sim$0.07 dex due to unconstrained age variations marks a significant improvement on existing techniques for estimating the stellar M/L with shorter wavelength data. A single $\\Upsilon_{3.6}$=0.6 (assuming a Chabrier IMF), independent of [3.6]-[4.5] color, is also feasible as it can be applied simultaneously to old, metal-rich and young, metal-poor populations, and still with comparable (or better) accuracy ($\\sim$0.1 dex) as alternatives. We expect our $\\Upsilon_{3.6}$ to be optimal for mapping the stellar mass distributions in S$^4$G galaxies, for which we have developed an Independent Component Analysis technique to first isolate the old stellar light at 3.6 $\\mu m$ from non-stellar emission (e.g. hot dust and the 3.3 PAH feature). Our estimate can also be used to determine the fractional contribution of non-stellar emission to global (rest-frame) 3.6 $\\mu m$ fluxes, e.g. in WISE imaging, and establishes a reliable basis for exploring variations in the stellar IMF. ", "introduction": "} Accurate maps of the stellar mass distribution in nearby galaxies are essential for charting the dynamical influence of stellar structures over time. They provide a key perspective on the evolution in the amount and distribution of baryons from high redshift to the local universe, both as the result of internal, secular processes (e.g. Elmegreen et al. 2007; Zhang \\& Buta 2007; Haan et al 2009; Foyle et al. 2010), and given `bottom up' assembly and external (heirarchical accretion) events within the context of CDM cosmology ( e.g., Balcells et al. 2003 and Courteau et al. 1996; Florido et al. 2001; and see Navarro \\& White 1994; Somerville 2002; Governato et al. 2007; Guo \\& White 2008).\\\\ \\indent NIR bands are often though to be optimal windows on the old stars that dominate the baryonic mass in galaxies, with images at these wavelengths serving as good relative proxies for stellar mass (Elmegreen \\& Elmegreen 1984; Rix \\& Zaritsky 1995; Grosbol 1993). Yet even in these bands the light is affected by the presence of dust extinction and young stars (albeit less than in the optical), requiring the application of a M/L ratio that incorporates dependencies on metallicity, star formation history, and especially the age of the stellar population (Rix \\& Rieke 1993; Rhoads 1998). Such estimates have emerged only relatively recently, following the demonstrated correlation between optical colors and the mass-to-light ratio in stellar population synthesis models \\citep{bdJ01}. The state-of-the art, which uses more than one color and incorporates a model of the dust, is optimal for use in 2D and has greatly reduced uncertainties \\citep{z09}. But, at present few such maps exist. \\indent We have begun an effort to map the stellar mass distribution in nearby galaxies in the Spitzer Survey of Stellar Structure in Nearby Galaxies (S$^4$G), an un-paralleled inventory of stellar mass and structure in $\\sim$2300 nearby galaxies ($D$$<$40 Mpc; Sheth et al. 2010). The survey consists of deep imaging at 3.6 $\\mu m$ and 4.5 $\\mu m$ where the light traces the old stars, and where the effects of dust extinction and star formation activity (e.g. young stars, HII regions; see \\citealt{sheth} and references therein) are at a minimum compared to ground-based images in either optical or NIR wavebands. Our method begins by using Independent Component Analysis (ICA) to isolate the old stellar light (from K and M giants) from the `contaminating' emission from several other sources: the 3.3 $\\mu m$ PAH emission feature in channel 1, accessory non-stellar continuum emission (i.e. from hot dust and PAH), and lower-M/L asymptotic giant branch (AGB) and red supergiant (RSG) stars. These three main contaminants can lead to an over-estimation of stellar mass when using uncorrected 3.6 $\\mu m$ and/or 4.5 $\\mu m$ images (\\citealt{meidtpaperI}; hereafter Paper I). Corrections for this `contaminating' emission are especially critical for mass estimates, since this emission is not incorporated into SED libraries at the core of now-standard techniques for estimating the stellar M/L (e.g. \\citealt{bdJ01}; \\citealt{z09}). Once contaminants are removed with ICA we obtain maps of the underlying distribution of old stellar light in S$^4$G images, with [3.6]-[4.5] colors that are consistent with those of M and K giants (\\citealt{pahre}; \\citealt{hunter}). To generate 2D maps of the stellar mass distribution in principle thus requires applying a M/L that is appropriate for an old, dust-free population of stars. Tight constraints on the stellar M/L at 3.6$\\mu m$ are possible using stellar population synthesis (SPS) models to link, e.g., optical and/or NIR colors to metallicity and age (or star formation history; SFH) variations in the underlying stellar population. Paired with this information, contaminant-corrected S$^4$G images should be readily convertible into high-quality mass maps. But even in the latest generation of models there are lingering uncertainties: 1) convergence on the TP-AGB phase of stellar evolution that dominates NIR light at $\\sim$1 Gyr has yet to be reached among different models (cf. Charlot \\& Bruzual 2007, \\citealt{maraston05} and \\citealt{lancon}), 2) differences in the molecular line opacities assumed in template spectra lead to a variety of model predictions at mid-IR wavelengths (e.g. Peletier et al. 2012), and 3) realistic models of dust extinction (as well as the reflection and scattering of stellar light by dust) are difficult to implement at the pixel-by-pixel level, given the variety of dust geometries \\citep{z09}. In this paper we explore estimating the M/L at 3.6$\\mu m$ in a way that is least susceptible to the uncertainties that plague state-of-the-art mass estimation techniques. The key is our focus on mid-IR wavelengths where the emission is dominated by the oldest stars and dust extinction is minimal. In $\\S$ \\ref{sec:ICA} we describe our ICA technique to isolate the old stellar light from non-stellar emission at 3.6 $\\mu m$. Then in $\\S$ \\ref{sec:MLdescription} we demonstrate that the variation in stellar $\\Upsilon_{3.6}$ is intrinsically quite low and, in this case, we argue that the IRAC [3.6]-[4.5] colors of the old stars provide sufficient constraint on the M/L ($\\S$ \\ref{sec:withcolor}). To demonstrate this we use SPS models, setting out first to improve their predictive power at mid-IR wavelengths. Using a new calibration of the relation between the NIR and mid-IR colors of GLIMPSE giants (Appendix A), we link SPS models to realistic [3.6]-[4.5] colors in $\\S$ \\ref{sec:iraccolor}. This allows us to explore the dependence of age/SFH (including bursts) and metallicity on the relation between $\\Upsilon_{3.6}$ and [3.6]-[4.5] color, which we present in $\\S$ \\ref{sec:therelation}. Our approach for constructing 2D mass maps, namely, using information exclusively from 3.6 and 4.5 $\\mu m$ images revealed by ICA, is an approach that can be applied to the greatest number of galaxies in the S$^4$G sample without relying on ancillary optical and/or NIR data. ", "conclusions": "We have shown that the information contained in IRAC images at 3.6 and 4.5 $\\mu m$ is optimal for estimating stellar masses of nearby galaxies. At 3.6 $\\mu m$ we combine an intrinsically modestly varying stellar M/L with the ability to constrain variations in the properties of the stellar population via the [3.6]-[4.5] colors of old stars. We achieve an uncertainty as low as 0.06 dex by using the [3.6]-[4.5] color to constrain the metallicity-dependence of the stellar M/L; more modest unconstrained age variations at these long wavelengths contribute an uncertainty that is comparable to, if not below, that which can be achieved at shorter wavelengths, where the effects of dust extinction/reddening and young stars are present. Much of the uncertainty normally introduced with the use of SPS models at long wavelengths ($\\lambda\\gtrsim$ 2.2 $\\mu m$) is reduced by using the observed colors of GLIMPSE giants to relate the NIR colors of SPS models to realistic [3.6]-[4.5] colors. Our preferred method of assigning the stellar M/L uses the [3.6]-[4.5] color to account for the increase in the brightness of giant stars with increasing metallicity (traced by [3.6]-[4.5] color). We define a relation between [3.6]-[4.5] color and $\\Upsilon_{3.6}$ that extends across the full range in [3.6]-[4.5] colors exhibited by galaxies across the Hubble sequence. This relation is designed to be applied to the light from old stars, alone, i.e. when [3.6]-[4.5] colors reflect genuine variation in the properties of the old stellar population. This makes it applicable to early-type galaxies, with little on-going star formation (and minimal dust extinction/reddening), or in late-type galaxies, where non-stellar contaminating emission is isolated from the old stellar light with ICA (Paper I). Remarkably, the existence of an age-metallicity relation in old stellar populations may even lead to a preference for a single, color-independent $\\Upsilon_{3.6}$=0.6, good to an accuracy of 0.06 dex. We recommend adopting a more conservative 0.1 dex uncertainty on $\\Upsilon_{3.6}$=0.6 given a suspected atypical evolutionary history, or when it is not possible to correct for the presence of non-stellar emission (which leads to characteristically red [3.6]-[4.5] colors). In this case, additional uncertainty in the stellar mass estimate can arise as a result of the contribution from the non-stellar emission, which can be as large as 30\\% locally (i.e. an uncertainty comparable to that associated with the stellar M/L). Either of the two estimates for $\\Upsilon_{3.6}$ presented here (with or without [3.6]-[4.5] color information) mark a significant improvement in our view of how the stellar mass is distributed in and among galaxies. With greater confidence in the stellar M/L we also gain much-needed leverage on the dark matter content of galaxies, as well as the opportunity to explore variations in the IMF with global galaxy properties. We anticipate successfully implementing our preferred approach, one that yields the smallest uncertainty in stellar mass, to galaxies in S$^4$G; future work will first remove the contribution from non-stellar emission in S$^4$G 3.6 $\\mu m$ images with ICA and then use the [3.6]-[4.5] color of the old stellar light to assign $\\Upsilon_{3.6}$ (Querejeta et al. 2014). \\\\ \\newline \\newline Thanks to Mariya Lyubenova for fruitful discussion and the entire S$^4$G team. E.A., A.B., J.K., G.vdV., M.Q., S.M. and E.S. acknowledge financial support of the DAGAL network from the People Programme (Marie Curie Actions) of the European Union's Seventh Framework Programme FP7/2007-2013/ under REA grant agreement number PITN-GA-2011-289313. K.S., J.-C.M.-M., and T.K. acknowledge support from the National Radio Astronomy Observatory, which is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. % \\newpage \\appendix" }, "1402/1402.4668.txt": { "abstract": "\\noindent Infrared emission is an invaluable tool for quantifying star formation in galaxies. Because the 8 $\\mu$m polycyclic aromatic hydrocarbon (PAH) emission has been found to correlate with other well-known star formation tracers, it has widely been used as a star formation rate (SFR) tracer. There are, however, studies that challenge the accuracy and reliability of the 8 $\\mu$m emission as a SFR tracer.\\\\ \\indent Our study, part of the Herschel\\footnote{Herschel is an ESA space observatory with science instruments provided by European-led Principal Investigator consortia and with important participation from NASA.} M33 Extended Survey (HERM33ES) open time key program, aims at addressing this issue by analyzing the infrared emission from the nearby spiral galaxy M33 at the high spatial scale of $\\sim$ 75 pc. Combining data from the \\it{Herschel} \\rm Space Observatory and the \\it{Spitzer} \\rm Space Telescope we find that the 8 $\\mu$m emission is better correlated with the 250 $\\mu$m emission, which traces cold interstellar gas, than with the 24 $\\mu$m emission. Furthermore, the L(8)/L(250) ratio is more tightly correlated with the 3.6 $\\mu$m emission, a tracer of evolved stellar populations and stellar mass, than with a combination of H$\\alpha$ and 24 $\\mu$m emission, a tracer of SFR. The L(8)/L(24) ratio is highly depressed in 24 $\\mu$m luminous regions, which correlate with known HII regions.\\\\ \\indent We also compare our results with the dust emission models by \\citet{draine07}. We confirm that the depression of 8 $\\mu$m PAH emission near star-forming regions is higher than what is predicted by models; this is possibly an effect of increased stellar radiation from young stars destroying the dust grains responsible for the 8 $\\mu$m emission as already suggested by other authors. We find that the majority of the 8 $\\mu$m emission is fully consistent with heating by the diffuse interstellar medium, similar to what recently determined for the dust emission in M31 by \\citet{draine13}. We also find that the fraction of 8 $\\mu$m emission associated with the diffuse interstellar radiation field ranges between $\\sim$60\\% and 80\\% and is 40\\% larger than the diffuse fraction at 24 $\\mu$m. ", "introduction": "\\noindent In order to understand the evolution of galaxies in the Universe, accurate measures of star formation (SF) within galactic structures need to be obtained. A large number of tracers of star formation in a galaxy have been defined in the literature, using different regions of the electromagnetic spectrum. Among these, the infrared (IR) radiation is a classical tracer of activity in galaxies \\citep{kennicutt98, ken12, calz12}. In star-forming galaxies, young, massive stars are responsible for most of the ultraviolet (UV) radiation and, in the presence of dust a smaller or larger fraction of their light may be absorbed and reradiated in the IR regime. By observing galaxies in IR light, determinations about the rate and areas of star formation can be surmised.\\\\ \\indent The mid-IR wavelength region ($\\approx$7-40 $\\mu$m) in general, and the emission around $\\sim$8 $\\mu$m in particular, are among the favored tracers of recent star formation, because of their ready detectability in galaxies at high redshift (e.g., Daddi et al. 2005, Reddy et al. 2010, 2012). The emission in the $\\sim$8 $\\mu$m region is mainly contributed by a combination of stellar photospheric emission, the featureless continuum of hot dust emission, and the Polycyclic Aromatic Hydrocarbon (PAH) spectral features (e.g., \\citet{smith07}). As the PAH features contribute about 70\\% or more of the emission in the Spitzer $\\sim$8 $\\mu$m band \\citep{smith07}, we call the emission in this band `PAH emission' henceforth. \\citet{roussel01} and \\citet{forster04} have suggested, using data from ISO, that the PAH emission is closely related to the hot dust emission at 15 $\\mu$m and other tracers of star formation. Other studies (e.g. \\citet{bos04,haas02}) and follow up investigations, utilizing the \\it{Spitzer} \\rm Space Telescope \\citep{werner04}, found the relation to be more complex than previously inferred. The PAH emission does not immediately correlate with star formation as traced by the 24 $\\mu$m emission, as the L(8)/L(24) luminosity ratio is depressed in regions of known star formation relative to the ratio in the diffuse starlight field \\citep{helou04, calz05, bendo08, pov07}. \\citet{bendo06, bendo08}, furthermore, showed that the PAH emission is more closely related to tracers of cool dust emission, heated by the diffuse starlight field, than tracers of warm dust, heated by recent star formation. This is confirmed by \\citet{verley09}, who used a complementary approach to our own to determine that at least 60\\% of the 8 and 24 $\\mu$m emission in M33 is diffuse. Conversely, a recent study by \\citet{crock13} determined that the 8 $\\mu$m emission in the galaxy NGC 628 has only a 30\\% - 43\\% fraction unassociated with recent star formation. These contradictory results call for further studies to assess the validity of the PAH emission as a SFR tracer.\\\\ \\indent The powerful \\it{Herschel} \\rm Space Observatory \\citep{pilbratt10} has targeted nearby galaxies with unprecedented resolution at the infrared wavelengths ($\\geq 70 \\mu$m) where the warm dust emission, powered primarily by massive stars, is progressively supplanted by the cold dust emission powered by low mass stars \\citep{bendo10,bendo12,boq11}. The synergy between the \\it{Spitzer} \\rm and \\it{Herschel} \\rm imaging data is such that the infrared spectral energy distribution of individual star forming regions from $\\sim$3 $\\mu$m to the sub-mm can be separated from that of the diffuse starlight in nearby galaxies, permitting perusal of the emission originating in these regions. With this level of detail, the properties of dust emission and its relation to the SFR can be better investigated. This paper uses data from both \\it{Spitzer} \\rm and \\it{Herschel} \\rm to observe the nearby galaxy, M33, in a range of IR wavelengths.\\\\ \\indent M33 is a member of the Local Group and a spiral galaxy, with an inclination of 56$^\\circ$ \\citep{regan94} and a distance of 840 kpc \\citep{freedman91}. Because of its proximity, M33 is an ideal site for the investigation of the properties of dust emission: 1$^{\\prime\\prime}$ subtends a spatial scale of $\\sim$4 pc. M33 harbors a large number of HII regions \\citep{bou74,hodge99,verley10, rel13}, which are easily identifiable in the \\it{Spitzer} \\rm and \\it{Herschel} \\rm images. M33 has an oxygen abundance of about half solar and a shallow--to--negligible metallicity gradient as a function of galactocentric distance \\citep{ros08, magrini09,bre11}; this characteristic enables us to investigate the PAH emission with less attention to the effect of metallicity on the strength of the PAH features \\citep{hunt05, eng05, mad06,draine072, eng08}. The \\it{Herschel} \\rm data for this project came from the HERM33ES open time key program \\citep{kramer10} and \\it{Spitzer} \\rm data were obtained through the Guaranteed Time observations for the IRAC and MIPS instruments \\citep{hinz04, mcquinn07}. We also make use of the H$\\alpha$ image from \\citet{hoopes00}, taken with the 0.6 meter Burrell-Schmidt telescope at Kitt Peak National Observatory. Details about the imaging and its reduction can be found in that paper.\\\\ \\indent Relationships among the PAH 8 $\\mu$m, 24 $\\mu$m, 250 $\\mu$m, and total infrared (TIR) emission are the primary subjects of this study, under the reasonable assumption that the 24 $\\mu$m emission is mainly tracing current SF \\citep{calz05,calz07} and that the 250 $\\mu$m emission is mainly tracing cold dust.\\\\ \\indent Our analysis plan is similar to that used by \\citet{bendo08} for the SINGS galaxies \\citep{ken03}: we will investigate the relationship of the 8 $\\mu$m emission with tracers of the warm (24 $\\mu$m) and cool (250 $\\mu$m) dust emission, and with the total infrared (TIR) emission. The advantage of our analysis over that of Bendo et al. is twofold: (1) M33 is 3 to 20 times closer than the SINGS galaxies, thus enabling exquisite spatial resolution; and (2) the 250 $\\mu$m emission from {\\it Herschel} traces more closely the cool dust and has more than twice the spatial resolution of the 160 $\\mu$m {\\it Spitzer} data used by Bendo et al.\\\\ \\indent In what follows, L($\\lambda$) refers to monochromatic luminosity derived as \\begin{equation} L(\\lambda) = [\\nu L_{\\nu}]_{\\lambda} \\end{equation} in units of erg s$^{-1}$ kpc$^{-2}$. In this work, each pixel will be $\\sim$73.3 pc in size, or about the size of a large HII region. ", "conclusions": "\\noindent In this paper, we have investigated the relationships between the 8 $\\mu$m emission and emission in other wavelengths that are correlated to old stellar populations, star formation, and cold dust. We have also compared our data with the predictions of the models from \\citet{draine07}. Ratios with 24 $\\mu$m and TIR luminosity show the 8 $\\mu$m emission to become underluminous in areas of strong 24 $\\mu$m emission. Taking the 24 $\\mu$m emission to be the product of heating of dust by young stellar radiation, this implies that 8 $\\mu$m emission is not a close tracer of young stellar populations, and the carriers responsible for its emission are possibly being destroyed by the intense radiation fields of star forming regions. Furthermore, the behavior of most ($\\ge$ 80\\%) 8 $\\mu$m emission can be explained using only the U$_{min}$ = U$_{max}$ models, which is valid for diffuse ISRF. We also find that the values of U$_{min}$ that account for the observed 8~$\\mu$m emission are in the range 0.1 - 25, consistent with the range found by Draine et al. (2014) for the galaxy M31. When applying a model that includes heating from HII regions, the observed 8 $\\mu$m values are lower than expected, supporting other authors' conclusions that the carriers responsible for 8 $\\mu$m emission are destroyed in intense stellar radiation fields. An alternative explanation to the destruction of the 8 $\\mu$m carriers is to make the PAHs more neutral, suppressing the 8 $\\mu$m emission and shifting power to the 11-12 $\\mu$m features\\citep{sand12}. Mid-IR spectroscopy may discriminate among these two scenarios, although we tentatively give preference to the PAH destruction interpretation.\\\\ \\indent Ratios of 8 $\\mu$m with 250 $\\mu$m emission show a strong correlation with both the TIR and 3.6 $\\mu$m luminosities. Both 3.6 $\\mu$m emission and L(8)/L(250) show a similar galactocentric radial trend, suggesting the 8 $\\mu$m emission may be more connected to old stellar populations than star formation. Further, the L(8)/L(250) as a function of 3.6 $\\mu$m luminosity shows the tightest relation among those presented in this paper, especially at high luminosity.\\\\ \\indent We also derive the fraction of 8 $\\mu$m and 24 $\\mu$m emission heated either by the ISRF or HII regions. The fraction of L(24) associated with diffuse emission is 60 - 43\\%, while the fraction of L(8) is nearly 33 - 42\\% higher, with 80 - 59\\% of the emission coming from heating by the ISRF.\\\\ \\indent The results of this study, which support findings by other authors, but at the exquisite spatial resolution enabled by the proximity of M33, suggest that the 8 $\\mu$m luminosity should not be used as a proxy for measuring and locating star formation in galaxies. Emission in this wavelength is shown in M33 to be more correlated with cold dust or old stellar populations than with star formation." }, "1402/1402.4136_arXiv.txt": { "abstract": "{We present a study of the overdensity of X-ray-selected active galactic nuclei (AGN) in 33 galaxy clusters in the XMM-LSS field (The XMM-Newton Large Scale Structure Survey), up to redshift $z=1.05$ and further divided into a lower ($0.14\\leq z\\leq 0.35$) and a higher redshift ($0.43\\leq z\\leq1.05$) subsample. Previous studies have shown that the presence of X-ray-selected AGN in rich galaxy clusters is suppressed, since their number is significantly lower than what is expected from the high galaxy overdensities in the area. In the current study we have investigated the occurrence of X-ray-selected AGN in low ($\\langle L_x, bol\\rangle=2.7\\times10^{43}$ erg/s) and moderate ($\\langle L_x, bol\\rangle=2.4\\times10^{44}$ erg/s) X-ray luminosity galaxy clusters in an attempt to trace back the relation between high-density environments and nuclear activity. Owing to the wide contiguous XMM-LSS survey area, we were able to extend the study to the cluster outskirts. We therefore determined the projected overdensity of X-ray point-like sources around each cluster out to $6r_{500}$ radius, within $\\delta r_{500}=1$ annulus, with respect to the field expectations based on the X-ray source $\\log N -\\log S$ of the XMM-LSS field. To provide robust statistical results we also conducted a consistent stacking analysis separately for the two $z$ ranges. We investigated whether the observed X-ray overdensities are to be expected thanks to the obvious enhancement of galaxy numbers in the cluster environment by also estimating the corresponding optical galaxy overdensities, and we assessed the possible enhancement or suppression of AGN activity in clusters. We find a positive X-ray projected overdensity in both redshift ranges at the first radial bin, which however has the same amplitude as that of optical galaxies. Therefore, no suppression (or enhancement) of X-ray AGN activity with respect to the field is found, in sharp contrast to previous results based on rich galaxy clusters, implying that the mechanisms responsible for the suppression are not as effective in lower density environments. After a drop to roughly the background level between 2 and $3 r_{500}$, the X-ray overdensity exhibits a rise at larger radii, significantly greater than the corresponding optical overdensity. The radial distance of this overdensity \\textquotedblleft bump\", corresponding to $\\sim 1.5 - 3$ Mpc, depends on the richness of the clusters, as well as on the overall X-ray overdensity profile. Finally, using the redshift information, photometric or spectroscopic, of the optical counterparts, we derive the spatial overdensity profile of the clusters. We find that the agreement between X-ray and optical overdensities in the first radial bin is also suggested in the 3-dimensional analysis. However, we argue that the X-ray overdensity \\textquotedblleft bump\" at larger radial distance is at least partially a result of flux boosting by gravitational lensing of background QSOs, confirming previous results. For high-redshift clusters the enhancement of X-ray AGN activity in their outskirts appears to be intrinsic. We argue that a spatial analysis is crucial for disentangling irrelevant phenomena affecting the projected analysis, but we are still not able to report statistically significant results on the spatial overdensity of AGN in clusters or their outskirts because we lack the necessary numbers.} ", "introduction": "As one of the most powerful extragalactic phenomena, active galactic nuclei (AGN) are a valuable tool in the study of the universe, since they can be used as cosmological probes, provide answers to various problems of galaxy evolution, and shed light on the innermost regions of galaxies, where the super massive black hole resides and highly energetic processes take place. However, the triggering mechanism of the omnipresent, but not always active, black hole is still elusive. Although major merging of gas-rich galaxies seems to be a highly probable mechanism for the triggering of nuclear activity (e.g., Sanders et al. 1988; Barnes \\& Hernquist 1991; Hopkins et al. 2006), recent studies of the morphology of AGN hosts find no evidence of a major merging-AGN connection (e.g., Cisternas et al. 2011; Kocevski et al. 2012). On the other hand, minor merging and interactions are still strongly disputed, while secular evolution also seems able to feed the central engine since many AGN are found to be isolated and undisturbed (e.g., Hopkins \\& Hernquist 2006; Cisternas et al. 2011; Kocevski et al. 2012). Therefore, the effect of the environment on the activity of the nucleus and vice versa is still fairly undetermined, but nevertheless crucial. Galaxy clusters represent the one end of the density spectrum in our universe, and as such it is an ideal place to investigate the effects of dense environment in the triggering of AGN, especially since an excessive number of X-ray point-like sources are undoubtedly found there (e.g., Cappi et al. 2001; Molnar et al. 2002; Johnson et al. 2003; D'Elia et al. 2004; Gilmour et al. 2009). Specifically, for the XMM-LSS field investigated in the current study, Melnyk et al. (2013) have found that 60\\% of X-ray-selected AGN reside in the overdense regions of group-like environment. Theoretically the feeding of the black hole can only be achieved by means of a non-axisymmetric perturbation that induces mass inflow. This kind of perturbation can be provided by interactions and merging between two galaxies, and the result of the inflow is the feeding of the black hole and activation of the AGN phase (e.g., Umemura 1998; Kawakatu et al. 2006; Koulouridis et al. 2006a, 2006b, 2013, Ellison et al. 2011; Silverman et al. 2011; Villforth et al. 2012; Hopkins \\& Quataert 2011). Thus, the cluster environment, where the concentration of galaxies is very high relative to the field, would also seem favorable to AGN. However, the rather extreme conditions within the gravitational potential of a galaxy cluster can work in the opposite direction as well. Ram pressure from the inter cluster medium (henceforth ICM) able to strip/evaporate the cold gas reservoir of galaxies (Gunn \\& Gott 1972; Cowie \\& Songaila 1977; Giovanelli et al. 1985) can strongly affect the feeding of the AGN. Nevertheless, other studies have argued that ram pressure stripping cannot be as effective in transforming blue sequence galaxies to red (e.g., Larson et al. 1980; Balogh et al. 2000, 2002; Bekki et al. 2002; van den Bosch et al. 2008; Wetzel et al. 2012), especially in lower density clusters where other processes should take place as well. Large velocity dispersion of galaxies within clusters could also prevent the effective interactions (Aarseth \\& Fall 1980), particularly mergers, while the fast \\textquotedblleft grazing\" bypassing galaxies may also cause gas stripping by \\textquotedblleft harassment\" (e.g., Natarajan et al. 2002, Cypriano et al. 2006). However, the efficiency of this phenomenon has once more been questioned (e.g., Giovanardi et al. 1983). A combination of the above mechanisms, in addition to the possible prevention of accretion of halo mass into cluster galaxies (\\textquotedblleft strangulation\"; e.g., Larson et al. 1980; Bekki et al. 2002; Tanaka et al. 2004) may, in fact, suppress the AGN activity in clusters despite the number of potentially merging and interactive galaxies. When using only optical data, the results seem to remain inconclusive. Early studies reported that AGN are less frequent in galaxy clusters than in the field (Osterbrock 1960; Gisler 1978; Dressler, Thompson \\& Schectman 1985) and recent large-area surveys support this suggestion (Kauffmann et al. 2004; Popesso \\& Biviano 2006; von der Linden et al. 2010; Pimbblet et al. 2013). Other studies, however, have found no differences between clusters and field (e.g., Miller et al. 2003), at least when selecting the weak AGN (e.g., Martini et al. 2002; Best et al. 2005; Martini et al. 2006; Haggard et al. 2010). We should note here that considering only the optical wavelengths is not the optimal way of finding AGN since they suffer greatly from absorption. Especially if gas depletion is at play and low accretion rates are expected, then most of the spectral signatures of the AGN could be \\textquotedblleft buried\" in the host galaxy. Radio-loud AGN on the other hand, seem to be more clustered than any other type of galaxy (Hart, Stocke \\& Hallman 2009) and are often associated with BCGs (brightest cluster galaxies)(e.g., Best 2004; Best et al. 2007). In addition, the fraction of X-ray AGN in BCGs is higher than in other cluster galaxies (e.g., Hlavacek-Larrondo et al. 2013, and references therein). These findings can be attributed to hot gas accretion from the hot X-ray cluster halo, although gas from any other source fueling the black hole at low accretion rates would also have the same effect. If the hot gas accretion is a possible fueling mechanism for the X-ray AGN, as well, then we should expect them to reside primarily within clusters. Undoubtedly, the best way to detect active galaxies is through X-ray observations (e.g., Brandt \\& Alexander 2010). However, during the previous decade, only a small fraction of X-ray point-like sources in clusters had positive confirmation as true cluster members (see Martini et al. 2002; Davis et al. 2003; Finoguenov et al. 2004; Arnold et al. 2009), leaving the question of whether the positive X-ray overdensities found in galaxy clusters represent enhancement or suppression of the nuclear activity unanswered. More recent studies, however, report more conclusive results by comparing X-ray to optical data. Koulouridis \\& Plionis (2010) demonstrate the significant suppression of X-ray-selected AGN in 16 rich Abell clusters (Abell et al. 1958) by comparing the X-ray point source overdensity to the optical galaxy overdensity. Ehlert et al. (2013; 2014) argue that the X-ray AGN fraction in the central regions of 42 of the most massive clusters known to date is about three times lower than the field value using the same technique. More importantly, after having complete spectroscopy for their X-ray point source sample, Haines et al. (2012) argue that X-ray AGN found in massive clusters are an infalling population, which is \\textquotedblleft extinguished\" later, and confirm the suppression in rich clusters. On the other hand, Martini et al. (2013) argue that this trend is not confirmed for a sample of high-redshift clusters ($1.010^{43}$ erg/s) is consistent with the field. We note, however, that the high-redshift regime studied and the large AGN photometric redshift uncertainties ($\\sigma_z=0.12(1+z)$, double that of normal galaxies) introduce some level of uncertainty to the results. Neverteless, they agree with findings from the DEEP2 Redshift Survey\\footnote{http://deep.ps.uci.edu/} that show that only below $z$=1.3 does the fraction of blue galaxies in groups drop rapidly and become constant below $z$=1 (Gerke et al. 2007), while the red fraction correlates weakly with overdensity above $z$=1.3 (Cooper et al. 2007). In the present study we only deal with clusters $z<$1.05, where the cluster's population is dominated by early type red galaxies. Finally, we should also mention that an indirect way to address the issue is by X-ray clustering analyses, but still their results also remain inconclusive (see relevant discussion in Haines et al. 2012 \\S5.2). Considering the above, there is still the need to clarify the influence of the environment on the AGN phenomenon. And while the majority of the above studies are dealing with the most massive and rich clusters, the population of moderate-to-poor clusters is still overlooked.\\footnote{We should note that the categorization of galaxy clusters to different richness classes is not explicit, and it is safer to be used statistically. Nevertheless, the relation between mass and X-ray properties is well studied (e.g., Edge \\& Stewart 1991a, b; Finoguenov, Reiprich \\& Bohringer, 2001) and, also considering more recent studies of X-ray luminous clusters (e.g., Ebeling et al. 2010), we can infer that massive clusters have X-ray luminosities higher than $\\sim5\\times10^{44}$ erg/s and temperatures higher than $kT>5$ keV.} If the reason for the deficiency of X-ray AGN in rich clusters is the strong gravitational potential, which provides the necessary conditions for the suppression (whichever these may be: gas stripping, strangulation, tidal stripping, evaporation, high velocity dispersion, etc.), one would expect the AGN presence to rise in shallower gravitational potentials. A similar relation between the strength of the gravitational potential and star formation quenching (Popesso et al. 2012) supports the above expectation (see also Wetzel et al. 2012). Another issue is the radial extent of the search for X-ray AGN around clusters. An enhancement of AGN activity is observed far from the cluster's center (e.g., Fassbender et al. 2012), and it could be due to an infalling population (Haines et al. 2012) coming from the \\textquotedblleft outskirts\" of the clusters where the concentration of galaxies is still high. The question is where should we place the \\textquotedblleft outskirts\" and to what extent. Most studies could not reach farther than a 2$r_{500}$ radius, although the overdensity profile of optical galaxies remains higher than the field level even beyond that radius (e.g., Ehlert et al. 2013). Finally, what is also overlooked is the background overdensity of X-ray sources in the area of clusters. In Koulouridis et al. (2010), we used SDSS (Sloan Digital Sky Survey) optical data for all the detected X-ray point-like sources within a 1 Mpc radius, and argue that their positive overdensity values were associated with background QSOs rather than cluster members. A possible cause is the gravitational lensing of background sources, which is unimportant when compared to the large number of optical galaxies in clusters but can become very important for X-ray sources and affect the assessment of their clustering. In the current study, our aim is to investigate the AGN phenomenon in the environment of moderate and poor clusters located in the XMM-LSS contiguous field of 11.1 deg$^2$. We identify all possible X-ray AGN candidates, which we define as sources with $L_{(0.5-2{\\rm\\; keV})}>10^{42}$ erg/s at the redshift of the cluster, and compare their overdensity in the area of the clusters to the respective overdensity of optical galaxies, available by the CFHT legacy survey. Such a large contiguous area gives us the unique opportunity not only to use a large cluster sample but also to extend our search for X-ray AGNs around clusters at great distances, reaching homogeneously up to a 6$r_{500}$ radius. In addition, we make use of photometric redshift data calculated specifically for X-ray-selected AGN hosts, in an attempt to assess the true number of X-ray AGNs in our clusters and clarify the effect of the excessive overdensity of background X-ray sources (e.g., Koulouridis \\& Plionis 2010). We describe our samples and methodology in \\S2, while our results and conclusions are presented in \\S3 and \\S4, respectively. Throughout this paper we use $H_0=72$ km/s/Mpc, $\\Omega_m=0.27$, and $\\Omega_{\\Lambda}=0.73$. ", "conclusions": "We conducted a statistical study of 33 clusters of poor and moderate richness, within the XMM-LSS field that covers $\\sim$20$\\%$ of the XXL survey, by comparing the density of X-ray sources within multiples of the $r_{500}$ radius with the expected field density, calculated from the $\\log N - \\log S$ for the same area. We compared this projected overdensity with the respective optical galaxy overdensity in an attempt to estimate the suppression or the enhancement of X-ray-selected AGN. In addition, we calculated the spatial overdensities, using the available spectroscopic and photometric redshifts in an attempt to identify and quantify the true cluster members and explain the results of our previous projected analysis. The conclusions that can be drawn from the above analysis are the following: $ \\bullet$ The projected analysis of X-ray versus optical overdensity within the two central $r_{500}$ annuli, corresponding to $\\sim 1$ Mpc radius, results in a strong positive signal showing that the environment of the low and moderate X-ray-luminosity clusters of our samples does not suppress the X-ray AGN activity. This result is in sharp contrast to the outcome of many studies of rich clusters (e.g., Koulouridis \\& Plionis 2010; Ehlert et al. 2013; Haines et al. 2012), which implies that lower richness cluster environments do not suppress X-ray AGN activity. Interestingly, in even lower density environments (galaxy groups), an enhancement of X-ray AGN may be present (Melnyk et al. 2013). $ \\bullet $ After calculating the projected overdensities at large radial distances from the center of the cluster ($3^{rd}-5^{th}$ bins, corresponding to $\\sim 1.5$-3 Mpc, depending on the redshift), a significant rise in the X-ray source overdensity is observed. This excess has also been reported in previous studies (Haines et al.2012, Fassbender et al. 2012) and has been attributed to an infalling population of galaxies from the outskirts of the clusters that interact and merge, producing the observed overdensity of X-ray AGN. This surplus is confirmed for both our low- and high-redshift clusters. $ \\bullet $ Using spectroscopic and photometric redshifts, we discovered that the X-ray \\textquotedblleft bump\" at a large radial distance vanishes completely from the poor low-$z$ sample, and we argue therefore that this density excess may be produced by flux boosting of background sources due to gravitational lensing, sometimes even enhanced by additional background galaxy clusters along the same line of sight. On the other hand, a high X-ray source overdensity persists in the last annulus of the moderate X-ray luminosity high-$z$ sample, implying that for intermediate \\textquotedblleft -richness\" clusters, additional triggering of X-ray AGN in the outskirts is still possible. In a nutshell, the projected overdensity analysis produces statistically significant results, but at the same time these results are contaminated by projection effects of background-lensed QSOs. On the other hand, although the spatial analysis performed is free of these effects, it is not able to reach definite results owing to the small numbers involved, making it necessary to study larger samples of galaxy clusters. Especially the area included in the annuli closer to the cluster center is so small that does not allow us to reach any definitive conclusions about the suppression of X-ray AGN. The stacking of clusters proves very useful, but splitting the total sample into two redshift subsamples again reduces the numbers greatly. However, the division is crucial since not only do we select a population of more X-ray luminous clusters in higher redshifts, but we can also detect only higher luminosity X-ray AGN. We should stress that the large contiguous area of the XMM-LSS has allowed us to study the overdensity of X-ray AGN within large radial distances from the cluster center for the first time. This proved to be essential for exploring the relation between the dense environment of clusters and the X-ray AGN activity in detail. To fully understand this relation, we need to trace its evolution as a galaxy approaches the cluster's gravitational potential, enters the hot ICM, and crosses the cluster. At the same time, we need to disentangle irrelevant effects such as the gravitational lensing of background sources, probably enhanced by the presence of additional clusters along the line of sight. A photometric variability study of these sources may also shed some light on whether the lensing amplification could be due to micro-lensing and/or convergence by matter in the clusters. We believe that the analysis of the full XXL field, which is almost five times larger than the XMM-LSS (reaching 50 deg$^2$), together with a detailed spectroscopic follow-up of the optical counterparts of all X-ray point sources, detected in the XXL clusters, may provide reliable and robust results as to the origin (true enchancement, lensing, presence of background clusters, etc) of the excess X-ray sources detected in the outer $\\sim 3-5 r_{500}$ annuli of either low and high-redshift clusters." }, "1402/1402.0073_arXiv.txt": { "abstract": "Variations in stellar flux can potentially overwhelm the photometric signal of a transiting planet. Such variability has not previously been well-characterized in the ultraviolet lines used to probe the inflated atmospheres surrounding hot Jupiters. Therefore, we surveyed 38 F-M stars for intensity variations in four narrow spectroscopic bands: two enclosing strong lines from species known to inhabit hot Jupiter atmospheres, \\ion{C}{2} $\\lm\\lm$1334,1335 and \\ion{Si}{3} $\\lm$1206; one enclosing \\ion{Si}{4} $\\lm\\lm$1393,1402; and 36.5 \\AA\\ of interspersed continuum. For each star/band combination, we generated 60~s cadence lightcurves from archival {\\it HST} COS and STIS time-tagged photon data. Within these lightcurves, we characterized flares and stochastic fluctuations as separate forms of variability. {\\bf Flares:} We used a cross-correlation approach to detect 116 flares. These events occur in the time-series an average of once per 2.5 h, over 50\\% last 4 min or less, and most produce the strongest response in \\siv. If the flare occurred during a transit measurement integrated for 60 min, 90/116 would destroy the signal of an Earth, 27/116 Neptune, and 7/116 Jupiter, with the upward bias in flux ranging from 1-109\\% of quiescent levels. {\\bf Fluctuations:} Photon noise and underlying stellar fluctuations produce scatter in the quiescent data. We model the stellar fluctuations as Gaussian white noise with standard deviation $\\s_x$. Maximum likelihood values of $\\s_x$ range from 1-41\\% for 60~s measurements. These values suggest that many cool stars will only permit a transit detection to high confidence in ultraviolet resonance lines if the radius of the occulting disk is $\\gtrsim$ 1 R$_J$. However, for some M dwarfs this limit can be as low as several $R_\\earth$. ", "introduction": "\\label{sec:intro} Transit observations in the far-ultraviolet (FUV, $1200\\leq\\lambda\\leq1400$ \\AA) have revealed the existence of inflated atmospheres surrounding the hot Jupiters HD209458b \\citep{madjar03,madjar04,linsky10} and HD189733b \\citep{lecavelier10,bourrier13,jaffel13}. During the transit of these planets, the UV resonance transitions of several species in their atmospheres -- \\ion{H}{1}, \\cii, \\ion{O}{1}, and \\siii\\ -- produce detectable absorption against the background stellar emission. The depth of the absorption indicates that these species occupy a volume overfilling the planets' Roche lobes, suggesting atmospheric escape. Furthermore, such UV transit spectrophotometry can constrain the atmospheric mass loss rate (e.g. \\citealt{madjar04,lecavelier10}) and characterize the atmospheric response to changes in the stellar radiation and particle flux \\citep{lecavelier12,bourrier13}. However, variability in the background flux source itself -- the host star -- presents additional, instrument-independent challenges for all transit observations. In a lightcurve, the transit signal can be spuriously weakened, sometimes completely obliterated, by flares buried within it, or amplified by flares flanking it. Outside of flares, the transit signal can be obscured by stochastic fluctuations in stellar luminosity that act as additional noise, compounding that of photon statistics and instrumental sources. These forms of stellar variability fundamentally limit transit observations. Therefore, evaluating the possibilities for future transit work necessitates measurements of this variability for potential host stars. The list of potential targets is diverse, encompassing stars beyond those just on the main sequence, such as HD209458b and HD189733b. Such candidate targets could include stars with transitional or debris disks (encompassing weak-line T-Tauri stars, WTTS) where planets might be in the process of coalescing from disk material. For example, \\citet{eyken12} report on the transit signature of a super-Jupiter orbiting a WTTS star in the Orion-OB 1a/25-Ori region. High contrast imaging of stars with disks has also revealed (proto)planetary objects or evidence for these objects through disk gaps, such as LkCa15 \\citep{kraus12}, PDS 70 \\citep{hashimoto12}, RX J1633.9-2442 \\citep{cieza12}, and TW Hya \\citep{debes13}. Candidate targets also include post main-sequence stars. The evolved F5-F7 stars WASP-76, WASP-82, and WASP-90 host transiting hot Jupiters \\citep{west13}, as does the G8III giant HIP 63242 \\citep{jones13}. In addition, observations of remnant debris orbiting white dwarfs (e.g. \\citealt{farihi13}) hint at the possibility of detecting extant or decomposing planets around stars nearing the end of their lives. To begin characterizing the range of background fluctuations faced by FUV transit observations, we have conducted a survey of stellar flares and stochastic fluctuations for the largest possible sample of stars with archival FUV photon event data from {\\it HST} covering \\cii\\ $\\lm\\lm$1334,1335 and \\siii\\ $\\lambda$1206, two key lines for probing outer atmospheres of close-in exoplanets. We also include \\siv\\ $\\lambda\\lambda$1394,1403 and a composite of interspersed continuum bands. We did not attempt to include \\ion{H}{1} $\\lambda$1216 (Ly$\\alpha$) or \\ion{O}{1} $\\lambda$1302 due to the correction for geocoronal emission that is required. This survey relies on UV data acquired by the two powerful UV spectrographs on board the {\\it Hubble Space Telescope (HST)}: the Cosmic Origins Spectrograph (COS), with its G130M and G140L gratings, and the Space Telescope Imaging Spectrograph (STIS) with its E140M grating. To enable an analysis of temporal variability, we traded spectral resolution for temporal resolution. Thus, we employed short, 60~s time bins and summed photon counts over roughly the full width ($\\sim$ 1-2 \\AA) of each line and 36.5 \\AA\\ of interspersed continuum to create lightcurves. In comparison, transit observations have achieved resolutions of $\\sim0.1$ \\AA\\ (roughly 20-30 km s$^{-1}$) when coadding an entire transit dataset. However, these same observations commonly integrate over the full line-widths for increased signal to noise (e.g. \\citealt{madjar03,lecavelier10}). We present the results of an analysis of 153 lightcurves of 60~s cadence, including (1) 116 flares we identified by cross-correlating with a flare-like kernel and (2) estimates or upper limits for the standard deviation of stochastic fluctuations in the quiescent portions of the lightcurve. In the remainder of this introduction, we expand upon the implications and physical sources of stellar variability and provide pointers to previous variability surveys. In Section~\\ref{sec:stars}, we describe the stellar sample and process of generating lightcurves from the FUV time-tagged photon data. In Section~\\ref{sec:analysis} we outline the variability analysis, treating flares and stochastic fluctuations separately. In Section~\\ref{sec:results}, we present the results, followed in Section~\\ref{sec:discussion} by a discussion, including the implications for observing transits. We summarize the work in Section~\\ref{sec:summary}. Throughout this paper, we will treat 3.5~h as the typical transit timescale. This is near the average transit duration of 3.6 h (sample standard deviation 1.9 h) for exoplanets listed in the Exoplanet Data Explorer \\citep{wright11} as of 2013 October. \\subsection{A Brief Discussion of Stellar Variability} \\label{sec:varintro} Stellar variability can be divided into three categories based on the differing implications for transit observations: \\begin{enumerate} \\item Periodicities: Oscillations of the stellar flux by phenomena that can be sufficiently characterized to predict the level of modulation over the course of a transit. \\item Flares: Bursts of brightness isolated in time with respect to the cadence of the data, and well above the quiescent scatter in the lightcurve. \\item Stochastic Fluctuations: Variations in the stellar flux that are too chaotic to be accurately predicted on transit timescales. \\end{enumerate} Predictable periodic signals are surmountable obstacles. Any periodic signal strong enough to be even modestly detected would allow an accurate model fit, bootstrapped over many cycles, such that removing or accounting for the signal would not obscure an otherwise detectable transit. Strong flares complicate the lightcurve analysis. While it is possible to describe the distribution of solar flares in both strength and frequency \\citep{cassak08}, a statistical model cannot predict the onset, magnitude, and evolution of flares for any specific timeline, such as during a transit. Strong flares thus pose the risk of overwhelming, and weak flares of attenuating, a transit signal. Flares occurring near a transit could augment estimates of the out-of-transit flux, spuriously deepening the transit signal. Beyond interfering with transit observations, flares might also impact the atmospheres of exoplanets, such as the stripping of atomic hydrogen from HD189733b conjectured by \\citet{bourrier13} to be caused by an observed host-star flare. Stochastic fluctuations, however, represent the greatest barrier to transit photometry. Because these fluctuations cannot be predicted deterministically over the course of a transit, they must be treated as noise. As with photon noise, they pose the risk of obscuring a true signal or causing a false one. This ``noise\" can be overcome by averaging measurements until the uncertainty is within tolerable limits, but with a caveat: Unlike photon noise, stochastic fluctuations are probably not white noise. For example, in 60~s cadence, broadband optical photometry from {\\it Kepler}, stochastic fluctuations of stellar flux (attributed to granulation and magnetic activity) have a power spectrum that, unlike white noise, is not flat \\citep{gilliland10}. Although {\\it Kepler} measures broadband optical flux, not the chromosphere and transition region FUV emission line flux we analyzed, the fluctuations of the chromospheric near UV flux from YZ CMi also show a frequency-dependent power spectrum \\citep{robinson99}. Therefore, it seems probable that stochastic fluctuations in stellar flux will not behave as white noise in any band. What appears as true noise at one cadence would resolve into smooth variations at some faster cadence. Below this threshold cadence, lightcurve points will be highly correlated. Binning adjacent flux measurements in this regime will not average out the scatter. The emission line flux we analyzed samples regions with temperatures of $10^4-10^5$ K in the outer atmospheres of stars. Specifically, the peak formation temperatures of the lines in solar conditions are estimated by \\citet{dere09} to be $3.2\\sn4$ K for \\cii\\ $\\lm\\lm$1334,1335, $5\\sn4$ K for \\siii\\ $\\lambda$1206, and $8\\sn4$ K for \\siv\\ $\\lambda\\lambda$1393,1402. Reference models of the solar atmosphere place emission from these lines in the thin transition region between the chromosphere and the corona. In contrast to the lines, the solar FUV continuum (shortward of 1500 \\AA) forms at slightly lower altitudes, primarily in a region of initial chromospheric temperature rise above the photospheric temperature minimum \\citep{linsky12a}. Variability in these regions of stellar atmospheres can result from several phenomena. Although drawing conclusions about these underlying physical phenomena is not our objective, the generally accepted origins of periodicities, flares, and stochastic fluctuations bear mentioning. Periodic variability can be the result of a pulsational instability in the star \\citep{gautschy95, gautschy96} and/or rotation of long-lived, localized brightness variations (starspots, faculae, etc.) through the observer's field of view \\citep{vaughan81}. A periodic signal can result from extrinsic phenomena as well. Gradual oscillations in flux are produced by phase changes of an orbiting planet (e.g. \\citealt{borucki09}). Isolated, but nonetheless periodic, dips in flux occur when an orbiting object, such as an exoplanet or stellar companion, transits the host star (e.g. \\citealt{wilson71}). Flares are generally thought to be the result of magnetic reconnection events in the corona that abruptly convert magnetic energy into plasma kinetic energy. Some of this energy is deposited in the chromosphere and photosphere and radiated away \\citep{haisch91,gershberg05}. Flares commonly produce a sharp rise in flux followed by an exponential decay lasting from hours to minutes, possibly even seconds \\citep{pettersen89,gershberg05}. Within a single flare, multiple peaks and changes in the decay rate are possible. Some researchers have identified as flares events in which the stellar flux rises and fades more gradually \\citep{houdebine03,tovmassian03}. The strength and frequency of flares typically exhibit an inverse power-law relationship (e.g. \\citealt{shakhovskaia89,audard00} for stellar flares, \\citealt{lin84,nita02} for solar flares). This implies that weaker flares are more prevalent than conspicuous events, such that many flares will occur in observations that cannot be clearly resolved as such. In fact, if the power law is steep enough, the lowest energy flares, often termed microflares, might inject enough heat into the corona of a star to explain the high temperatures present there \\citep{hudson91,audard00}. Low energy ``microflares'' or even ``regular'' flares, if the data is not of sufficient quality to resolve them, will contribute to the observed stochastic fluctuations of a target. For instance, \\citet{robinson99} suggest microflaring as an explanation for the quiescent stochastic fluctuations in near UV flux that they observed from YZ CMi . They simulate the production of such stochastic fluctuations with a microflare model and find that it closely resembles the YZ CMi quiescent data. Ultimately, the extent of the contribution of flares to the observed stochastic fluctuations of any target is determined by the level of stellar magnetic activity, the photometric quality of the data, and the threshold set for identifying a lightcurve anomaly as a flare rather than fluctuation. The remaining proportion of stochastic fluctuations in transition region line emission could be explained by several phenomena. Transition region explosions, smaller events possibly associated with magnetic restructuring at the edges of newly emerging flux loops \\citep{gershberg05}, could introduce variability while also serving as a dominant heating source for the transition region. \\citet{wood97} suggested such events might explain broad components of \\siv\\ and \\ion{C}{4} emission in the FUV spectra of 11 late-type stars. However, \\citet{peter06} suggests magnetic flux braiding and consequent Joule dissipation might be the dominant heat source for the transition region. Both braiding of surface field and the emergence of field loops produce pockets of rapid heating in the three-dimensional MHD models of \\citet{hansteen10} that could explain much of the temporal variability of line emission originating in the transition region. In the \\citet{hansteen10} model, the injected energy results from work done on the magnetic field by photospheric motions, tying transition region variability to the convective cells and p-mode oscillations within the star. These convective cells and p-mode oscillations also affect the transition region environment by initiating high altitude shock waves \\citep{wedemeyer09}. Variability of a planet-hosting star could be influenced by the planet itself. Planets orbiting close enough to a star will interact tidally and, possibly, magnetically with the host \\citep{cuntz00}. Magnetic interactions could lead to flares from the reconnection of planetary and stellar fields \\citep{rubenstein00,lanza08}, increased stochastic fluctuations from overall magnetic activity enhancements \\citep{cuntz00}, or periodicities from enhanced plages and faculae surrounding the sub-planetary point on the star \\citep{lanza08,cohen09,kopp11}. Tidal interactions could produce flows and turbulence associated with the tidal bulge \\citep{cuntz00}. They could also spin up the star \\citep{aigrain08}, indirectly increasing overall stellar magnetic activity. These interactions are supported by some evidence (beginning with \\citealt{shkolnik03}), but more definitive conclusions require future, dedicated observations \\citep{lanza11}. \\subsection{A Selection of Relevant Flare and Variability Studies} \\label{sec:context} There is a long history of research into the frequency and intensity of flares on the Sun and other stars. Especially relevant is recent work by \\citet{hilton10} and \\citet{davenport12}, and references therein, examining large (several $10^4$) samples of M dwarf stars using multi-epoch data in the optical from the Sloan Digital Sky Survey (SDSS) and in the infrared from the Two Micron All Sky Survey (2MASS). \\citet{tofflemire12} specifically assessed the impact of M dwarf flares on exoplanet observations in the infrared using three such stars. Recently, \\citet{kowalski13} conducted a detailed spectrophotometric study in the near UV and optical of 20 M dwarf flares in order to probe the various mechanisms responsible for flare emission. Previous studies in the far and extreme UV are scarcer. \\citet{welsh07} leveraged data from the {\\it Galaxy Evolution Explorer} in the broadband near and far-ultraviolet to find 49 variable sources exhibiting 52 flares. The {\\it Galaxy Evolution Explorer} FUV band data contain the \\siv\\ line we analyzed for variability. In addition, \\citet{mullan06} examined 44 F-M stars in broadband extreme UV time-series data from the {\\it Extreme Ultraviolet Explorer}. The band they utilized is dominated by emission lines of \\ion{Fe}{18} -- \\ion{Fe}{22} formed at coronal temperatures upwards of $10^7$ K, expected in magnetically active regions. Several previous studies have quantified the stochastic variability of large samples of stars in the optical, most notably employing {\\it Kepler} results to place the Sun's well-characterized variability in the context of other stars (see \\citealt{basri10} and \\citealt{mcquillan12} for examples using {\\it Kepler} data and \\citealt{eyer97} for one using {\\it Hipparcos} data). In addition, it is standard practice to quantify the variability of the exoplanet host star complimentary to radial-velocity or transit measurements, so many individual measurements of stellar variability exist (e.g. \\citealt{dragomir12, kane11, berta11}). However, to the knowledge of the authors this paper presents the first analysis, focusing specifically on the implications for transit observations, of stellar variability in UV line emission flux. ", "conclusions": "\\label{sec:discussion} \\subsection{The Range of Flare Behavior} \\label{sec:flarebehave} \\subsubsection{Strong Flares} Several flares appear in the data that peak at tens of times the quiescent flux. The strongest of these is a flare on Prox Cen, displayed in Figure~\\ref{fig:flarepalette}a (see also \\citealt{christian04}). This flare raises the continuum-subtracted flux in each band, normalized by the quiescent mean, to values at the peak data point of $27.5\\pm0.2$ in \\cii, $66.7\\pm1.2$ in \\siii, $91.6\\pm0.3$ in \\siv, and $13.5\\pm0.6$ in the continuum. The $\\nfe$ and duration of this flare in each band can be compared with the rest of the flare sample in Figure~\\ref{fig:flaredist}, where these data points are labeled with 39, Prox Cen's dataset number from Table~\\ref{tbl:obs}. Very strong flares also appear in the AD Leo (dataset no. 36; see also \\citealt{hawley03}) and IL Aqr (dataset no. 37; see also \\citealt{france12}) data and are also labeled in Figure~\\ref{fig:flaredist}. The AD Leo flare peaks at $8.2\\pm0.1$ in \\cii, $14.0\\pm0.2$ in \\siii, $33.7\\pm0.1$ in \\siv, and $45.9\\pm0.3$ in the continuum (Figure~\\ref{fig:flarepalette}b). The IL Aqr flare peaks at 15.4$\\pm0.3$ in \\cii, $36.9\\pm0.3$ in \\siii, $46.5\\pm0.3$ in \\siv, and $20.1\\pm1.4$ in the continuum (not included in Figure~\\ref{fig:flarepalette}). \\subsubsection{Symmetric Flares} Some of the flares in the sample show roughly equal rise and decay times, in contrast to the impulse-decay shape of many flares. The clearest example of such a flare is that which appears in the EG Cha data, plotted in Figure~\\ref{fig:flarepalette}c. Note that this event was not flagged in the continuum or \\cii\\ data because the short span of the data, 0.82 hours, resulted in a poor determination of the quiescent scatter. The continuum and \\cii\\ did not sufficiently exceed the estimated scatter to result in an identification. Other flares that show a relatively clear symmetric photometric profile (though not nearly as clear as that of EG Cha) appear in the \\siv\\ data of EV Lac at mean Julian Date 52172.709 and the AD Leo data at mean Julian Date 51616.119. Many flares in the data appear as though they might be symmetric, but their rise and fall are not adequately resolved. \\subsubsection{F Star Anomalies} The flare identification algorithm flagged two events on F stars, one in \\cii\\ on $\\delta$ Cep and another in the continuum on $\\beta$ Dor. The latter is displayed in Figure~\\ref{fig:flarepalette}d. Both events are gradual elevations of the flux in a single band relative to the other three. The $\\beta$ Dor flare is particularly curious, as the other three bands clearly do not show the same consistent decline in flux that the elevated continuum flux does. These events seem likely to be true anomalies rather than spurious detections. \\subsubsection{Multi-peak Flares} Many of the flares in the data exhibit complicated shapes, and might be a superposition of nearby small flares. Figure~\\ref{fig:flarepalette}e depicts a flare on AD Leo that exhibits two peaks, separated by 240 s. This was the clearest specimen of a multi-peak flare. In most such flares, the data are not as clearly resolved above the surrounding scatter and/or different bands show different behavior. \\subsubsection{Response in Different Lines} We found that flare signals generally vary significantly between bands. In addition, these variations are not consistent from flare to flare. The difference in the response between bands is clear in Figure~\\ref{fig:flarepalette}. For example, the Prox Cen, EG Cha, and AD Leo flares of subplots a, c, and e peak highest in \\siv, with a roughly similar shape in each band. Alternatively, the AD Leo flare depicted in subplot b is the strongest in the continuum, and that of Prox Cen depicted in subplot f is the strongest in \\siii. For the latter, the continuum shows essentially no response, though better signal to noise might reveal otherwise. (Note in Figure~\\ref{fig:flarepalette}f that some \\siii\\ points are negative because the subtracted signal from the Ly$\\alpha$ wing sometimes exceeds the low \\siii\\ signal.) Differences between lines in emitted flux during a flare have also been observed on the Sun (e.g. \\citealt{brekke96}) and other low mass stars (e.g. GJ876, \\citealt{ayres10,france12}). \\begin{figure} \\includegraphics[width=\\columnwidth]{f14.pdf} \\caption{A palette of flare behaviors, discussed in Section~\\ref{sec:flarebehave}. Each subplot is labeled with the star and the mean Julian Date of $t = 0$. The fluxes are continuum-subtracted, high-pass filtered, and normalized to the mean of the quiescent points. The x symbols denote points flagged for removal before computing excess noise values (\\emph{not} equivalent to the flare duration, see Section~\\ref{sec:flareresults}. Note that the y-axis scales differ.} \\label{fig:flarepalette} \\end{figure} \\subsection{Risks Flares Pose to Transit Measurements} \\label{sec:flarerisks} All of the detected flares are short lived, with one lasting 27 min, the others lasting less than 20 min, and over half lasting 4 min or less. These could easily be hidden by longer cadence data. Typical cadences in UV exoplanet transit observations are $\\dt \\approx$ 30-60 min, based on recent literature. While hour or longer cadences might often be unavoidable due to instrument or signal-to-noise limitations, a flare hidden in such data could bias a measurement of transit depth (see Section~\\ref{sec:varintro}). Of the 116 flares we identified (again, excluding FK Com), 57 (roughly one event per 5 h of data) would boost a 60 min integrated flux measurement by $\\gtrsim10\\%$, and all flares (roughly one event per 2.5 h of data) would boost a 60 min flux measurement by $\\gtrsim\\ $1\\%. Flares produced the largest $\\nfe$ values in the \\siv\\ band in roughly 2/3 of the 60~separate events and 7/9 of the events that registered a detection in all four bands. Thus, it appears that transit observations in \\siv\\ are somewhat likely to be more strongly affected by flaring. The flare stars (namely AD Leo and Prox Cen) account for most of the flares we detected, 92 of 116. However, 24 flares were identified on objects not classified as ``flare stars.'' Such flares are of particular interest because these stars are more likely to be targeted in exoplanet search programs. These occurred on 5/8 M, 1/8 K, 3/11 G, and 2/5 F stars, and occurred roughly once per 5 h in the time all 32 stars were observed. Of these 24 flares, 9 (roughly one event per 13 h) would boost a 60 min integrated flux measurement by $\\gtrsim10\\%$. Most such flares occurred on M stars (7/9), while the remaining 2/9 are the same event observed in two different bands on the K star EG Cha. A 10\\% boost in flux exceeds the transit depth in \\cii\\ or \\siii\\ of HD209458b \\citep{madjar04,linsky10}. However, not all stars in the sample are similar in size to HD209458, and it is the relative size of the star and planet that determines the transit depth and thus the impact of a flare. In other words, a strong flare will have less of an impact on a measurement of a Jupiter transiting and M star than a Jupiter transiting an F star. We used the stellar radii from the literature to determine the size of an object that would produce a transit signal of the same amplitude as the boost in flux from a flare in a 60 min integration. These range from 0.02 $R_J$ to 12.5 $R_J$ for the 116 flares identified. Of these, 90 (one flare per 3 h of data) would boost a 60 min flux integration by an amount larger than the signal of an Earth transiting the flaring star, 23 (one flare per 13 h) would boost flux by an amount exceeding a Neptune transit, and 7 (one flare per 42 h) would boost flux by an amount exceeding a Jupiter transit. Limiting the sample to the 24 events on non-flare stars only, 5 flares (one per 31 h) boost flux beyond an Earth signal, 2 (one per 76 h) beyond a Neptune signal, and 1 (one per 153 h) beyond a Jupiter signal. We did not explore trends in flare rates with respect to stellar properties because the rates were only well constrained on Prox Cen and AD Leo. In addition, the small size, high diversity, and, most importantly, range of flare detection limits in the sample pose problems to such an analysis. Our general conclusion is that all low-mass stars likely pose a risk of flaring near or within a transit observation. Because flares might be so easily hidden in long-cadence data, we recommend that, when possible, transit observers employ minute-scale cadences to inspect their data prior to employing longer cadences for noise suppression. This would enable a sweep of the lightcurve for obvious flare events below the transit timescale. Flares could then be excised and the data, if desired, binned to a longer cadence, correcting for the ``dead time.\" Cadences of arbitrary length are possible with data from the photon-counting detectors common in UV work, provided the data are recorded as a time-tagged event list, rather than time-integrated counts. \\subsection{Flare Statistics on AD Leo} We detected enough flares on AD Leo to examine their distribution in $\\nfe$. As such, Figure~\\ref{fig:adleoflares} plots the frequency of flares, $\\nu$, with $>\\nfe$ versus $\\nfe$. We fit a power law to these distributions of the form \\begin{equation} \\nu = \\alpha\\nfe^\\beta \\end{equation} using the maximum likelihood method of \\citet{crawford70}. The resulting values of $\\beta$ are $-0.90\\pm0.29$ for \\cii, $-0.92\\pm0.27$ for \\siii, and $-0.82\\pm0.17$ for \\siv. In comparison, previous values include \\begin{itemize} \\item{$-0.82\\pm0.27$ from 21 h of visible and near-ultraviolet observations \\citep{lacy76}} \\item{$-0.62\\pm0.09$ from 111.5 h (spread over $>5$ years) of {\\it U} band observations \\citep{pettersen84}} \\item{$-1.01\\pm0.28$ from $<72$ h of extreme-ultraviolet observations \\citep{audard00}} \\item{$-0.68\\pm0.16$ from 139.7 h of visible observations \\citep{hunt12}.} \\end{itemize} The above values all agree with those we computed for each band. This agreement is consistent with the response of each band tracing common energy deposition events, even for emission resulting from different regions of the stellar atmosphere. To explore further how impulsive energy deposition affects differing regions of a stellar atmosphere, simultaneous, panchromatic flare observations would be desirable. \\begin{figure} \\includegraphics[width=\\columnwidth]{f15.pdf} \\caption{A cumulative flare-frequency distribution for the flares identified in data for AD Leo. Power-law fits are overplotted. While the data are binned for display, following the methodology of \\citet{crawford70} they were not binned when fitting the power laws.} \\label{fig:adleoflares} \\end{figure} \\subsection{Size of Detectable Transiting Objects} This work aims to explore the boundaries placed on transit observations by stellar variability in the ultraviolet. Stochastic fluctuations in the host star determine the minimum transit depth that will stand out from these fluctuations, consequently limiting the minimum detectable size of transiting objects. As a metric for this limitation set by each star + band's stochastic fluctuations, we computed the radius of an occulting disk that would produce a transit signal equivalent to $\\s_x$ projected to 3.5~h, $R_{\\s_x}$ -- in essence, the object size needed for a 1-$\\s$ detection of a single transit in the absence of photometric noise. Explicitly, we compute $R_{\\s_x}$ from \\begin{equation} R_{\\s_x}^2 = \\s_x R_\\star^2, \\end{equation} where $R_\\star$ is the radius of the host star. This $R_{\\s_x}$ does not represent an actual detection limit. The true minimum detectable object size depends on the instrument and observing time available. Instead of a true limit, $R_{\\s_x}$ is an instrument independent means of comparing the suitability of stars for transit measurements. This metric is thus free from any assumptions about the number of photons an instrument will collect from the star or what other noise the instrument will add. Figure~\\ref{fig:disksizes} shows the results, grouped by spectral type. These suggest that, in the absence of photometric uncertainties, roughly Jovian-size disks would produce the smallest detectable transit signal (for a reasonable quantity of data) in a typical system. Indeed, HD209458b, HD189733b, and WASP-12b (see Section~\\ref{sec:intro}) are all of Jovian dimensions. However, $R_{\\s_x}$ spans around an order of magnitude within each spectral type, and more than two orders of magnitude overall. \\begin{figure} \\plotone{f16.pdf} \\caption{The size in Jupiter radii of an occulting disk that would cause a dip equivalent to the 1-$\\s$ scatter in the stellar flux at $\\dt = 3.5$ h. Filled circles represent confident measurements of $\\s_x$ whereas open circles represent upper limits. Squares represent post-main-sequence stars, circles represent main sequence stars, and diamonds represent WTTS. We group values on the x-axis by spectral type (labels below axis), then by band (symbol color).} \\label{fig:disksizes} \\end{figure} The results in Figure~\\ref{fig:disksizes} are grouped by spectral type to examine the tradeoff between the smaller $R_\\star$ but higher $\\s_x$ (see Section~\\ref{sec:varcor}) of less massive stars. Smaller stellar disks imply deeper transit signals, but higher levels of stochastic fluctuations better hide these signals. From Figure~\\ref{fig:disksizes}, it appears stellar size trumps $\\s_x$: the smallest stars also permit the smallest objects to produce detectable transits. However, this apparent trend is significantly weakened when the F stars, all post main-sequence giants or sub-giants, are removed. Of the stars with data robustly sampling their stochastic fluctuations on transit timescales (mean flux over $5\\sn{-16}$ erg s$^{-1}$ cm$^{-2}$ and total accumulated observations $\\geq3.5$ h), Prox Cen has the most generous $R_{\\s_x}$ limits in each band. These are $1.8\\pm0.1~ R_\\earth$ in \\cii, $<1.5~R_\\earth$ in \\siii, $2.1\\pm0.1~R_\\earth$ in \\siv\\, and $<1.1~R_\\earth$ in the FUV continuum. Interestingly, Prox Cen is classified as a flare star, and the chances that it could flare near or within a transit are significant. Like Prox Cen, most of the stars in this sample have received flare, variable, or WTTS classifications. The sample is biased: The targets were preferentially selected for variability by the various individuals that commissioned the observations. It seems probable that a less-biased sample would produce detection limits clustered lower than those in Figure~\\ref{fig:disksizes}. Therefore, the prospects for FUV transit measurements of planets the size of Neptune or super-Earths seem promising. Furthermore, if the observed planet hosts an atmosphere inflated to several times the area of the solid disk, like HD209458b \\citep{madjar03}, then in an ideal case even an Earth-size planet might produce a detectable FUV signal within a few folded transits. These prospects are exciting, but it bears remembering that, to attain such limits, photometric noise must be suppressed to below the level of stochastic fluctuations. \\subsection{Correlations with Stellar Properties} \\label{sec:varcor} We explored correlations between the mean-normalized excess noise measurements in \\cii\\ and \\siv\\ with properties of the sample stars (see Table~\\ref{tbl:props}). Excess noise values in the remaining bands, \\siii\\ and the FUV continuum, were too poorly constrained (essentially, there were too few detections and too many upper limits) to support correlations with any value. As a means of visualizing potential correlations, Figures~\\ref{fig:varcorCII} and~\\ref{fig:varcorSiIV} graph $\\s_x$ values (both detections, black, and upper limits, blue) against each of the stellar properties, excluding mass and luminosity, for the \\cii\\ and \\siv\\ bands respectively. These figures show that the error bars of the mean-normalized excess noise values and stellar properties often overlap. Thus, changes in the point values even under the 1-$\\s$ error bars could change the Spearman Rank-Order correlation coefficient for the data and, more importantly, confidence that the correlation cannot be explained by randomly scattered points. To account for the uncertainty in how the true values of the points fall, we constructed a Monte-Carlo simulation, generating $10^4$ possible arrangements of the data points given the uncertainties. For each trial in the Monte-Carlo simulation, we randomly drew stellar parameter (e.g., age, $M$, $P_{rot}$) values from Gaussians matching the means and 1-$\\s$ uncertainties. When there was no uncertainty accompanying the measurement we found in the literature, we assigned an uncertainty of 10\\%. For the mean-normalized excess noise values, we did not use Gaussians. Instead we randomly drew values from the actual likelihood distributions of $\\s_x$ for each star/band (Appendix~\\ref{sec:likelihood}). For each of the $10^4$ such realizations of the data, we recorded probability to exceed (PTE) for the null hypothesis of randomly distributed points (i.e. no correlation) using the Spearman Rank-Order test. We multiplied the PTE by $10^{-4}$ to compute the joint probability of the data producing the correlation coefficient \\emph{and} such a correlation coefficient resulting from random point scatter. The integral of these values over the range of possible correlation coefficients (-1 to 1), provided our overall estimate of the PTE for random scatter in light of the data and uncertainties. We then recorded the significance of the correlation as $1-$PTE. We carried out this process with and without the $\\s_x$ upper limits and, for both cases, quote the significance of the correlation in Figures~\\ref{fig:varcorCII} and~\\ref{fig:varcorSiIV}. The $\\s_x$ upper limits provide useful constraints on correlations when they fall below the surrounding $\\s_x$ detections. As with the $\\s_x$ detections, when we included upper limits we generated $\\s_x$ values in the Monte-Carlo simulation from the likelihood distribution that produced each upper limit. For instance, an upper limit of 0.01 on $\\s_x$ for a point meant that we randomly drew a value between 0 and 0.01 with roughly uniform probability for each trial in the Monte-Carlo simulation. The results computed without including the upper limits suggest weak correlations in all cases except $\\s_x$-age and, in \\cii\\, $\\s_x$-$P_{rot}$. The $\\s_x$ upper limits further constrain these correlations, both quantitatively and by eye, and bring \\cii\\ and \\siv\\ into closer agreement. The subsections below address each of these in light of previous literature. \\begin{figure*} \\plotone{f17.pdf} \\caption{Excess noise versus stellar properties in the \\cii\\ band. Symbols differentiate between spectral types (see legend in bottom right plot). Black symbols represent excess noise detections, while blue symbols represent upper limits. The black numbers above each plot give the significance of the correlation (probability it is not produced by uncorrelated points) using only excess noise detections while the blue number gives the significance including upper limits (see Section~\\ref{sec:varcor}).} \\label{fig:varcorCII} \\end{figure*} \\begin{figure*} \\plotone{f18.pdf} \\caption{Excess noise versus stellar properties in the \\ion{Si}{4} band, following the same format as Figure~\\ref{fig:varcorCII}.} \\label{fig:varcorSiIV} \\end{figure*} \\subsubsection{Temperature, Radius, Mass, and Luminosity} \\label{sec:sptcorr} Possible correlations of excess noise with temperature and radius (as well as mass and luminosity, though these are left out of Figures~\\ref{fig:varcorCII} and~\\ref{fig:varcorSiIV}) become strong when upper limits are included. This indicates a more general correlation between excess noise and spectral type: Stars of later spectral type typically exhibit higher $\\s_x$. Correlations between stellar variability and spectral type were explored by \\citet{mcquillan12} for all {\\it Kepler} stars save those with known eclipsing companions (stellar or planetary) and lightcurve discontinuities. They found that cooler, later type stars exhibited high levels of stochastic variability by their metric, in agreement with our findings. Similarly, \\citet{ciardi11} found a relationship between variability and temperature in their analysis of {\\it Kepler} field stars, in both dwarfs and giants. We note, however, that {\\it Kepler} utilizes a broad optical bandpass ($\\sim$4000-9000 \\AA). Therefore, variability in the {\\it Kepler} data reflects processes occurring in the photosphere, rather than the transition region, and there is no guarantee that the two are directly related. \\subsubsection{Rotation Period, Equatorial Velocity, and Age} Excess noise does not correlate with age, and the hint of a correlation between $\\s_x$ and rotation period is all but eliminated with the inclusion of $\\s_x$ upper limits. A possible anticorrelation seems to exist between $\\s_x$ and \\vsini; however, when $\\s_x$ upper limits are included, the anticorrelation is probably weak enough to dismiss outright. Furthermore, because $v = 2\\pi R/P_{rot}$, the combination of a lack of a $\\s_x$-$P_{rot}$ correlation and the presence of a strong $\\s_x$-$R$ anticorrelation are capable of producing the $\\s_x$-\\vsini\\ trend. Thus, it is the $\\s_x$-$R$ anticorrelation (or, rather, the relationship of $\\s_x$ to spectral type) that drives the $\\s_x$-\\vsini\\ trend. Indeed, $\\s_x$ and the quotient $R/P_{rot}$ (excluding the 8 stars for which we computed $P_{rot}/\\sin{i}$ from \\vsini\\ and $R$) are anticorrelated to roughly the same confidences as $\\s_x$-\\vsini. Interpreting the lack of correlations with age and rotation in light of previous work is difficult, as we did not find any studies exploring correlations between these properties and stellar stochastic fluctuations. Both age and rotation, however, have been tied to stellar activity as quantified by chromospheric emission (e.g., \\citealt{wilson63,skumanich72,noyes84,mamajek08}). In turn, chromospheric emission likely has a direct relationship with stochastic fluctuations at visible wavelengths \\citep{hall09}. This suggests that younger, faster rotating stars might exhibit higher levels of UV emission line variability. Given these past results, the lack of $\\s_x$-age and $\\s_x$-$P_{rot}$ correlations in this sample could be explained if either the age-rotation-activity or the variability-activity relationships do not hold for this sample. That either might not hold would not be particularly alarming given that this sample includes stars approaching, dwelling on, and departing the main sequence, whereas the above-mentioned studies only analyze stars on the main sequence. The pre and post main-sequence stars in this sample might exhibit magnetic behavior not in line with that of main-sequence stars, such that relationships between magnetic activity (chromospheric emission) and rotation, age, and/or variability for main-sequence stars are not extensible to this broader sample. \\subsubsection{Surface Gravity} A strong correlation exists between excess noise and surface gravity in the sample, particularly in \\cii. This is in conflict with the recent results of \\citet{bastien13}, who find an inverse relationship between variability and surface gravity in {\\it Kepler} field stars. However, compared to this work, the stellar sample employed by \\citet{bastien13} is far more restrictive. They limit their study to stars with $4500$ K $< T_{eff}<6500$ K, $2.5< \\log_{10}{g} <4.5$ (cgs units), and relative brightness variations (measured using a 30 min cadence) of $< 0.003$. Applying the same cuts in $T_{eff}$ and $\\log_{10}{g}$ to this stellar sample eliminates any correlation (null hypothesis can produce the observed or stronger correlation with $\\sim$50\\% probability). The strong correlation present in this data when all stars are included is driven primarily by the F giants, as these have both very low $g$ and $\\s_x$. However, high $\\s_x$, high $g$ M stars also promote the correlation. \\subsection{Star-Planet Interactions} Close-in planets might interact with their host star magnetically and tidally, producing enhanced activity in the stellar upper atmosphere \\citep{cuntz00}. Evidence for these interactions could be found by monitoring a single star with a transiting planet to search for signs of elevated activity near transit, when the portion of the star interacting with the planet is in view. Indeed, this has been attempted for two systems with transiting planets in this sample, HD209458 \\citep{shkolnik04} and HD189733 \\citep{shkolnik08,fares10}. However, these studies found no clear evidence of enhanced activity correlated with orbital phase for either of these systems. Alternatively, evidence for star-planet interactions might result from comparing the overall activity or variability of the stars with close-in planets to a sample of control stars. \\citet{shkolnik13} recommended investigations into star-planet interactions in the form of time-resolved observations of UV flux variability rather than activity (i.e. mean line flux), such as those presented in this paper. The small size and high diversity of the 32 stars in this sample that do not host planets make for a poor control. Nevertheless, we looked for increased $\\s_x$ in the six known planet hosts. These hosts are identified in the $\\s_x$ detections and upper limits presented in Figures~\\ref{fig:longvar} and~\\ref{fig:longvar} as open points. The hosts with $\\s_x$ detections in Figure~\\ref{fig:longvar} do appear clustered at higher $\\s_x$. However, these are all M dwarfs, shown in the previous section to have the highest $\\s_x$ values in the sample. The M dwarf planet hosts do not stand out when grouped with the other M dwarfs in the sample. The G and K planet hosts (HD209458 and HD189733) have insufficient data for a $\\s_x$ detection, allowing only 95\\% upper limits. Because only upper limits are possible, these stars are not constrained to be more variable than other sample stars of the same spectral type. Thus, star-planet interactions, manifested as increased variability, are not supported by the limited volume of this data. We have examined stellar luminosity fluctuations and stellar flares in a sample of 38 cool stars using 60~s cadence lightcurves constructed from narrow spectroscopic bands containing the \\ion{Si}{3} $\\lm$1206, \\ion{C}{2} $\\lm\\lm$1334,1335, and \\ion{Si}{4} $\\lm\\lm$1393,1402 resonance lines and a combined 36.5 \\AA\\ of interspersed FUV continuum bands. In the high pass filtered lightcurves, we detected 116 flares, occurring roughly once per 2.5 h. Flares commonly radiated more energy in \\siv, relative to quiescent levels, than the other bands. Shorter flares are more prevalent, with over half lasting 4 min or less. Most (90 of 116) of the detected flares could annihilate the signal of an Earth transit, while 7 of the 116 could annihilate the signal of a Jupiter transit. These results highlight the usefulness of minute-scale cadences for finding and removing flares prior to estimating transit depths. To quantify stochastic fluctuations, we found the maximum-likelihood ``excess noise\" that, in addition to the photometric noise, accounts for the scatter of the high-pass filtered, mean-normalized lightcurves excluding flares. Values of the excess noise, relative to the mean flux, range from about 1-41\\% in \\cii, 8-18\\% in \\siii, 0.9-26\\% in \\siv, and 1-22\\% in the FUV continuum. Where the likelihood distribution of the excess noise was one-sided, we instead quote a 95\\% upper limit on the excess noise. These upper limits on excess noise are often strong enough to be lower than the values for stars where excess noise was detected. We found significant anticorrelations of excess noise with mass, radius, temperature, and luminosity in \\cii\\ and \\siv. An additional, weaker anticorrelation of excess noise with \\vsini\\ could be an artifact of the strong underlying correlation between radius and rotation rate in the sample. There was no correlation with age or rotation period. The median levels of stochastic stellar fluctuations we estimated, integrated over a typical transit timescale of 3.5~h, would impose a rough 1-$\\s$ transit detection limit of $\\sim1~R_J$ occulting disks for many of the stars in the sample. However, the range in these limits is broad, spanning from tenths of $R_J$ to tens of $R_J$. M dwarfs might permit the FUV observation of transiting objects as small as Neptunes or even super-Earths in the absence of photometric noise. While the large fluctuations of some stars might stymie transit spectroscopy in the FUV for any but the largest planets, these results suggest that many planetary host stars might be found for which the limits of FUV transit spectroscopy can be pushed well below the hot Jupiters observed thus far. {\\it Acknowledgments:} The authors wish to acknowledge the anonymous referee, whose familiarity with the field and resulting thoughtful and constructive comments greatly improved this work. The authors also wish to thank Tom Ayres for his careful reading and helpful suggestions that similarly improved this work. Kevin France acknowledges support through a NASA Nancy Grace Roman Fellowship during a portion of this work. This research has made use of the Exoplanet Orbit Database and the Exoplanet Data Explorer at exoplanets.org \\citep{wright11} as well as extensive use of the SIMBAD database and VizieR catalog access tool, operated at CDS, Strasbourg, France (\\citet{wenger00}, \\citet{ochsenbein00}). The authors wish in particular to acknowledge \\citet{glebocki05} for the excellent stellar rotation catalog that we used in this work. Some of the data presented in this paper were obtained from the Mikulski Archive for Space Telescopes (MAST). STScI is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. This work was supported by NASA grant NNX08AC146 to the University of Colorado at Boulder. {\\it Facility:} \\facility{{\\it HST} (COS, STIS)} \\appendix" }, "1402/1402.0245_arXiv.txt": { "abstract": "It is suggested that many $\\gamma$-ray bursts (GRBs) are cloaked by an ultra-relativistic baryonic shell that has high optical depth when the photons are manufactured. Such a shell would not fully block photons reflected or emitted from its inner surface, because the radial velocity of the photons can be less than that of the shell. This avoids the standard problem associated with GRBs that the thermal component should be produced where the flow is still obscured by high optical depth. The radiation that escapes high optical depth obeys the Amati relation. Observational implications may include a) anomalously high ratios of afterglow to prompt emission, such as may have been the case in the recently discovered PTF 11agg, and b) ultrahigh-energy neutrino pulses that are non-coincident with detectable GRB. It is suggested that GRB 090510, a short, very hard GRB with very little afterglow, was an {\\it exposed} GRB, in contrast to those cloaked by baryonic shells. ", "introduction": "A longstanding puzzle concerning GRBs % is that their highly super-Eddington luminosities would seemingly imply that the very photons that comprise the GRB should drag out enough baryonic material to obscure themselves. The problem is compounded by the discovery that long GRB occur inside massive stars, meaning that the GRB fireball must plow up material from its host star as it pushes its way out, not all of which can move out of its way. Neutrons drifting into the fireball ( Eichler \\& Levinson, 1999, Levinson \\& Eichler, 2003), and highly opaque pre-collapse winds from the host stars (Ofek et al., 2013) may also cloak GRBs with an optically thick layer out to $10^{3.5}$ lightseconds. Naively, this would seem to be enough to completely hide the GRB, which typically lasts less than $10^2$ s. One suggestion to solve the first problem is that the energy powering the GRB originates on magnetic field lines that thread a black hole that forms at the center of the collapsing host star (e.g. Levinson and Eichler, 1993), which should be nearly devoid of baryonic material, but this does not address the issue of the plowed-up host star material, which cloaks the fireball even on horizon-threading field lines. One solution to the cloaking problem is that many GRBs are seen by observers just off the (angular) edge of the jetted fireball (Eichler \\& Levinson, 2004, Levinson \\& Eichler 2006), and that this accounts for the Amati and Ghirlanda relations (Amati, 2002, Ghirlanda et al 2004 ). The hypothesis successfully predicted an intermediate flat phase for GRB afterglows (Eichler, 2005, Eichler \\& Jontof-Hutter, 2005) that was later confirmed by Swift observations. The \"off-edge observer\" hypothesis also solves the problem\\footnote{See Vurm, Lyubarsky \\& Piran (2013) for a detailed, carefully quantified discussion of this point. They do not, however, consider scattering within the host star envelope off collimating walls, which could lower the photon energies in the observer frame via Compton recoil and pair production.} that because blue-shifted $\\gamma$-ray photospheres should have spectral peaks that are or order $3 \\Gamma kT_{ann}\\gtrsim 1$ MeV, where $kT_{ann}$, the temperature at which pairs annihilate and permit optical transparency, is of order 10 KeV, and $\\Gamma$, the bulk Lorentz factor of the flow, is $\\gtrsim 10^2$, the spectral peak should then be $\\sim 3$ MeV, whereas most are observed to have spectral peaks at $\\lesssim 300$ KeV. Taking into account the possibility that pair production opacity limits the photon energy to $\\sim 1$ MeV, there is still some discrepancy with the fact that the distribution of $E_{peak}$ for {\\it long} GRB (in contrast to short, hard GRB) peaks comfortably below the pair production threshold. Because the hypothesis predicts that the observed spectral peak is softened relative to that seen by a head-on observer ($\\theta=0$) by a factor of $S(\\theta)\\equiv E_{peak}(0)/E_{peak}(\\theta) ={(1-\\beta \\cos \\theta)\\over(1-\\beta)}=(1 + {\\beta(1-\\cos \\theta)\\over(1-\\beta)})\\sim 1+(\\theta\\Gamma)^2$, where $\\theta$ is the viewing offset angle and $\\Gamma$ is the bulk Lorentz factor, the problem disappears if $ \\theta \\Gamma \\sim 3$. The \"off-edge\" viewing hypothesis, however, does not yet explain why {\\it most} GRBs are viewed from such an offset angle, especially considering that the opening angle for GRB, $\\sim 0.1$, is probably larger than $1/\\Gamma$. (Indeed, about half of all classical GRB, although they show spectral peaks of $\\lesssim 300$ KeV, do {\\it not } display a noticeable flat phase in their afterglow.) Neither would it solve the cloaking problem if the cloak has a larger solid angle than the jet, as in the case of a thick stellar wind. ", "conclusions": "" }, "1402/1402.5656_arXiv.txt": { "abstract": "We study the oscillations of relativistic stars, incorporating key physics associated with internal composition, thermal gradients and crust elasticity. Our aim is to develop a formalism which is able to account for the state-of-the-art understanding of the complex physics associated with these systems. As a first step, we build models using a modern equation of state including composition gradients and density discontinuities associated with internal phase-transitions (like the crust-core transition and the point where muons first appear in the core). In order to understand the nature of the oscillation spectrum, we carry out cooling simulations to provide realistic snapshots of the temperature distribution in the interior as the star evolves through adolescence. The associated thermal pressure is incorporated in the perturbation analysis, and we discuss the presence of $g$-modes arising as a result of thermal effects. We also consider interface modes due to phase-transitions and the gradual formation of the star's crust and the emergence of a set of shear modes. ", "introduction": "Neutron stars are seismically complex, with particular classes of oscillation modes associated with specific parts of the involved physics. This makes the construction of a truly realistic model of the star's oscillation spectrum a complicated task. Nevertheless, we are reaching the point where much of the physics involved is understood at the level required to build moderately realistic models. The aim of this work is thus quite natural; we want to take important steps towards realism by accounting for processes that are relevant as a neutron star matures. We achieve this by tracking the cooling of a given star from a minute or so after birth through the first several hundred years, paying particular attention to the changes in the thermal pressure and the formation of that star's elastic crust. The output from our state-of-the-art cooling code provides input for the seismology analysis. The results provide a sequence of snapshots of how the star's oscillation spectrum evolves as the star ages. This is an important advance on previous work in this area. The problem of relativistic star seismology has been considered since pioneering work by Thorne \\& Campolattaro in the late 1960s \\cite{thorne_campolattaro_1967}. Like much of the subsequent work, the initial focus was on the mathematical formulation of the problem with the complex supranuclear physics playing a secondary role. Thus, the bulk of the literature is focussed on perfect fluid stars, often without any consideration of the interior composition and state of matter. In fact, many studies have made use of rather ad hoc polytropic models which only capture the rough properties of a realistic equation of state. Nevertheless, this has led to an understanding of the basic nature of the stellar spectrum, like the fundamental $f$-mode and the pressure restored $p$-modes \\cite{lindblom_detweiler_1983, detweiler_lindblom_1985}, and some discoveries, like the existence of the gravitational-wave $w$-modes, \\cite{kokkotas_schutz_1992}. However, the understanding of the fine details of the problem has so far been developed one piece at a time. Gravity $g$-modes arising because of composition gradients have been considered \\cite{finn_1986, finn_1987, strohmayer_1993, miniutti_etal_2003}, the crust elasticity has been accounted for (especially for shear modes) \\cite{schumaker_thorne_1983}, the roles of core superfluidity \\cite{comer_etal_1999, andersson_etal_2002, lin_etal_2008} and the star's magnetic field \\cite{sotani_etal_2008, colaiuda_etal_2009, gabler_etal_2012, gabler_etal_2013} have been investigated and oscillations of proto-neutron stars and their detectability~\\cite{ferrari_etal_2003} have been studied, as well. In the last decade, much attention has been focussed on the Coriolis restored $r$-modes~\\cite{andersson_1998, andersson_kokkotas_2001, lockitch_etal_2001, haskell_etal_2009}, which are relevant because they may be driven unstable due to the gravitational radiation they generate. In this context, much of the complex microphysics related to dissipative processes has been discussed, progressing our understanding of the involved processes considerably. Neutron star oscillations may impact on a range of observations, involving in particular radio and X-ray timing and gravitational waves. At the present time, the most promising connection between observations and theory is provided by the observed quasiperiodic oscillations seen in the X-ray tail of magnetar giant flares \\cite{watts_strohmayer_2006, samuelsson_andersson_2007}. The inferred oscillation frequencies match those expected for the elastic crust reasonably well, and provide the first credible example of actual neutron star asteroseismology. There are, of course, issues to be resolved, especially concerning the dynamical role of the interior magnetic field in these systems \\cite{colaiuda_kokkotas_2011}. A more indirect example comes from X-ray timing of fast spinning accreting neutron stars in Low-Mass X-ray Binaries \\cite{strohmayer_mahmoodifar_2014, andersson_etal_2014}. The spin of these stars appears to be limited by some process. One of the leading contenders for the underlying mechanism is the $r$-mode instability which leads to the star losing angular momentum at a rate that could balance the spin-up due to accretion \\cite{andersson_etal_1999, bondarescu_etal_2007}. Finally, as we are getting closer to the advent of gravitational-wave astronomy, there has been searches for neutron star oscillations from a range of systems \\cite{ligo_search1, ligo_search2, ligo_search3, ligo_search4}. At present, these studies may only be providing relatively uninteresting upper limits on the possible signals. With an advanced generation of detectors coming online in the next few years, this could turn to actual detections. The existing body of work allows us to piece together a picture of neutron star dynamics. This understanding should be valid provided the star is in a regime where the different ingredients remain distinct. Unfortunately, this is unlikely to be the case. Hence, there is a pressing need to develop a new generation of models that account for as much of the relevant physics as possible. The present work should be seen in that context. The physics associated with realistic neutron star dynamics is daunting. Many aspects are rather poorly understood, in particular concerning the deep core (at several times the nuclear saturation density). Nevertheless, a focussed effort on this problem is timely. First of all, gravitational-wave astronomy should become reality in the next few years. This promises to provide us with observational data (perhaps most likely from the merger of compact binaries) that need to be matched against the best possible theoretical models. Secondly, our understanding of the principles associated with neutron star superfluidity has improved considerably in the last decade. In particular, we have constraints on the superfluid parameters (the superfluid pairing gaps or, equivalently, the critical temperatures) from the observed real-time cooling of the remnant in Cassiopeia A \\cite{shternin_etal_2011}. In order to be able to make productive use of future observational results, we need to improve our level of modelling. The basic requirement for progress on this problem is a realistic equation of state which accounts for the two-fluid nature of a neutron star's core, including information about the superfluid gap energies and the entrainment parameters. We also need a comprehensive formalism for modelling dynamics which accounts for all the desired parameters. Progress on the first of these issues was recently made by Chamel \\cite{chamel_2008}, who provided the first ever consistent equation of state including entrainment. The second problem has been explored in basic models \\cite{andersson_etal_2013}, to the point where there are no technical stumbling blocks preventing more realistic studies. In this work we take the first steps towards true realism, starting from the classic fluid formalism of Detweiler \\& Lindblom \\cite{detweiler_lindblom_1985} and extending it to account for density discontinuities associated with distinct phase-transitions, interior composition gradients, thermal pressure, and the elastic crust which will form and grow in thickness as the neutron star cools. We use this model to study the evolution of the star's oscillation modes as it matures. To do this we couple the oscillation mode calculation to the long-term cooling of the star. At different points in the thermal evolution we output the temperature profile and feed it into the mode calculation. The results of this exercise shed light on the influence of thermal effects on the various oscillation modes of the star. As expected, we find that the fundamental $f$-mode and the various $p$-modes are only weakly affected by the changes in temperature. Meanwhile, the gravity $g$-mode spectrum changes completely as the thermal pressure weakens and we demonstrate how only a few interface modes (arising from density ``discontinuities\") remain (in the considered frequency range above 10~Hz or so) when the star reaches maturity. Finally, the presence of the elastic crust enriches the spectrum by shear modes, which also evolve as the star cools and the crust region grows. This paper is the first in a series where we aim to build a truly realistic model of a dynamical neutron star. Our initial focus is on issues relating to composition and thermal effects. This will lead naturally on to the role of superfluidity, which will be considered in a follow-up paper. The basic reason for the division is that the problems we consider here can be modelled within the standard single-fluid framework, while superfluidity adds degrees of freedom that make the problem richer (and obviously more complicated). The same is true for the star's magnetic field, which would eventually have to be included in the model (involving issues associated with the presence of a superconducting core \\cite{glampedakis_etal_2011}). The layout of the paper is as follows: Section \\ref{sec:ns_model} is devoted to the physics which we take into account for the background model; in Section \\ref{sec:perturbations}, we discuss the perturbation equations and explain how these are modified to account for additional physics (the perturbation equations for the elastic crust are given in Appendix \\ref{sec:app_eq_crust_polar}). A general numerical strategy for solving the perturbation equations in the interior of a multi-layered star is presented in Section \\ref{sec:num_strategy} and Appendix \\ref{sec:app_numerics}; Section \\ref{sec:num_strategy} also covers the reasoning for a new set of equations for perturbations of the perfect fluid, which aids in solving the perturbation problem at low frequencies; the actual equations can be found in Appendix~\\ref{sec:app_eq_low_freq}. Finally, Section \\ref{sec:results} contains the results and Section \\ref{sec:summary} summarises our work. Unless stated otherwise, we use units in which $G=c=1$ and Misner, Thorne and Wheeler (MTW) \\cite{MTW1973} conventions throughout the paper. ", "conclusions": "\\label{sec:summary} The aim of this paper was to establish a formalism that allows us to study neutron star oscillations in full general relativity, accounting for as much realistic microphysics as possible. So far we extended the standard perfect fluid formalism to account for thermal pressure in the star's core and derived a new set of equations which govern perturbations in the solid crust. We provided sample results for a particular realistic equation of state. However, our numerical code is generic in the sense that we can easily use a different equation of state straightaway. We focus our attention on the low-frequency part of the oscillation spectrum (above 15~Hz or so due to computational limitations). In this regime we find a set of interface modes, firstly due to the (artificial) density discontinuity at the crust-core interface and secondly there are interface modes associated with the crust region, due to sharp changes in the low-density equation of state. The composition gradient shifts the frequency of the artificial interface mode to slightly higher frequencies, and we find a set of composition $g$-modes in the low frequency regime. The frequencies of these $g$-modes are slightly lower than literature values from pervious studies---this is because the composition gradient is rather small throughout nearly the entire core; only close to the crust-core transition is the composition gradient reasonably pronounced. When we account for thermal pressure due to neutrons and protons in the core, we find that a number of thermal $g$-modes enter the low frequency part of the spectrum. Meanwhile, the high frequency interface modes are shifted to even higher frequencies whereas the composition $g$-modes stay unaffected. Tracking the thermal evolution of the neutron star for the first few hundred years, we investigate how the frequencies of the various modes evolve as the star matures through adolescence. After approximately 10 years, all thermal $g$-modes have dropped below $15\\unit{Hz}$ and the thermal effect on the interface modes has almost vanished entirely. During the next 100 years, as the star continues to cool, the frequencies of the interface modes change slightly, finally leaving us with the spectrum of a cold star. Finally, we considered the crystallization of the crust and investigated the associated changes in the spectrum. From our thermal evolution, we determined the solid region in the outer parts of the star. We quantified how the crust elasticity affects the interface modes associated with the outer layers of the star and we discovered that they are suppressed by shear stress. We discussed how the spectrum is enriched by shear modes. As expected, our results for the shear modes are in good agreement with results from previous work \\cite{mcdermott_etal_1988}. In summary, we have taken the first step towards a comprehensive computational technology to study quasinormal modes of realistic compact stars. We plan to extend the model to account for a multifluid core where the neutrons are superfluid and the protons are superconducting. In doing so we will consider recent equation of state data, in particular concerning the entrainment effect, both in the core and the crust. We will also update the cooling sequence to account for superfluid effects. We expect to report on results in these directions in the not too distant future." }, "1402/1402.2289_arXiv.txt": { "abstract": "Direct $N$-body simulations of globular clusters in a realistic Milky Way-like potential are carried out using the code \\texttt{NBODY6} to determine the impact of the host galaxy disk mass and geometry on the survival of star clusters. A relationship between disk mass and star cluster dissolution timescale is derived. These $N$-body models show that doubling the mass of the disk from $5\\times 10^{10}$~M$_{\\odot}$ to $10\\times10^{10}$~M$_{\\odot}$ halves the dissolution time of a satellite star cluster orbiting the host galaxy at 6 kpc from the galactic center. Different geometries in a disk of identical mass can determine either the survival or dissolution of a star cluster orbiting within the inner 6 kpc of the galactic center. Furthermore, disk geometry has measurable effects on the mass loss of star clusters up to 15 kpc from the galactic center. $N$-body simulations performed with a fine output time step show that at each disk crossing the outer layers of star clusters experience an increase in velocity dispersion of $\\sim$5\\% of the average velocity dispersion in the outer section of star clusters. This leads to an enhancement of mass-loss -- a clearly discernable effect of disk shocking. By running models with different inclinations we determine that star clusters with an orbit perpendicular to the Galactic plane have larger mass loss rates than both clusters evolving in the Galactic plane or in an inclined orbit. ", "introduction": "In the current paradigm of galaxy formation smaller structures merge into larger ones from the Big Bang up to the present day (White \\& Rees 1978). Galaxies grow through two main processes: the hierarchical merging with smaller galaxies and the accretion of fresh gas fuelling new star formation. These different mechanisms contribute to the growth of the disk, bulge and halo. As the first significant stellar structures to form, globular clusters witness the entire evolution of their host galaxy as satellite systems. Indeed, globular clusters are believed to follow galaxies during galaxy mergers and close encounters (e.g.\\ West et al.\\ 2004). The gravitational potential of their host galaxy has a direct influence on the survival of globular clusters: they lose stars through tidal stripping and disk shocking. In turn, these lost stars contribute to the build-up of the galaxy's stellar halo. The evolution of host galaxy and globular clusters are clearly connected. The aim of this work is to derive, using numerical simulations, the importance of the host galaxy disk mass and size on the evolution and survival of star clusters. How are satellite stellar systems affected by disks and bulges of changing mass and with different geometries? And reciprocally, how do satellite stellar systems contribute, through their dissolution, to the formation of the halo of the host galaxy? D'Onghia et al.\\ (2010) show that halo and disk shocking efficiently deplete the satellite population of dark matter halos within 30 kpc of the Milky Way center. The results of that study cannot be directly applied to globular clusters because of their much higher densities compared to dark matter halos. In their important paper, Gnedin \\& Ostriker (1997) model the dynamical evolution of the Galactic globular cluster system using a Fokker-Planck code. These authors build on earlier analytical work by Aguilar, Hut \\& Ostriker (1988) and Kundic \\& Ostriker (1995) among others. Gnedin \\& Ostriker (1997) give analytical expressions to estimate the impact of disk shocking and also call for numerical simulations to be carried out. With the recent progress in computational capacity, large $N$-body simulations can now be carried out in order to determine the physical mechanisms that govern the dynamical evolution of globular clusters in a galactic potential. Recently, Renaud \\& Gieles (2013) found that after a merger event of the host galaxy the mass loss rate of a satellite star cluster increases. Interestingly, Renaud \\& Gieles (2013) also find that even if the tidal forces reach a maximum during the merger itself they are too short-lived to have a significant impact on the long term survival of star clusters. Previous work on the mass loss of star clusters focused on internal dynamical effects and stellar evolution. Vesperini \\& Heggie (1997) carried out numerical simulations of star clusters and determined their mass loss rates through a Hubble time. Baumgardt \\& Makino (2003) established an analytical dissolution time scale for star clusters and showed that one third of the cluster mass is lost due to stellar evolution alone. A recent review by of recent $N$-body studies can be found in Portegies Zwart et al.\\ (2010). The approach in our work is purely numerical and different in nature to Gnedin \\& Ostriker (1997) given that no assumptions or explicit expressions to treat the impact of the disk are used. $N$-body models of star clusters were run where the properties of the star cluster remain identical but the mass and geometry of the galactic disk change. The effect of the disk on the star cluster is computed as a part of the numerical calculations of the gravitational force experienced by each star. The above is carried out with the code \\texttt{NBODY6} used for the study of the dynamics of star clusters through $N$-body simulations. \\texttt{NBODY6} now includes a detailed model of the host galaxy where star clusters evolve as satellites (Aarseth 2003). The current set-up of \\texttt{NBODY6} includes the tools to model a Milky-Way type galaxy with three distinct components: disk, bulge, and halo. The gravitational force for each star of the cluster is computed at each time step by taking into account the effect of all other stars, and of the disk, bulge, and halo. We make use of this new capacity of \\texttt{NBODY6} to run several models of star clusters where the mass and physical size of the disk are different between models. Throughout this work, a Hubble time of 13.5 Gyr is adopted, in agreement with the results of the Wilkinson Microwave Anisotropy Probe (Spergel et al.\\ 2003). ", "conclusions": "" }, "1402/1402.5526_arXiv.txt": { "abstract": "{We present a study of the 3-dimensional environment for a sample of 386 galaxies in the \\textbf{C}atalogue of \\textbf{I}solated \\textbf{G}alaxies (CIG, Karachentseva 1973) using the Ninth Data Release of the Sloan Digital Sky Survey (SDSS-DR9).} {We aim to identify and quantify the effects of the satellite distribution around a sample of galaxies in the CIG, as well as the effects of the Large Scale Structure (LSS).} {To recover the physically bound galaxies we first focus on the satellites which are within the escape speed of each CIG galaxy. We also propose a more conservative method using the stacked Gaussian distribution of the velocity difference of the neighbours. The tidal strengths affecting the primary galaxy are estimated to quantify the effects of the local and LSS environments. We also define the projected number density parameter at the 5$^{\\rm th}$ nearest neighbour to characterise the LSS around the CIG galaxies.} {Out of the 386 CIG galaxies considered in this study, at least 340 (88\\% of the sample) have no physically linked satellite. Following the more conservative Gaussian distribution of physical satellites around the CIG galaxies leads to upper limits. Out of the 386 CIG galaxies, 327 (85\\% of the sample) have no physical companion within a projected distance of 0.3\\,Mpc. The CIG galaxies are distributed following the LSS of the local Universe, although presenting a large heterogeneity in their degree of connection with it. When present around a CIG galaxy, the effect of physically bound galaxies largely dominates (usually by more than 90\\%) the tidal strengths generated by the LSS.} {The CIG samples a variety of environments, from galaxies with physical satellites to galaxies with no neighbours within 3\\,Mpc. A clear segregation appears between early-type CIG galaxies with companions and isolated late-type CIG galaxies. Isolated galaxies are in general bluer, with likely younger stellar populations and rather high star formation with respect to older, redder CIG galaxies with companions. Reciprocally, the satellites are redder and with an older stellar populations around massive early-type CIG galaxies, while they have a younger stellar content around massive late-type CIG galaxies. This suggests that the CIG is composed of a heterogeneous population of galaxies, sampling from old to more recent, dynamical systems of galaxies. CIG galaxies with companions might have a mild tendency (0.3-0.4 dex) to be more massive, and may indicate a higher frequency of having suffered a merger in the past.} ", "introduction": "\\label{Sec:intro} Isolated galaxies are located, by definition, in low-density regions of the Universe, and should not be significantly influenced by their neighbours. Does a separate population of isolated galaxies exist, or are isolated galaxies simply the least clustered galaxies of the Large Scale Structure (LSS)? It is assumed that over the past several billion years the evolution of these objects has largely been driven by internal processes. A significant population of isolated galaxies is of great interest for testing different scenarios of the origin and evolution of galaxies. In this sense, isolated galaxies are an ideal sample of reference for studying the effects of environment on different galaxy properties. Such a sample would represent the most nurture-free galaxy population. Studies of isolated galaxies can be argued to begin with the publication of the Catalogue of Isolated Galaxies \\citep[CIG;][]{1973AISAO...8....3K}. The AMIGA (\\textbf{A}nalysis of the interstellar \\textbf{M}edium of \\textbf{I}solated \\textbf{GA}laxies\\footnote{\\texttt{http://amiga.iaa.es}}) project \\citep{2005A&A...436..443V} is based upon a re-evaluation of the CIG. It is a first step in trying to identify and better characterise isolated galaxies in the local Universe. \\citet{2005A&A...436..443V} argued that 50\\% or more galaxies in the CIG show a homogeneous redshift distribution. \\citet{2006A&A...449..937S}, and more recently \\citet{2012A&A...540A..47F}, found that 2/3 of the CIG are Sb-Sc late-type galaxies, and 14\\% are early-type. This implies an extremely high late-type fraction and extremely low early-type population. At intermediate redshift, \\citet{2012MNRAS.419.3018C} found that early-type systems in higher density regions tend to be more extended than their counterparts in low density environments. Taking into account the effect of the local environment, \\citet{2013MNRAS.434..325F} show that the number of satellites around a galaxy affects its size. CIG galaxies have larger sizes than galaxies in the \\citet{2010ApJS..186..427N} sample with zero or one satellite, which are also larger than galaxies in \\citet{2010ApJS..186..427N} with two or more satellites. The distribution of satellites (faint companions) around isolated primary galaxies provides important information about galaxy formation, as well as a critical test of the $\\Lambda$CDM model on small scales \\citep{1987MNRAS.226..543E,2007AAS...21112602C,2010ApJ...709.1321A,2012MNRAS.425.2817F,2013JCAP...03..014A,2013ApJ...772..109B}. This explains the growing interest for studying the satellite distribution \\citep{2003ApJ...598..260P,2005MNRAS.356.1045S,2012MNRAS.427..428G}, and for exploring the link between galaxy properties and the satellite population \\citep{2007ApJ...658..898P,2011MNRAS.417..370G,2011AstBu..66..389K,2012MNRAS.424.1454E,2013ApJ...770...96G}. According to a previous study \\citep{2013A&A...560A...9A}, the criteria proposed by \\citet{1973AISAO...8....3K} to remove fore- and background galaxies are not fully efficient. About 50\\% of the neighbours, considered as potential companions, have very high recession velocities with respect to the central CIG galaxy: the condition is too restrictive, and may consider as not isolated galaxies slightly affected by their environment. On the other hand, about 92\\% of neighbour galaxies showing recession velocities similar to the corresponding CIG galaxy are not considered as potential companions by the CIG isolation criteria, and may have a non negligible influence on the evolution of the central CIG galaxy. This motivates us to extend the study, taking into account nearby and similar redshift companions to identify physical satellites affecting the evolution of the central CIG galaxy, so as to provide a more physical estimation of the isolation degree of the CIG. About 60\\% of the CIG galaxies have no major (similar-size) companion in the SDSS, according to the CIG (purely photometric) isolation criteria. Nevertheless, considering the third dimension, only 1/3 of the sample has no similar redshift neighbours \\citep{2013A&A...560A...9A}. In this context the CIG represents an excellent sample to study the relation of galaxy properties on both local and large-scale environments. In the present work, we aim to identify and quantify the effects of the satellite distribution around a sample of CIG galaxies, as well as the effects of the Large Scale Structure. This study is organised as follows: in Sect.~\\ref{Sec:data}, the sample and the data used are presented. The method to identify the potential satellite galaxies is described in Sect.~\\ref{Sec:physical}. In Sect.~\\ref{Sec:isolparam}, we describe the parameters used to quantify the environment. We present our results in Sect.~\\ref{Sec:results} and the associated discussion in Sect.~\\ref{Sec:discussion}. Finally a summary and the main conclusions of the study are presented in Sect.~\\ref{Sec:con}. Throughout the study, a cosmology with $\\Omega_{\\Lambda 0} = 0.7$, $\\Omega_{\\rm{m} 0} = 0.3$, and $H_{0}=70$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$ is assumed. ", "conclusions": "\\label{Sec:con} We present a study of the 3-dimensional environment for a sample of 386 galaxies in the \\textbf{C}atalogue of \\textbf{I}solated \\textbf{G}alaxies \\citep[CIG;][]{1973AISAO...8....3K}, using the Ninth Data Release of the Sloan Digital Sky Survey (SDSS-DR9). We identify and quantify the effects of the satellite distribution around a sample of galaxies in the CIG. To recover the physical satellites around the CIG galaxies, we first focus on the satellites which are within the escape speed of each CIG galaxy. We also propose a more conservative method based on the stacked Gaussian distribution of the velocity difference of the neighbours, which gives an upper limit to the influence of the local environment. In comparison to a previous study \\citep{2013A&A...560A...9A} we can estimate the effect of the physical associations that were not taken into account by the CIG isolation criteria, which could also have a non negligible influence on the evolution of the central CIG galaxy. The tidal strengths affecting the primary galaxy are estimated to quantify the effects of the local and Large Scale Structure (LSS) environments. To characterise the LSS around the CIG galaxies, we define the projected number density parameter at the 5$^{\\rm th}$ nearest neighbour. Our main conclusions are the following: \\begin{enumerate} \\item Out of the 386 CIG galaxies considered in this study, at least 340 (88\\% of the sample) have no physically linked satellite. Following the more conservative Gaussian distribution method to identify physical satellites around the CIG galaxies leads to upper limits: out of the 386 CIG galaxies, 327 galaxies (85\\% of the sample) would have no physical companion within a projected distance of 0.3\\,Mpc. \\item Consequently, about 12\\% (and up to 15\\%) of the CIG galaxies have physically bound satellites. CIG galaxies with companions might have a mild tendency (0.3-0.4 dex) to be more massive, which suggests a higher frequency of having suffered a merger in the past. Satellites are in general redder, brighter, and bigger for more massive central CIG galaxies. Also, massive elliptical and lenticular CIG galaxies tend to have satellites with earlier types than similar mass spiral CIG galaxies. \\item Although 15\\% at most of the CIG galaxies in the sample have physically bound satellite, almost all galaxies (97\\%) can be directly related to a LSS. The very large scatter in the quantification of the LSS shows that the CIG includes both galaxies dominated by their immediate environment as well as galaxies almost free from any external influence. \\item The continuous distributions of the $\\eta_{k, \\rm LSS}$ and $Q_{\\rm LSS}$ isolation parameters show that the CIG spans a variety of environments. The connection of the CIG galaxies with the LSS is obvious due to the excess of similar redshift galaxies between 0.3 and 3\\,Mpc. The CIG galaxies are distributed following the LSS of the local universe, although presenting a large heterogeneity in their degrees of connection with it. \\item To evaluate the role of the physically bound satellites with respect to the large scale environment, we compare the magnitudes of the sum of the tidal strengths produced by the physical companions to the sum of the tidal strengths created by all the galaxies in the LSS. When at least one physical companion is present near a CIG galaxy, its effect largely dominates (usually more than 90\\%) the tidal strengths generated by the LSS. \\item To delimit the role of the environment on the physical properties of the galaxies we compare, within the CIG, the most isolated galaxies to the galaxies with companions. We find a clear segregation between CIG galaxies with companions and isolated CIG galaxies. Isolated galaxies are in general bluer, with a younger stellar population and rather high star formation with respect to the older, redder galaxies with companions. These $(u-r)$ colours, in combination with the morphological trends and the $(g-r)$ colours, lead to a coherent view where isolated star forming galaxies are separated from older elliptical galaxies with companions. \\item Conjointly, we find that the satellites are redder and with older stellar populations around massive early-type CIG galaxies while they have a younger stellar content around massive late-type CIG galaxies. This means that if the local environment has an influence on the evolution of the CIG galaxies, reciprocally, the satellites around the CIG galaxies may also be affected by the nature of the primary galaxy. This suggests that the CIG is composed of a heterogeneous population of galaxies, sampling old systems of galaxies but also spanning more recent, dynamical systems of galaxies. \\item As mentioned, the CIG samples a variety of environments, from galaxies in interaction with physical satellites to galaxies with no neighbours in the first 3\\,Mpc around them. Hence, in the construction of catalogues of galaxies in relation to their environments (isolated, pairs, triplets, groups of galaxies), redshift surveys are required in order to distinguish small, faint, physically bound satellites from a background projected galaxy population and reach a more comprehensive 3-dimensional picture of the surroundings. \\end{enumerate}" }, "1402/1402.6713_arXiv.txt": { "abstract": "The conservation of energy, linear momentum and angular momentum are important drivers for our physical understanding of the evolution of the Universe. These quantities are also conserved in Newton's laws of motion under gravity \\citep{Newton:1687}. Numerical integration of the associated equations of motion is extremely challenging, in particular due to the steady growth of numerical errors (by round-off and discrete time-stepping, \\cite{1981PAZh....7..752B,1993ApJ...415..715G,1993ApJ...402L..85H,1994LNP...430..131M}) and the exponential divergence \\citep{1964ApJ...140..250M,2009MNRAS.392.1051U} between two nearby solution. As a result, numerical solutions to the general N-body problem are intrinsically questionable \\citep{2003gmbp.book.....H,1994JAM....61..226L}. Using brute force integrations to arbitrary numerical precision we demonstrate empirically that ensembles of different realizations of resonant 3-body interactions produce statistically indistinguishable results. Although individual solutions using common integration methods are notoriously unreliable, we conjecture that an ensemble of approximate 3-body solutions accurately represents an ensemble of true solutions, so long as the energy during integration is conserved to better than 1/10. We therefore provide an independent confirmation that previous work on self-gravitating systems can actually be trusted, irrespective of the intrinsic chaotic nature of the N-body problem. ", "introduction": "Newton's law of gravitation is one of the fundamental laws in the Universe that holds everything together. Although formulated in the 17th century, scientists today still study the consequences, in particular those of many-body systems, like the solar system, star clusters and the Milky Way galaxy. General analytic solutions to the N-body problem only exist for configurations with one mass, commonly referred to as $N=1$ solutions, and for two masses (equivalently named $N=2$, \\cite{Kepler:1609, Newton:1687}). Problems for $N\\rightarrow \\infty$ can be reduced via Liouville's theorem for Hamiltonian systems to the collisionless Boltzmann equation \\citep[][but see also \\cite{1860dMaxwell}]{1868Bolzmann,1968SvPhU..10..721V}, and therefore analytic solutions for the global distribution function exist. Solutions for $N$ in between these two limits are generally realized by computer simulations. These so-called $N$-body simulations have a major shortcoming in that the solution to any initial realization can only be approximated. The main limiting factors in numerically obtaining a true solution include errors due to round-off and approximations both in the integration and in the time-step strategy \\citep{1996magr.meet..167K}. These generally small errors are magnified by the exponentially sensitive dependence on the 6N-dimensional phase-space coordinates, position and velocity \\citep{1964ApJ...140..250M}. As a consequence, the solution for a numerically integrated self-gravitating system of $N$ masses diverges from the true solution. This error can be controlled to some degree by selecting a phase-space volume preserving or a symplectic algorithm \\citep{1991AJ....102.1528W} and by reducing the integration time step \\citep{1975ARA&A..13....1A,2008MNRAS.386..295H}. The latter however, cannot be reduced indefinitely due to the accumulation of numerical round-off in the mantissa, which is generally limited to 53 bits (64 bits in total, but 11 bits are reserved for the exponent, resulting in only about 15 significant digits). The exponential divergence subsequently causes this small error to propagate to the entire system on a dynamical time scale \\citep{1993ApJ...415..715G}, which is the time scale for a particle to cross the system once. The result of these errors together with the exponential divergence, is the loss of predicting power for a numerical solution to a self gravitating system with $N>2$ after a dynamical time scale. One can subsequently question the predicting qualities of $N$-body simulations for self-gravitating systems, and therewith their usefulness as a scientific instrument. We address this question for $N=3$ by brute-force numerical integration to arbitrary precision. The choice of $N=3$ is motivated by the realization that this represents the first fundamental irregular configuration with the smallest possible number of objects that cannot be solved analytically and cannot be addressed with collisionless theory. In addition, 3-body encounters form a fundamental and frequently occurring topology in any large $N$-body simulation, and therefore also drive the global dynamics of these larger systems. ", "conclusions": "The properties of the binary and the escaper of a three-body system can be described in a statistical way. This is consistent with the findings in previous analytic \\citep{1976MNRAS.176...63M} and numerical \\citep{2004ASPC..316...45V} studies. This behavior was named quasi-ergodicity by \\cite{1976MNRAS.176...63M}. We confirm that this behavior remains valid also for converged 3-body solutions. Based on the symmetry of the distribution in dissolution times (see Fig.\\,\\ref{Fig:CumulativeError}), the final parameters of the binary and escaper, as well as the consistency of the mean and median values of the inaccurate simulations when compared to the converged solution (see Fig.\\,\\ref{Fig:CumulativeError} and Fig.\\,\\ref{Fig:fdissolved}) we argue that global statistical distributions are preserved irrespective of the precision of the calculation as long as energy is preserved to better than 1/10th of the initial energy of the system. Although we have tested only three algorithms for solving the equations of motion we conjecture that the statistical consistency may be preserved also for some other direct $N^2$ methods, and these may also require that energy and angular momentum are preserved to $\\leq 1/10$th. If such direct $N$-body methods comply to the same statistical behavior for collision-less ($N\\gg 3$) systems, it will be interesting to investigate how also other --non-$N^2$-- algorithms, like the hierarchical tree-method \\citep{1986Natur.324..446B} or particle-mesh methods \\citep{Hockney1988} behave in this respect. In studies of self-gravitating systems which adopt the 4th order Hermite integrator, energy and angular momentum are generally conserved up to $\\aplt 10^{-6}$ per dynamical time. Only those simulations in which this requirement is met are often considered reliable and suitable for scientific interpretation. Proof for this seemingly conservative choice has never been provided, and it is unknown whether or not the numerical error and the exponential divergence are not preventing certain parts in parameter space to be accessed, or new physically inaccessible parts in parameters space being explored. We argued that for the resonant 3-body problem the error made during the integration of the equations of motion poses no problem for obtaining scientifically meaningful results so long as energy is conserved to better than about one-tenth of the initial total energy of the system. In that case resonant 3-body interactions should be treated as an ensemble average, and individual results only contribute statistically. By means of numerical integration until a converged solution is obtained we find that the statistical properties of the binary and the escaper resulting from a 3-body resonant encounter are deterministic. This behavior is not guaranteed to propagate to larger $N$ (see also \\cite{1992MNRAS.259..505Q}); $N>3$ requires independent testing, because these introduce more complex solutions in the form of, for example binary-binary outcomes and hierarchical triples. The more extended parameter space for increasing $N$ from 3 to $N=4$ is quite dramatic, in particular for solving the system until a converged solution is reached. {\\bf Acknowledgements} We would like to thank Douglas Heggie for many in-depth discussions, but also Alice Quillen, Piet Hut, Jun Makino, Steve McMillan, Vincent Icke and Inti Pelupessy for discussions and comments on the manuscript, as well as the anonymous referee for careful reading and detailed comments. This work was supported by the Netherlands Research Council NWO (grants \\#643.200.503, \\#639.073.803 and \\#614.061.608) and by the Netherlands Research School for Astronomy (NOVA). Part of the numerical computations were carried out on the Little Green Machine at Leiden University and on the LISA cluster at SURFSara in Amsterdam." }, "1402/1402.1075_arXiv.txt": { "abstract": "A significant portion of transients measured by spacecraft at 1~AU does not show the well-defined properties of magnetic clouds (MCs). % Here, we propose a new class of complex, non-MC ejecta resulting from the interaction of two CMEs with different orientation, which differ from the previously studied multiple-MC event. At 1~AU, they are associated with a smooth rotation of the magnetic field vector over an extended duration and do not show clear signs of interaction. We determine the characteristics of such events based on a numerical simulation and identify and analyze a potential case in the long-duration CME measured {\\it in situ} in 2001 March 19--22. Such events may result in intense, long-duration geo-magnetic storms, with sawtooth events, and may sometimes be misidentified as isolated CMEs. ", "introduction": "Coronal mass ejections (CMEs) are the major driver of intense geo-magnetic activity \\citep[]{Richardson:2001} and have been studied extensively for the past 40 years. CMEs are observed remotely by coronagraphs and heliospheric imagers and measured {\\it in situ} by spacecraft such as ACE and {\\it Wind}. With the launch of SOHO and STEREO, the availability of white-light imagers with wide fields-of-view has made it possible to associate eruptions observed in the corona to CMEs measured {\\it in situ} at 1~AU \\citep[see for example the list by][]{Richardson:2010}. CMEs measured {\\it in-situ} may be divided into three broad categories: magnetic clouds (MCs), non-MC isolated ejecta, and complex ejecta \\citep[similar to the categories of][]{Zurbuchen:2006}. MCs have well-defined properties \\citep[see][]{Burlaga:1981}. Non-MC isolated ejecta typically have some but not all the properties of MCs, and may be sometimes referred to as MC-like ejecta \\citep[]{Lepping:2005}. They may correspond to a distorted CME or to the crossing through the ``leg'' of a CME. Lists of MCs and MC-like ejecta measured at 1~AU by the Wind and ACE spacecraft are maintained \\citep[]{Lepping:2005,Jian:2006,Richardson:2010}. Complex ejecta result from the interaction of successive CMEs \\citep[]{Burlaga:1987}. Some consist of many individual eruptions and it is impossible to relate {\\it in situ} measurements to coronagraphic observations of CMEs \\citep[]{Burlaga:2002}. Others are made up of two clearly distinct MCs separated by an interaction region \\citep[multiple-MC events, see][]{Wang:2003, Lugaz:2005b}. Complex ejecta tend to have long duration and may drive the magnetosphere for an extended period. \\citet{Xie:2006}, for example, studied long (3 days or more) and intense (peak Dst $\\le -100$ nT) geomagnetic storms and found that 24 out of 37 such storms were associated with multiple CMEs. While a typical CME passes over Earth in $\\sim 20$ hours, some events have duration well in excess of 30 hours \\citep[]{Marubashi:2007}. It is possible that some of these long-duration events, believed to be associated with a single, isolated CME are in fact the results of the interaction of two CMEs, a possibility raised for the 2005 May 15 CME by \\citet{Dasso:2009}. Here, we identify a new type of complex ejecta due to the interaction of two CMEs, which results in a long-duration event with a smooth rotation of the magnetic field vector. In section \\ref{simu}, we present the result at 1~AU of two simulations, one of an isolated CME and one of two interacting CMEs, and we discuss the expected geo-effectiveness of such events. In section \\ref{geo}, we present measurements of the 2001 March 19--22 period, which may be associated with the interaction of two CMEs in a way similar to that of the simulation. We discuss our findings and conclude in section \\ref{conclusion}. ", "conclusions": "By combining numerical simulations and the analysis of {\\it in situ} measurements, we have identified a new class of complex ejecta resulting from the interaction of two CMEs. Due to different orientations of the ejections, measurements at 1~AU appear to indicate the passage of a long magnetic cloud, but are in fact, due to two successive and interacting CMEs. With an appropriate orientation of the two CMEs, such an event may result in long-duration geomagnetic storm and be associated with sawtooth event. In details, we have presented the results at 1~AU of a simulation of the interaction of two CMEs with different orientations. We have shown that the resulting complex ejecta is very similar to a MC from an isolated CME, except for the presence of a long ``tail'' in the magnetic field and the hotter temperature throughout the ejecta. We have estimated the expected Dst index for this complex ejecta, and we have found that, while the peak Dst is not as low as that from a well-oriented isolated CME, the tail in the magnetic field results in the Dst to be below $-100$~nT for more than a day, or about 50\\% longer than for the isolated CME. We have also presented the analysis of one long-duration magnetic ejecta observed at 1~AU in 2001 March 19--22. This event resulted in a long, intense geomagnetic storm with a peak Dst of only $-149$~nT but the Dst stayed below $-50$~nT for more than 2 days. There were also a number of sawteeth in March 20 in the first half of the ejecta. Most studies have identified this as an isolated magnetic cloud with a duration in excess of 2 days. We have presented some potential evidence that this ejecta is in fact a complex ejecta associated with two CMEs. A more complete investigation of combined {\\it in situ} and remote-sensing database will be required to assess how common this type of complex ejecta is. This could be helped in the future by the availability of remote-sensing observations of CMEs as they propagate and interact on their way to Earth \\citep[]{CShen:2012, Lugaz:2012b}. The other event that we have tentatively identified in 2004 April 3--6 was also associated with an extended sawtooth event, although the Dst index peaked only at $-117$~nT and was below $-50$~nT for only 15 hours. Further studies are also required to determine how this type of complex ejecta affects Earth's magnetosphere and how the interaction differs from that with an isolated MC or a multiple-MC event." }, "1402/1402.3593_arXiv.txt": { "abstract": "To shed light on the fundamental problems posed by Dark Energy and Dark Matter, a large number of experiments have been performed and combined to constrain cosmological models. We propose a novel way of quantifying the information gained by updates on the parameter constraints from a series of experiments which can either complement earlier measurements or replace them. For this purpose, we use the Kullback-Leibler divergence or relative entropy from information theory to measure differences in the posterior distributions in model parameter space from a pair of experiments. We apply this formalism to a historical series of Cosmic Microwave Background experiments ranging from Boomerang to WMAP, SPT, and Planck. Considering different combinations of these experiments, we thus estimate the information gain in units of bits and distinguish contributions from the reduction of statistical errors and the `surprise' corresponding to a significant shift of the parameters' central values. For this experiment series, we find individual relative entropy gains ranging from about 1 to 30 bits. In some cases, e.g. when comparing WMAP and Planck results, we find that the gains are dominated by the surprise rather than by improvements in statistical precision. We discuss how this technique provides a useful tool for both quantifying the constraining power of data from cosmological probes and detecting the tensions between experiments. ", "introduction": "% \\label{sec:introduction} \\noindent Over recent decades, observational evidence in support of the $\\Lambda$CDM model has grown steadily. Though some of the key ingredients of the model, including Dark Matter and Dark Energy, are not fully understood, an impressive array of new experiments show findings consistent with predictions of the model. Chief among the datasets are high-precision measurements of the Cosmic Microwave Background (CMB) \\cite{Wright:1996um,MacTavish:2008tm,Bennett:2012wy,Hinshaw:2012vg,Story:2012vh,Das:2011gd}. This area has received significant attention recently with the release of the first cosmological analysis of data from the Planck satellite \\cite{Collaboration:2013ww,Collaboration:2013uv}. This experiment can be seen as the latest in a long line of measurements that have targeted the CMB. At each step, data has been used to place constraints on the parameters of the $\\Lambda$CDM model using Bayesian inference to represent the constraints as a probability density in parameter space called the posterior distribution. To judge the progress made between successive measurements, a framework for comparing probability distributions is needed. One method to quantify the difference between the constraints from different surveys is the relative entropy or Kullback-Leibler divergence \\cite{Kullback:1951va} between the respective distributions. Initially motivated from information theory, relative entropy has been proposed in the cosmology literature for forecasting and experiment design \\cite{Paykari:2013dd,Amara:2013wv,March:2011ij,0004-637X-771-1-12} as well as for parameter estimation and model selection \\cite{Verde:2013hp,Kunz:2006de}. In this paper, the relative entropy is introduced as a new tool for measuring the information gained from individual experiments by applying it to their posteriors on the full cosmological parameter space. Two distinct cases of data combinations are analyzed: adding complementary data to existing constraints and replacing data with a more accurate but correlated measurement. The relative entropy between two posteriors measures gains in statistical precision and shifts of confidence regions at the same time. Disentangling these contributions is of great interest for detecting tensions between datasets. In the limit of linear models and Gaussian likelihoods, it is shown that the relative entropy can indeed be separated into an expected part measuring the improvements in precision and a contribution from shifts in the distribution means that is named `surprise'. Explicit expressions for the relative entropy and its decomposition into expected relative entropy and surprise are derived in this limit and can be evaluated from moments of the posteriors. These concepts are then applied to the posteriors of the $\\Lambda$CDM parameters from the Balloon Observations of Millimetric Extragalactic Radiation and Geophysics (BOOMERANG) \\cite{MacTavish:2008tm}, the Wilkinson Microwave Anisotropy Probe (WMAP) \\cite{Bennett:2012wy,Hinshaw:2012vg}, the South Pole Telescope (SPT) \\cite{Story:2012vh}, and the Planck \\cite{Collaboration:2013ww,Collaboration:2013uv} CMB surveys. Using the Monte Carlo Markov chain framework \\verb|CosmoHammer| \\cite{Akeret:2012uk}, estimates for the relative entropy, its expected, and its surprise contributions are given for different combinations of these datasets. The concepts can be easily applied to other probes, too. This paper is organized as follows. In section \\ref{sec:information_in_parameter_estimation} the connection between relative entropy and parameter estimation is discussed. The results for disentangling expected relative entropy and surprise in the Gaussian limit are derived in section \\ref{sec:the_gaussian_limit}. Applying those concepts to CMB surveys, numerical results for the relative entropy between BOOMERANG, WMAP, SPT, and Planck data are shown in section \\ref{sec:information_gains_from_cmb}. The conclusions are summarized in section \\ref{sec:discussion}. ", "conclusions": "% \\label{sec:discussion} \\noindent In order to compare the cosmological parameter constraints from different experiments, a tool for quantifying changes in posterior distributions on the full parameter space is needed. Motivated from information theory, the concept of relative entropy measures differences between distributions in a parametrization independent way and is therefore able to quantify the information gained from new data. In this work, relative entropy is used to develop a new tool for comparing the parameter constraints of the $\\Lambda$CDM model from different CMB surveys. Two ways of combining data from different experiments are discussed: complementary datasets that can be analyzed sequentially and correlated measurements that replace earlier datasets. Relative entropy captures both changes in confidence volumes and location of the regions of the posteriors. In the regime of Gaussian likelihoods and linear models, these contributions can even be distinguished as an expected relative entropy measuring differences in confidence volume and a surprise coming from shifts in parameter space. This Gaussian regime is furthermore at least a good approximation for CMB data analysis. The notions of expected relative entropy and surprise turn the relative entropy into a powerful diagnostic for the consistency of datasets. The relative entropy gains in units of bits from BOOMERANG, WMAP, SPT, and Planck surveys range from about 1 to 30. In general, the numbers are driven by the contributions from the expected relative entropy, $\\ere$, but in three cases the surprise is found to dominate the results. In terms of expected relative entropy, the step from Boomerang to WMAP is the biggest ($\\ere \\sim 18$ bits), followed by the update of WMAP by Planck data ($\\ere \\sim 7$ bits). The addition of SPT data to the WMAP constraints leads to an expected relative entropy gain of $\\ere \\sim 2$ bits. Looking at the total relative entropy gains, inclusion of Planck data shows the biggest gains ($D=29.8$ bits from WMAP 9 and $D=27.8$ bits from WMAP 9 and SPT). When these numbers are decomposed, the relative entropy is found to be dominated by the surprise ($S=21.9$ and $S=21.2$ bits, respectively). These are very significant surprise values since they are $6.5$ standard deviations from expectations. Note that the expected distribution of $D$ is non-Gaussian; for the corresponding p-values see Table \\ref{tab:relent}. This indicates that the shifts in the confidence intervals of the posteriors are large compared to the shifts expected from the increased precision. This conclusion is further supported by the changes in the marginalized posterior plots and points to possible tensions when fitting predictions from $\\Lambda$CDM to different measurements, in line with other findings \\cite{Verde:2013hp,Spergel:2013va}. Other updates considered here also show significant surprise. In particular the update from WMAP 3 to WMAP 5 shows a large surprise of $S=5.5$ which is $5.3$ standard deviations away from expectations. This result might be caused by both the deviations of the WMAP 3 posterior from a normal distribution and the adjustments of the likelihood function for the low-$\\ell$ temperature power spectrum by the WMAP team. It is also interesting to note that while it is possible and expected for the surprise to be both positive and negative, the findings presented here typically show positive surprise. The reason for this is unclear, but if measurement errors are systematically underestimated this in itself would tend to bias the results towards positive surprise. To conclude, the relative entropy is found to be a valuable diagnostic to compare constraints from different measurements. In cases where the likelihood is close to Gaussian and the model is effectively linear, the contributions from shifts in the confidence regions can be separated from the gains in precision. The resulting quantities are easy to estimate and are capable of describing the overall changes in multidimensional constraints in an efficient way." }, "1402/1402.3503.txt": { "abstract": "\\baselineskip = 11pt \\leftskip = 0.65in \\rightskip = 0.65in \\parindent=1pc {\\small Protoplanetary disks composed of dust and gas are ubiquitous around young stars and are commonly recognized as nurseries of planetary systems. Their lifetime, appearance, and structure are determined by an interplay between stellar radiation, gravity, thermal pressure, magnetic field, gas viscosity, turbulence, and rotation. Molecules and dust serve as major heating and cooling agents in disks. Dust grains dominate the disk opacities, reprocess most of the stellar radiation, and shield molecules from ionizing UV/X-ray photons. Disks also dynamically evolve by building up planetary systems which drastically change their gas and dust density structures. Over the past decade significant progress has been achieved in our understanding of disk chemical composition thanks to the upgrade or advent of new millimeter/Infrared facilities (SMA, PdBI, CARMA, Herschel, {\\dbf e-VLA, ALMA).} %which provide more accurate molecular observations. Some major breakthroughs in our comprehension of the disk physics and chemistry have been done since PPV. This review will present and discuss the impact of such improvements on our understanding of the disk physical structure and chemical composition. \\\\~\\\\~\\\\~} %leave this in to get the correct vertical space after the abstract ", "introduction": " ", "conclusions": "" }, "1402/1402.0251_arXiv.txt": { "abstract": "Neutron stars and black holes are the most compact astrophysical objects we can think of and as a consequence they are the main sources of gravitational waves. There are many astrophysically relevant scenarios in which these objects are immersed in or endowed with strong magnetic fields, in such a way that gravitational perturbations can couple to electromagnetic ones and can potentially trigger synergistic electromagnetic signatures. In a recent paper we derived the main equations for gravito-electromagnetic perturbations and studied in detail the case of polar electromagnetic perturbations driven by axial gravitational perturbations. In this paper we deal with the case of axial electromagnetic perturbations driven by polar black-hole or neutron stars oscillations, in which the energy emitted in case is considerably larger than in the previous case. In the case of neutron stars the phenomenon lasts considerably longer since the fluid acts as an energy reservoir that shakes the magnetic field for a timescale of the order of secs. \\PACS{04.30.-w, 04.40.Nr, 95.85.Sz} ", "introduction": "\\label{sec:Intro} Gravitational wave observations will enable an unprecedented view of the Cosmos thanks to interferometric detectors such as the forthcoming second-generation ground-based detectors LIGO, Virgo, and KAGRA~\\cite{LIGO,VIRGO,KAGRA}, third-generation future ones like the Einstein Telescope~\\cite{ET}, and space-based ones like eLISA and DECIGO~\\cite{eLISA,DECIGO}. In addition, there are also good prospect of detection of gravitational waves using several Pulsar Timing Arrays that are organized in the International Pulsar Timing Array collaboration~\\cite{IPTA}. When the second generation of ground-based detectors becomes fully operational and reaches the designed sensitivity, these instruments will be truly tools of astronomical discovery. For compact objects (neutron stars and black holes), we will be able to determine from gravitational-wave observations properties such as their masses, radius, rotation rates, and, for the case of neutron stars, the structure of matter itself (i.e. equation of state)~\\cite{AK1998,KAA2001,Sotani2004,Erich2011,Sotani2011a}. Another remarkable capability of gravitational-wave observations is their potential to validate whether General Relativity is indeed the correct classical theory of gravity and to put constraints to alternative theories~\\cite{SK2004,S2009,YYT2012}. In summary, the detection and characterization of gravitational waves will provide insights on a plethora of physical phenomena, insights that will complement our knowledge obtained from particle accelerators and electromagnetic observations. Processes in compact objects associated with strong magnetic fields are an example of physical phenomena for which only multi-messenger observations provide a complete picture. For instance, in the case of the giant flares in soft gamma-ray repeaters, evidence points to magnetars (strongly magnetized neutron stars) as the objects where these events takes place. % Moreover, in the three giant flares detected so far, theoretical studies~\\cite{L2007,SKS2007,CBK2009,CSF2009,AGS2009,S2011,HoLevin11,GCFES2012,CK2012,SNIO2012a,SNIO2012b,HoLevin12} suggest that the quasi-periodic oscillations frequencies, measured during the afterglow of the flare activity~\\cite{WS2006}, are associated with crustal oscillations in the neutron star and/or oscillations in the magnetic field of the neutron star. Furthermore, recent studies~\\cite{A2011} have speculated that the flare activity could couple with the dynamical gravitational fields of the star to trigger the production of gravitational waves. In addition, the importance of electromagnetic counterparts to gravitational-wave signals has been also discussed in the context of recent numerical simulations of binary black hole systems~\\cite{PL2010,PT2010}, of collapsing hypermassive neutron stars~\\cite{LP2012}, and of binary neutron star systems~\\cite{PL2013a,PL2013b}. Further studies are needed to investigate the conditions under which electromagnetic and gravitational perturbations couple to generate emission of radiation through both channels. We have taken a first step in recent work~\\cite{SKLS2013} (Paper I). We studied the coupling of axial gravitational and electromagnetic perturbations for both, black holes and neutron stars, neglecting the back reaction of the electromagnetic waves since the energy budget in the system is dominated by gravity. Our study found that the emitted energy in electromagnetic waves driven by the gravitational waves is proportional to not only the emitted energy in gravitational waves but also to the square of the strength of the magnetic field of the central object. That is, $E_{\\rm EM}/E_{\\rm GW} = \\alpha (B/10^{15} {\\rm G})^2$, where $B$ is the magnetic field strength. For the case of a black hole background it was found that $\\alpha \\approx 8\\times 10^{-6}$, while for neutron stars this factor depends on the compactness of the neutron star. For a compactness of $M/R = 0.162$ we obtained $\\alpha \\approx 1.61\\times 10^{-5}$ while for a compactness of $M/R = 0.237$ the result was $\\alpha \\approx 4.37\\times 10^{-6}$. Additionally, it is suggested that there can be resonances between the two types of waves under certain conditions \\cite{KT2013}. In this paper, we investigate electromagnetic emission driven by polar gravitational perturbations. Unlike axial gravitational perturbations, the expectation is that polar perturbations are able to drive longer emission in both gravitational and electromagnetic waves. Here again, we ignore back-reaction effects from the magnetic field on the spacetime geometry. Following the previous study in Paper I, we consider both the cases of black holes and neutron stars. We examine in particular the dependence of the gravitational-electromagnetic coupling for neutron stars on the stellar mass and the equation of state. Then, we find that $\\alpha=1.78\\times 10^{-5}$ for a black hole background, a value roughly twice the one found in Paper I for electromagnetic waves driven by the axial gravitational waves. On the other hand, we also find that $\\alpha$, for a neutron star background, can be written as a function of ${\\cal \\chi} \\equiv (1-2M/R)^{-1}(M/R^3)^{-1}$ as $\\alpha \\times 10^7 = 0.922 - 0.903({\\cal \\chi}/{500}) + 0.552 ({\\cal \\chi}/{500})^2$, where $M$ an $R$ denote the stellar mass and radius. This paper is organized as follows. In the next section (Section~\\ref{sec:Background}), we briefly describe the equilibrium configurations that provide the gravitational background (spacetime background geometry) as well as the the perturbative equations. In Section~\\ref{sec:Results}, we show and discuss the numerical results for both the black hole and neutron star cases. We make conclusions in Section~\\ref{sec:conclusion}. Throughout this paper we adopt geometric units, $c=G=1$, where $c$ and $G$ denote the speed of light and the gravitational constant, respectively, and use the metric signature $(-,+,+,+)$. ", "conclusions": "\\label{sec:conclusion} This paper presents results of the emission of gravitational and electromagnetic waves from the strongly magnetized compact objects, extending the study presented in Paper I~\\cite{SKLS2013}. In Paper I we formulated the problem and examined in detail the case of ``electric-type\" (polar) electromagnetic waves driven by ``magnetic-type'' (axial) gravitational perturbations. In the present work, we focused on the case in which ``magnetic-type\" (axial) electromagnetic waves are driven by ``electric-type'' (polar) gravitational waves. As we have pointed out before, for the case of neutron stars, this situation is observationally more interesting because such type of gravitational waves are associated with fluid oscillations, and thus they will tell us about properties of the structure of the star. In the case of black holes, we find that the emission of ``magnetic-type\" (axial) electromagnetic waves driven by ``electric-type'' (polar) gravitational perturbations is very similar to that of ``electric-type\" (polar) electromagnetic waves driven by ``magnetic-type'' (axial) gravitational perturbations studied in Paper I. More specifically, we obtained the same relationship as in Paper I between the radiated energy in electromagnetic waves and the energy emitted in gravitational waves, namely $E_{\\rm EM} = \\alpha B_{15}^2E_{\\rm GW}$. On the other hand, for neutron stars, we observe that the polar gravitational waves last longer than in the axial case, and thus also the radiation of ``magnetic-type\" (axial) electromagnetic waves. This is due to the coupling with the fluid which acts as a storage reservoir of energy which slowly leaks energy to the electromagnetic sector during a longer time period. This process is revealed via the study of the spectra of the perturbations, where we can see that the oscillation frequencies of gravitational and electromagnetic waves are the same as those of the fluid perturbations. Our results also show that the larger the compactness of the star, the stronger is the coupling between the two types of perturbations leading to an enhancement in the energy transfer and the emission of electromagnetic waves. At the same time, the emitted radiation get damped out faster for large compactness models than for those with smaller values. Unfortunately, at frequencies of a few kHz, the emitted radiation in electromagnetic waves is easily absorbed by the interstellar medium and/or the plasma around the compact objects. To be detectable, one would have to explore reprocessing mechanisms. We are currently investigating situations/mechanism where this electromagnetic emission may be detectable and the type of information that we could extract from them." }, "1402/1402.6850_arXiv.txt": { "abstract": "{Towards the high galactic latitude sky, the far-infrared (FIR) intensity is tightly correlated to the total hydrogen column density which is made up of atomic ($\\ion{H}{i}$) and molecular hydrogen (H$_{2})$. Above a certain column density threshold, atomic hydrogen turns molecular.} {We analyse gas and dust properties of intermediate-velocity clouds (IVCs) in the lower galactic halo to explore their transition from the atomic to the molecular phase. Driven by observations, we investigate the physical processes that transform a purely atomic IVC into a molecular one.} {Data from the Effelsberg-Bonn $\\ion{H}{i}$-Survey (EBHIS) are correlated to FIR wavebands of the Planck satellite and IRIS. Modified black-body emission spectra are fitted to deduce dust optical depths and grain temperatures. We remove the contribution of atomic hydrogen to the FIR intensity to estimate molecular hydrogen column densities.} {Two IVCs show different FIR properties, despite their similarity in $\\ion{H}{i}$, such as narrow spectral lines and large column densities. One FIR bright IVC is associated with H$_{2}$, confirmed by $^{12}$CO $(1\\rightarrow0)$ emission; the other IVC is FIR dim and shows no FIR excess, which indicates the absence of molecular hydrogen.} {We propose that the FIR dim and bright IVCs probe the transition between the atomic and molecular gas phase. Triggered by dynamical processes, this transition happens during the descent of IVCs onto the galactic disk. The most natural driver is ram pressure exerted onto the cloud by the increasing halo density. Because of the enhanced pressure, the formation timescale of H$_{2}$ is reduced, allowing the formation of large amounts of H$_{2}$ within a few Myr.} ", "introduction": "\\label{sec:introduction} The Infrared Astronomical Satellite (IRAS) showed that, up to a certain threshold, the $\\ion{H}{i}$ $21\\,\\mathrm{cm}$ emission correlates linearly with the far-infrared (FIR) dust continuum at high galactic latitudes \\citep{Low1984,Boulanger1988,Boulanger1996}. Despite this linear relationship, previous studies towards high galactic latitudes \\citep[e.g.][]{Desert1988,Reach1998} find excess FIR radiation associated with molecular gas, which is traced by carbon monoxide (CO) emission \\citep{Magnani1985}. Several of these objects show radial velocities relative to the local standard of rest (LSR) between $30\\,\\mathrm{km}\\,\\mathrm{s}^{-1} \\leq |v_{\\mathrm{LSR}}| \\leq 90\\,\\mathrm{km}\\,\\mathrm{s}^{-1}$ that are difficult to account for with a simple model of galactic rotation. These clouds are classified as intermediate-velocity clouds (IVCs) and, as a subcategory, intermediate-velocity molecular clouds (IVMCs). Intermediate-velocity clouds are thought to originate from a galactic fountain process \\citep{Bregman2004}. Typically they show up with metal abundances close to solar, a traceable dust content, and distances below $5\\,\\mathrm{kpc}$ \\citep{Wakker2001}. Commonly, IVCs emit in the FIR \\citep[e.g.][]{Planckcollaboration2011XXIV}. Towards many IVCs a diffuse H$_{2}$ column density of $N_{\\mathrm{H}_{2}} = 10^{14}-10^{16}\\,\\mathrm{cm}^{-2}$ is inferred by interstellar absorption line measurements \\citep{Richter2003,Wakker2006}; however, only a few IVMCs are known \\citep{Magnani2010}. Dynamical processes resulting from the motion of IVCs through the halo and their descent onto the galactic disk are thought to play a major role in the process of H$_{2}$ formation in IVCs \\citep{Odenwald1987,Desert1990a,Gillmon2006b,Guillard2009}. The transition between the atomic and the molecular phase happens at a total hydrogen column density of $N_{\\mathrm{H}} = 2.0-5.0 \\times 10^{20}\\,\\mathrm{cm}^{-2}$ \\citep{ Savage1977,Reach1994,Lagache1998,Gillmon2006a,Gillmon2006b,Planckcollaboration2011XXIV}. In this paper we study the gas and dust properties of IVCs and their transition from the atomic to the molecular phase by correlating the $\\ion{H}{i}$ distribution to the FIR dust emission at high galactic latitudes. We are interested in the conditions required for the transition from $\\ion{H}{i}$ to H$_{2}$ in IVCs at the disk-halo interface. We use data from the new Effelsberg-Bonn $\\ion{H}{i}$-Survey \\citep[EBHIS,][]{Winkel2010,Kerp2011}, the Planck Satellite \\citep{Planckcollaboration2013I}, and IRIS \\citep{MivilleDeschenes2005b}. Our field of interest has a size of $25 \\times 25\\,\\mathrm{deg}^{2}$ centred on galactic coordinates ($l$, $b$) $=$ ($235^{\\circ}$, $65^{\\circ}$). This area is located between the Intermediate-Velocity (IV) Arch and IV Spur which are two large structures in the distribution of IVCs \\citep{Wakker2004}. In \\citet{Planckcollaboration2013XXX} this field is used also in the analysis of the cosmic infrared background. Section \\ref{sec:data} gives the properties of our data in $\\ion{H}{i}$ and the FIR. Section \\ref{sec:methods} describes the $\\ion{H}{i}$-FIR correlation in more detail. Section \\ref{sec:entire-field} presents the observational results inferred from the $\\ion{H}{i}$, the FIR, and their correlation for the entire field, while Sect. \\ref{sec:individual-clouds} gives results for two particular IVCs within the field. Section \\ref{sec:discussion} discusses a dynamically driven $\\ion{H}{i}$-H$_{2}$ transition with respect to the two IVCs and Sect. \\ref{sec:conclusion} summarises our results. ", "conclusions": "\\label{sec:conclusion} We correlate $\\ion{H}{i}$ emission to the brightness at various wavelengths in the FIR dust continuum using new data from the Effelsberg-Bonn $\\ion{H}{i}$ Survey (EBHIS) and the Planck satellite complemented by IRIS data. We study in detail two IVCs that show many similarities in their $\\ion{H}{i}$ properties, such as narrow spectral lines with $\\mathrm{FWHM}\\simeq5\\,\\mathrm{km}\\,\\mathrm{\\mathrm{s}}^{-1}$ and $\\ion{H}{i}$ column densities of $N_{\\ion{H}{i}} = 3 - 4 \\times 10^{20}\\,\\mathrm{cm}^{-2}$. Despite their similarity in $\\ion{H}{i}$, their FIR emission exhibits large differences: one cloud is FIR bright while the other IVC is FIR faint. From the quantitative correlation of $\\ion{H}{i}$ and FIR emission, we calculate maps of molecular hydrogen column density revealing large amounts of H$_{2}$ in the field of interest for which no existing CO surveys of the region has detected a CO counterpart. How much of this H$_{2}$ is actually CO-dark gas we cannot tell since the CO survey data that is available today is not sensitive enough. We do, however, know that the molecular IVC contains CO. The $\\ion{H}{i}$ emission traces only a part of the total gas distribution. Together with the inferred H$_{2}$ column densities, the relation between gas and dust is consistent. Based on our findings we describe a scenario of a dynamical transition from atomic to molecular IVCs in the lower galactic halo. During the descent of the IVCs through the galactic halo, they are compressed as a result of external pressure and ram pressure. Once the ram pressure exceeds the thermal pressure, shocks are created which enhance the pressure locally and accumulate gas and dust. An increased pressure reduces the formation timescale of H$_{2}$ in condensations of the cold atomic medium. The only physical distinctions between the two IVCs are the amount of dust within the clouds, measured by the dust surface mass density, and the amount of H$_{2}$. The molecular IVC has a factor of $2-3$ more dust within its central region than the atomic IVC, which is accompanied by a $50\\%$ higher total hydrogen column density. On the other hand, IVC\\,1 has more $\\ion{H}{i}$ mass in total. Apparently, this mass is not distributed so as to allow efficient H$_{2}$ formation. According to the data presented in this paper, we expect that the atomic IVC will also turn molecular in a few Myrs. Processes on spatial scales that are not resolved by our data govern the evolution of an IVC from an atomic to a molecular cloud. The $10.8\\,\\arcmin$ resolution of EBHIS corresponds to a spatial resolution of $1\\,\\mathrm{pc}$ at a distance of $0.4\\,\\mathrm{kpc}$. However, the accumulation and condensation of smaller and denser clumps regulate the H$_{2}$ formation on sub-parsec scales. Radio-interferometric observations should reveal a different distribution of $\\ion{H}{i}$ in the two IVCs on sub-parsec scales, for example compact cores in the molecular IVC. Our approach appears to open a way to search for dark H$_{2}$ gas across the entire sky. Globally, this new search may reveal other clouds in transition from the atomic to the molecular gas phase. Compared to the low angular resolution of former large-scale $\\ion{H}{i}$ single-dish surveys, the few detections of molecular IVCs so far may be due to the small-angular extent of the molecular cores." }, "1402/1402.4408_arXiv.txt": { "abstract": "Atmospheric spectroscopy of extrasolar planets is an intricate business. Atmospheric signatures typically require a photometric precision of $1 \\times 10^{-4}$ in flux over several hours. Such precision demands high instrument stability as well as an understanding of stellar variability and an optimal data reduction and removal of systematic noise. In the context of the \\echo~mission concept, we here discuss the data reduction and analysis pipeline developed for the \\echo~end-to-end simulator \\echosim. We present and discuss the step by step procedures required in order to obtain the final exoplanetary spectrum from the \\echosim~`raw data' using a simulated observation of the secondary eclipse of the hot-Neptune 55 Cnc e. ", "introduction": "\\label{intro} The field of extrasolar planets is innovative as it is new. Recent successes in characterisation of extrasolar planets are also always tales of characterising the instrument response function to an unprecedented detail. Always being at the edge of technical feasibility means that instrument calibration, observing strategy as well as data analysis and modelling are interdependent. In the light of the \\echo~ESA-M3 mission concept \\cite{tinetti12}, such interdependence becomes important in the study of engineering decisions and instrument trade-offs. In other words, one needs to simulate the full observational and data analysis chain in order to gauge the impact the instrument concept has on the achievable error bar of the detection. Such a feat requires an advanced mission end-to-end simulator as well as an advanced data analysis pipeline. In this paper, we discuss the data analysis pipeline which is used in conjunction to the mission simulator, \\echosim~ \\cite{pascale13}. The \\echosim~data pipeline (from here on \\echodp) is a stand-alone software custom built for \\echosim~but with easy adaptability to other instruments and data sets in mind. The method by which the \\echo~mission will characterise the nature of extrasolar planets is by time resolved spectroscopy of their atmospheres, in particular of transiting extrasolar planets. Briefly, when an exoplanet transits in front of its host star (in our line of sight) we observe a diminishing of the stellar flux due to the obscuration of the planet. The depth of the resulting lightcurve allows us to estimate the planetary radius (given the stellar radius is known). This we refer to as transit (or primary eclipse) observation. Should the exoplanet feature an extended atmosphere, we expect some of the stellar light to filter through the terminator region of the planetary atmosphere. Here we are sensitive to molecules absorbing the stellar light at specific wavelengths. We hence perceive a variation of transit depths depending on the wavelength range observed. These variations constitute the signatures of an exoplanetary absorption spectrum. Similarly, we can observe the occultation (or secondary eclipse) where the thermal contribution of the exoplanet's day-side is lost to the observer as the planet passes behind its host star. The study of transmission and emission spectroscopy is now a well established field for both space and ground based observations of exoplanetary atmospheres (e.g. \\cite{beaulieu10,beaulieu11,charbonneau08,brogi12,bean11,swain08,swain08b,swain09a,crouzet12,deming13,grillmair08,thatte10,tinetti07,pont08,swain12,knutson11,sing11,tinetti10,mooij12,bean11b,stevenson10} also see \\cite{tinetti13} for a comprehensive review). \\subsection{\\echosim} \\echosim~ is the \\echo~ mission end-to-end simulator. \\echosim~ implements a detailed simulation of the major observational and instrumental effects, and associated systematics. It also allows the influence of individual instrumental and astrophysical parameters to be studied and thus represents a key tool in the optimisation of the instrument design. Observation and calibration strategies, data reduction pipelines and analysis tools can all be designed effectively using the realistic outputs produced by \\echosim \\cite{pascale13,waldmann13}. The simulation output closely mimics standard STSci\\footnote{$http://archive.stsci.edu/hst/$} FITS files, allowing for a high degree of compatibility with standard astronomical data reduction routines. \\subsection{Examples} We illustrate individual steps in \\echodp~using diagrams. Unless specified otherwise, we follow a single data processing run of \\echosim~simulated data of the hot-Neptune 55 Cnc e. \\echosim~was run to simulate the Chemical Census mode of \\echo, in which we co-add (in the case of 55 Cnc e) five eclipse observations to obtain a minimal signal-to-noise (S/N) of the final spectrum of S/N $\\sim$ 5. For this we assume spectra reconstructed with a resolving powers of 50,50,30,30,30 for the VNIR, SWIR, MWIR-1, MWIR-2, and LWIR channels. With each channel having a larger native resolving power, this allows us to increase the SNR of the detection for this particular observing mode. See \\cite{tinetti12} for a review of the proposed \\echo~observing modes. ", "conclusions": "\\echodp~is a custom built data reduction and analysis pipeline for the \\echosim~end-to-end mission simulator of the \\echo~mission concept. Despite its customised nature, we have developed the pipeline with easy adaptability (through its fully object-orientated programming ) to other instruments and data-sets in mind. The pipeline features state of the art data de-correlation algorithms as well as a full Bayesian analysis implementation via adaptive MCMC. Both these aspects, the de-trending as well as the exploration of stellar variability are not required for the current version of \\echosim~(version 3.x) but included with future releases. These releases will have special emphasis on realistic stellar noise simulations \\cite{ballerini12} as well as more advanced non-Gaussian instrument systematics." }, "1402/1402.1707_arXiv.txt": { "abstract": "Atom interferometry is an exciting tool to probe fundamental physics. It is considered especially apt to test the universality of free fall by using two different sorts of atoms. The increasing sensitivity required for this kind of experiment sets severe requirements on its environments, instrument control, and systematic effects. This can partially be mitigated by going to space as was proposed, for example, in the Spacetime Explorer and Quantum Equivalence Principle Space Test (STE-QUEST) mission. However, the requirements on the instrument are still very challenging. For example, the specifications of the STE-QUEST mission imply that the Feshbach coils of the atom interferometer are allowed to change their radius only by about $260\\,$nm or $2.6\\cdot 10^{-4}\\,$\\% due to thermal expansion although they consume an average power of $22\\,$W. Also Earth's magnetic field has to be suppressed by a factor of $10^5$. We show in this article that with the right design such thermal and magnetic requirements can indeed be met and that these are not an impediment for the exciting physics possible with atom interferometers in space. ", "introduction": "One of the biggest challenges in current theoretical physics is that of finding a valid theory of quantum gravity. Although many theories were proposed, ultimately this question has to be resolved experimentally. Thus, many experiments regarding tests of the fundamental assumptions of gravity and its basic effects were carried out and suggested including tests of the Universality of Free Fall (UFF). This basic principle is a cornerstone of Einstein's theory of gravity and implies that the trajectories of freely falling, structureless test particles in a gravitational field only depend on their initial position and velocity. In particular, the path of such test particles is independent of their composition. A violation of this principle measured by the E\\\"otv\\\"os ratio would help to discriminate between the different proposed theories of quantum gravity and it would point in the right direction for further theoretical and experimental development. If the UFF holds to the tested accuracy, bounds will be placed on the viable alternative theories. A very thriving tool to test the UFF using quantum matter is dual species atom interferometry,\\cite{Peters_2001,Fray_2004,Bonnin_2013} where two atomic clouds propagate freely and their trajectories are compared. Although the advances along this avenue are tremendous, very stringent requirements have to be satisfied to reach down to an accuracy of the UFF test comparable to the classical tests like lunar laser ranging\\cite{Williams2012CQG} and torsion balance experiments.\\cite{Schlamminger08PRL} In the present paper, we show that the requirements regarding the magnetic field and the thermal requirements, which are closely tied to each other, can indeed be satisfied. As a source for specific requirements we refer to the Spacetime Explorer and Quantum Equivalence Principle Space Test (STE-QUEST) mission, which is a medium-size candidate mission in the Cosmic Vision Program of the European Space Agency. Different aspects of the STE-QUEST mission are described in Ref.\\ \\onlinecite{Hechenblaikner_2013,Aguilera_2013,Tino_2013,Schuldt_2014}. The specific requirements for STE-QUEST are derived in Ref.\\ \\onlinecite{Schubert_2013}. The STE-QUEST atom interferometer was designed with heritage from projects such as the DLR project QUANTUS\\cite{Zoest_2010} (QUANTengase Unter Schwere\\-losig\\-keit) and the CNES project I.C.E.\\cite{Nyman_2006} (Inter\\-f\\'erom\\'etrie Coh\\'erente pour l'Espace). In QUANTUS, experiments with Rubidium (Rb) Bose-Einstein condensates are carried out in the drop tower at ZARM in Bremen and are currently prepared for a sounding rocket by the end of 2014.\\cite{MAIUS} A dual species interferometer is already included in I.C.E., where the experiment is performed in parabolic zero-g flight. The experimental sequence foreseen for STE-QUEST, is to cool two ensembles of \\textsuperscript{85}Rb and \\textsuperscript{87}Rb atoms down to Bose-Einstein condensation in two steps. First atoms are loaded from a magneto-optical trap into a magnetic trap on an atom chip for evaporative pre-cooling. Then, they are loaded into an optical dipole trap for final evaporative cooling. During this step a strong homogeneous magnetic field of $160\\,$G is applied in order to tune atomic interactions using a so called Feshbach resonance. The magnetic field is generated by two coils in Helmholtz configuration with a middle distance of $105\\,$mm, a coil's mean diameter of $206\\,$mm, and equal electrical current of $4.8\\,$A flowing in the same direction. This way, two Bose-Einstein condensates of $10^6$ atoms shall be generated and released from the trap to fall freely in Earth's gravitational field. A detailed description of the atom interferometer design of STE-QUEST can be found in Ref.\\ \\onlinecite{Schuldt_2014}. The article is organized as follows: In Sec.\\ \\ref{sec:requirements}, the thermal requirements and the requirements for the magnetic shielding are summarized for the example of the STE-QUEST atom interferometer. Subsequently, the methods and models used for the analysis of the thermal control system and the magnetic shielding are discussed in detail. The results are summarized in Sec.\\ \\ref{sec:results}. ", "conclusions": "While dual species atom interferometry in space is a powerful and promising tool for fundamental physics science, the technological challenges that have to be overcome in order to realize a sensor offering a high measurement accuracy are considerable. Among these, the thermal management and the magnetic shielding of the sensor hardware have a strong impact on the achievable sensor accuracy. In this paper, we have analyzed the current design of the thermal control system and the magnetic shielding with respect to the STE-QUEST mission specification as a test case scenario. Both systems are designed and verified by means of complex 3D FE models, which have been used to simulate the evolution of heat flows, temperatures and magnetic field gradients for the dynamic boundary condition throughout the STE-QUEST orbit and the measurement cycle. The numeric results show that the current design of the atom interferometer allows to satisfy thermal and magnetic requirements by a thorough design of both the thermal control system and the magnetic shielding. Furthermore a more advanced design of the magnetic coils could help to relax the requirements on thermal expansion. The design and the FE models of both the thermal control system and the magnetic shielding will be used for future optimization or verification of overall design changes arising from other subsystem or mission requirements." }, "1402/1402.5312_arXiv.txt": { "abstract": "{ Most Class~II sources (of nearby star forming regions) are surrounded by disks with weak millimeter continuum emission. These ``faint'' disks may hold clues to the disk dissipation mechanism. However, the physical properties of proto-planetary disks have been directly constrained by imaging only the brightest sources. } { We attempt to determine the characteristics of such faint disks around classical T Tauri stars, and to explore the link between disk faintness and the proposed disk dispersal mechanisms (accretion, viscous spreading, photo-evaporation, planetary system formation). } { We performed high-angular resolution ($0.3''$) imaging of a small sample of disks (9 sources) with low 1.3\\, mm continuum flux (mostly $<30$\\,mJy) with the IRAM Plateau de Bure interferometer and simultaneously searched for $^{13}$CO (or CO) J=2-1 line emission. Using a simple parametric disk model, we determine characteristic sizes of the disks, in dust and gas, and we constrain surface densities in the central 50\\,AU. } { All disks are much smaller than the bright disks imaged so far, both in continuum and $^{13}$CO lines (5 detections). In continuum, half of the disks are very small, with characteristic radii less than 10\\,AU, but still have high surface density values. Small sizes appear to be the main cause for the low disk luminosity. Direct evidence for grain growth is found for the three disks % that are sufficiently resolved. Low continuum opacity is attested in two systems only, but we cannot firmly distinguish between a low gas surface density and a lower dust emissivity resulting from grain growth. Finally, we report a tentative discovery of a $\\sim 20$\\,AU radius cavity in DS Tau, which with the (unresolved) ``transition'' disk of CX~Tau, brings the proportion of ``transitional'' disks to a similar value to that of brighter sources. The existence of cavities cannot explain by itself their observed low mm flux. } { This study highlights a category of very compact dust disks, still exhibiting high surface densities, which may represent up to 25\\% of the whole disk population. While its origin is unclear with the current data alone, it may be related to the compact planetary systems found by the Kepler mission.} ", "introduction": "The many known exoplanet systems display tremendous variety in the masses of their stars, the masses of planets they host, and their architectures. Study of the early evolution of protostellar disks will contribute to our understanding of the origins of this diversity. \"Faint\" disks (which we defined as having 1.3 mm continuum flux $\\simless 60$ mJy at the Taurus distance, 140 pc) play a key role in this respect. First of all, most disks around Class II sources are ``faint''. \\citet{Andre+Montmerle_1994} find that 50 \\% of the Class II sources in $\\rho$ Oph (at 120 pc) have 1.3\\,mm flux $< 40$ mJy, while only 25 \\% have flux $> 125$~mJy. Similar statistics were obtained in the Taurus region by \\citet{Beckwith+etal_1990}. Second, young stars of mass $< 0.3$\\,M$_\\odot$ (the most numerous in our Galaxy) are expected to be surrounded by low-mass, intrinsically faint disks from theoretical considerations. Until the work of \\citet{Andrews+etal_2013}, this trend was not clearly observed because of incompleteness of the samples towards low masses (sensitivity). The large dispersion of disk continuum brightness as a function of stellar mass also loosen the expected correlation. Finally, low-mass stars with faint disks play a critical role in validating stellar evolution models. Providing accurate stellar masses in the 0.2-0.4\\,$\\Msun$ range would allow critical tests of early stellar evolution models, hopefully offering better age estimates for all young stars to ultimately establish a more reliable clock for planetary system formation. Disks can be ``faint'' for many different reasons. \\textit{Low disk surface density} is the simplest reason. Temperature also affects continuum flux, but varies over a much more limited range. \\textit{Dust settling} when seen at high inclination \\citep[$> 80^\\circ$, e.g. the edge-on objects HK Tau B, HV Tau C in][]{Guilloteau+etal_2011} leads to smaller flux \\citep[see the study of][]{Boehler+etal_2013}, as the warmer parts are hidden by opacity. Disks may also \\textit{just be small}, as for example \\object{BP Tau} \\citep{Dutrey+etal_2003}, sometimes as a result of \\textit{outer disk truncation}, as happens in binaries \\citep[e.g. Haro 6-10, UY Aur,][]{Guilloteau+etal_2011,Duvert+etal_2000}. On another hand, \\textit{inner disk clearing}, which suppresses the densest parts of the disk, is also observed in many objects \\citep[e.g. AB Aur, LkCa 15, MWC 758,][]{Pietu+etal_2005,Pietu+etal_2006,Isella+etal_2010a}. Finally, \\textit{grain growth}, which reduces the opacity per unit mass, can lead to small (sub)mm continuum flux densities. These causes are not mutually exclusive: \\object{HH 30} \\citep{Guilloteau+etal_2008} combines small size, inner hole and grain growth, as well as a somewhat lower than average surface density. The various possible causes have different observational consequences. \\citet{Ricci+etal_2010a} found a correlation between the spectral index $\\alpha$ ($S(\\nu) \\propto \\nu^\\alpha)$ with S(1.3\\,mm), with $\\alpha \\simeq 2$ for low flux densities. $\\alpha = 2$ indicates either optically thick emission, or large grains (``pebbles''), as the dust emissivity index $\\beta = \\alpha-2$ in the optically thin case. However, the observed correlation fails to be reproduced by current models of disk evolution including grain growth and viscous evolution \\citep[][their Fig.2]{Birnstiel+etal_2010}. A trend for smaller sizes at low flux \\citep[][their Fig.7]{Andrews+etal_2010} supports the compact disk interpretation. Furthermore, since \\citet{Guilloteau+etal_2011} showed that larger grains are preferentially found in the inner 60 - 100 AU of disks, small disks are also expected to have lower $\\alpha$, producing the observed correlation between flux and spectral index. Whatever the cause of the faintness, ``faint'' disks are essential objects for understanding the planetary system formation process. Grain growth and dust settling are major steps in this respect. Inner cavities can be signposts of planetary system formation, while disks with low surface density may be in the dust (and gas) dissipation stage. To separate between the possible causes, resolved images are essential. Because of the resolution and sensitivity limitations of the existing mm arrays, most studies have focussed on the brighter objects \\citep[e.g.][]{Andrews+Williams_2007,Isella+etal_2009,Guilloteau+etal_2011}, so that resolved images of faint disks are rare. A first attempt was performed by \\citet{Andrews+etal_2010}, who studied the continuum emission from some disks in the $\\rho$ Oph regions, 6 of which would qualify as ``faint'', showing that on average these disks appeared indeed smaller than brighter ones. However, the sensitivity was insufficient to study the CO line emission. We report here on a high angular resolution (0.4-0.6$''$) study in line and continuum emission of 9 disks in the Taurus complex, 8 of them having $S_\\nu < 30$ mJy at 1.3\\,mm, and the last one, CW Tau, being at the boundary between ``faint'' and ``bright'' disks with $S_\\nu \\approx 60$ mJy. ", "conclusions": "We report the first high angular resolution study of disks with low level of continuum emission at mm wavelengths. We find that these disks are substantially smaller than the brighter ones imaged so far. A majority of them remain essentially unresolved at 0.4$''$ resolution, leading to radii $< 15$ AU. These disks are small enough to be optically thick at 1.4 mm, which provides a simple explanation for their apparent spectral index $\\alpha \\sim 2$. $^{13}$CO emission has been detected in five disks. Although the gas distribution extends further out than the apparent size of the dust emission (as in all other disks around T Tauri and HAe stars), its outer radius is also small compared to previously studied gas disks. Only two disks (V836 Tau, an intermediate case between Class II and Class III sources, and HO Tau) have apparently low surface densities simultaneously attested by their low opacity in continuum and by their low molecular content. These stars may be caught in the act of dissipating their disks. Despite its higher $^{13}$CO content, DS Tau, which forms a common proper motion pair with a Class III/wTT star, is another good candidate for a dissipating disk because of the possible existence of an inner cavity. This study reveals a population of small disks whose surface density remains large in the central 10 AU region. This population had been ignored so far because of the limited sensitivity and angular resolution of previous observations. It may represent up to 25 \\% of the whole disk population in Taurus. These disks apparently only differ from their larger siblings by their small sizes. The origin for such small sizes remains unknown, and requires higher resolution observations with ALMA and the JVLA to be unveiled. Such observations would also reveal the ability of these disks to form planets." }, "1402/1402.7317_arXiv.txt": { "abstract": "Recent studies suggest that binary neutron star (NS-NS) mergers robustly produce the heavy $r$-process nuclei above the atomic mass number $A \\sim 130$ because of their ejecta consisting of almost pure neutrons (electron fraction of $Y_\\mathrm{e} < 0.1$). However, little production of the lighter $r$-process nuclei ($A \\approx 90$--120) conflicts with the spectroscopic results of $r$-process-enhanced Galactic halo stars. We present, for the first time, the result of nucleosynthesis calculations based on the fully general-relativistic simulation of a NS-NS merger with approximate neutrino transport. It is found that the bulk of the dynamical ejecta are appreciably shock-heated and neutrino-processed, resulting in a wide range of $Y_\\mathrm{e}$ ($\\approx 0.09$--0.45). The mass-averaged abundance distribution of calculated nucleosynthesis yields is in reasonable agreement with the full-mass range ($A \\approx 90$--240) of the solar $r$-process curve. This implies, if our model is representative of such events, that the dynamical ejecta of NS-NS mergers can be the origin of the Galactic $r$-process nuclei. Our result also shows that the radioactive heating after $\\sim 1$~day from the merging, giving rise to $r$-process-powered transient emission, is dominated by the $\\beta$-decays of several species close to stability with precisely measured half-lives. This implies that the total radioactive heating rate for such an event can be well constrained within about a factor of two if the ejected material has a solar-like $r$-process pattern. ", "introduction": "\\label{sec:intro} The astrophysical site of the $r$-process, the rapid neutron-capture process that makes half the elements heavier than iron, remains a long-standing mystery of nucleosynthesis. Recently, compact binary mergers (CBMs) of double neutron star (NS-NS) and black hole--neutron star (BH-NS) systems have received considerable attention as possible sources of the $r$-process nuclei \\citep{Lattimer1974, Symbalisty1982, Eichler1989, Meyer1989, Freiburghaus1999} according to the following reasons. First, radioactively powered ``kilonova'' emission from the $r$-processed ejecta can be a promising electro-magnetic counterpart to the gravitational-wave signal from a CBM event \\citep{Li1998, Metzger2010, Goriely2011, Kasen2013, Barnes2013, Tanaka2013, Grossman2014}. The possible identification of a kilonova associated with the \\textit{Swift} GRB~130603B \\citep{Berger2013, Tanvir2013} also indicates that CBMs are the progenitors of short-duration gamma-ray bursts and the sources of $r$-process elements \\citep{Hotokezaka2013a, Tanaka2014}. Another reason is that core-collapse supernovae (CCSNe; in particular proto-NS wind), the site that has been believed to be the promising sources of the $r$-process nuclei, are found to provide only marginal conditions for making the elements beyond iron \\citep{Martinez2012, Roberts2012, Fischer2012}. Nucleosynthesis studies with such physical environments confirm that CCSNe produce the elements only up to the atomic mass number $A \\sim 110$ \\citep{Wanajo2011, Wanajo2013}. One possible exception could be the scenario of (still hypothetical) rapidly rotating, strongly magnetized CCSN cores \\citep{Winteler2012}. Recently, \\citet{Goriely2011} and \\citet{Bauswein2013} have explored nucleosynthesis based on the approximate (conformally flat spatial metric) general-relativistic (GR) simulations of NS-NS mergers. They found that the ejecta had extremely low electron fractions ($Y_\\mathrm{e} < 0.1$), which led to fission recycling and thus robust production of only heavy $r$-process nuclei with $A \\gtrsim 130$. Similar results were obtained from the Newtonian simulations of NS-NS and BH-NS mergers by \\citet{Roberts2011, Korobkin2012, Rosswog2014}. Little production of the lighter $r$-process nuclei ($A \\approx 90$--120) conflicts, however, with the recent spectroscopic results of Galactic halo stars \\citep{Sneden2008, Siqueira2014}. That is, the so-called ``universality'' of the (solar-like) $r$-process pattern, first identified for $Z \\gtrsim 56$ ($A \\gtrsim 140$), persists down to $Z \\sim 38$ ($A \\sim 90$) within about a factor of two. There has been no sign of nucleosynthetic events making the nuclei exclusively with $A \\gtrsim 130$. Contribution from, e.g., the subsequent BH accretion-torus wind \\citep{Surman2008, Wanajo2012, Fernandez2013} might cure this problem. In this Letter, we report our first result of nucleosynthesis study based on the full-GR, approximate neutrino transport simulation of a NS-NS merger. The GR effects, being crucial for the dynamical evolutions of merger ejecta as pointed out by \\citet{Hotokezaka2013b}, were not fully taken into account in the previous studies. Moreover, neutrino transport that can affect the ejecta $Y_\\mathrm{e}$ is neglected in all previous studies \\citep[except for the 2D Newtonian simulation by][without nucleosynthesis calculations]{Dessart2009}. Our NS-NS merger model is described in Section~\\ref{sec:model}. The subsequent nucleosynthesis result is presented in Section~\\ref{sec:nucleosynthesis}. The radioactive heating rates (relevant for kilonova emission) are also obtained from the nucleosynthesis calculations (Section~\\ref{sec:heating}). ", "conclusions": "\\label{sec:summary} We examined $r$-process calculations based on the full-GR, approximate neutrino transport simulation of the NS-NS merger with the equal masses ($= 1.3\\, M_\\odot$) of NSs. Different from previous studies, the merger ejecta exhibited a wide range of $Y_\\mathrm{e} \\approx 0.09$--0.45 that led to the nucleosynthetic abundance distribution being in good agreement with the solar $r$-process pattern. Given that the model is representative, our result (with the present estimate of the Galactic event rate) implies that NS-NS mergers can be the major origin of all the $r$-process elements in the Galaxy. Our result also indicates that the radioactive heating (that powers a kilonova transient) after $\\sim 1$~day from the merging is dominated by the $\\beta$-decays of a small number of species with measured half-lives. The total heating rates are thus well approximated by the $\\beta$-decays of the solar $r$-process-like abundances as well as by the approximation of $\\propto t^{-1.3}$. Detailed multi-dimensional information of nucleosynthesis abundances should be, however, taken into account when we consider the spatial dependences of kilonova emission. Our result implies that the previous thought of NS-NS merger events, dynamically ejecting almost pure NS matter, should be reconsidered. The shock-heated and neutrino-processed ejecta from a HMNS are in fact modestly neutron-rich: the phenomenon similar to the early stage of a CCSN (a proto-NS instead of a HMNS). Much more works will be needed to test if similar results are obtained with full-3D nucleosynthetic analyses, with different NS masses and their ratios, with other (reasonable) EOSs, with higher spatial resolution, etc. Nucleosynthetic contributions from BH-NS mergers, as well as from the BH-accretion tori subsequent to NS-NS/BH-NS mergers, should be also explored to draw conclusions on the role of CBMs to the Galactic chemical evolution of the $r$-process nuclei." }, "1402/1402.3759_arXiv.txt": { "abstract": "% High energy emissions from supernovae (SNe), originated from newly formed radioactive species, provide direct evidence of nucleosynthesis at SN explosions. However, observational difficulties in the MeV range have so far allowed the signal detected only from the extremely nearby core-collapse SN 1987A. No solid detection has been reported for thermonuclear SNe Ia, despite the importance of the direct confirmation of the formation of $^{56}$Ni, which is believed to be a key ingredient in their nature as distance indicators. In this paper, we show that the new generation hard X-ray and soft $\\gamma$-ray instruments, on board {\\em Astro-H} and {\\em NuStar}, are capable of detecting the signal, at least at a pace of once in a few years, opening up this new window for studying SN explosion and nucleosynthesis. ", "introduction": "Supernova (SN) explosions trigger (or are triggered by) explosive nucleosynthesis, and they are believed to be main production sites of heavy elements in the Universe. The resulting yields are sensitive to explosion mechanism(s), and thus studying nucleosynthesis products is important to uncover the still-debated explosion mechanism. Especially important is the production of $^{56}$Ni -- this is the origin of Fe (as a result of the radioactive decay chain $^{56}$Ni $\\to$ $^{56}$Co $\\to$ $^{56}$Fe), and the decay is believed to provide a source of emissions from (many classes of) SNe through thermalization of emitted $\\gamma$-rays and positrons. In type Ia supernovae (SNe Ia), about half of an exploding white dwarf in mass is processed into $^{56}$Ni, supporting their huge luminosities as distance indicators. However, the most direct evidence in this scenario is still missing -- there has been no solid detection of the decay $\\gamma$-rays from SNe except for SN 1987A (e.g., Dotani et al., 1987; Sunyaev et al., 1987). Especially, no solid detection has been reported for SNe Ia (see Milne et al., 2004 for a review). From a theoretical point of view, studying this high energy emission has been restricted to one-dimensional models (see Milne et al., 2004, for a review) despite the importance of multi-dimensional structures of the explosion both in theory and observation (e.g., Kasen et al., 2009; Maeda et al., 2010a). Most previous studies also focused on the emission in the MeV range. In this paper, we present our radiation transfer simulations of the high energy emission based on the state-of-the-art SN Ia explosion models. We extend our analysis to hard X-ray and soft $\\gamma$-ray regimes, for which dramatic improvement is expected in the observational sensitivities thanks to new generation observatories like {\\em NuStar} (Koglin et al., 2005) or {\\em Astro-H} (Takahashi et al., 2010). We predict that these telescopes are capable of detecting the radioactive decay signals from SNe Ia, at a rate of once in a year or at least once in a few years. We also briefly comment on perspectives for core-collapse SNe. ", "conclusions": "\\begin{table*} [!ht] % \\begin{center} \\caption{Expected Detectability (for an exposure of $10^6$ s centered at the peak date in each band pass). Shown here are limiting distance and the expected number of SNe Ia within the distance (shown in parenthesis). `cons' and `opt' are conservative and optimistic estimates, respectively. See Maeda et al. (2012) for details.} \\bigskip \\begin{tabular}{lllll} \\hline\\hline & & DD2D\\_asym\\_04 & W7 & DD2D\\_iso\\_04\\\\\\hline & $M$($^{56}$Ni)/$M_{\\odot}$ & 1.02 & 0.64 & 0.42\\\\\\hline Band (keV) & Instrument & Mpc (SNe year$^{-1}$) & &\\\\\\hline 60--80 & HXI & 13.9 (0.43) & 17.7 (0.96) & 10.5 (0.09)\\\\ & NuStar (cons.) & 13.0 (0.43) & 16.5 (0.70) & 9.7 (0.09)\\\\ & NuStar (opt.) & 18.4 (1.13) & 23.3 (2.52) & 13.8 (0.43)\\\\ 158 & SPI & 4.6 ($<$0.09) & 2.9 ($<$0.09) & 2.3 ($<$0.09)\\\\ & SGD (cons.) & 22.2 (2.09) & 14.2 (0.43) & 11.4 (0.09)\\\\ & SGD (opt.) & 38.5 (6.70) & 24.6 (2.96) & 19.7 (1.57)\\\\ 200--460 & SPI & 3.7 ($<$0.09) & 2.7 ($<$0.09) & 2.3 ($<$0.09) \\\\ & SGD (cons.) & 11.6 (0.09) & 8.6 (0.09) & 7.1 (0.09) \\\\ & SGD (opt.) & 20.2 (1.74) & 14.8 (0.43) & 12.3 (0.26) \\\\ 812 & SPI & 4.3 ($<$0.09) & 2.6 ($<$0.09) & 2.0 ($<$0.09)\\\\ & GRIPS & 16.8 (0.87) & 10.0 (0.09) & 7.6 (0.09)\\\\ 847 & SPI & 7.7 (0.09) & 5.4 ($<$0.09) & 4.6 ($<$0.09)\\\\ & GRIPS & 29.8 (4.52) & 21.0 (2.00) & 18.0 (1.04)\\\\\\hline \\end{tabular} \\end{center} \\end{table*} \\begin{figure*}[ht] \\centerline{ \\begin{minipage}[]{0.3\\textwidth} \\resizebox{30mm}{!}{\\includegraphics{maeda_FRAWS_2013_01_fig03a_REVISED.ps}} \\end{minipage} \\begin{minipage}[]{0.3\\textwidth} \\resizebox{30mm}{!}{\\includegraphics{maeda_FRAWS_2013_01_fig03b_REVISED.ps}} \\end{minipage} \\begin{minipage}[]{0.3\\textwidth} \\resizebox{30mm}{!}{\\includegraphics{maeda_FRAWS_2013_01_fig03c_REVISED.ps}} \\end{minipage} } \\caption{Detector response simulations for an exposure of $10^{6}$ seconds for selected models (Tab. 1), for HXI (black) and SGD (red) on board {\\em Astro-H}. The model spectra at 20 days after the explosion are used as input, placed at distances of 15 Mpc. The sensitivity curves are adopted from Kokubun et al. (2010), Tajima et al. (2010), and Takahashi et al. (2010). Note that the photon count is very low in the HXI band for all the models at this distance, thus an apparent detection by HXI (left panel) just comes from the statistical fluctuation. } \\label{fig1} \\end{figure*} According to our simulations of radioactive decay signals from SNe Ia, the new generation hard X-ray and soft $\\gamma$-ray observatories (either {\\em NuStar} or {\\em Astro-H}) are expected to be capable of detecting these signals from SNe Ia up to $\\sim 20$ Mpc with $10^6$ s exposure. This will hopefully lead to nearly annual detections, dramatically changing the field. We thus propose follow-up of nearby SNe Ia by these telescopes in a ToO mode. With a standard set up with a few $10^{5}$ s exposure, the detection will be limited to extremely nearby objects (i.e., up to $\\sim $8 - 12 Mpc for SNe Ia with average brightness), but still there is a good chance of first solid detection of the signal from SNe Ia. Once detected, it will provide various diagnostics on explosive nucleosynthesis and explosion mechanisms, and here a combination of hard X-ray and soft $\\gamma$-ray will be essential. A similar argument applies for core-collapse SNe. By combining results from similar simulations for core-collapse SNe (Maeda, 2006) and those obtained for SNe Ia (Maeda et al., 2012), we find the following. SNe IIp (an explosion of a red supergiant) is not a promising target in the soft bands, since the thick H envelope is still opaque in the early phase where the $^{56}$Ni decay can provide the strong emission in these band passes. Among different types of core-collapse SNe, SNe IIb/Ib/Ic (an explosion of a He or C+O star) are most promising. We estimate that the peak date in the high energy emission will be similar to (or a bit delayed than) that of SNe Ia (i.e., 2 - 3 weeks since the explosion). The smaller amount of $^{56}$Ni here (typically $\\sim 0.1 M_{\\odot}$), taking into account the delayed peak date, results in the expected horizon as $5 - 8$ Mpc (conservative and optimistic). Such nearby objects are rare, but expected to be discovered at a rate of a few in a decade according to the past statistics. \\thanks KM thanks Franco Giovannelli and the organizers of Frascati Workshop 2013 for creating the friendly and stimulating atmosphere. The work by KM has been supported by WPI initiative, MEXT, Japan. The authors acknowledge financial support by Grant-in-Aid for Scientific Research from MEXT (22684012, 23340055, 23740141)." }, "1402/1402.3874_arXiv.txt": { "abstract": "We present the results of our analysis of the RR Lyrae (RRL) variable stars detected in two transition-type dwarf galaxies (dTrans), ESO294-G010 and ESO410-G005 in the Sculptor group, which is known to be one of the closest neighboring galaxy groups to our Local Group. Using deep archival images from the Advanced Camera for Surveys (ACS) onboard the Hubble Space Telescope (HST), we have identified a sample of RR Lyrae candidates in both dTrans galaxies [219 RRab (RR0) and 13 RRc (RR1) variables in ESO294-G010; 225 RRab and 44 RRc stars in ESO410-G005]. The metallicities of the individual RRab stars are calculated via the period-amplitude-[Fe/H] relation derived by Alcock et al. This yields mean metallicities of $\\langle [Fe/H] \\rangle_{ESO294} = -1.77 \\pm 0.03$ and $\\langle [Fe/H] \\rangle_{ESO410} = -1.64 \\pm 0.03$. The RRL metallicity distribution functions (MDFs) are investigated further via simple chemical evolution models; these reveal the relics of the early chemical enrichment processes for these two dTrans galaxies. In the case of both galaxies, the shapes of the RRL MDFs are well-described by pre-enrichment models. This suggests two possible channels for the early chemical evolution for these Sculptor group dTrans galaxies: 1) The ancient stellar populations of our target dwarf galaxies might have formed from the star forming gas which was already enriched through ``prompt initial enrichment'' or an ``initial nucleosynthetic spike'' from the very first massive stars, or 2) this pre-enrichment state might have been achieved by the end products from more evolved systems of their nearest neighbor, NGC 55. We also study the environmental effects of the formation and evolution of our target dTrans galaxies by comparing their properties with those of 79 volume limited ($D_{\\odot} < $2 Mpc) dwarf galaxy samples in terms of the luminosity-metallicity relation and the {{\\sc H~i}\\/} gas content. The presence of these RRL stars strongly supports the idea that although the Sculptor Group galaxies have a considerably different environment from the Local Group (e.g. no giant host galaxies, loosely bound and very low local density), they share a common epoch of early star formation with the dwarf satellite galaxies in the Local Group. ", "introduction": "The chemical evolution of galaxies during their earliest stages provides important insights for the initial physical and environmental conditions of the galaxy formation process. The star formation history (SFH) of galaxies is one of the important factors determining the path of their chemical evolution because the number of stars formed (i.e. the star formation rate) with a given initial mass function (IMF) controls the rate of chemical enrichment (Matteucci 2012). One key observable signature of the SFH of a galaxy is the metallicity distribution function (MDF) of its stars. Thus, we can test possible scenarios for the early epochs of galaxy formation by comparing the observed MDF of the oldest stellar population in the galaxy with several analytical models of chemical evolution. In this context, the Sculptor group dwarf galaxies provide an excellent test bed for investigating the early chemical evolution of dwarf satellite galaxies formed in low density environments (i.e. outskirts of dense galaxy clusters, Local Group-like group environment, or field-like environment; see Section 4.4 of the present study). The Sculptor group, also known as the Sculptor filament, is one of the closest groups of galaxies to our Local Group (i.e. distance to the group center, $d_{gc} \\sim $3.9 Mpc; Karachentsev et al. 2003). Five bright late-type galaxies (NGC 55, 247, 253, 300, and 7793) and at least 16 dwarf satellites form a very loosely bound, field-like system stretched along the line of sight over $\\sim$5 Mpc. Our target galaxies, ESO294-G010 and ESO410-G005, are Phoenix-like transition-type dwarf galaxies (dTrans) that morphologically resemble the gas-poor dwarf spheroidal galaxies (dSphs) but contain detectable amounts of neutral hydrogen (H I) gas (i.e. $\\sim10^5 M_{\\odot}$; Bouchard et al. 2005; Jerjen, H., Freeman, K. C., \\& Binggeli, B. 1998), and show signs of recent star formation activity. Most recently, from their detailed analysis of the color-magnitude diagrams obtained using deep HST/ACS imaging for the five Sculptor group dwarf galaxies, Lianou et al. (2013; hereafter L13) found population gradients in ESO294-G010 and ESO410-G005 in the sense that young blue main sequence stars (age $<$ 100 Myr) and intermediate age asymptotic giant branch stars (age $\\sim$ 1-2 Gyr) are more centrally concentrated, while old horizontal branch stars (age $>$ 10 Gyr) tend to be more spatially extended. The high quality of those HST/ACS images also allowed the first detection of a significant population of RR Lyrae variable star candidates beyond the Local Group. Da Costa et al. (2010) were able to discover numerous RR Lyrae candidates in both dTrans galaxies using a template light curve technique (Layden 1998). The presence of the RR Lyrae stars in ESO294-G010 and ESO410-G005 directly confirms the existence of ancient stellar populations with ages $>$ 10 Gyr and indicates that the Sculptor group dwarf galaxies share a common epoch of the earliest star formation with our Local Group galaxies even though these two groups of galaxies have quite different local density environments. In addition to being a direct probe of ancient stellar populations, RR Lyrae stars provide a variety of utilities to investigate a number of important astrophysical applications; these stars are well-known reliable population II distance indicators. There is a correlation between the metallicities and the pulsation properties (i.e. periods and amplitudes) of the fundamental mode RR Lyrae stars (RR0 or RRab). Using this relation (Alcock et al. 2000; Sarajedini et al. 2006), we can calculate the metallicities of individual RRab stars, and thus construct the MDF for a purely old stellar population. Lastly, the intrinsic colors of RRab stars at their faintest luminosity are largely independent of their other physical properties [$(V-I)_{0,min}=$0.58$\\pm$0.02; Guldenschuh et al. 2005], therefore using their observed colors at minimum light we can also estimate the line-of-sight reddening. The main goal of the present study is to examine the MDFs of legitimate RR Lyrae stars in the two Sculptor group dTrans galaxies, ESO294-G010 and ESO410-G005, in order to investigate the early chemical evolution of these systems, especially in light of the low density environments in which they reside. This paper is organized as follows. Section 2 provides a description of the data set and photometry process. In section 3 we report the results of our analysis including the general trends in the color-magnitude diagrams (CMDs) of the galaxies (Sec 3.1), the details of our RR Lyrae detection method and the pulsation properties of the RR Lyrae candidates (sec 3.2), the calculation of the metallicities of individual RR Lyraes (sec 3.3), and the distance measurements (sec 3.4). In section 4 based on our estimates in the previous sections, we present our in depth discussion on the star formation histories, early chemical evolutions, the luminosity-metallicity relations of the our target dTrans, and the environmental effects on the evolution of the near field dwarf galaxies. Finally in section 5 we present a summary of our results. \\vskip 1cm ", "conclusions": "We have presented an extensive analysis of the RR Lyrae stars in two transition type Sculptor group dwarf galaxies, ESO294-G010 and ESO410-G005. Based on the properties of the RR Lyrae stars we present the following results. \\begin{enumerate} \\item We have detected numerous RR Lyrae candidates in both ESO294-G010 ($N_{ab}=$219) and ESO410-G005 ($N_{ab}=$225). Based on the Bailey diagrams, the characteristics of the RR Lyrae stars in these two Sculptor group dTrans appear to follow the typical trend of the RR Lyrae populations of Local Group dwarf satellite galaxies. \\item We construct the MDFs of the RRab stars using the period-amplitude-[Fe/H] relationship of Alcock et al. (2000). The mean metallicities estimated from the best-fit Gaussians to the MDFs are $<[Fe/H]>=$--1.77 $\\pm$ 0.03 (sem) for ESO294-G010, and $<[Fe/H]>=$-1.64 $\\pm$ 0.03 (sem) for ESO410-G005 respectively. Our results represent the metallicity values for purely ancient stellar populations in both dTrans galaxies. \\item The distance of each dTran was calculated using the absolute V magnitudes of the RRab stars. We find $(m-M)_{0}=$ 26.40 $\\pm$ 0.07 mag for ESO294-G010 and $(m-M)_{0} =$ 26.33 $\\pm$ 0.07 mag for ESO410-G005. We have also calculated the distances of the two dTrans using the I-band magnitudes of TRGBs in order to check the validity of our RR Lyrae distance estimates [$(m-M)_{0,ESO294}=$26.37; $(m-M)_{0,ESO410}=$26.38]. Our distance estimates for ESO294-G010 using two independent methods (RRLs and TRGB) agree very well with each other. In the case of ESO410-G005, there is a $\\sim$ 0.1 mag difference but this is consistent within the margin allowed by the errors. \\item We have compared the RR Lyrae MDFs for ESO294-G010 and ESO410-G005 with several chemical evolution models. For both galaxies, the shapes of the RR Lyrae MDFs are nicely described by pre-enrichment models. This suggests two possible channels for the early chemical evolution of these Sculptor group systems: 1) The ancient stellar populations of our target dwarf galaxies might have formed from the star forming gas which was already enriched through ``prompt initial enrichment'' or an ``initial nucleosynthesis spike'' caused by the very first massive stars, or 2) This pre-enrichment state might have been achieved by the end products from more evolved systems of their nearest neighbor, NGC 55. We also fit a simple accretion model to the RR Lyrae MDFs of each galaxy. The gas infall hypothesis seems to explain the metal-poor tail of the RR Lyrae MDFs indicating that we cannot completely rule out the role of gas accretion on the early chemical evolution of the dTrans. \\item The L-M relation of our target dTrans ESO294-G010 and ESO410-G005 follow the typical trend of other Local Group dTrans galaxies, such as Phoenix, Tucana, Cetus, LGS 3, KKR 25, DDO 210, and Antlia. Within a luminosity range of 5 $< log (L_{K}/L_{\\odot}) <$ 7.5, the L-M relation of the sample dTrans appear to show a closer resemblance to that of the canonical dSphs than dIrrs. This suggests that the bulk of the dTran galaxies can be considered as ``$present-day$ $progenitors$'' of the dSph galaxies if the recent star formation of dTrans galaxies has been largely suppressed by rapid gas loss (G03). \\item Our examination of the {{\\sc H~i}\\/} gas content as a function of the tidal index $\\Theta_{5}$ for 79 volume limited ($D_{\\odot} < $2 Mpc) dwarf galaxies reveals a clear morphological segregation, in the sense that gas-deficient dSphs tend to be located in dense areas while most gas-rich dIrrs (except LMC/SMC pairs) are isolated. Similar to the dIrr galaxies, most dTran galaxies are also found to be isolated but in general they have lower {{\\sc H~i}\\/} gas fractions as compared to the dIrrs. Our analysis supports the idea that the morphology-density relation appears to be a ubiquitous phenomena in cluster environments regardless of their shapes and richness (van der Wel et al. 2010, and references therein), and the environmental effects on galaxy evolution operate in a similar fashion even in such low density environments of the Local Group or its neighboring groups. \\end{enumerate}" }, "1402/1402.6726_arXiv.txt": { "abstract": "Pulsating stars in eclipsing binary systems play an important role in asteroseismology. The combination of their spectroscopic and photometric orbital solutions can be used to determine, or at least to constrain, the masses and radii of components. To successfully perform any seismic modelling of a star, one has to identify at least some of the detected modes, which requires precise time-series photometric and spectroscopic observations. This work presents a progress report on the analysis of two $\\beta$ Cephei-type stars in eclipsing binaries: HD\\,101794 (V916 Cen) and HD\\,167003 (V4386 Sgr). ", "introduction": "HD\\,101794 and HD\\,167003 have been observed at the South African Astronomical Observatory (SAAO) between May 2 and 19, 2009. The $UBVRI$ time-series photometry has been acquired using the UCT CCD detector at the 1.0-m telescope. Spectroscopic observations were carried out using the GIRAFFE {\\'e}chelle spectrograph at the 1.9-m telescope. We obtained spectra in the wavelength range between 4200 {\\AA} and 6900 {\\AA} with a resolution $R \\approx$ 39000. Stellar magnitudes were calculated using the program DAOPHOT II (\\cite[Stetson 1987]{stet87}). Spectra were wavelength-calibrated and extracted using the IRAF software package (\\cite[Tody 1993]{tody93}). The {\\'e}chelle orders were merged and normalised with a program of our own, and the radial velocities were calculated by cross-correlating the observed spectra with non-LTE models of \\cite[Lanz \\&~Hubeny (2007)]{lahu07} using the method of \\cite[Tonry \\&~Davis (1979)]{toda79}. ", "conclusions": "The results for HD\\,167003 are very encouraging. First of all, we have confirmed it is indeed a pulsating star in a binary system. In addition, we have confirmed that the orbital period amounts to 10.79824 d and is not twice as long, as initially suspected by \\cite[Pigulski \\&~Pojma{\\'n}ski (2008)]{pipo08}. This star seems to be a single-lined spectroscopic binary, and our modelling suggests that the orbit is close to circular. \\cite[Pigulski \\&~Pojma{\\'n}ski (2008)]{pipo07} detected only the primary eclipse in the ASAS-3 photometry of HD\\,167003. Our initial hypothesis was that the lack of the secondary eclipse is caused by a highly eccentric orbit. In light of the results of our modelling, we now know this cannot be the case. This suggests that the secondary eclipse is very shallow, and that the contribution of the secondary component to the total flux is small. While our present photometry is more accurate than the ASAS-3 photometry used by the previous investigators, we are also unable to find the secondary minimum in our observations. This could be because the orbital period of HD\\,167003 is quite long, and our phase coverage is incomplete. However, we managed to detect hints of a minute reflection effect. Once the orbital inclination is known from the light curve modelling, we will be able to use the mass function to constrain the masses of components. From the fact that at least two of the photometrically detected modes are also seen in the radial velocity data, it seems probable that the attempts to identify mode degrees from amplitude ratios and phase differences will be successful. Therefore, this star seems to be a very good candidate for asteroseismic analysis." }, "1402/1402.0793_arXiv.txt": { "abstract": "We have evaluated the electron capture rates on $^{20}$Ne, $^{20}$F, $^{24}$Mg, $^{24}$Na and the $\\beta$ decay rates for $^{20}$F and $^{24}$Na at temperature and density conditions relevant for the late-evolution stages of stars with $M=8$--12~M$_\\odot$. The rates are based on recent experimental data and large-scale shell model calculations. We show that the electron capture rates on $^{20}$Ne, $^{24}$Mg and the $^{20}$F, $^{24}$Na $\\beta$-decay rates are based on data in this astrophysical range, except for the capture rate on $^{20}$Ne, which we predict to have a dominating contribution from the second-forbidden transition between the $^{20}$Ne and $^{20}$F ground states in the density range $\\log \\rho Y_e (\\mathrm{g~cm}^{-3}) = 9.3$--9.6. The dominance of a few individual transitions allows us to present the various rates by analytical expressions at the relevant astrophysical conditions. We also derive the screening corrections to the rates. ", "introduction": "Electron captures on nuclei play a crucial role in the high-density environment of late-stage stellar evolution~\\cite{Bethe:1990,Langanke.Martinez-Pinedo:2003} with three important consequences. It reduces the pressure which the degenerate relativistic electron gas can supply against the gravitational contraction of the stellar core. Furthermore it cools the core environment as the neutrinos produced in the capture process can leave the star virtually unhindered (as long as the density is less than about $10^{11}$~g~cm$^{-3}$ and carry away energy. Finally electron captures change protons in the nucleus into neutrons and hence drive the stellar composition more neutron rich. Improving on the pioneering work by Fuller, Fowler and Newman (FFN)~\\cite{Fuller.Fowler.Newman:1980,*Fuller.Fowler.Newman:1982a,*Fuller.Fowler.Newman:1982b,Fuller.Fowler.Newman:1985} and making use of advances in nuclear modeling and in computational hard- and software development, electron capture rates have been determined for sd-shell nuclei ($A=17$--39)~\\cite{Oda.Hino.ea:1994} and for pf-shell nuclei ($A=45$--64) \\cite{Caurier.Langanke.ea:1999,% Langanke.Martinez-Pinedo:2000,Langanke.Martinez-Pinedo:2001} based on large-scale shell-model diagonalization calculations. The reliability of the calculations benefitted also strongly from experimental data for the Gamow-Teller (GT$_+$) distribution in nuclei (e.g.~\\cite{frekers:2006,Fujita.Rubio.Gelletly:2011,Sasano.Perdikakis.ea:2011}) which determine the electron capture rates at the stellar conditions for which nuclei in the mass range $A=17$--64 dominate the stellar matter composition. Indeed a detailed comparison of stellar capture rates derived from experimental GT$_+$ distributions for all $pf$-shell nuclei, for which data exist, with modern shell-model rates convincingly validated the use of the latter in late-stage stellar evolution studies~\\cite{Cole.Anderson.ea:2012} (for applications and consequences see~\\cite{Janka.Langanke.ea:2007}). We note that diagonalization shell-model calculations are yet not globally feasible for the very neutron-rich nuclei with $A>64$ which dominate the electron capture at densities in excess of a few $10^{10}$~g~cm$^{-3}$~\\cite{Juodagalvis.Langanke.ea:2010} and the respective rates must be determined based on other approaches such as the Random Phase Approximation (RPA) with occupation numbers from Shell Monte Carlo~\\cite{Langanke.Martinez-Pinedo.ea:2003}, thermofield dynamics approach~\\cite{Dzhioev.Vdovin.ea:2010}, and finite-temperature Quasiparticle RPA~\\cite{Paar.Colo.ea:2009,Niu.Paar.ea:2011}. While stellar electron captures usually occur on an ensemble of nuclei present in the matter composition, capture on the specific nuclei $^{20}$Ne and $^{24}$Mg has been identified as crucial for the core collapse of 8--12~M$_\\odot$ stars \\cite{Nomoto:1984,*Nomoto:1987,hillebrandt.nomoto.wolff:1984}. Stars in this mass range develop degenerate ONe or ONeMg cores which are driven towards collapse in a process dubbed electron capture supernova triggered by the loss of electron pressure support due to electron captures, mainly on the very abundant nuclear species $^{20}$Ne and $^{24}$Mg \\cite{Nomoto:1984,*Nomoto:1987,hillebrandt.nomoto.wolff:1984,Huedepohl.Mueller.ea:2010}. We note that 8--12~M$_\\odot$ stars crucially contribute to the nucleosynthesis of specific nuclides. Its role for the synthesis of r-process elements is currently controversially discussed~\\cite{Ning.Qian.Meyer:2007,Janka.Mueller.ea:2008}. Simulations of late-stage evolution of 8--12~M$_\\odot$ stars and electron-capture supernovae usually adopt the weak-interaction rates, including those for electron capture on $^{20}$Ne and $^{24}$Mg, from the work of Oda et al. \\cite{Oda.Hino.ea:1994}. These authors made available rate tabulations for an extensive set of nuclei in the mass range $A=17-39$, however, on a rather sparse temperature-density grid which is argued to be insufficient for detailed studies of the evolution stage for which the weak rates are essential~\\cite{Jones.Hirschi.ea:2013}. Shell-model rates for electron captures on $^{20}$Ne and $^{24}$Mg had previously to the work by Oda \\emph{et al.}~\\cite{Oda.Hino.ea:1994} been calculated by Takahara \\emph{et al.}~\\cite{Takahara.Hino.ea:1989}. Importantly the calculations of Takahara \\emph{et al.} and of Oda \\emph{et al.} had been performed before the Gamow-Teller strength distributions for $^{20}$Ne and $^{24}$Mg have been determined by charge-exchange experiments. Due to the isospin symmetry of the two nuclei, this goal could not only be achieved by techniques which determine the GT$_+$ strength distribution using $(d,{}^2\\textrm{He})$ and $(t,{}^3\\mathrm{He})$ reactions, but also by those measuring the GT$_-$ distribution by $(p,n)$ and $(^3\\textrm{He},t)$ reactions. (In the latter a neutron is changed into a proton supplying the information required for $\\beta^-$ decays). The availability of these data calls for a reevaluation of the electron capture rates which we will present in this manuscript. Besides the incorporation of recent experimental GT$_+$ data, we improve the previous rates also in two other important aspects. At first we point to the relevance of the ground-state-to-ground-state transition in the capture on $^{20}$Ne which, although it is of forbidden nature, is likely to dominate the capture rate in the astrophysically relevant temperature-density range for the core evolution of 8-12 M$_\\odot$ stars. Secondly, we correct the capture rates for screening effects in the dense environment (which decrease electron capture rates, but increase the competing $\\beta$-decays). Our study is completed by a reevaluation of the rates for electron captures and $\\beta$ decays of $^{20}$F and $^{24}$Na, which are the daughters of the electron capture processes on $^{20}$Ne and $^{24}$Mg, respectively. We note that the electron capture rate on $^{20}$Ne as well as the $^{20}$F $\\beta$-decay rate is dominated by a few transitions which are experimentally determined, except for the forbidden ground-state-to-ground-state transition for which only an upper limit is known. The dominance of a few transitions allows us to present the rates by an analytical expression for the relevant temperature-density region which removes uncertainties associated with extrapolations required for rate tabulations provided on a grid. ", "conclusions": "We have calculated the rates for electron captures on $^{20}$Ne, $^{20}$F, $^{24}$Mg, $^{24}$Na and $\\beta$ decays of $^{20}$F and $^{24}$Na which are key quantities for studies of the late-time evolution of 8--12~M$_\\odot$ stars. So far such late-time studies are based on the rate evaluations of~\\cite{Takahara.Hino.ea:1989} and~\\cite{Oda.Hino.ea:1994}. We have improved these rates in three important aspects. First, we have incorportated experimental data from either $\\beta$ decay or charge-exchange experiments which have not been available at the time when Takahara \\emph{et al.}~\\cite{Takahara.Hino.ea:1989} and Oda \\emph{et al.}~\\cite{Oda.Hino.ea:1994} did their work. In our study the recent experimental data are supplemented by Gamow-Teller transitions derived from large-scale shell model calculations, similar to the procedure in~\\cite{Takahara.Hino.ea:1989,Oda.Hino.ea:1994}. Importantly we find that nuclear physics input into the astrophysically relevant rates for electron captures on $^{20}$Ne and $^{24}$Mg and the competing beta decays of the respective daughters is completely based on experimental data. The exception is the electron capture on $^{20}$Ne in the density regime $\\log \\rho Y_e = 9.3$--9.7. As our second improvement we point out that at temperatures $T < 0.7 \\times 10^9$ the capture rate is likely to be dominated by the second-forbidden transition from the $^{20}$Ne ground state to the $^{20}$F. Experimentally only an upper limit exists, which, however, is of the order of typical second-forbidden transition strengths. While we have used the upper limit as an estimate for this transition in our present work a calculation of the transition with an appropriate method like the shell model or an experimental determination is highly desirable. As the third improvement, we have corrected the various rates for medium-induced effects. Here we followed the formalism discussed in ref.~\\cite{Juodagalvis.Langanke.ea:2010} for electron captures and extended it to the treatment for $\\beta$-decays. The environment reduces the electron chemical potential and enhances (reduces) the reaction $Q$-value for electron captures ($\\beta$ decays). As a consequence electron capture rates are lower in dense astrophysical environments than for bare nuclei, while $\\beta$ decay rates are larger. For the astrophysical conditions here, the medium corrections change the rates typically by a factor of order two. The effect is, of course, significantly larger at such densities where the rates change from dominance of a certain transition to another (as is the case in the weak processes here) as these transitions are extremely sensitive to the effective $Q$ values. We note that medium effects should also have a significant effect on the densities at which so-called URCA pairs operate and influence the late-stage evolution of the stars. As $\\beta$ decay rates are enhanced and the competing electron capture rates are lowered, the medium modifiactions will move the operation of the URCA pairs to somewhat higher densities. As the shifts of the electron chemical potential and of the $Q$ values are of order 100~keV under the relevant density (and temperature) conditions encountered, we expect that the URCA pairs operate at densities which are about $0.1 \\times 10^9$~g~cm$^{-3}$ larger than found in calculations which do not consider medium corrections on the rate. Stellar evolution studies which investigate the impact of screening on the URCA pairs are needed. We have presented analytical expressions for both electron capture and beta-decay rates that allow for an accurate description of these processes for conditions at which URCA pairs operate in both intermediate mass stars~\\cite{Tsuruta.Cameron:1970} and neutron stars~\\cite{Schatz.Gupta.ea:2014}. Rate tables on fine grids in temperature and density in the ranges $\\rho Y_e = 10^8 - 10^{10}$~g~cm$^{-3}$ and $T=10^8 - 10^{10}$~K can be obtained by request from the authors." }, "1402/1402.4045_arXiv.txt": { "abstract": " ", "introduction": "Scalar field models have played a vital role in cosmological theoretical studies in nearly half a century. Those assumed scalar fields appeared in different cosmological research aspects to settle different cosmological problems \\cite{1}, such as to drive inflation, to explain a time variable cosmological $'$constant$'$ and so on. The scalar fields have played another important essential role in the past nearly fifteen years as a candidate of dark energy after the discovery of the accelerating expansion of universe. There are so many phenomenological dark energy models of scalar fields, such as quintessence, phantom, quintom and the scalar fields with non-canonical kinetic energy term (for a review, see\\cite{2,3}). \\par To study the dynamical evolution of those scalar fields models and their cosmological implications with a phase-plane analysis is a very useful and common method(see Ref\\cite{add2} and recent papers, e.g. \\cite{Saridakis1, Saridakis2, Saridakis3, Saridakis4}). However, most of those works only focus on the quintessence models(including phantom quintessence and quintom) with unique exponential potential and tachyon models(including phantom tachyon) with inverse square potential, and correspondingly, the dynamical systems are two dimensional autonomous system with those special form of potentials(see the references cited in \\cite{6, 12}). Using a method which considers the potential related variable $\\Gamma$ as a function of another potential related variable $\\lambda$ (see Eq.(\\ref{eqsadd4}) for the definition of $\\Gamma$ and $\\lambda$)\\cite{26, 12, 6}, we are able to analyze the phase-plane of the dynamical systems of the quintessence and tachyon models with many different potentials. When the potentials are beyond the special type such as exponential or inverse square potentials, the dynamical systems consequently become a three-dimensional autonomous systems. This method is quite effective and powerful, it therefore has been generalized to several other cosmological contexts\\cite{cite61, cite62, cite63, cite65, cite66, cite67, cite68, cite69, 25, cite611}. However, there is very few work focusing on the dynamical behavior of the scalar field with a general modified kinetic term, such as k-essence ($L=V(\\phi)F(X)$) and general non-canonical scalar field ($L=F(X)-V(\\phi)$). Recently, Josue De-Santiago et. al analyzed the dynamical system of general non-canonical scalar field with the lagrangian $L=F(X)-V(\\phi)$ and studied the phase plane after a suitable choice of variables\\cite{20}. They obtained the three-dimensional autonomous system of this non-canonical scalar field after specifying the kinetic term as $F(X)=AX^{\\eta}$ and choosing the potential as $V(\\phi)=V_0(\\phi-\\phi_0)^{1/(1-\\Gamma)}$(i.e., the special case that potential related parameter $\\Gamma$ is a constant) and studied the critical points as well as their stability. \\par Motivated by the work, in this paper we try to extend our works in\\cite{6, 12} to give the three-dimensional autonomous dynamical systems for most of the popular scalar field dark energy models including (phantom) quintessence, (phantom) tachyon, k-essence and general non-canonical scalar field models. We will show that the three-dimensional autonomous systems of general non-canonical scalar field and k-essence will reduce to the quintessence and tachyon scalar field respectively. Not like the previous works, here we express the three dimensional autonomous systems from the trivial variables $(x, y, \\lambda)$ to the observable related variables $(w_{\\phi}, \\Omega_{\\phi}, \\lambda)$. It will be very convenient to investigate the dynamical properties of the autonomous system based on the observable related variables $w_{\\phi}$ and $\\Omega_{\\phi}$(see \\cite{7, 8, 9, 10, 12, 13} and a recent paper about the general property of dynamical quintessence field\\cite{add1}). Since the definition of the trivial variables $x$ and $y$ could vary with different scalar field models, while the observable related variables such as the equation of state $w_{\\phi}$ and the dark energy density parameter $\\Omega_{\\phi}$ are the same for different dark energy models. The paper is organized as follows. We firstly present the basic theoretical framework for (phantom) quintessence, (phantom) tachyon, k-essence and general non-canonical scalar field models in section 2, and try to give the relationships between those different scalar fields in this section. We then give the three dimensional autonomous dynamical systems for those scalar fields and switch the dynamical variables from $(x, y)$ to $(w_{\\phi}, \\Omega_{\\phi})$ in each subsection of section 3. Additionally, using the dynamical systems, we give the exact solution of $w_{\\phi}$ and $\\Omega_{\\phi}$ for a special case of tachyon model when the potential is chosen to be a constant in subsection 3.2. We show that the dynamical autonomous system of k-essence can reduce to tachyon model, and investigate another special case called kinetically driven quintessence with the lagrangian $p(X, \\phi)=f(\\phi)(-X+X^2)$ detailedly in subsection 3.3. In subsection 3.4, we show that the dynamical autonomous system of general non-canonical scalar field can reduce to quintessence and tachyon model respectively for some special cases. We also studied two special cases of purely kinetic united model $L = F(X)$ in detailed in this subsection. We try to give the cosmological implications of the three-dimensional dynamical autonomous system and present the conclusion in section 4. We proved in section 4 that the dark energy density parameter $\\Omega_{\\phi}$ would obey the same differential equation for all the non-coupled dark energy model under the GR frame. We also raise a question about the possibility of the chaotic behavior in the spatially flat single scalar field FRW cosmological models in the presence of ordinary matter. ", "conclusions": "\\par The main purpose of the paper is not to analyze the dynamical behavior about different scalar fields in detail, so we will not investigate the detailed critical points and their stable properties for each dynamical system. What we want to focus on is about the dynamical system itself. \\par First thing we want to emphasize is that, the dynamical variables ($x, y$) in the previous papers are about the scalar field and its first derivative. these dynamical variables have no direct cosmological meaning, and the autonomous system for $dx/dN$ and $dy/dN$ also varies from different models. However, it would be more convenient if we change the variables from $(x, y)$ to observable quantities $(\\gamma_{\\phi}, \\Omega_{\\phi})$. Firstly, $(\\gamma_{\\phi}, \\Omega_{\\phi})$ are directly related to the observable quantities and also about the the properties of dark energy. Analyzing the system based on $(\\gamma_{\\phi}, \\Omega_{\\phi})$, we can figure out how the equation of state of dark energy $w_{\\phi}$ and the density parameter $\\Omega_{\\phi}$ evolve. Secondly, though the form of autonomous system for $dx/dN$ and $dy/dN$ are completely different for different models, but we found that the function for $d\\Omega_{\\phi}/dN$ has the same expression in quintessence, tachyon, k-essence and general non-canonical scalar field model(i.e., Eq.(\\ref{eqs14}),Eq.(\\ref{eqs23}),Eq.(\\ref{eqs35}),Eq.(\\ref{eqs44})) as follows: \\begin{equation}\\label{eqs47}\\frac{d \\Omega_{\\phi}}{dN}=3(\\gamma_b-\\gamma_{\\phi})\\Omega_{\\phi}(1-\\Omega_{\\phi})\\end{equation} The only difference are the form of function $d\\gamma_{\\phi}/dN$. In fact, we can prove that Eq.(\\ref{eqs47}) holds for all non-coupled dark energy models as long as they satisfy the following equations: \\begin{equation}\\label{eqs48} H^2=\\frac{1}{3M^2_{pl}}(\\rho_b+\\rho_{de}), ~ \\dot{\\rho}_i+3H(p_i+\\rho_i)=0\\end{equation} where subscript $i$ denotes each energy component such as dark energy, matter or radiation. If we set $\\gamma_{de}=w_{de}+1=p_{de}/\\rho_{de}+1$ and $\\Omega_{de}=\\rho_{de}/3M^2_{pl}H^2$, we can obtain following equation from Eqs.(\\ref{eqs48}): \\begin{equation}\\label{eqs49}\\frac{d \\Omega_{de}}{dN}=3(\\gamma_b-\\gamma_{de})\\Omega_{de}(1-\\Omega_{de})\\end{equation} Eq.(\\ref{eqs49}) is the same with Eq.(\\ref{eqs47}), and also the same with Eq.(\\ref{eqs14}), Eq.(\\ref{eqs23}), Eq.(\\ref{eqs35}) and Eq.(\\ref{eqs44}). So Eq.(\\ref{eqs49}) is a very common form for the evolution of $\\Omega_{de}$ of various different models. For example, the authors got the same equation even for the purely kinetic coupled gravity model which modified the standard general relativity action through the addition of a coupling between functions of the metric and kinetic terms of a free scalar field\\cite{24}(Eq.(28) in this paper and $\\gamma_b$ is taken as 1). \\par Eq.(\\ref{eqs49}) holds for the different scalar field models mentioned in this paper and even all the non-coupled dark energy models which satisfy Eq.(\\ref{eqs48}), so we can conclude that there are only three possible destinies(three types of critical points) for $\\Omega_{de}$ to be in these models. The value of $\\Omega_{de}$ will be determined by the stable properties of these three types of critical points, and the stable properties depend on different models and the value of parameters in different models. Two of them are the cases that $\\Omega_{de}=0$ or $\\Omega_{de}=1$ which are completely opposite destinies. These two solutions correspond to the universe completely dominated by the scalar field or by the barotropic fluid. The last destiny corresponds to the case of $\\gamma_{de}=\\gamma_b$, and the value of $\\Omega_{de}$ will vary from different models and be determined by other equations in the dynamical system. Generally we can obtain $0\\leq\\Omega_{de}\\leq1$ and so this solution is called the scaling solution. However, for the scaling solution, the equation of state of dark energy $w_{de}$ is the same as the equation of state of barotropic fluid $w_b$, so there is no accelerating expansion. Since the observation suggested that we are living in a accelerated expanding universe with $\\Omega_{de}\\sim0.7$, none of these three destinies correspond to the present universe we observed. If we try to construct a successful theoretical model which has a dynamical stable solution corresponding to present observational universe(namely a scaling attractor with $\\Omega_{de}\\sim 0.70$ and $\\gamma_{de}\\sim 0.1$ in agreement with observations) to solve or at least alleviate the cosmological coincidence problem without fine-tunings, we must consider the interaction between dark energy and other barotropic fluids(see \\cite{25} for such model). This result is valid for not only all the non- coupled dark energy models, but also for many modified gravity models as long as the energy density and the pressure of dark energy or effective dark energy satisfy the continuity equation Eq.(\\ref{eqs48}). \\par Another important thing we want to emphasize is that, it is more reasonable and more scientific to investigate the dynamical behaviors of a dark energy model under the three-dimensional autonomous system rather than the two-dimensional system. Firstly, the two-dimensional dynamical autonomous system is just a specific case when the potential takes a special form. if we want to completely study the general dynamical properties of a dark energy model, we need to study the system beyond a special potential. Then we can find more critical points than the ones found in a two-dimensional system. We therefore are able to analyze which critical points are possessed by a class of dark energy models and which ones exist only due to the concrete potentials. The method studying the three-dimensional dynamical autonomous system beyond one special potential is originated for the quintessence \\cite{6, 26} and then developed to other dark energy models \\cite{cite61, cite62, cite63, cite65, cite66, cite67, cite68, cite69, 25, cite611}. Here we extend this method to the more general scalar field models in Section 3. Secondly, from the viewpoint of chaos theory, the dynamical properties of three-dimensional autonomous system is more fruitful than the two-dimensional system. According to the Poincar$\\acute{e}$ -Bendixson theorem, chaos does not exist in any two-dimensional autonomous dynamical system\\cite{27, 28} but could be possible in three-dimensional autonomous dynamical system. For a number of three-dimensional systems, such as the famous three-dimensional Lorenz equations which is a model describing the atmospheric convection \\cite{29}, there exist chaos for certain values of the parameters. \\par The studies of chaotic dynamics in cosmological models has a long story. Chaotic properties had reported in spatially closed scalar field FRW cosmological models \\cite{30, 31, 32, 33, 34, 35}, spatially flat FRW cosmological model with two or more scalar fields \\cite{36,37}, Bianchi IX universe \\cite{38, 39}, Bianchi I universe\\cite{40} and the mixmaster universe\\cite{41}. It would be very interesting and also a big challenge for the theoretical study of dark energy if the dynamical systems we consider here(i.e., Eqs.(\\ref{eqs14}-\\ref{eqs16}),Eqs.(\\ref{eqs23}-\\ref{eqs25}),Eqs.(\\ref{eqs35}-\\ref{eqs37}) and Eqs.(\\ref{eqs44}-\\ref{eqs46}) ) exist the chaotic properties. Then the evolution of $\\Omega_{\\phi}$ and $\\gamma_{\\phi}$ will be very sensitive to the initial condition, and therefore predicting their evolution in future becomes totally impossible. However, it is proved that there is no chaotic behavior in spatially flat single scalar field FRW cosmological models\\cite{42, 43}. Since for the spatially flat case with $k=0$, the dynamical system can be described by a three-dimensional autonomous system with a set of variables $(H, \\phi, \\dot\\phi)$ under a Hamiltonian constraint, so the dynamical system is actually a two-dimensional autonomous system($a$ and $\\dot a$ appear only in the combination $H=\\dot a/ a$)\\cite{44}. We know that for the two-dimensional autonomous systems, there are no enough degrees of freedom to exist chaos, so this proved no-chaotic dynamics in the spatially flat scalar field FRW cosmological model. However, we noted that this result is obtained in the absence of matter and radiation. This may be the case in the very early time when our universe is undergoing an inflation era and completely dominated by the scalar field. However, for the study of dark energy of late-time cosmic acceleration, the component of matter is comparable with the density of dark energy and should not be ignored when we investigate the dynamical behavior of scalar field. In the presence of matter, scale factor $a$ will reappear in the dynamical system beside the variables $(H, \\phi, \\dot\\phi)$, and then the system can not be reduced to two-dimensional autonomous dynamical system any more. So here we argue that it is still possible for the chaotic behavior in spatially flat single scalar field FRW cosmological models in the presence of matter. It is very like the case of spatially non-flat($k\\neq 0$) single scalar field FRW cosmological models where the dynamical system can not be reduced to two-dimensional autonomous dynamical system too. What we argued here is also supported by the equations in section 3( i.e., Eqs.(\\ref{eqs14}-\\ref{eqs16}), Eqs.(\\ref{eqs23}-\\ref{eqs25}), Eqs.(\\ref{eqs35}-\\ref{eqs37}) and Eqs.(\\ref{eqs44}-\\ref{eqs46})), which described three-dimensional autonomous dynamical systems. However, we are not sure whether there truly exist the chaotic behavior in spatially flat scalar field FRW cosmological models now, to find the chaotic behavior is beyond the scope of this paper, we will investigate it in future." }, "1402/1402.3948_arXiv.txt": { "abstract": "The Galactic microquasar GRS 1915+105 exhibits various types of light curves. There is, however, no understanding of when a certain type of light curve will be exhibited and only in a handful of cases, the transitions from one type to another have actually been observed. We study the detailed spectral properties in these cases to show that that different classes have different ratio of the power-law photon and the blackbody photon. Since the power-law photons are from the Compton cloud, and the intensity of the power-law photon component depends on the degree of interception of the soft photons by the Compton cloud, we conclude that not only the accretion rate, but the accretion flow geometry must also change during a class transition. ", "introduction": "Indian X-ray Astronomy Experiments (IXAE) payload on board IRS-P3 satellite data of the enigmatic stellar mass black hole binary GRS 1915+105 (Harlaftis \\& Greiner, 2004) clearly indicated that GRS 1915+105 can have various types of light curves (Yadav et al., 1999; Rao, Yadav \\& Paul, 2000; Chakrabarti \\& Nandi, 2000 and references therein; Naik et al. 2002a). So far, a total of thirteen types of light curves have been detected. However, there are only a handful of cases when an actual transition from one type to another has been reported (Chakrabarti et al., 2004; Chakrabarti et al. 2005, hereafter Paper I and Paper II respectively), and those too using the IXAE data which had only two energy channels. It was possible to see the transitions because the IXAE payload observed the same object, namely, GRS 1915+105, for several orbits. Specifically, Papers I and II showed that in a matter of two to three hours indicating that it is not the viscous time scale of the Keplerian disk, but the free-fall time scale of the low-angular momentum halo which decides the variability class transition. In the present paper, we study the spectra of the classes which showed the transitions and found that one of the parameters which could be responsible for the class transition is the geometry of the Compton cloud. There are various models of this cloud, including a Keplerian disk with corona (Haardt \\& Maraschi, 1993) and the unstable inner edge of the standard disk (Kobayashi et al. 2003). However, we shall use a physically reasonable solution of the two component advective flow (TCAF) model (Chakrabarti \\& Titarchuk, 1995, hereafter CT95), where, the Compton cloud is made up of the post-shock region of the sub-Keplerian component which includes the jet and the outflows as well. In the literature many works are present which discusses the quasi-periodic oscillations of black hole candidates (e.g., Strohmayer, 2001; Wagoner, Silbergleit, Ortega-Rodr\u00edguez, 2001; Stella \\& Vietri, 1999; Abramowicz, Kluzniak, 2001; Stuchl\u00edk, Slan\u00fd, T\u00f6r\u00f6k, 2007; Kato, 2008; Blaes, Sramkova, Abramowicz, Kluzniak \\& Torkelson, 2007). However, we shall concentrate only on the spectral properties of GRS 1915+105 in the current paper. The plan of this {\\it letter} is the following: we briefly present the data analysis procedure in our next section. In \\S 3, we study the nature of the components of the spectra, especially the black body and the power-law components. We compute the efficiency of Comptonization by taking the ratio of the photon numbers in those components. We show that indeed, the average ratio varies from class to class. Finally, in \\S 4, we discuss the physics behind such changes in relation to the accretion flow dynamics. ", "conclusions": "In this paper, we have analysed the classes of GRS 1915+105 which have demonstrated class transitions. Purely in a model independent way, we computed the ratios of the power-law photons and the soft photons coming from the Keplerian disk in all these classes. We ignore very high energy photons from HXETE since its contribution is less than a percent and excluding this does not change our conclusion. Since the power-law photons are are believed to be formed due to inverse Comptonization of the injected soft photons, the only way the ratio can change is to change the geometry of the Compton cloud. In CT95 model, the puffed-up part of the low angular momentum flow (CENBOL) remains big in harder (burst-off) states and collapses in softer (burst-on) states. Thus we naturally see that the average Comptonizing efficiency in burst-off states is higher than that in burst-on states. Not only that, all the observed transitions are found to reside side-by-side in the class vs. CE plane. In other words, for a transition, the average CE changes smoothly. This implies that the geometry of the Compton cloud also varies from class to class. The geometry of a Compton cloud can depend on several physical conditions. In CT95, the CENBOL collapses when the accretion rate of the Keplerian disk is increased. This in softer state, CE is low. When the disk rate is very low, the CENBOL cannot be cooled down. It remains big, as in an ion-pressure supported torus (Rees et al. 1982), and thus the interception of soft photons could be large, thereby increasing CE. We believe that this interpretation can be used even for classes not discussed in this paper. This work is being done and the results would be presented shortly. Recently, Remillard and McClintock. (2006) presented evidances that perhaps GRS 1915+105 is an extremely rotating Kerr black hole. However, results discussed in this paper are directly the analysis of the observational data and the interpretation of variation of the Comptonizing efficiency is totally model independent. In other words, our conclusion of the variation of the geometry from one class to another remains valid even when the spin of the black hole is extreme. The work of P. S. Pal is supported by CSIR fellowship." }, "1402/1402.1456_arXiv.txt": { "abstract": "Far-infrared and submillimeter wavelength surveys have now established the important role of dusty, star-forming galaxies (DSFGs) in the assembly of stellar mass and the evolution of massive galaxies in the Universe. The brightest of these galaxies have infrared luminosities in excess of 10$^{13}$ L$_{\\odot}$ with implied star-formation rates of thousands of solar masses per year. They represent the most intense starbursts in the Universe, yet many are completely optically obscured. Their easy detection at submm wavelengths is due to dust heated by ultraviolet radiation of newly forming stars. When summed up, all of the dusty, star-forming galaxies in the Universe produce an infrared radiation field that has an equal energy density as the direct starlight emission from all galaxies visible at ultraviolet and optical wavelengths. The bulk of this infrared extragalactic background light emanates from galaxies as diverse as gas-rich disks to mergers of intense starbursting galaxies. Major advances in far-infrared instrumentation in recent years, both space-based and ground-based, has led to the detection of nearly a million DSFGs, yet our understanding of the underlying astrophysics that govern the start and end of the dusty starburst phase is still in nascent stage. This review is aimed at summarizing the current status of DSFG studies, focusing especially on the detailed characterization of the best-understood subset (submillimeter galaxies, who were summarized in the last review of this field over a decade ago, \\citealt{blain02a}), but also the selection and characterization of more recently discovered DSFG populations. We review DSFG population statistics, their physical properties including dust, gas and stellar contents, their environments, and current theoretical models related to the formation and evolution of these galaxies. ", "introduction": " ", "conclusions": "" }, "1402/1402.3723_arXiv.txt": { "abstract": "The Survey of \\HI\\ in Extremely Low-mass Dwarf galaxies (SHIELD) is an on-going multi-wavelength program to characterize the gas, star formation, and evolution in gas-rich, very low-mass galaxies. The galaxies were selected from the first $\\sim10$\\% of the \\HI\\ ALFALFA survey based on their inferred low \\HI\\ mass and low baryonic mass, and all systems have recent star formation. Thus, the SHIELD sample probes the faint end of the galaxy luminosity function for star-forming galaxies. Here, we measure the distances to the 12 SHIELD galaxies to be between $5-12$ Mpc by applying the tip of the red giant method to the resolved stellar populations imaged by the Hubble Space Telescope. Based on these distances, the \\HI\\ masses in the sample range from $4\\times10^6$ to $6\\times10^7$ \\msun, with a median \\HI\\ mass of $1\\times 10^7$ \\msun. The TRGB distances are up to 73\\% farther than flow-model estimates in the ALFALFA catalog. Because of the relatively large uncertainties of flow model distances, we are biased towards selecting galaxies from the ALFALFA catalog where the flow model underestimates the true distances. The measured distances allow for an assessment of the native environments around the sample members. Five of the galaxies are part of the NGC~672 and NGC~784 groups, which together constitute a single structure. One galaxy is part of a larger linear ensemble of 9 systems that stretches 1.6 Mpc from end to end. Two galaxies reside in regions with $1-4$ neighbors, and four galaxies are truly isolated with no known system identified within a radius of 1 Mpc. ", "introduction": "} Historically, optical surveys have cataloged the majority of dwarf galaxies in the nearby universe. However, optical identification of systems populating the extremely low-mass end of the galaxy luminosity function are hampered by the intrinsic faintness and small angular size of such systems. These low luminosity systems dominate the local population of galaxies and are important tracers of the distribution of mass. Outside of a dense group or cluster environment where low mass galaxies do not repeatedly experience ram pressure and/or tidal stripping of their ISM, populations of such extremely low mass dwarfs are expected to be predominantly gas-rich. Thus, large \\HI\\ surveys offer an opportunity to identify and catalogue such systems in a statistically complete volume. The Arecibo Legacy Fast ALFA (ALFALFA) survey \\citep{Giovanelli2005, Haynes2011} is a blind extragalactic \\HI\\ survey to map the nearby \\HI\\ universe over 7000 sq. degrees of high Galactic latitude sky. With an \\HI\\ mass detection limit as low as $10^6$ \\msun\\ for galaxies in the local universe and $10^{9.5}$ \\msun\\ at the survey velocity limit of $z\\sim0.06$, the ALFALFA survey was designed to populate the faint end of the \\HI\\ mass function over a cosmologically significant volume. The low mass end of the \\HI\\ mass function has been sampled by previous studies including the wide area \\HI\\ Parkes All-Sky Survey \\citep[HIPASS;][]{Barnes2001, Meyer2004, Wong2006}. However, ALFALFA has higher angular and spectral resolution and higher sensitivity. Thus, the ALFALFA catalogue provides a representative sampling of the \\HI\\ mass function to lower masses in a much larger volume. The Survey of \\HI\\ in Extremely Low-mass Dwarf systems \\citep[SHIELD;][]{Cannon2011} was designed to probe the early release of the ALFALFA dataset in a systematic investigation of nearby galaxies with HI masses $\\ltsimeq 10^7$ \\msun\\ outside the Local Group. From the first $\\sim10$\\% of the processed ALFALFA survey data, 12 systems were selected that had gas mass estimates between $10^6-10^7$ \\msun\\ based on \\HI\\ line widths and flow-model distances, and overall low masses based on an \\HI\\ full width at half maximum $< 65$ km s$^{-1}$ which discriminated against gas-poor massive galaxies. The high positional accuracy of the ALFALFA survey (i.e., better than 20$\\arcsec$) facilitated the identification of optical counterparts in SDSS imaging, revealing young, blue stellar populations in each galaxy. These low mass systems have retained \\HI\\ mass reservoirs of $10^6-10^7$ \\msun\\ over a Hubble time, apparently evolving in relative quiescence, and all have recent star formation activity, consistent with all previous observations of dwarf irregular galaxies \\citep[e.g.,][]{Hunter1982}. Once the presence of stellar components was confirmed with SDSS imaging, follow-up observations became possible in order to obtain accurate distance measurements and to probe the evolutionary histories of the galaxies. Accurate distances to these very low mass galaxies are important for a number of reasons. First, accurate distances to galaxies reduce uncertainties in the \\HI\\ mass function made possible by the ALFALFA survey. Second, distance measurements enable distance dependent analyses of the individual systems to be placed on an absolute scale. For example, the distance measurement of 1.72$^{+0.14}_{-0.40}$ to a newly discovered, extremely metal-deficient galaxy from the ALFALFA survey, Leo~P, allows interpretation of the star formation and evolutionary processes to be placed on solid ground \\citep{Giovanelli2013, Rhode2013, Skillman2013, McQuinn2013}. Third, combining \\HI\\ velocity measurements and accurate distances allows an assessment of environment. Most dwarf galaxies are thought to exist in some form of association with other galaxies \\citep{Tully2006}, with truly isolated galaxies the exception even in low density environments. The Nearby Catalog of Galaxies \\citep{Karachentsev2013} compiles measurements of 869 galaxies within 11 Mpc, including calculations of tidal indices that parameterize the likelihood of neighboring galaxies being kinematically bound in groups. Clearly, such work depends critically on accurate distance measurements. Finally, \\HI\\ velocity measurements and accurate distances enable low mass galaxies to be used to help trace the distribution of matter in larger low-density galaxy volumes. Significant effort has been made to map the distribution of galaxies locally. The very local Hubble flow ($0.7$ Mpc $<$ D$_{MW} < 3.0$ Mpc) has been mapped with increasing accuracy, measuring the expansion of the universe outside the boundaries of the Local Group \\citep{Karachentsev2009}. The Extragalactic Distance Database \\citep{Tully2009}, the Cosmicflow program \\citep{Courtois2011a, Courtois2011b, Tully2012}, and the Cosmicflows$-$2 program \\citep{Tully2013} are working to expand this analysis out to larger distances. The population of low mass galaxies with accurate distances from the richly populated ALFALFA data set can make an important contribution to mapping the mass in the Local Volume (LV; defined by an approximate radius of 11 Mpc). Here, we present the distance measurements to the 12 SHIELD galaxies based on Hubble Space Telescope (HST) optical imaging and the tip of the red giant branch (TRGB) standard candle method \\citep[e.g.,][]{Mould1986}. The paper is organized as follows. \\S2 describes the sample and summarizes the observations and data processing. \\S3 describes the distance determination methods and results. \\S4 maps the native environment around the SHIELD galaxies including all known galaxies within a radius of 1 Mpc of each system. \\S5 summarizes our conclusions. Future work on the SHIELD sample will include an investigation of their star formation histories (McQuinn et~al., in preparation), metallicity measurements based on nebular abundance analyses (Haurberg et~al., in preparation), and a study of the gas kinematics from VLA observations (Cannon et~al., in preparation). ", "conclusions": "} We have used optical imaging of resolved stellar populations obtained from the $HST$ to measure the TRGB distance to 12 low-mass, gas-rich galaxies in the SHIELD program. The distances to the galaxies range from $5 - 12$ Mpc. The CMDs show a population of young main sequence stars indicating these systems have experienced recent star formation. A subset of the sample also shows red and blue helium burning sequences suggesting the recent star formation has been on-going for $\\gtsimeq$ 100 Myr. The star formation histories will be investigated in detail in a future paper (McQuinn et al. in prep). Five galaxies are part of two galaxy groups, NGC~784 and NGC~672, which are likely a single, bound kinematic structure. AGC~174585 is located $\\sim1.0$ Mpc away from the previously identified association of galaxies, $14+19$, but is likely unbound to this structure. Similarly, AGC~731457 lies 0.9 Mpc from another system, DDO~83. AGC~749241 is situated in a more populated volume, forming part of 1.6 Mpc linear structure with six other galaxies. Similarly to the NGC~672 and NGC~784 groups, these galaxies are likely bound, lying along a dark matter filament. The nearest neighbor identified is $\\sim0.5$ Mpc away. Four galaxies, AGC~174605, AGC~182595, AGC~748778, and AGC~749237 appear to be truly isolated with no known neighbors identified within a radius of 1 Mpc, although three of these galaxies are at relatively large distances (i.e., $\\gtsimeq10$ Mpc) where our knowledge of the galaxy population is incomplete. Based on the distances and the \\HI\\ measurements from the ALFALFA survey, the SHIELD galaxies have \\HI\\ masses ranging from $4\\times10^6$ to $6\\times10^7$ \\msun, with a median \\HI\\ mass of $1\\times 10^7$ \\msun\\ and 5 galaxies with an \\HI\\ mass below $10^7$ \\msun. Thus, the SHIELD program probes the low mass regime of gas-rich galaxies in a larger cosmological volume than previous surveys. As these 12 galaxies were selected from the first $\\sim10$\\% of the reduced ALFALFA database, it suggests that a large sample of low luminosity galaxies in this mass range exists in the nearby universe. Nominally, one might expect the total yield from the ALFALFA survey to be 10$\\times$ greater. However, investigation of the first 40\\% of the ALFALFA catalogue \\citep{Haynes2011} suggests the true yield of low luminosity galaxies with low \\HI\\ velocities and inferred masses is likely more than 12$\\times$ greater." }, "1402/1402.4429_arXiv.txt": { "abstract": "{ Our understanding of radiative feedback and star formation in galaxies at high redshift is hindered by the rarity of similar systems at low redshift. However, the recently identified Green Pea (GP) galaxies are similar to high-redshift galaxies in their morphologies and star formation rates and are vital tools for probing the generation and transmission of ionizing photons. The GPs contain massive star clusters that emit copious amounts of high-energy radiation, as indicated by intense \\oiii~$\\lambda$5007 emission and \\heii~$\\lambda$4686 emission. We focus on six GP galaxies with high ratios of \\oiii~$\\lambda\\lambda$5007,4959/\\oii~$\\lambda$3727 $\\sim$10 or more. Such high ratios indicate gas with a high ionization parameter or a low optical depth. The GP line ratios and ages point to chemically homogeneous massive stars, Wolf-Rayet stars, or shock ionization as the most likely sources of the \\heii~emission. Models including shock ionization suggest that the GPs may have low optical depths, consistent with a scenario in which ionizing photons escape along passageways created by recent supernovae. The GPs and similar galaxies can shed new light on cosmic reionization by revealing how ionizing photons propagate from massive star clusters to the intergalactic medium. } ", "introduction": " ", "conclusions": "" }, "1402/1402.5934_arXiv.txt": { "abstract": "\\noindent We study structure and dynamics of turbulent photospheric magnetic field in active region NOAA 11158 by characterizing spatial and temporal scaling properties of the line-of-sight (LOS) component. Using high-resolution high-cadence LOS magnetograms from SDO/HMI, we measured power-law exponents $\\alpha$ and $\\beta$ describing Fourier power spectra in wavenumber ($k$) and frequency ($f$) domains and investigated their evolution during the passage of the active region through the field of view of HMI. The flaring active region NOAA 11158 produces a one-dimensional spatial power spectral density that follows approximately a $k^{-2}$ power law -- a spectrum that suggests parallel MHD fluctuations in an anisotropic turbulent medium. In addition, we found that the values of $\\alpha$ capture systematic changes in the configuration of LOS photospheric magnetic field during flaring activity in the corona. Position-dependent values of the temporal scaling exponent $\\beta$ showed that, on average, the core of the active region scales with $\\beta >$ 3 surrounded by a diffusive region with an approximately $f^{-2}$-type spectrum. Our results indicate that only about 1 - 3 \\% of the studied LOS photospheric magnetic flux displays $\\beta\\approx\\alpha$, implying that Taylor's hypothesis of frozen-in-flow turbulence is typically invalid for this scalar field in the presence of turbulent photospheric flows. In consequence, both spatial and temporal variations of the plasma and magnetic field must be included in a complete description of the turbulent evolution of active regions. ", "introduction": "Solar active regions (ARs) are the central building blocks in the path to understanding the drivers of space weather. Major solar flares and coronal mass ejections (CME) originate from active regions, where strong ($\\approx 10 ^{3}$ gauss (G)) and complex magnetic field structures can accumulate sufficient free energy to power energetic eruptions. The high degree of complexity in terms of topology and spatial distribution present in the photospheric magnetic field appears to emerge from a turbulent photospheric plasma state (\\cite{2005ApJ...629.1141A}, \\cite{2010ApJ...720..717A}). In this state, field emergence, fragmentation, and dissipation associated with turbulent flows lead to highly irregular spatio-temporal distribution of the magnetic field. In a simple way, we can view an AR as a system that takes the magnetic field and evolves it into an unstable non-potential configuration by non-linear shear and stress. For this system to return to a lower-energy state, the excess free energy must be released in a bursty event in the corona while electrical currents are dissipated and potential field configuration is restored (\\cite{2011LRSP....8....6S}). Non-linear dynamical processes, such as turbulence, are often studied using a description that involves statistical momenta of the turbulent field. For instance, in hydrodynamics (HD) and magnetohydrodynamics (MHD), kinetic and/or magnetic energy injection, transfer, and dissipation processes in turbulent flows are understood in terms of the scale-free behavior of their Fourier spectrum (\\cite{1993noma.book.....B}, \\cite{2011soca.book.....A}), which follows a power-law distribution in space and time. Kolmogorov's 5/3 law is a classic example of these phenomena \\cite{1941DoSSR..30..301K}. In the case of the photospheric magnetic field, statistical parametric analyses have been performed with the aim of quantifying the complexity present in the field (see \\cite{2005ApJ...629.1141A}; \\cite{2010AdSpR..45.1067M}). However, only recently, when better and more accurate measurements of the photospheric magnetic field have become available, a more coherent picture of its complexity has started to emerge (see \\cite{2005ApJ...629.1141A}). Previous studies on the complexity in the photospheric magnetic field in ARs can be divided into two categories: (1) analysis of physical and statistical parameters of the magnetic field such as the effective connected magnetic field strength (\\cite{2007ApJ...661L.109G}, \\cite{2008GeoRL..3506S02G}), the strong gradient length (\\cite{2002ApJ...569.1016F}, \\citeyear{2003JGRA..108.1380F}), or the statistical momenta of the field spatial distribution (\\cite{2003ApJ...595.1277L}, \\citeyear{2003ApJ...595.1296L}; \\cite{2008ApJ...688L.107B}), and (2) description of magnetic structures based on transformations of the LOS component such as the spatial power scaling exponent (Fourier analysis; \\cite{2005ApJ...629.1141A}; \\cite{2010ApJ...720..717A}) or fractal dimension (wavelet analysis; \\cite{2010AdSpR..45.1067M}). For example, \\cite{2007ApJ...661L.109G} defined the AR effective connected magnetic field strength $B_{\\rm eff}$ as a measure of magnetic field complexity. The $B_{\\rm eff}$ parameter accounts for the connectivity of individual photospheric magnetic flux concentrations; therefore its value depends on the spatial distribution of the flux concentrations. Values of $B_{\\rm eff}$ were measured using LOS magnetograms averaged over 12 h -- a cadence too low in order to capture transient phenomena of magnetic concentrations, which encompass a wide range of temporal scales (\\cite{2012ApJ...748...60U,uritsky13}). \\cite{2005ApJ...629.1141A} analyzed a sample of ARs using photospheric magnetic data from the {\\it Michelson Doppler Imager} (MDI; \\cite{1995SoPh..162..129S}) instrument onboard SoHO and the Digital Magnetograph (DMG) located at the Big Bear Solar Observatory. This study was focused on measuring power-law scaling exponents of the spatial power spectral densities. \\cite{2005ApJ...629.1141A} concluded that the derived exponents described the scale-free behavior of the magnetic field and served as indicators for differentiating between ARs that are prone to produce flaring activity and those that are flare-quiet. A common approach in the studies mentioned above was to quantify the complexity present in the instantaneous spatial distribution of the photospheric field and then to observe subsequent time evolution of the spatial parameter. Consequently, spatial and temporal domain analyses have been conducted for the most part in an independent fashion. A question that then naturally arises concerns the coupling between these two domains: is there a way to link the spatial and temporal variations? The first step to study this coupling is to verify the possible validity of Taylor's hypothesis of frozen-in-flow turbulence (\\cite{Taylor18021938}) for the photospheric plasma. If the hypothesis is valid, determining the (temporal) spatial scaling ({\\it e.g.} \\cite{2002ApJ...577..487A}) is sufficient since the (spatial) temporal scaling is constrained to be identical. On the other hand, if the hypothesis is not valid, both spatial and temporal scaling must be considered in the analysis in order to provide a complete picture of the state of the photospheric magnetic field and plasma. In this report, we will demonstrate that the latter is indeed the case for NOAA AR 11158. In this paper, we extend previous studies of solar AR magnetic field complexity by addressing both spatial and temporal variability of the LOS photospheric magnetic field across a wide range of scales. By constructing a more comprehensive picture of the turbulent spatio-temporal dynamics in the AR photospheric magnetic field, we will provide new information that can help to better understand, for example, magnetic energy release signatures in the photosphere and the coupling between the photosphere and corona. In Section 2, we describe the analyzed set of LOS magnetograms and the active region to which they belong, NOAA 11158. Section 3 explains the data analysis and discusses the results. We explain the method of measuring the power-law exponents in two separate subsections: the spatial scaling analysis (Section 3.1) and the temporal scaling analysis (Section 3.3), both based on the Fourier transform of LOS magnetic field. Our main results are reported and discussed in Sections 3.2 and 3.4, while in Section 4 we draw conclusions and outline future work. ", "conclusions": "We have investigated the turbulent state of the line-of-sight photospheric magnetic field by characterizing its Fourier spectral density, both in spatial and temporal domains. By measuring power-law spectral exponents $\\alpha$ (spatial) and $\\beta$ (temporal), we studied the photospheric plasma dynamics and its possible implications for the photosphere-corona coupling. In this investigation we used high spatial resolution, high cadence LOS magnetograms and maps of coronal emission from SDO/HMI and SDO/AIA, correspondingly. The utilized data represent NOAA AR 11158. We determined the spatio-temporal scaling in two stages. First, the scaling was studied by measuring $\\alpha$ for the spatial power spectral density. In the second stage we carried out the temporal analysis, in which we determined temporal scaling of time series corresponding to the net LOS magnetic field. In both stages, average and time-dependent values were measured. Time-averaged and time-dependent values of the spatial scaling exponent were measured for the turbulence inertial range of scales, which was determined here to be the scales with linear sizes $l=k^{-1}\\approx$ 2 - 20 Mm. Average power spectral density displayed a power law $\\overline{E}(k)\\sim k^{-\\overline{\\alpha}}$ with scaling exponent $\\overline{\\alpha}\\approx 2$. In addition, the time evolution of the power-law exponent shows values greater than 5/3 during the stable phase of the AR, in agreement with \\cite{2010ApJ...720..717A} for flaring ARs. On the other hand, the power law $k^{-2}$ seems to be a characteristic spectrum for MHD turbulence in which the presence of dynamically uniform strong magnetic field favors kinetic and magnetic fluctuations along the field. Temporal spectral analysis of the data showed that the time series of net signed LOS magnetic flux density displays a power-law spectrum which can be approximated by the Kolmogorov exponent $E(f)\\propto f^{-5/3}$ for an inertial range of temporal scales from several minutes to several days. Time series presenting a power-law spectra with $\\beta\\neq$ 2 are described in terms of the fBm model. In particular, for a fBm time series, $\\beta\\approx 5/3$ implies an anti-persistent behavior with weakly-anticorrelated increments. In the context of photospheric magnetic field evolution this fBm behavior could be an indication of the system seeking for a balance between injection and dissipation of the photospheric magnetic flux. We believe this balance is a signature of fully-developed turbulence that controls the photospheric magnetic flux dynamics and it is present in order to maintain the statistically steady state of the global photospheric magnetic network (\\cite{2001ApJ...561..427S}). Position-dependent average values of temporal scaling exponent ($\\overline{\\beta}$) indicate that regions with high average magnetic flux densities are typically associated with higher spectral slopes. Exponents $\\overline{\\beta}$ were measured for a range of frequencies corresponding to the spatial inertial range and consistent with Taylor's approximation. We found that only 1 - 3 \\% of the studied image area satisfies the condition $\\overline{\\beta}\\approx\\overline{\\alpha}$. This implies that Taylor's frozen-in-flow turbulence hypothesis is invalid for most of the field of view, including the AR. Consequently, a linear mapping between spatial and temporal behavior using Taylor's hypothesis seems questionable, and a full spatio-temporal characterization of the photospheric magnetic field is required for a complete description of the system turbulent dynamics. We have taken the first initial steps towards such spatio-temporal characterization in this work. Short-term evolution (minutes to a few hours) of spatial scaling exponent $\\alpha(t)$ captures systematic changes in the spatial distribution of LOS photospheric magnetic field associated with flaring activity. Flare-related changes manifest themselves in both transient (9 - 12 min) and persistent ($\\approx$ 1 h or longer) variations of $\\alpha(t)$ at the time of the flare and immediately afterwards, respectively. Transient variations in $\\alpha(t)$ are most likely associated with artifacts in the magnetic field data, while persistent changes suggest a change in the state of the photospheric field. NOAA AR 11158 produced six M-class flares during its passage through the field of view of the instrument. We detected such systematic changes for two of these flares, in addition to the X-class flare. Although our results support the idea of a back reaction from the corona to the photosphere right after a flare occurs, careful analysis of a larger data set is required to confirm what types of flares are capable of influencing the post-flare state of the photospheric magnetic field. Careful spatio-temporal analysis of high-resolution photospheric and coronal images such as the one conducted in this study can improve our understating of the physics of solar ARs and the links between the photosphere and corona during flaring activity. Furthermore, photospheric parameters such as the scaling exponents $\\alpha$ and $\\beta$ may also contain advanced information about the coronal flaring activity. We will expand our studies initiated in this paper by inclusion of new ARs into the analysis and by using full vector photospheric magnetic field data. Our ultimate goal is to better understand the physical properties of flaring ARs and to seek for new precursors for pending major solar eruptions." }, "1402/1402.5333_arXiv.txt": { "abstract": "Understanding the phases of water ice that were present in the solar nebula has implications for understanding cometary and planetary compositions as well as internal evolution of these bodies. Here we show that amorphous ice formed more readily than previously recognized, with formation at temperatures $<$70 K being possible under protoplanetary disk conditions. We further argue that photodesorption and freeze-out of water molecules near the surface layers of the solar nebula would have provided the conditions needed for amorphous ice to form. This processing would be a natural consequence of ice dynamics, and would allow for the trapping of noble gases and other volatiles in water ice in the outer solar nebula. ", "introduction": "Whether water was present in the outer solar nebula as amorphous or crystalline ice remains a topic of debate. If amorphous ice formed in the nebula, then it would have been capable of trapping other gases present and could possibly explain the noble gas contents of comets and Jupiter \\citep{barnun85,owen99,yokochi12}. However, \\citet{kouchi94} demonstrated that as water ice is expected to condense at temperatures of 120-180 K in the solar nebula, water molecules would have been able to arrange themselves into a crystalline structure. Because the D/H ratios of water throughout the solar nebula is significantly less than that found in the interstellar medium, it is thought that much of the ice in the nebula saw high temperatures near the young Sun then was carried outward to cooler regions, condensing at these temperatures along the way \\citep{mousis00,yang13}. For example, bringing the D/H ratio of water from the interstellar ratio of $\\sim$0.001-0.01 to the cometary values of 1.5-3$\\times$10$^{-4}$ \\citep{hartogh11} requires that that 70-97\\% of water equilibrated with molecular hydrogen (solar D/H $\\sim$10$^{-5}$) at high temperatures ($>$500 K) in the solar nebula, before cooling and condensing as ice. This has led many to argue that amorphous ice formation was precluded in the solar nebula, and that all water ice was crystalline requiring another means of trapping volatiles such as clathrates \\citep{kellerjorda01,hersant04}. The solar nebula is expected to have been a dynamic object, in which gas and dust were pushed around as mass and angular momentum were transported throughout the disk \\citep{armitage11}. As a consequence of this evolution, gas and dust would be transported throughout the disk and exposed to a variety of physical and chemical environments. Upon passing through such environments, the dust and gas could be altered by chemical reactions, irradiation, or phase transitions. This was demonstrated in \\citet{cieslasandford12}, as icy grains in the outer solar nebula were found to be lofted to high altitudes above the disk midplane and exposed to intense radiation fluxes which would break molecular bonds in the ices and allow the resulting ions and radicals to react to form more complex species. Thus just because a compound or phase is formed in a protoplanetary disk, it does not mean that original phase is preserved throughout the lifetime of the solar nebula. In this Letter we revisit the conditions under which the various phases of water ice would form in the solar nebula. In particular, the conditions for amorphous ice formation in the solar nebula are derived assuming that the ice forms on a substrate other than hexagonally crystalline ice. We then show that the conditions for amorphous ice formation would have been met as icy grains were irradiated by UV photons at the surface layers of the solar nebula and the resulting photodesorbed water froze-out again at greater depths inside the protoplanetary disk. ", "conclusions": "The results presented here show that even if water ice condensed in the solar nebula at temperatures in the 120-180 K range, where crystalline ice is expected to form, the dynamical evolution of the icy grains in the outer solar nebula would lead this ice to be lost and reformed as amorphous ice. Formation of amorphous ice in this manner may allow for the trapping of other gaseous species, such as N- and C- bearing species as well as noble gases. The trapping of such species in ices would allow volatile-rich planetesimals to form in the outer nebula, and may explain the enhanced abundances of these elements relative to hydrogen in the Jovian atmosphere compared to a gas of solar composition \\citep{owen99,atreya03}. An issue that requires future attention is the detailed fate of the liberated water at the upper layers of the protoplanetary disk. If water molecules are desorbed sufficiently high in the disk, they may be photodissociated by the same UV photons that liberate them from the grains they are on. The extent to which materials are photodissociated will depend on the amount of time materials spend in the extreme upper layers of the disk. As shown here, the amount of water surrounding a grain will constantly be changing, with molecules freezing out and then being liberated over and over again as the conditions they exposed to are constantly changing. Thus there will be a time component to determining the precise fraction of water molecules that are liberated around particular grains that gets photodissociated compared to the bulk water present in a given region. However, estimates of the amount of water vapor present in irradiated regions of the disk under steady-state conditions with no dynamic transport give $n_{H_{2}O}$/$n_{H} \\sim$2$\\times$10$^{-7}$, or 1/1000 the maximum value reached here \\citep{hollenbach09}. Taking this as an upper limit for the amount of water vapor present in the regions of the disk considered would imply freeze-out fluxes that are ~1000$\\times$ smaller than calculated here, or 10$^{2}$-10$^{6}$ cm$^{-2}$ s$^{-1}$. These values still exceed the $F_{c}(I_{p})$ (Fig. 1) for temperatures $<$60 K by orders of magnitude, meaning amorphous ice would still form. Further, even with photodissociation, the resulting species will freeze-out again and reform H$_{2}$O on grain surfaces at very rapid rates \\citep{ioppolo08,dulieu10}. Given the low surface diffusivities of the molecules on $I_{p}$ at the temperatures considered here, when such molecules form they would not be able to reorder themselves, resulting in amorphous ice. The formation of amorphous ice alone would aid in the continued formation of this phase of water ice: the surface diffusivity of amorphous ice is orders of magnitude smaller than $I_{p}$ \\citep{kouchi94}, meaning any water molecules formed or frozen out on it would remain locked in place and unable to rearrange into a crystalline structure. This would be true whether the amorphous ice formed via direct freeze-out, surface processes, or through irradiation damage. \\citet{letobaratta03} found that crystalline water ice would be amorphized after receiving a dose of UV photons of a few eV per molecule, or a few UV photons per molecule. The dosages calculated here and found by \\citet{cieslasandford12} are beyond those needed for amorphization. Thus any water ice that forms as a result of photodesorption in the outer solar nebula would likely form on a very rough substrate with limited surface diffusivity, leading the water to freeze out in the amorphous phase. Should we thus expect all water ice in the solar nebula and protoplanetary disks to become amorphous? Not necessarily. Just as we have shown that cycling of this type could lead crystalline ice to be transformed to amorphous ice, amorphous ice could be transformed to crystalline as it migrates through the protoplanetary disk. At temperatures near $\\sim$100 K, amorphous ice undergoes a spontaneous phase transition, where molecules rearrange themselves to form crystalline ice. Thus as ice particles migrate radially in a protoplanetary disk and reach higher temperatures, the amorphous ice may be lost. Trapped volatiles which resided on the grain surface when ice molecules froze out on top of them may be retained, despite the phase transition \\citep{viti04, collings04}, meaning that crystalline ices with trapped volatiles may also be present in such disks. Determining the relative abundances of each phase of ice requires a detailed investigation of the radial transport of icy grains throughout the evolution of the solar nebula, which should be the goal of future work." }, "1402/1402.2380_arXiv.txt": { "abstract": "PSR~J0218$+$4232 is a millisecond pulsar (MSP) with a flux density $\\sim$0.9 mJy at 1.4~GHz. It is very bright in the high-energy X-ray and $\\gamma$-ray domains. We conducted an astrometric program using the European VLBI Network (EVN) at 1.6~GHz to measure its proper motion and parallax. A model-independent distance would also help constrain its $\\gamma$-ray luminosity. We achieved a detection of signal-to-noise ratio S/N~$>37$ for the weak pulsar in all five epochs. Using an extragalactic radio source lying 20 arcmin away from the pulsar, we estimate the pulsar's proper motion to be $\\mu_{\\alpha}\\cos\\delta=5.35\\pm0.05$~mas\\,yr$^{-1}$ and $\\mu_{\\delta}=-3.74\\pm 0.12$\\,mas\\,yr$^{-1}$, and a parallax of $\\pi=0.16\\pm0.09$\\,mas. The very long baseline interferometry (VLBI) proper motion has significantly improved upon the estimates from long-term pulsar timing observations. The VLBI parallax provides the first model-independent distance constraints: $d=6.3^{+8.0}_{-2.3}$\\,kpc, with a corresponding $3\\sigma$ lower-limit of $d=2.3$\\,kpc. This is the first pulsar trigonometric parallax measurement based solely on EVN observations. Using the derived distance, we believe that PSR~J0218$+$4232 is the most energetic $\\gamma$-ray MSP known to date. The luminosity based on even our 3$\\sigma$ lower-limit distance is high enough to pose challenges to the conventional outer gap and slot gap models. ", "introduction": "PSR~J0218$+$4232 is a pulsar with a spin period of 2.3~millisecond and a period derivative of 8.0$\\times$10$^{-20}$~s\\,s$^{-1}$. This millisecond pulsar (MSP) was first discovered by \\citet{navarro95} using the Lovell telescope. The radio timing observations also found that it has a low-mass companion ($M\\gtrsim$~0.16~$M_\\odot$) with an orbital period of 2 days. Optical observations using the Keck telescope revealed that the companion is a helium-core white dwarf with a temperature of $T_\\mathrm{eff}=8060\\pm150$~K, and a distance constraint of 2.5--4~kpc is given by white-dwarf modeling \\citep{bassa03}. Note that the distance uncertainty derived by this method is difficult to quantify exactly, because it is dependent on the white-dwarf's mass and optical luminosity, which are both correlated with the cooling age that is highly uncertain for observations and theoretical white-dwarf models \\citep{bassa03}. PSR~J0218$+$4232 is also an energetic pulse emitter in X-rays and $\\gamma$-rays. It was a $\\gamma$-ray MSP candidate detected by Energetic Gamma Ray Experiment Telescope \\citep{kuiper00}. Soon after \\textit{Fermi Gamma-ray Space Telescope} was launched, its was confirmed as a $\\gamma$-ray MSP \\citep{msp-sci}. The X-ray pulsed emission has also been well detected and monitored by many X-ray telescopes \\citep[][]{webb04}. Currently, there are approximately 150 rotation-powered pulsars detected in the X-ray band, and about 50 of those have millisecond spin periods \\citep{becker09}. PSR~J0218$+$4232 is one of a few pulsars with a flux $>10^{-5}$ photon cm$^{-2}$\\,s$^{-1}$ in the 2~--~10~keV band \\citep[e.g.,][]{kuiper02}. An absolute X-ray timing accuracy of $\\sim$200~$\\mu$s was achieved by \\emph{Chandra} \\citep{kuiper04} and 40 $\\mu$s by \\emph{XMM-Newton} \\citep{webb04}. Astrometric parameters (e.g. position, proper motion, parallax) can be determined from pulsar timing observations over a time span of several years. It is still a challenge to measure the times of arrival of its pulses with a high precision since PSR~J0218$+$4232 has broad profile and significant ($\\sim$50\\%) non-pulsed emission \\citep{navarro95}. The best timing solution published to date was derived from observations at Effelsberg \\citep{laz09}. The proper motion derived from these observations is $\\mu_{\\alpha}\\cos\\delta=+5.1\\pm0.3$~mas\\,yr$^{-1}$ and $\\mu_{\\delta}=-2.3\\pm0.7$\\,mas\\,yr$^{-1}$, whose uncertainties are better than the previously published values in \\citet{hobbs05}. Furthermore, due to the occurrence of sources of timing noise in the time-of-arrival data (e.g., the pulsar's intrinsic spin-down noise, noise induced by the stochastic gravitational wave background, dispersion measure (DM) variations) as well as a relatively short timespan of timing observations that can lead to covariances with other parameters of the timing model, the pulsar-timing method can lead to significant errors in the estimated astrometric parameters. The proper motion and trigonometric parallax of a pulsar can also be independently measured with VLBI observations. The high-precision VLBI astrometry has been applied to many bright slow pulsars \\citep[e.g.,][]{cam96, brisken02, brisken03, chatterjee09}, and MSPs, such as PSR~B1937$+$21 \\citep{1937}, PSR~J0437$-$4715 \\citep{0437}, and PSR~B1257$+$12 with three planets \\citep{yan13}. A pulsar distance measured to 0.4\\% accuracy has been recently achieved in the VLBI astrometry of PSR~J2222$-$0137 \\citep{del13}. Astrometric parameters derived from VLBI observations can further improve the estimation of parameters from long-term pulsar timing observations, by providing a prior constraint on astrometric parameters to which the timing analysis is insensitive, but which may themselves be highly covariant with other parameters uniquely approachable via timing. Combining VLBI- and timing-derived astrometry can contribute to frame ties between the International Celestial Reference Frame and the dynamical solar-system frame, which underlie VLBI and pulsar timing, respectively \\citep{madison13}. \\begin{figure*}[!tb] \\centering \\includegraphics[angle=-90,width=0.98\\textwidth]{f1.eps} \\caption{EVN astrometry: PSR J0218+4232 positions, $1\\sigma$ error ellipses, and the modeled track in the sky plane. A star symbol denotes the calculated pulsar position at the epoch of each observing session. The origin gives the position derived at the reference epoch J2011.8793. Some results from the least-squares fit are annotated in each panel. Table~\\ref{tbl_2} lists the results for the fit that includes parallax. } \\label{fig1} \\end{figure*} In this Letter, we present the results of the first VLBI observations of PSR~J0218$+$4232. We summarize the strategy of the VLBI observations and the post-correlation data reduction in Section 2. We discuss the estimation of the pulsar's astrometric parameters in Section 3. Finally, we address the model constraints on the $\\gamma$-ray luminosity of the pulsar in Section 4. ", "conclusions": "" }, "1402/1402.7264_arXiv.txt": { "abstract": "We present ages and masses for 601 star clusters in M31 from the analysis of the six filter integrated light measurements from near ultraviolet to near infrared wavelengths, made as part of the Panchromatic Hubble Andromeda Treasury (PHAT). We derive the ages and masses using a probabilistic technique, which accounts for the effects of stochastic sampling of the stellar initial mass function. Tests on synthetic data show that this method, in conjunction with the exquisite sensitivity of the PHAT observations and their broad wavelength baseline, provides robust age and mass recovery for clusters ranging from $\\sim10^2 - 2\\times10^6\\msun$. We find that the cluster age distribution is consistent with being uniform over the past $100 \\Myr$, { which suggests a weak effect of cluster disruption within M31}. The age distribution of older ($>100\\Myr$) clusters fall towards old ages, consistent with a power-law decline of index $-1$, { likely from a combination of fading and disruption of the clusters}. We find that the mass distribution of the whole sample can be well-described by a single power-law with a spectral index of $-1.9 \\pm 0.1$ over the range of $10^3-3\\times10^5\\msun$. However, if we subdivide the sample by galactocentric radius, we find that the age distributions remain unchanged. However, the mass spectral index varies significantly, showing best fit values between $-2.2$ and $-1.8$, with the shallower slope in the highest star formation intensity regions. We explore the robustness of our study to potential systematics and conclude that the cluster mass function may vary with respect to environment. ", "introduction": "It has become clear that a significant fraction of star formation occurs in stellar clusters. However, deriving galaxy histories from observations of clusters is complicated by significant uncertainties. Controversial questions have been raised regarding cluster properties in various environments as different analyses could lead to different conclusions. Claims exists for interesting variations or trends in cluster colors, their lifetimes as gravitationally bound objects, and age or mass distributions within or between galaxies, and the evolution from initial to current cluster mass functions \\citep[\\eg,][]{Zepf1993, Kumai1993, Girardi1995, Elmegreen1997, Bastian2011}. Such studies rely on our ability to estimate intrinsic properties of stellar clusters, in particular, their ages and masses. Much observational effort has therefore been invested in determining the distributions of star cluster ages and masses \\citep{Searle1980, Larsen2000, Billett2002, Hunter2003, Fall2005, Rafelski2005, Dowell2008, Larsen2009, Chandar2010, Bastian2012}, and using the resulting data to determine the dominant mechanisms of cluster formation and disruption \\citep{Kroupa2002,Boutloukos2003, Lamers2005, Whitmore2007, Parmentier2008, Fall2009, Elmegreen2010, Converse2011}. However, most of the existing work deals with observations of relatively massive clusters (a few $10^4-10^5\\msun$), which are the least affected by disruption processes and thus are the most stable in various environments. As a result, divergences between disruption models \\citep[\\eg][among many others]{Baumgardt2003, Fall2009} are still subject to debate in the literature \\citep[\\eg][most recently in M83]{Whitmore2011, SilvaVilla2014arXiv}. Only a few studies have been able to probe the smaller clusters that are the most sensitive to environmental effects (mainly in the Galaxy, \\eg, \\citealt{Borissova2011}, or the Magellanic Clouds, \\eg, \\citealt{Popescu2012}). The Panchromatic Hubble Andromeda Treasury (PHAT; \\citealt{Dalcanton2012a}) is an ongoing multi-cycle Hubble Space Telescope (HST) program that is ideal for studying stellar clusters in M31. The survey imaged one-third of the M31 disk at high spatial resolution with wavelength coverage from the ultraviolet through the near-infrared. The sensitivity of the latest HST instruments allow us to detect clusters in M31 down to a regime in which cluster luminosities overlap those of individual bright stars, and hence to very low masses \\citep{Johnson2012a}. This survey spans a wide range of environments in both star formation intensities and gas densities. This diversity is an advantage for addressing how environment affects cluster formation. The PHAT survey has already significantly increased the number of clusters known in M31. \\citet{Johnson2012a} identified $601$ stellar clusters using the first quarter of the total PHAT coverage. This new cluster catalog contains more than a factor of four increase in the number of known clusters within the survey area. Moreover, the uniform photometric coverage from the UV to near-IR allows accurate age-dating of the clusters. Even this preliminary sample breaks new ground for studying clusters outside the Milky Way and the Magellanic Clouds, probing about two orders of magnitudes fainter in the luminosity function \\citep[their figure 11]{Johnson2012a}. This paper is part of a series utilizing the PHAT dataset for studies of stellar clusters. \\citet{Johnson2012a} presented the first installment of a HST-based cluster catalog, which serves as the basis for an extensive study of Andromeda's cluster population. In this paper, we focus on the determination of ages and masses of the first year sample, looking forward to the final product of this four year Treasury program. Our estimates of the properties of the clusters are derived from integrated photometry in six broad bands and we especially focus our attention on the characterization of the lowest-mass clusters. Additional studies, including analysis of structural parameters, resolved star content, and integrated spectroscopy of the cluster sample will follow in subsequent work. \\medskip This paper is organized as follows. \\S \\ref{sec:data} presents the cluster sample and the key elements of their photometry. \\S \\ref{sec:Analysis} describes the analysis and the cluster models used to derive the properties of the clusters, and briefly highlights the possible artifacts of the method using synthetic data. \\S \\ref{sec:yr1properties} describes our results for the entire sample and for individual regions across M31. Finally, we discuss those results in \\S \\ref{sec:discussion} before drawing our conclusions. ", "conclusions": "\\label{sec:conclusions} We have derived ages, masses and exintction for the Year 1 PHAT cluster sample \\citep{Johnson2012a}, by comparing the cluster integrated 6-filter fluxes with an extended version of the stochastically sampled model clusters presented in \\citet{Fouesneau2010}. The locus of the collection of stochastic models in color space (\\eg, Figure 2) shows excellent agreement with that of the collection of cluster observations. Clusters with broadband colors either bluer or redder than those of the traditional continuous models find a natural match with the models we used in this paper. We generated the full joint probability distribution function of the age, mass, and extinction for each of the 601 individual clusters in the sample. We then combined their individual distributions into global cluster age and mass distributions, noting limits at which completeness issues in the sample become severe. The sample of clusters spans the entire length of the age sequence and includes a significant number of clusters with masses well below $10^3\\msun$. Only a few datasets have the ability to sample objects across a variety of stages in cluster evolution over such a large, uninterrupted mass range. We find that the cluster age distribution shows a constant number of clusters over the last $\\sim 100\\Myr$, with a power-law decline at older ages (see Figs\\,\\ref{fig:marginal_age_dist} \\& \\ref{fig:dndt_dist_per_brick}). At least above the mass of $10^{3.2}\\msun$, these results are consistent with M31 producing a constant number of clusters from $100\\Myr$ ago to present, with little significant cluster disruption over this timescale. The mass distribution derived from the analysis closely resembles the power-law distributions obtained from many other galaxies. Specifically, the overall power-law index of the mass distribution is consistent with the canonical value of $-$2. However, the current cluster sample suggests a possible radial variation of this distribution across the disk, with the shallowest power-law found in the region with the highest star formation rate. When we study the entire PHAT survey, including lower masses and a larger sample of fainter clusters, the improved accuracy and time resolution achievable with the new stochastic methods will allow us to address new questions. Future work will account for the challenging determination of completeness and selection effects. In particular, the expected number of clusters in PHAT will eventually provide 5 times more clusters over a broad range of local environments, which will open the possibility to study local variations among cluster populations beyond our current the initial assessment in this study." }, "1402/1402.2488_arXiv.txt": { "abstract": "{The structure of protoplanetary disks is thought to be linked to the temperature and chemistry of their dust and gas. Whether the disk is flat or flaring depends on the amount of radiation that it absorbs at a given radius, and on the efficiency with which this is converted into thermal energy. The understanding of these heating and cooling processes is crucial to provide a reliable disk structure for the interpretation of dust continuum emission and gas line fluxes. Especially in the upper layers of the disk, where gas and dust are thermally decoupled, the infrared line emission is strictly related to the gas heating/cooling processes.}{We aim to study the thermal properties of the disk in the oxygen line emission region, and to investigate the relative importance of X-ray (1-120 \\AA) and far-UV radiation (FUV, 912-2070 \\AA) for the heating balance there.}{We use \\on\\,63 \\mic\\,line fluxes observed in a sample of protoplanetary disks of the Taurus/Auriga star forming region and compare it to the model predictions presented in our previous work. The data were obtained with the PACS instrument on board the Herschel Space Observatory as part of the Herschel Open Time Key Program GASPS (GAS in Protoplanetary diskS), published in Howard et al. (2013).}{Our theoretical grid of disk models can reproduce the \\on\\,absolute fluxes and predict a correlation between \\on\\,and the sum \\lx+\\luv. The data show no correlation between the \\on\\,line flux and the X-ray luminosity, the FUV luminosity or their sum.}{The data show that the FUV or X-ray radiation has no notable impact on the region where the \\on\\,line is formed. This is in contrast with what is predicted from our models. Possible explanations are that the disks in Taurus are less flaring than the hydrostatic models predict, and/or that other disk structure aspects that were left unchanged in our models are important. Disk models should include flat geometries, varying parameters such as outer radius, dust settling, and the dust-to-gas mass ratio, which might play an equally important role for the \\on\\,emission. To improve statistics and draw more robust conclusions on the thermal processes that dominate the atmosphere of protoplanetary disks surrounding T\\,Tauri stars, more \\luv\\,and \\lx\\,measurements are needed. High spatial and spectra resolution data is required to disentangle the fraction of \\on\\,flux emitted by the disk in outflow sources.} ", "introduction": "Planet formation is strongly linked to the physical properties of the parent disk. Important constraints on the timescale for the gas accretion of giant planets are posed by photoevaporation models. The results of such models are essential in order to estimate the mass loss rates, and hence the survival time of gas in disks (\\citealt{Ale06,Erc08,GDH09}). The stellar radiation, especially in the high energy regime (E $>$ 6 eV), is responsible for the thermo-chemical conditions in the disk atmosphere, as it provides most of the energy that causes the gas temperature to exceed the dust temperature there \\citep{Kam04,Jon04,Gla04}. However, the thermal processes that heat and shape protoplanetary disks are poorly constrained and can only be indirectly measured through cooling lines. One of the dominant cooling lines that can be used to understand these processes is the 63 micron line of neutral oxygen \\citep{Gor08,Mei08,Woi09,Are12}. T\\,Tauri stars emit radiation at high energies, due both to chromospheric activity and accretion of disk material onto the stellar surface. The FUV luminosity between 7 and 10 eV ($\\Delta\\lambda$=1240-1770 \\AA), has been measured by \\citet{Yan12} for a sample of accreting sources in Taurus: they found values between 10$^{30}$ and few times 10$^{32}$ erg/s. The emission is in excess when compared to the stellar emission in the same energy band for non-accreting young stars of the same spectral type. This suggests that accretion is responsible for this emission, in which case it is caused by shocks created by the magnetic field that channels disk material toward the stellar surface \\citep{Cal98,Val00}. EUV ($\\Delta\\lambda$= 120-912 \\AA, $\\Delta$E=13.6-100 eV) radiation is believed to mainly affect the upper disk surface at small radii, as the high cross section for absorption only allows penetration of small columns of $N_{\\rm H}\\sim 10^{19}$ cm$^{-2}$. The XEST survey \\citep{Gue07} has shown that young stars are also active X-ray emitters, mainly due to chromospheric activity, and can reach luminosities between 10$^{29}$ and 10$^{31}$ erg/s. The high energy depositions ($>$0.01 $L_{*}$) and heating efficiencies ($\\sim$ 30\\%) of X-rays cause the tenuous disk atmosphere to heat up to temperatures of the order of a few thousand Kelvin \\citep{Gla07,Nom07,Gor08,Erc08,Are11}. Recent observations, carried out with the Herschel Space Observatory toward the Taurus forming region, offer the chance to test model predictions on the thermal structure of the region where the \\on\\,63.2 \\mic\\,line is emitted. This line is predicted to arise from the disk atmosphere in the radial region between a few 10 AU and 200 AU \\citep{Woi09,Are12}. The emission region is directly exposed to the stellar radiation and models suggest that FUV and X-ray radiation are the main heating agents there. PAH and dust photoelectric heating as well as Coulomb heating, cause the gas temperature to be of the order of $\\sim$ 200-300 K \\citep{Gor08,Mei08,Mei12}. In this work, we explore possible correlations of the \\on\\,emission with X-ray luminosity and FUV luminosity, and compare the \\citet{Are12} model predictions for the \\on\\,63 \\mic\\,emission with data collected within the GASPS (GAS in Protoplanetary DiskS, P.I. Dent) Open Time Key Program, taken with the PACS instrument on board the Herschel Space Observatory \\citep{Den13}. In the following, we make the hypothesis that most of the \\on\\,emission is produced in the disk. Outflow sources, that have on average higher accretion rates, will then produce more FUV radiation and thus stronger FUV illumination of the disk surface and stronger line emission. % In Sect.\\ref{obser} we present the collected observational data set and in Sect.\\ref{models} we explain the main findings of the models studied in \\citet{Are12}. In Sect.\\ref{results} we show the results of the comparison between model predictions and observations, these will be discussed in Sect.\\ref{disc}. Conclusions and remarks about future work are summarised in Sect.\\ref{fut}. ", "conclusions": "\\label{fut} In this work we studied the impact of FUV and X-ray radiation on the thermal balance in the oxygen emission region for protoplanetary disks surrounding T\\,Tauri stars. We compared disk model predictions with observations of the \\on\\,63\\mic\\,line toward protoplanetary disks that do not show outflow emission in the Taurus region obtained with the PACS instrument on board Herschel. The observations show no correlation between the \\on\\,63 \\mic\\,line emission and the X-ray luminosity or the FUV luminosity or with their sum. Our thermo-chemical disk models calculated with ProDiMo, show that our predictions on the \\on\\,fluxes qualitatively agree with the observations. There is no correlation between \\on\\,and \\lx\\,or \\luv, as the data suggest. Nevertheless, the models predict a correlation between \\on\\,and the sum \\lx+\\luv, which is not seen in the data. The reason can be the limited set of parameters varied in our model (\\lx, \\luv, grain minimum size, power law of the grain size distribution and power law of the surface density distribution) grid to understand the relative importance of \\lx\\,and \\luv.However, other parameters can affect the \\on\\,line, causing the correlation we predict to vanish when a more complete grid is used. To include all the disk parameters that influence the \\on\\,line, a different set of models should be used. Flatter disk geometries should be included as well as a proper treatment of X-ray and FUV physics and dust settling (local variations of the gas-to-dust mass ratio). Moreover, such models should be compared to a higher number of observations: high spatial and spectral resolution data is required to disentangle the location of the emission region of the line. In many sources that drive outflows, the contribution of the disk to the total flux of the line remains unclear. More measurements of \\luv\\,would also be necessary. To test the threshold mechanisms proposed in our previous work, observations of \\on\\,of sources with \\lx $>$ \\luv, if any, are essential. The understanding of the [OI] dependence on the FUV and X-ray radiation gives the possibility to investigate the gas surface layers above the H/H2 transition. Such studies are very interesting for understanding the photoevaporation mechanism and how it may drive disk evolution across the transition from optically thick to debris disk. \\begin{figure}[h!] \\centering \\includegraphics[scale=0.4,angle=-90]{fit03.pdf}% \\caption{\\small \\on\\,flux versus \\lx+\\luv. Model points are black while data points are red. In our models we considered a sun-like star surrounded by a disk of 0.02 $M_{\\odot}$ which extends from 0.5 to 500 AU.} \\label{fit3} \\end{figure} \\tiny \\emph{Acknowledgements}. We thank Aki Roberge for her comments on the FUV analysis which helped to improved the paper and Glenn White for thoroughly reading the paper. WFT, PW, FM, MG and IK acknowledge funding from the EU FP7-2011 under Grant Agreement nr. 284405. L.P. acknowledges the funding from the FP7 Intra-European Marie Curie Fellowship (PIEF-GA-2009-253896)." }, "1402/1402.1507.txt": { "abstract": " ", "introduction": "One of the most exciting open questions at the interface between particle physics and cosmology is the nature of Dark Matter (DM). The first person who provided evidence and inferred the presence of DM was a Swiss-American astrophysicist, Fritz Zwicky. He applied the virial theorem to the Coma cluster of galaxies and obtained evidence of unseen mass. Roughly 40 years following the discoveries of Zwicky and others, Vera Rubin and collaborators conducted an extensive study of the rotation curves of isolated spiral galaxies. They announced the discovery that the rotational curves of stars in spiral galaxies exhibit a characteristic flat behaviour at large distance in contrast with Kepler's law. Many other evidence of unseen mass on distance scales of the size of galaxies and clusters of galaxies appeared throughout the years, but the most precise measurement of the total amount of DM comes from cosmological scales. In particular, the measurements of modern precision cosmology (the Cosmic Microwave Background (CMB) and the surveys of the Large Scale Structure (LSS) of the Universe), provide the current most relevant evidence. Apart from the qualitative agreement, it is the quantitative fitting of the wealth of available data that allows the amount of DM to be one of the cosmological parameters now most precisely measured ($\\Omega_\\chi h^2=0.1199\\pm0.0027$, see Tab.~2 of~\\cite{Ade:2013zuv}). Therefore we have compelling evidence of unseen mass, but the microscopic features of this new kind of matter remain unknown yet. Direct and Indirect searches may shed light on the nature of DM, and therefore a careful study of their phenomenology is fundamental. For a pedagogical review on this subject, see e.g.~\\cite{Bertone:2004pz}. \\medskip Direct searches for DM aim at detecting the nuclear recoils arising from scattering between DM particles and target nuclei in underground detectors. DM direct detection experiments are providing exciting results in terms of measured features which have the right properties to be potentially ascribed to a DM signal. For example in addition to the long-standing \\DAMA\\ results, nowadays there are other experiments, such as \\CoGeNT, \\CRESST\\ and \\CDMSSi\\, that start to see some anomalies in their counting rates. On the other hand ,the situation in this field is extremely unclear and confusing, because on top on these positive result experiments, the constraints coming from null results, like \\XENONhundred, \\COUPP, \\PICASSO\\ and very recently \\LUX, are very stringent and put the interpretation of the anomalies in terms of a DM interaction in serious trouble. Nevertheless, there are at least two main caveats when the results from the experiments commented upon above are interpreted. The first is that one has to treat with great care the fine experimental details associated with the results quoted by each experiment. The second caveat is instead associated with the interpretation of the data within a very simple-minded DM model. For instance, the DM-nucleus spin independent contact interaction is just a benchmark example. Upon relaxing some of these assumptions, the current complicated experimental puzzle can probably be solved. \\medskip The {\\bf scope of this work} is to present the {\\em status} of direct DM detection with specific attention to the experimental results and their phenomenological interpretation in terms of DM interaction. In particular in Sec.~\\ref{Basics} I review a new and more general approach to study signals in this field based on non-relativistic operators which parameterize most efficiently the DM-nucleus interactions in terms of a very limited number of relevant degrees of freedom. In Sec.~\\ref{DDStatus} I review the experimental results and their interpretation in terms of the ``standard'' spin independent (SI) interaction. I list then the main uncertainties that affect the theoretical interpretation of the data: this is a very promising area of research since only major advancements here can probably reconcile the complicated puzzle showed by the experiments up. Finally in Sec.~\\ref{LRInteraction}, I pose my attention on the uncertainties coming from the nature of the interaction. In particular the phenomenology of a class of models in which the interaction between DM particles and target nuclei is of a long range type is discussed. ", "conclusions": "Direct DM searches is now characterized by tantalizing results and hints that make this field very active both from the theoretical and experimental side. In particular, in addition to the long standing \\DAMA\\ results, nowadays there are other experiments, like \\CoGeNT, \\CRESST\\ and \\CDMSSi\\ that are starting to observe anomalies in their counting rates. On the other hand, the increasingly stringent constraints coming from null result experiments put in serious trouble the theoretical interpretation of the data, at least in terms of the simple-minded SI contact interaction. In this work I discussed the {\\em status} of direct DM detection with specific attention to the experimental results and their phenomenological interpretation in terms of DM interaction. In particular, in the first part I presented a new and more general approach to study signals in this field based on non-relativistic operators. Then I reviewed the experimental results and their interpretation in terms of the ``standard'' SI interaction pointing out all the uncertainties which enter in this field. In the last part of this work, I investigated a fermionic Dark Matter particle carrying a small milli-charged and analyzed its impact on direct detection experiments. I showed that this kind of long range interaction can accommodate the positive experimental results. By assuming a conservative choice for the lower threshold of the \\XENONhundred\\ and \\LUX\\ experiments I have demonstrated that this candidate is not ruled out. I also determined the complementary class of constraints which are relevant for milli-charged DM particles with long range forces. \\medskip Finally, I would like to propose a possible direction to pursue in order to make sense of the current exciting experimental panorama based on the formalism of non-relativistic operators. Indeed, as we have seen in the first part of this work, it allows us to describe the DM-nucleus interactions in terms of a very limited number of relevant degrees of freedom. In this way, it is possible to parametrize the model-dependent part of the rate from the model-independent one encapsulated in a sort of \\emph{integrated form factors} that encode all the dependences on the astrophysics, nuclear physics and experimental details. Therefore, since one is ignorant or agnostic about the underlying theory, I would like to encourage a synergy between nuclear physicists and experimentalists in order to provide a complete set of \\emph{integrated form factors} defined in Eq.~\\eqref{integratedFF}. It would be extremely useful for the community, because in this way, one can compute the expected number of events for any kind of interaction (e.g.~including isospin-violating interactions, momentum-dependent form factors, velocity-dependent form factors) and compare directly with the experimental results. Providing the \\emph{integrated form factors} will thus be the first step towards a model-independent analysis in direct DM searches. \\small \\paragraph{Acknowledgements} We thank Marco Cirelli, Eugenio Del Nobile and Joe Silk for useful discussions. \\bigskip \\noindent The author declare that there is no conflict of interests regarding the publication of this article. \\bigskip \\appendix \\footnotesize \\begin{multicols}{2}" }, "1402/1402.0194.txt": { "abstract": "{}{}{}{}{} % 5 {} token are mandatory \\abstract % context heading (optional) {Galactic winds are a common phenomenon in starburst galaxies in the local universe as well as at higher redshifts. Their sources are superbubbles driven by sequential supernova explosions in star forming regions, which carve out large holes in the interstellar medium and eject hot, metal enriched gas into the halo and to the galactic neighborhood. } % {} leave it empty if necessary %{To investigate the physical nature of the `nuc\\-leated instability' of %proto giant planets, the stability of layers %in static, radiative gas spheres is analysed on the basis of Baker's %standard one-zone model.} % aims heading (mandatory) {We investigate the evolution of superbubbles in exponentially stratified disks. We present advanced analytical models for the expansion of such bubbles and calculate their evolution in space and time. With these models one can derive the energy input that is needed for blow-out of superbubbles into the halo and derive the break-up of the shell, since Rayleigh-Taylor instabilities develop soon after a bubble starts to accelerate into the halo. } % In our paper, we analyze the behavior of blow-out superbubbles. % methods heading (mandatory) {The approximation of Kompaneets is modified in order to calculate velocity and acceleration of a bubble analytically. Our new model differs from earlier ones, because it presents for the first time an analytical calculation for the expansion of superbubbles in an exponential density distribution driven by a time-dependent energy input rate. The time-sequence of supernova explosions of OB-stars is modeled using their main sequence lifetime and an initial mass function. } % results heading (mandatory) {We calculate the morphology and kinematics of superbubbles powered by three different kinds of energy input and we derive the energy input required for blow-out as a function of the density and the scale height of the ambient interstellar medium. The Rayleigh-Taylor instability timescale in the shell is calculated in order to estimate when the shell starts to fragment and finally breaks up. Analytical models are a very efficient tool for comparison to observations, like e.g. the Local Bubble and the W4 bubble discussed in this paper, and also give insight into the dynamics of superbubble evolution.} %% %% DB: Satz ge\u00e4ndert %% % % conclusions heading (optional), leave it empty if necessary {} ", "introduction": "% Most massive stars are born in OB-associations in coeval starbursts on timescales of less than 1-2 Myr \\citep{m99}. % These associations can contain a few up to many thousand OB-stars, so-called super star clusters, but typically have 20-40 members \\citep{mk87}. % Energy and mass are injected through strong stellar winds and subsequent supernova (SN) explosions of stars with masses above 8 M$_{\\odot}$. The emerging shock fronts sweep-up the ambient interstellar medium (ISM) and, as the energy input in form of SN-explosions continues, superbubbles (SBs) are produced, which may reach dimensions of kiloparsec-size \\citep[e.g.][]{tsm03}. The swept-up ISM collapses early in the evolution of the SB into a cool, thin, and dense shell \\citep{cmw75} present in HI and H${\\alpha}$ observations. % The bubble interior contains hot ($> 10^6 \\, $K), rarefied material, usually associated with extended diffuse X-ray emission \\citep{sta05}. % % Due to the stratification of the ISM in disk galaxies, the superbubbles can accelerate along the density and pressure gradient and blow out into the halo, appearing as elongated structures. Examples of huge bubbles and supergiant shells are the Cygnus SB with a diameter of 450 pc in the Milky Way \\citep{ccb80} and the Aquila supershell extending at least 550 pc into the Galactic halo \\citep{m96}. %% %% DB: n\u00e4chster Satz ge\u00e4ndert %% Our solar system itself is embedded in an HI cavity with a size of a few hundred parsecs called the Local Bubble (\\citeauthor{l03} \\citeyear{l03}), and is most likely generated by stellar explosions in a nearby moving group (\\citeauthor{bb02} \\citeyear{bb02}). % Also in external systems like in the LMC \\citep{cm90}, in NGC 253 \\citep{s06} and M101 \\citep{ksh91} such bubbles, holes and shells are observed.\\\\ % The acceleration of the shell promotes Rayleigh-Taylor instabilities and after it is fully fragmented, only the walls of the SB are observed. Through such a chimney the hot pressurized SN-ejecta can escape into the halo. The walls may be subject to the gravitational instability and interstellar clouds can form again, which triggers star formation \\citep[e.g.][]{mk87}. This is observed, for example, on the border of the Orion-Eridanus SB \\citep{lc09}.\\\\ % % The knowledge of SB evolution is crucial for understanding the so-called disk-halo connection and it also gives us information about the chemical evolution of the galaxies, the enrichment of the intergalactic as well as the intracluster medium. The thick extraplanar layer of ionized hydrogen seen in many galaxies has probably been blown out of the disk into the halo by photoionization of OB-stars and correlated SNe \\citep{t06}. % With the high star formation rate of starburst galaxies, the energy released by massive bursts of star formation can even push the gas out of the galactic potential, forming a galactic wind. Outflow rates of $0.1-10 \\, $M$_{\\odot}$/yr are common for starburst driven outflows \\citep{bvc07}. % %% %% DB: ich denke im folgenden Satz meinst Du \"metallicity\" anstatt \"abundance\" %% If the hot and metal enriched material is brought to the surrounding intergalactic medium, it will mix after some time, increasing its metallicity. Galactic winds are observed in nearby galaxies \\citep[e.g.][]{dwh98}, as well as up to redshifts of $z \\sim 5$ \\citep[e.g.][]{sw07,dss02}, thus being an ubiquitous phenomenon in star forming galaxies. % If the energy input is not high enough, the gas will fall back onto the disk due to gravity, after loss of pressure support, forming a galactic fountain \\citep{sf76,a00}. % %% %% DB: Satz erweitert %% In this case, the heavy elements released by SN-explosions are returned back to the ISM in the disk, presumably spread over a wider area, and future generations of stars will incorporate them. Spitoni et al. (2008), also using the Kompaneets approximation, have investigated the expansion of a superbubble, its subsequent fragmentation and also the ballistic motion of the fragments including a drag term, which describes the interaction between a cloud and the halo gas. In addition, these authors have analyzed the chemical enrichment of superbubble shells, and their subsequent fragmentation by Rayleigh-Taylor instabilities, in order to compare the [O/Fe]-ratios of these blobs to high velocity clouds (HVCs). They find that HVCs are not part of the galactic fountain, and even for intermediate velocity clouds (IVCs), which are in the observed range of velocities and heights from the galactic plane, the observed [O/Fe]-ratios can only be reproduced by unrealistically low initial disk abundances. The chemical enrichment of the intergalactic medium will be the subject of a forthcoming paper.\\\\ % This paper is structured as follows: Section 2 shows how the evolution of superbubbles can be described. In Sect.~3 we present the results of this work, which is mainly the expansion of a bubble in space and time and also the onset of Rayleigh-Taylor instabilities in the shell. In Sect.~4 the models are used to analyze two Milky-Way superbubbles. A discussion follows in Sect.~5 and summary and conclusions are presented in Sect.~6. % % % % %________________________________________________SECTION 2_________________________________________________________ % ", "conclusions": "% % %% %% DB: hier w\u00fcrde ich KA ausschreiben; viele Leute lesen ja leider nur die Summary %% We have developed analytical models based on the Kompaneets approximation (KA) in order to derive in a fairly simple and straightforward manner the physical parameters of observed superbubbles and their ambient medium and to gain physical insight into the blow-out phenomenon associated with star forming regions.\\\\ % %% %% DB: hier w\u00fcrde ich noch folgenden Satz einf\u00fcgen %% In this paper we have deliberately refrained from building more complex models, which e.g. include stellar wind and Wolf-Rayet wind phases, because we have put our focus on the important dynamical phenomena of blow-out and fragmentation of the outer shell, and their dependence on the energy input source over time. A more detailed description of superbubble evolution models including further stellar evolutionary phases will be the subject of a forthcoming paper.\\\\ % In our work the key aspect was to work out analytically the dynamics of (unfragmented) superbubbles for different energy input modes. We modified the KA to implement a more realistic way of energy input, i.e. modeling the time sequence of exploding stars in an OB association by including the main sequence lifetime of the massive stars and describing the numbers per mass interval by an initial mass function. We tested three different IMF slopes and also compared the IMF-model to a simple SN-model with an instantaneous release of energy and to a wind model with a constant energy input rate. Two different density distributions of the ISM were applied, a symmetric medium with parameters of the Lockman layer and a high-density, low-scale height pure exponential atmosphere with the star cluster dislocated from the galactic plane. Velocity and acceleration of the shock front can be calculated analytically and the question how many SNe are needed for blow-out into a galactic halo can be answered. The exact position of the outer shock in scale height units when the acceleration starts can be given. Furthermore, the timescale for the development of Rayleigh-Taylor instabilities in a SB shell is calculated, and thus, a fragmentation timescale can be derived. The overall pattern shows that at larger scale heights ($H > 400 \\,$pc), independent of the ISM density, the SN-model needs the highest energy input, followed by the wind-model, whereas an IMF with a steep slope is the most efficient one. The same ranking applies to the blow-out timescales of these models (Figs.~\\ref{time_SN}-\\ref{time_wind} $\\&$ Table \\ref{accel}). At low scale heights ($H \\sim 100 \\, $pc) and moderate or high densities ($n_0 \\ge 5 \\,$cm$^{-3}$), the picture changes completely, i.e. the SN-model requires the lowest energy input and the IMF-model with $\\Gamma_3=-1.7$ is the least efficient one. The explanation is that a single release of energy is more powerful in sweeping up a thin layer of ISM, whereas it is easier for the IMF model with an \\emph{increasing} energy input with time ($L_{\\rm{SB}} \\propto t^{\\delta}$, $0 < \\delta < 1$, Eq.~\\ref{eLSB2}) to sustain the supply of energy over a larger distance. % % % When comparing fragmentation timescales (Figs.~\\ref{frag_SN}-\\ref{frag_wind} $\\&$ Table \\ref{frag}), the IMF-models exhibit the lowest values, favouring flatter IMFs with $\\Gamma_1$ and $\\Gamma_2$. In terms of $\\tilde{y}$ and absolute time, fragmentation happens first for the $\\Gamma_2$-model for SBs driven by the minimum number of SNe for blow-out.\\\\ % % Still, the KA is a rather simple model. It does not account for magnetic fields, ambient pressure, inertia and evaporation of the shell. Also a galactic gravitational field and cooling of the shocked gas inside the cavity are neglected. \\\\ % In our model, we include galactic gravity in a rudimentary way: if a bubble fulfills the blow-out criterion and additionally the shell's acceleration at the top of the bubble at the time of fragmentation exceeds the vertical component of the gravitational acceleration in the disk, the SN-ejecta and fragmenting shell are expelled into the halo. We find that this is true for all bubbles created by an energy input of $N_{\\rm{blow}} \\cdot E_{\\rm{SN}} $.\\\\ % Further analysis of a fragmenting superbubble would also involve to calculate the motion of the shell fragments, and ultimately the dynamics of a galactic fountain like e.g. in \\citet{s08}. A simple ballistic treatment of the motion of blob fragments in a gravitational potential would, however, be too simplistic, as they would for some time experience a drag force due to the outflowing hot bubble interior. The drag would be proportional to the ram pressure of the hot gas and the cross section of a blob, which would, to lowest order, be proportional to the thickness of the fragmented shell; and finally it would depend on the geometry of the blob, which might be taken as spherical. However, due to compressibility, bow-shocks and head-tail structures might subsequently be formed, so that for a realistic treatment numerical simulations would be the best choice.\\\\ % According to MM88 cooling of the bubble interior can be neglected for Milky Way type parameters of the ISM, but should be taken into account for smaller OB-associations or a dense and cool ISM. %% %% DB: wieso wolltest Du den folgenden Satz nicht ins Paper nehmen? Er ist auf jeden Fall richtig. -- ok %% In general, cooling is not important as long as the timescale for radiative cooling is large compared to the characteristic dynamical timescale of the superbubble. % % % It was also shown by MM88 that galactic gravity does not need to be included, because the acceleration of high energy superbubbles exceeds the local gravity. % %. Including a magnetic field would be quite important, but this goes beyond the scope of our analytical model and is left to numerical simulations. \\citet{fmz91} find that the presence of a magnetic field could slow down the expansion of a superbubble. \\citet{s09} argue that the scale height and age of a bubble are underestimated by $\\sim 50 \\, \\%$ when using a Kompaneets model without magnetic fields. However, they cannot produce such narrow superbubbles like W4 with their MHD simulations. Moreover, 3D high resolution numerical simulations \\citep{ab05} show that magnetic tension forces are much less efficient in 3D than in 2D in holding back the expanding bubble.\\\\ % A slightly slower growth of a bubble would be also achieved by taking into account the inertia of the cold massive shell in the calculations. MM88 find a difference of $\\sim 10 \\, \\%$ in radius after comparison with the models of \\citet{s85}, which neglect inertia. Also due to the ambient pressure of the ISM superbubbles should expand more slowly as it was suggested by \\citet{og04}. % In our analytical calculations we find that these effects have to be compensated when we make comparisons to observed bubbles by including a rather high ISM density to prevent SBs from expanding too fast. %% %% DB: was ist damit gemeint? Wieso kann man KA nicht auf \"existierende\" SB anwenden? - gemeint ist, wenn schon eine fr\u00fchere Generation von Superbubbles in der Umgebung dagewesen ist -- lasse es weg %% Furthermore, a clumpy ambient medium cannot be considered by the KA.\\\\ % % We applied our models to the W4 superbubble and the Local Bubble, both in the Milky Way. It is certainly not easy to compare a simple model with an observed SB, which is not isolated. In the region of the W3/W4/W5 bubbles, several complexes and clouds are found and multiple epochs of star formation make it difficult to distinguish, which cluster has formed which bubble at what time. However, it is most likely that the cluster OCl 352 is responsible for driving the evolution of the bubble \\citep{w07}. \\citet{o05} suggest that winds or SNe of previous stellar generations are responsible for earlier clearing of this region, which could explain the low scale height of around 30 pc. Our calculations suggest that the bubble is younger than found by other authors, which is due to the offset of the association more than one scale height above the Galactic plane. %% %% DB: Satz ge\u00e4ndert, ok? ja %% Shifting to lower densities makes it easier to produce a blow-out superbubble in a shorter timescale. This is an important result and should be included in the models.\\\\ The Local Bubble is one of the rare cases for a double-sided bubble, which can be tested with our symmetric superbubble model. From geometrical properties, we estimate an ISM scale height of $\\sim 80 \\,$pc. In order to reproduce size and age of the bubble correctly, we deduce from our models that it was an intermediate density region ($n_0 \\sim 7 \\,$cm$^{-3}$) before the first SN-explosion around 14 Myrs ago. Also this place in the Milky Way is very complex (neighbouring Loop I superbubble), which can't be included in our modeling of superbubbles. \\\\ We conclude that blow-out energies derived in this paper are lower thresholds and might be higher if e.g. magnetic fields play a role. Accordingly, fitting the models to observed bubbles gives an upper limit for densities of the ambient ISM prior to the first SN-explosion. %% %% DB: Noch ein Schlusswort -- sehr gut, danke %% Observers are encouraged to use the model presented here for deriving important physical parameters of e.g. the energy input sources (number of OB stars, richness of cluster etc.), scale heights, dynamical time scales among other quantities. The solutions of the equations derived in detail here are easy to obtain by simple mathematical programs. Theorists may find it useful to compare our analytic results to high resolution numerical simulations in order to separate more complex effects, such as turbulence, mass loading, magnetic fields etc. from basic physical effects, incorporated in our model. % % % % % % %% %% DB: Du k\u00f6nntest auch noch der armen TU f\u00fcr ein paar Gastaufenthalte danken;-) -->> nat\u00fcrlich !!! %%" }, "1402/1402.6324_arXiv.txt": { "abstract": "We present \\textit{XMM-Newton} and \\textit{Chandra} observations of the low-mass X-ray binary XSS J12270--4859, which experienced a dramatic decline in optical/X-ray brightness at the end of 2012, indicative of the disappearance of its accretion disk. In this new state, the system exhibits previously absent orbital-phase-dependent, large-amplitude X-ray modulations with a decline in flux at superior conjunction. The X-ray emission remains predominantly non-thermal but with an order of magnitude lower mean luminosity and significantly harder spectrum relative to the previous high flux state. This phenomenology is identical to the behavior of the radio millisecond pulsar binary PSR J1023+0038 in the absence of an accretion disk, where the X-ray emission is produced in an intra-binary shock driven by the pulsar wind. This further demonstrates that XSS J12270--4859 no longer has an accretion disk and has transformed to a full-fledged eclipsing ``redback'' system that hosts an active rotation-powered millisecond pulsar. There is no evidence for diffuse X-ray emission associated with the binary that may arise due to outflows or a wind nebula. An extended source situated 1.5$'$ from XSS J12270--4859 is unlikely to be associated, and is probably a previously uncatalogued galaxy cluster. ", "introduction": "Since the original discovery of millisecond pulsars \\citep[MSPs;][]{Back82} the prevailing theory of their formation has centered on ``recycling'' by transfer of matter and angular momentum from a close companion star during a low-mass X-ray binary (LMXB) phase \\citep{Alp82,Rad82}. The discovery of the first accreting X-ray MSP, SAX J1808.4--3658 \\citep{Wij98}, provided compelling evidence for this evolutionary sequence \\citep[see also][for a review]{Pat12}. Additional observational support in favor of this hypothesis came with the discovery of the ``missing link'' radio pulsar PSR J1023+0038 (also known as FIRST J102347.6+003841). Optical observations revealed an accretion disk in the system in 2001 \\citep{Wang09}, which was absent after 2003 \\citep{Thor05} and at the time of the radio pulsar discovery in 2009 \\citep{Arch09}. The long-suspected evolutionary connection between LXMBs and radio MSPs was conclusively established when PSR J1824--2452I in the globular cluster M28 was seen to switch between rotation-powered (radio) and high-luminosity accretion-powered (X-ray) pulsations \\citep{Pap13}. Recent radio, X-ray, and $\\gamma$-ray observations revealed that in 2013 June PSR J1023+0038 had undergone another transformation \\citep{Stap14}. Complete cessation of radio pulsations was observed, accompanied by an extraordinary five-fold increase in the \\textit{Fermi} Large Area Telescope (LAT) flux and enhancement in UV and X-ray brightness \\citep{Pat14,Ten14}. These recent findings have revealed that rather than a one-time change from a LMXB to a permanent radio MSP state, for some systems this phase of neutron star compact binary evolution involves recurrent switching between the two states. In the absence of an accretion disk, PSRs J1023+0038 and PSR J1824--2452I belong to the family of so-called ``redbacks'' \\citep[see][and references therein]{Rob11}, namely, eclipsing radio MSPs with relatively massive ($\\gtrsim$0.2 M$_{\\odot}$) non-degenerate companions that are (nearly) Roche-lobe filling. These objects are distinct from ``black widow'' systems, which are bound to very low mass degenerate stars being ablated by the pulsar wind \\citep[][]{Fru88,Stap96}. Extensive X-ray studies of redbacks in globular clusters \\citep{Bog05,Bog10,Bog11a} and the field of the Galaxy \\citep{Arch10,Bog11b,Bog14,Gen13} have revealed that the dominant source of X-rays from these binaries is non-thermal radiation with $\\Gamma\\approx 1-1.5$ originating in an intra-binary shock \\citep{Arons93} generated by the interaction of the pulsar wind and matter from the companion star. The X-ray emission is strongly modulated at the orbital period in all redback systems, with a decline in flux when the secondary star is between the pulsar and the observer. This orbital-phase-dependent X-ray variability appears to be a distinguishing characteristic of these peculiar MSP binaries and provides a convenient way to identify additional redbacks even in the absence of radio pulsations. The Galactic X-ray source XSS J12270--4859 (1RXS J122758.8--485343) has posed a mystery since its discovery. Although initially classified as a cataclysmic variable (Masetti et al.~2006; Butters et al.~2008), subsequent multi-wavelength studies (Pretorius~2009; de Martino et al.~2010; Hill et al.~2011; de Martino et al.~2013; Papitto et al.~2014) revealed that this system closely resembles a quiescent LMXB. XSS J12270--4859 is exceptional in that it was the first LMXB putatively associated with a $\\gamma$-ray source, 1FGL J1227.9--4852 \\citep{Hill11}, now known as 2FGL J1227.7--4853 \\citep{Nolan12}. Deep searches have failed to detect radio pulsation from XSS J12270--4859 during its high optical/X-ray flux state \\citep{Hill11}. Prior to 2012 November/December, the X-ray spectrum of XSS J12270--4859 was well described by a pure powerlaw with $\\Gamma=1.7$ and a 0.1--10 keV luminosity of $\\sim$$6\\times10^{33}$ erg s$^{-1}$ \\citep{deM10,deM13}. The source exhibited occasional intense flares and peculiar, frequent drops in X-ray flux that are not correlated with orbital phase. This behavior is reminiscent of the erratic flux variations observed in PSRs J1023+0038 (Patruno et al.~2014; Tendulkar et al.~2014; S.~Bogdanov in prep.) and J1824--2452I (Papitto et al.~2013; Linares et al.~2014) in their accretion disk states, hinting at a connection with these systems. Recent optical and X-ray observations revealed that XSS J12270--4859 had undergone a substantial decline in brightness and no longer exhibits evidence for an accretion disk (Bassa et al.~2014). The abrupt change to a disk-free system appears to have occured between 2012 November 14 and 2012 December 21. Given the similarities with PSR J1023+0038 in both states, XSS J12270--4859 is presently consistent with hosting a rotation-powered millisecond pulsar. This prediction was confirmed with the recent detection of 1.69-millisecond radio pulsations from the system (Roy et al.~2014). Here we present recently acquired \\textit{XMM-Newton} European Photon Imaging Camera (EPIC) and \\textit{Chandra X-ray Observatory} Advanced CCD Imaging Spectrometer (ACIS) observations of XSS J12270--4859 in its new low flux state. These observations confirm that this source has undergone a metamorphosis to a redback, i.e.~a compact binary containing a rotation-powered millisecond pulsar. The analysis is presented as follows. In \\S 2, we provide details on the observations and data reduction procedures. In \\S3, we present an X-ray variability study. In \\S4 we summarize the phase-averaged and phase-resolved spectroscopy, while in \\S5 we report on an imaging analysis. We provide a discussion and conclusions in \\S6. ", "conclusions": "As reported in Bassa et al.~(2014), the recent decline in optical and X-ray brightness and the disappearance of the previously prominent optical emission lines are indicative of the disappearance of the accretion disk in XSS J12270--4859. The \\textit{XMM-Newton} and \\textit{Chandra} observations of XSS J12270--4859 in its new low flux state presented herein reveal further information regarding this binary: 1.~Previously absent large-amplitude modulation at the binary orbital period is now present. 2.~The time-averaged X-ray luminosity of $\\sim$$(1-7)\\times 10^{32}$ erg s$^{-1}$ is smaller by an order of magnitude relative to the accretion disk dominated state. 3.~The predominantly non-thermal spectrum with spectral photon index of $\\Gamma=1.2$ is significantly harder than the $\\Gamma=1.7$ observed in the high flux state. 4.~The spectral index of the power-law emission appears not to vary substantially even though the flux changes by a factor of $\\sim$1.5--1.7 over an orbital period. These X-ray properties, as well as the optical properties described in \\citet{Bassa14}, are identical to what is observed from ``redback'' radio MSPs, including PSRs J1023+0038 \\citep{Arch09,Arch10,Bog11b}, J1723--2837 \\citep{Faulk04,Craw13,Bog14}, J2215+5135 \\citep{Hes11,Gen13} in the field of the Galaxy, as well as PSR J0024--7204W in the globular cluster 47 Tuc \\citep{Camilo00,Freire03,Bog05}, PSR J1740--5340 in NGC 6397 \\citep{DAm01,Bog10}, and PSR J1824--2452H in M28 \\citep{Bog11a,Pal10}. Therefore, the results obtained herein, and the recent radio pulsar discovery by \\citet{Roy14} further demonstrate that XSS J12270--4859 is a redback, i.e. a compact binary with an active rotation-powered pulsar and no accretion disk. The non-thermal X-rays from redback systems are generated in an intra-binary shock formed by the interaction of the pulsar wind with material from the close companion star \\citep{Arons93}. As shown in \\citet{Bog11b}, the decline in X-ray flux at $\\phi_b\\approx0.25$ seen in PSR J1023+0038 can be reproduced by a simple geometric model in which the intra-binary shock is partially occulted by the bloated secondary star. In this scenario, the depth and phase extent of the X-ray eclipse require that the shock be situated at or very near the surface of the secondary star on the side exposed to the pulsar wind. There is no indication of extended X-ray emission associated with XSS J12270--4859 that could arise due to outflows in the LMXB phase or the interaction of the pulsar wind with the ambient medium. An adjacent diffuse source is likely not associated with the binary and is probably a background galaxy cluster. Given that only two MSPs, PSRs B1957+21 \\citep{Stap03} and J2124--3358 \\citep{Hui06} show evidence for X-ray bow shocks, the absence of nebular emission associated with XSS J12270--4859 does not imply the lack of a pulsar wind. The lack of knowledge of the orbital separation, the companion mass, and the pulsar spin-down luminosity ($\\dot{E}$) for XSS J12270--4859, do not allow meaningful constraints on the physics of the intrabinary shock. Nevertheless, given the nearly identical X-ray properties compared to other redbacks it is possible to place crude constraints on the energetics of the MSP in XSS J12270--4859. In particular, if we consider that the shock luminosity at X-ray maximum is $\\sim$$2\\times10^{-3}\\dot{E}$ for PSRs J1023+0038 \\citep{Arch10,Bog11b} and J1723--2837 \\citep{Bog14}, the implied spin-down of the pulsar in XSS J12270--4859 is $(0.7-5)\\times10^{35}$ erg s$^{-1}$ assuming a distance range $1.4-3.6$ kpc. This suggests that the pulsar in XSS J12270--4859 may be among the small number of energetic MSPs with $\\dot{E}\\gtrsim10^{35}$ erg s$^{-1}$. The 0.1--100 GeV $\\gamma$-ray luminosity of 2FGL J1227.7--4853, the putative $\\gamma$-ray counterpart of XSS J12270--4859, of $L_{\\gamma}=(0.8-5)\\times10^{34}$ erg s$^{-1}$ provides an independent hard lower limit on the $\\dot{E}$ of the pulsar, if we assume a $\\gamma$-ray conversion efficiency of 100\\%. For efficiencies typical of the MSP population \\citep[$\\lesssim$50\\%; see Table 10 in][]{Abdo13}), this yields $\\dot{E}\\gtrsim10^{35}$ erg s$^{-1}$, in general agreement with the limit obtained from the X-ray luminosity. Continued timing of the radio pulsar would allow a determination of the true $\\dot{E}$ of this pulsar. The redback nature of the XSS J12270--4859 binary determined from the X-ray analysis presented above implies that the pulsar should undergo radio eclipses of frequency-dependent duration around superior conjunction. These eclipses are caused by intra-binary plasma, presumably emanating from the companion star. For most redback and black widow systems that are detectable in radio pulsations, the radio emission is eclipsed for $\\lesssim$50\\% of the orbit, although there is indication that for some systems this fraction can be close to 100\\% \\citep[see][for the case of PSR J1311--3430]{Ray13}. The non-detections with the Parkes telescope \\citep{Bassa14} and only a single, brief detection with the GMRT \\citep{Roy14} despite deep pulsation searches, are indicative of a high radio eclipse fraction for XSS J12270--4859 that likely varies from orbit to orbit. The large $\\dot{E}$ estimated above is consistent with this scenario, since an energetic wind would drive off material from the surface of companion at a higher rate than an MSP with a lower $\\dot{E}$. This may result in occasional severe enshrouding by this stripped material, causing the pulsar to be eclipsed at radio frequencies for a larger fraction of the orbit. This would make XSS J12270--4859 a so-called ``hidden'' MSP, as postulated by \\citet{Tav91}. The X-ray properties of XSS J12270--4859 in both the present disk-free radio pulsar state and the accretion disk-dominated state \\citep[see][]{deM10,deM13} are virtually identical to those of PSRs J1023+0038 \\citep{Arch10,Bog11b,Pat14} and J1824--2452I \\citep{Pap13,Lin14}. This remarkable consistency in X-ray characteristics provides a powerful discriminant for identifying more such MSP binaries, especially ones that are heavily enshrouded and hence cannot be identified via radio pulsation searches." }, "1402/1402.1168_arXiv.txt": { "abstract": "We present a statistical study of velocities of Ly$\\alpha$, interstellar (IS) absorption, and nebular lines and gas covering fraction for Ly$\\alpha$ emitters (LAEs) at $z\\simeq 2$. We make a sample of 22 LAEs with a large Ly$\\alpha$ equivalent width (EW) of $\\gtrsim 50$\\,\\AA\\ based on our deep Keck/LRIS observations, in conjunction with spectroscopic data from the Subaru/FMOS program and the literature. We estimate the average velocity offset of Ly$\\alpha$ from a systemic redshift determined with nebular lines to be $\\Delta v_{\\rm Ly\\alpha}=234\\pm 9$ km s$^{-1}$. Using a Kolmogorv-Smirnov test, we confirm the previous claim of Hashimoto et al. (2013) that the average $\\Delta v_{\\rm Ly\\alpha}$ of LAEs is smaller than that of LBGs. Our LRIS data successfully identify blue-shifted multiple IS absorption lines in the UV continua of four LAEs on an individual basis. The average velocity offset of IS absorption lines from a systemic redshift is $\\Delta v_{\\rm IS}=204 \\pm 27$ km s$^{-1}$, indicating LAE's gas outflow with a velocity comparable to typical LBGs. Thus, the ratio, $ R^{\\rm Ly\\alpha}_{\\rm IS} \\equiv \\Delta v_{\\rm Ly\\alpha}/\\Delta v_{\\rm IS}$ of LAEs, is around unity, suggestive of low impacts on Ly$\\alpha$ transmission by resonant scattering of neutral hydrogen in the IS medium. We find an anti-correlation between Ly$\\alpha$ EW and the covering fraction, $f_c$, estimated from the depth of absorption lines, where $f_c$ is an indicator of average neutral hydrogen column density, $N_{\\rm HI}$. The results of our study support the idea that $N_{\\rm HI}$ is a key quantity determining Ly$\\alpha$ emissivity. ", "introduction": "Ly$\\alpha$ Emitters (LAEs) are an important population of high-$z$ star-forming galaxies in the context of galaxy formation. LAEs at $z=2-7$ and beyond $z=7$ are found by narrow-band (NB) imaging observations based on an NB excess resulting from their prominent Ly$\\alpha$ emission \\citep[e.g., ][]{2010ApJ...711..928C,2007ApJ...667...79G,2012ApJ...744..110C,2008ApJS..176..301O,2008ApJ...677...12O,2010ApJ...723..869O,2010ApJ...725..394H,2007ApJ...660.1023F,2011ApJ...734..119K,2006ApJ...648....7K,2012ApJ...752..114S}. Observational studies on a morphology and spectral energy distribution (SED) of LAEs reveal that such a galaxy is typically young, compact, less-massive, less-dusty than other high-$z$ galaxy populations, and a possible progenitor of Milky Way mass galaxies \\citep[e.g., ][]{2011ApJ...743....9G,2011ApJ...733..114G,2010MNRAS.402.1580O,2007ApJ...671..278G,2011ApJ...740...71D,2008ApJ...681..856R,2012MNRAS.424.1672D}. Additionally, LAEs are used to measure the neutral hydrogen fraction at the reionizing epoch, because Ly$\\alpha$ photons are absorbed by intergalactic medium (IGM). Ly$\\alpha$ emitting mechanism is not fully understood due to the highly-complex radiative transfer of Ly$\\alpha$ in the interstellar medium (ISM). Many theoretical models have predicted that the neutral gas and/or dust distributions surrounding central ionizing sources are closely linked to the Ly$\\alpha$ emissivity \\citep[e.g., ][]{1991ApJ...370L..85N, 2008ApJ...678..655F,2013ApJ...766..124L,2009ApJ...704.1640L,2007ApJ...657L..69L,2013arXiv1302.7042D,2013arXiv1308.1405Z,2010ApJ...716..574Z,2012arXiv1209.5842Y}. Thus, resonant scattering in the neutral ISM can significantly attenuate the Ly$\\alpha$ emission. Ly$\\alpha$ emissivity may not only depend on the spatial ISM distribution, but on the gas kinematics as well. The large-scale galactic outflows driven by starbursts or active galactic nuclei could allow Ly$\\alpha$ photons to emerge at wavelengths where the Gunn-Peterson opacity is reduced, and consequently enhance the Ly$\\alpha$ emissivity, particularly in the high-$z$ Universe \\citep[e.g., ][]{2010MNRAS.408..352D}. The outflow may also blow out the Ly$\\alpha$ absorbing ISM. The gas kinematics of LAEs has been evaluated from the Ly$\\alpha$ velocity offset ($\\Delta v_{{\\rm Ly\\alpha}}$) with respect to the systemic redshift ($z_{\\rm sys}$) traced by nebular emission lines (e.g, H$\\alpha$, $[$O {\\sc iii}$]$) from their H {\\sc ii} regions. In the past few years, deep NIR spectroscopic studies have detected nebular emission lines from $\\sim10$ LAEs at $z=2-3$, and measured their $\\Delta v_{{\\rm Ly\\alpha}}$ \\citep{2011ApJ...730..136M, 2013ApJ...765...70H,2013A&A...551A..93G,2011ApJ...729..140F,2013ApJ...775...99C}. The Ly$\\alpha$ emission lines for these LAEs are redshifted from their $z_{\\rm sys}$ by a $\\Delta v_{{\\rm Ly\\alpha}}$ of $200-300$ km s$^{-1}$. \\citet{2013ApJ...765...70H} find an anti-correlation between Ly$\\alpha$ equivalent width (EW) and $\\Delta v_{{\\rm Ly\\alpha}}$ in a compilation of LAE and LBG samples. This result is in contrast to a simple picture where Ly$\\alpha$ photons more easily escape in the presence of a galactic outflow. However, the Ly$\\alpha$ velocity offset is thought to increase with both resonant scattering in H {\\sc i} gas clouds as well as galactic outflow velocity \\citep[e.g., ][]{2006A&A...460..397V,2008A&A...491...89V}. The anti-correlation could result from a difference in H {\\sc i} column density ($N_{\\rm HI}$) rather than outflowing velocity. The gas kinematics can be investigated more directly from the velocity offset between interstellar (IS) absorption lines of the rest-frame UV continuum and $z_{\\rm sys}$ (IS velocity offset; $\\Delta v_{{\\rm IS}}$). The IS velocity offset traces the speed of outflowing gas clouds, and may help to distinguish the two effects on $\\Delta v_{{\\rm Ly\\alpha}}$. For UV-continuum selected galaxies, the $\\Delta v_{{\\rm IS}}$ has been measured for $>100$ objects \\citep[e.g.][]{2001ApJ...554..981P,2012MNRAS.427.1973C,2012ApJ...745...33K,2013ApJ...777...67S,2010ApJ...717..289S}. \\citet{2010ApJ...717..289S} find that LBGs have an average of $\\langle\\Delta v_{{\\rm IS}}\\rangle = -164$ km s$^{-1}$ in their sample of 89 LBGs at $z\\sim3$. This statistical study indicates ubiquitousness of galactic outflow in LBGs. However, there have been no NB-selected galaxies with a $\\Delta v_{{\\rm IS}}$ measurement to date except for a stacked UV spectrum in \\citet{2013ApJ...765...70H}. This is because it is difficult to estimate $\\Delta v_{{\\rm IS}}$ for individual LAEs, especially for galaxies with a large Ly$\\alpha$ EW of $\\gtrsim 50$\\,\\AA\\, due to their faint UV-continuum emission, while $\\Delta v_{{\\rm IS}}$ are measured for some UV-selected galaxies with EW(Ly$\\alpha) \\sim50$\\,\\AA\\, \\citep[e.g., ][]{2010ApJ...719.1168E}. A statistical investigation of Ly$\\alpha$ kinematics for LAEs could shed light on the physical origin of the anti-correlation and the underlying Ly$\\alpha$ emitting mechanism. This is the second paper in the series exploring the Ly$\\alpha$ emitting mechanisms\\footnote{The first paper presents a study on LAE structures \\citep{2014ApJ...785...64S}.}. In this paper, we present the results of our optical and NIR spectroscopy for a large sample of $z=2.2$ LAEs with Keck/LRIS and Subaru/FMOS to verify possible differences of $\\Delta v_{{\\rm Ly\\alpha}}$ and $\\Delta v_{{\\rm IS}}$ between LAEs and LBGs. These spectroscopic observations are in an extension of the project of \\citet{2013ApJ...765...70H} aiming to confirm the anti-correlation between Ly$\\alpha$ EW and $\\Delta v_{{\\rm Ly\\alpha}}$. The organization of this paper is as follows. In Section \\ref{sec_targets}, we describe the details of the LAEs targeted for our spectroscopy. Next, we show our optical and NIR spectroscopic observations in Section \\ref{sec_observation}. We present methods to reduce the spectra, and to measure kinematic quantities such as $\\Delta v_{{\\rm Ly\\alpha}}$ and $\\Delta v_{{\\rm IS}}$ in \\ref{sec_reduction}. We perform SED fitting to derive physical properties in Section \\ref{sec_sed_fitting}. We compare kinematic properties between LAEs and LBGs in Section \\ref{sec_results}, and discuss physical origins of possible differences in these quantities in Section \\ref{sec_discussion}. In the last section Section \\ref{sec_conclusion}, we summarize our findings. Throughout this paper, we adopt the concordance cosmology with $(\\Omega_m, \\Omega_\\Lambda, h)=(0.3, 0.7, 0.7)$, \\citep{2011ApJS..192...18K}. All magnitudes are given in the AB system \\citep{1983ApJ...266..713O}. ", "conclusions": "We carry out deep optical spectroscopy for our large sample of LAEs at $z=2.2$ in order to detect their Ly$\\alpha$ lines with Keck/LRIS. We compare redshifts of the Ly$\\alpha$ and nebular emission lines detected with Subaru/FMOS, and calculate $\\Delta v_{\\rm Ly\\alpha}$ for new 11 LAEs. This observation doubles the sample size of LAEs with a $\\Delta v_{\\rm Ly\\alpha}$ measurement in literatures. The conclusions of this study are summarized below. \\begin{itemize} \\item Almost all of our new LAEs have a $\\Delta v_{\\rm Ly\\alpha}$ of $\\sim200$ km s$^{-1}$ which is systematically-smaller than that of LBGs. Using 22 LAEs with $\\Delta v_{\\rm Ly\\alpha}$ measurements taken from our new observations and the literature, we definitively confirm the anti-correlation between Ly$\\alpha$ EW and $\\Delta v_{\\rm Ly\\alpha}$ suggested by previous work. \\item Long exposure times and the high sensitivity of LRIS at blue wavelengths enabled us to successfully detect IS absorption lines against faint UV continua from four individual LAEs. These IS absorption lines are found to be blueshifted from the systemic redshift by $200-300$ km s$^{-1}$, indicating strong gaseous outflows are present even in LAEs. \\item We estimate $R_{\\rm IS}^{\\rm Ly\\alpha}$ ($\\equiv\\Delta v_{\\rm Ly\\alpha}/\\Delta v_{\\rm IS}$) that would be a quantity sensitive to $N_{\\rm HI}$ for the four UV continuum-detected LAEs. We find the value of $R_{\\rm IS}^{\\rm Ly\\alpha}$ in LAEs to be smaller than that of LBGs, indicating a lower $N_{\\rm HI}$ in LAEs. We performed a test for correlations between $R_{\\rm IS}^{\\rm Ly\\alpha}$ and physical properties inferred from SED fitting. As a result, we tentatively conclude that SFR may be most closely related to $R_{\\rm IS}^{\\rm Ly\\alpha}$. The correlation may suggest that the star formation preferentially occurs in systems with large amounts of neutral hydrogen gas, which would have a larger value of $R_{\\rm IS}^{\\rm Ly\\alpha}$. \\item We estimate the covering fraction, $f_c$, of surrounding H {\\sc i} gas from the depth of LIS absorption lines the four LAEs. We identify a tentative trend for $f_c$ to decrease with increasing Ly$\\alpha$ EW, as suggested by a study for LBGs in \\citet{2013ApJ...779...52J}. A central source being covered by patchy H {\\sc i} gas clouds would lead to a small flux-averaged $N_{\\rm HI}$ corresponding to a small $R^{\\rm Ly\\alpha}_{\\rm IS}$. In this condition, Ly$\\alpha$ photons could easily escape less affected by resonant scattering in the clouds. \\item The results of our kinematic analyses support the idea that the H {\\sc i} column density is a key quantity determining Ly$\\alpha$ emissivity. \\end{itemize} In this kinematic study, we obtain $\\Delta v_{\\rm IS}$, $R_{\\rm IS}^{\\rm Ly\\alpha}$, and $f_c$ only for objects with a moderate Ly$\\alpha$ EW of $20-100$\\,\\AA\\, which overlaps with the Ly$\\alpha$ EW range of LBG samples in e.g., \\citet{2003ApJ...588...65S}. We need to estimate these quantities for objects with a higher Ly$\\alpha$ EW in order to check whether such objects follow the kinematic trends found in this study." }, "1402/1402.4051_arXiv.txt": { "abstract": "{Super-Alfv\\'enic shocks associated with coronal mass ejections (CMEs) can produce radio emission known as Type II bursts. In the absence of direct imaging, accurate estimates of coronal electron densities, magnetic field strengths, and Alfv\\'en speeds are required to calculate the kinematics of shocks. To date, 1D radial models have been used, but these are not appropriate for shocks propagating in non--radial directions.} {Here, we study a coronal shock wave associated with a CME and Type II radio burst using 2D electron density and Alfv\\'en speed maps to determine the locations that shocks are excited as the CME expands through the corona.}{Coronal density maps were obtained from emission measures derived from the Atmospheric Imaging Assembly (AIA) on board the \\emph{Solar Dynamic Observatory} ($SDO$) and polarized brightness measurements from the Large Angle and Spectrometric Coronagraph (LASCO) on board the \\emph{Solar and Heliospheric Observatory} ($SOHO$). Alfv\\'en speed maps were calculated using these density maps and magnetic field extrapolations from the Helioseismic and Magnetic Imager ($SDO$/HMI). The computed density and Alfv\\'en speed maps were then used to calculate the shock kinematics in non-radial directions.}{Using the kinematics of the Type II burst and associated shock, we find our observations to be consistent with the formation of a shock located at the CME flanks where the Alfv\\'en speed has a local minimum.}{The 1D density models are not appropriate for shocks that propagate non--radially along the flanks of a CME. Rather, the 2D density, magnetic field and Alfv\\'en speed maps described here give a more accurate method for determining the fundamental properties of shocks and their relation to CMEs.} ", "introduction": "Solar flares and coronal mass ejections (CMEs) are energetic manifestations of the restructuring coronal magnetic fields. As a CME travels through the corona, its velocity can become sufficiently larger than the background coronal Alfv\\'en speed, causing a shock wave to form along its leading edge and/or flanks \\citep{Cho2007}. It is within these shocks that electrons can be accelerated to near-relativistic energies to produce Type II radio signatures in low-frequency dynamic spectra \\citep[see,][]{Carley2013}. Although CMEs and Type II radio bursts have been studied for many decades \\citep[see,][]{Pick2006}, there remains unanswered questions in relation to where CME shocks are formed and how these phenomena are associated with the generation of Type II bursts. It has long been suggested that Type II radio bursts are signatures of coronal shocks \\citep{Wild1950,Uchida1960}. They appear as features slowly drifting toward lower frequencies at decimetric to kilometric wavelengths in dynamic radio spectra. This drift is the result of plasma emission generated by a super-Alfv\\'enic shock traveling upwards in the corona \\citep{Cane1981}, where density decreases with height. When direct low-frequency imaging is not available, a key problem is therefore the accurate calculation of the coronal density and Alfv\\'en speed distributions with height; This is to relate the Type II emission frequency to its height and to investigate the direction of the shock propagation. \\begin{figure}[h] \\begin{center} \\includegraphics[trim=1cm 1cm 1cm 1cm,clip=true,width=9cm,angle=0]{plot_density-eps-converted-to.pdf} \\caption{A variety of commonly used electron density models, which can give very different height estimates. For example, a density of $10^7$ cm$^{-3}$ can be located at $\\sim$1.4~$R_{\\odot}$ or $\\sim$2.0~$R_{\\odot}$, depending on the density model. Plotted for comparison, the set of density measurements obtained with Mars Express (MEX) between the years 2006--2008 \\citep{Verma2013}.} \\label{Fig_1_models} \\end{center} \\end{figure} The plasma frequency is related to the density of the emitting plasma by $f_\\mathrm{p}= C\\sqrt{n_\\mathrm{e}} $, where $C=8980$ Hz cm$^{3/2}$ is a constant. To derive the shock kinematics, electron density models are normally employed to relate the plasma density to its coronal height and velocity. Specifically, the shock radial velocity is related to the plasma frequency drift rate, $\\mathrm{d} f_\\mathrm{p}/\\mathrm{d} t$, and the electron density model, $ n_\\mathrm{e}(r)$, by, \\begin{equation} v= \\frac{2\\sqrt{n_\\mathrm{e}}}{C} \\left( \\frac{\\mathrm{d} n_\\mathrm{e}}{\\mathrm{d} r} \\right)^{-1} \\frac{\\mathrm{d} f_\\mathrm{p}}{\\mathrm{d} t} , \\end{equation} \\begin{figure} \\centering \\includegraphics[trim=0cm 0cm 0cm 0cm, clip=true, width=9cm, angle=0]{Fig_2_one_column-eps-converted-to.pdf} \\caption{\\emph{GOES--15} soft X--ray light curve showing an X1.4 flare starting at 10:29:00~UT on 2011 Septemeber 22 (top) and the associated RSTO dynamic spectrum showing a Type II radio burst starting at 10:39:06~UT (bottom). This burst shows fundamental (F) and harmonic (H) emission.} \\label{RSTO} \\end{figure} where $v$ is the shock velocity, $n_\\mathrm{e}$ is the coronal plasma electron density and $r$ is the heliocentric radial distance. Different coronal density models can therefore lead to differing kinematics. Several density models have been used to calculate the Type II shock position, such as the \\citet{Newkirk1961} model derived from the barometric height behavior of a gravitationally stratified corona, the \\citet{Saito1977} model obtained from measurements of coronal polarized brightness ($pB$), and the \\citet{Mann1999} model, derived from solutions of magnetohydrodynamic equations. The use of an arbitrary radial density model can lead to an inaccurate calculation of shock heights and hence velocities. This is due to the significant difference in the shock height derived from different density models. An example is shown in Fig.~\\ref{Fig_1_models}, where an electron density of $10^7$~cm$^{-3}$, occurs at a height of 1.4~$R_{\\odot}$ for the \\citet{Mann1999} model, while occurs at a height of 2.0~$R_{\\odot}$ for the Allen-Baumbach model \\citep{Allen1947}. Another reason for an inaccurate shock height and velocity calculation is the time variability of the coronal density distribution \\citep{Parenti2000,Bemporad2003}, which is not taken into account with typically employed density models. Finally, if a Type II radio source propagates non-radially, speeds derived from 1D radial density models underestimate the true shock speed. Knowledge of the coronal density in the 2D plane is therefore crucial for determining accurate shock kinematics, while the knowledge of the 2D Alfv\\'en speed is important to determine when the propagating shock reaches a super-Alfv\\'enic speed. A 2D analytic model of the Alfv\\'en speed was presented by \\citet{Warmuth2005}. However, due to the temporal variability of the density and magnetic field in the corona, it is important to use observational data specific to the radio emission time rather than a generic analytic model for the Alfv\\'en speed. \\begin{figure*} \\begin{center} \\includegraphics[trim=6.1cm 2.5cm 6.5cm 2.9cm,clip=true,width=6.7cm,angle=90]{sdo_running_diff_cme_new-eps-converted-to.pdf} \\caption{{\\it SDO}/AIA running--difference images of the erupting CME at 10:36:39--10:40:24~UT in the 211~\\AA\\ passband. The CME leading edge is outside the field--of--view of $SDO$ in panel (a), while the propagation of the coronal bright front is visible at the location of the CME flanks at 10:38:38~UT panel (b) and at 10:40:24~UT panel (c).} \\label{SDO_cme} \\end{center} \\end{figure*} \\begin{figure*} \\begin{center} \\includegraphics[trim=6cm 3.8cm 5.9cm 4cm,clip=true,width=7.9cm,angle=90]{Intensity_Density_SDO_new-eps-converted-to.pdf} \\caption{{\\it SDO}/AIA 211~\\AA\\ passband intensity at 10:39:24~UT (a) and the corresponding density map (b). The contours show the harmonic plasma emission frequencies related to the measured coronal densities at 150, 200 and 300~MHz. The contours have been smoothed by a boxcar average of 3.6 arcsecs for display purposes, while} dashed lines are at 1.0~$R_{\\odot}$ and 1.2~$R_{\\odot}$. \\label{Density_map_sdo} \\end{center} \\end{figure*} Here, a new method to calculate coronal densities, magnetic field strengths, and Alfv\\'en speeds in a 2D plane is presented. These 2D maps are then used to calculate the kinematics and the direction of propagation of a Type II radio burst observed on 2011 September 22. The Type II radio burst and coronal mass ejection (CME) are described in Section~\\ref{observations}. The observational method and the models used to produce the 2D density maps are described in Section~\\ref{density}, while the Alfv\\'en maps are shown in Section~\\ref{Alfven}. Results are presented in Section~\\ref{results}, and finally, conclusions are discussed in Section~\\ref{discussion}. ", "conclusions": "\\label{discussion} A new method to obtain semi--empirical 2D maps of coronal density and Alfv\\'en speed has been presented. Using these calculated maps, a decametric Type II burst and a CME that occurs on 2011 September 22, were analyzed to understand their relationship. The analysis of the Type II kinematics, using the 2D density maps, have been performed for non--radial profiles in the POS starting from the active region. Previously, analytical radial density models have been employed to calculate the radio burst kinematics \\citep[e.g., ][]{Robinson1985,Vrsnak2002}. Using a 2D density map at the specific time of the radio burst and performing a non--radial analysis, we were able to relate the Type II burst with the CME and discriminate which portion of the CME front was responsible for triggering the Type II emission. Comparing the CME kinematics calculated for different non-radial traces to the Type II burst we find evidence for the shock to be formed in the flank region of the CME. This agrees with previous interpretations \\citep[e.g.,][]{Classen2002,Mancuso2004}. By calculating the height of the upstream shock front (70~MHz) using the 2D density map and relating it to the position of the CME, we were able to locate the position of the shock in the CME flanks at a height of $\\sim$1.6~$R_{\\odot}$. In addition, we were able to compare the Type II burst shock speed with the local Alfv\\'en speed using the 2D Alfv\\'en speed maps . The Type II burst speed was found to be super-Alfv\\'enic \\citep[e.g.,][]{Vrsnak2008,Cho2005,Klein1999}). Furthermore, the decelerating phase of the Type II shock can be related with the fading of the Type II burst in the dynamic spectrum as the shock approaches the local Alfv\\'en speed. The method used to construct 2D density and Alfv\\'en speed maps presented in this work can be used for the study of all Type II radio bursts. These time specific semi-empirical maps improve the calculation of shock kinematics and represent a step forward from the standardized radial density profiles currently employed. Further investigation relating the kinematics of the limb event of Type IIs and CMEs using the 2D maps and radio imaging is necessary. This can be done with Type II radio bursts in the NRH frequency range (i.e., 150--445~MHz) where direct radio imaging can be used to locate the position of the burst, or with lower frequency imaging telescopes, such as the LOw Frequency ARray \\citep[LOFAR;][]{VanHaarlem2013}." }, "1402/1402.4267_arXiv.txt": { "abstract": "The dispersion of small amplitude, impulsively excited wave trains propagating along a magnetic flux tube is investigated. The initial disturbance is a localized transverse displacement of the tube that excites a fast kink wave packet. The spatial and temporal evolution of the perturbed variables (density, plasma displacement, velocity, \\ldots) is given by an analytical expression containing an integral that is computed numerically. We find that the dispersion of fast kink wave trains is more important for shorter initial disturbances (i.e. more concentrated in the longitudinal direction) and for larger density ratios (i.e. for larger contrasts of the tube density with respect to the environment density). This type of excitation generates a wave train whose signature at a fixed position along a coronal loop is a short event (duration $\\simeq 20$~s) in which the velocity and density oscillate very rapidly with typical periods of the order of a few seconds. The oscillatory period is not constant but gradually declines during the course of this event. Peak values of the velocity are of the order of 10~km~s$^{-1}$ and are accompanied by maximum density variations of the order of 10--15\\% the unperturbed loop density. ", "introduction": "In the last two decades abundant evidence about waves and oscillations in the solar atmosphere has been gathered. Events of various nature (standing and propagating waves) and in various environments (chromosphere, prominences, active regions, coronal holes) have been detected. Here we present some examples, while emphasizing that the following list is not exhaustive. In the chromosphere, propagating and standing transverse waves have been detected in spicules \\citep{zaqarashvili2007,okamoto2011}, in mottles \\citep{kuridze2012}, and in active region fibrils \\citep{pietarila2011}. Regarding solar prominences, \\citet{lin2007} used high-resolution H$\\alpha$ filtergrams and observed traveling transverse waves in thin filament threads. \\citet{okamoto2007} found transverse oscillations of flowing active region filament threads observed with {\\em Hinode} SOT. Oscillatory events of different nature have been observed in coronal loops: transverse oscillations of active region loops triggered by a disturbance that propagates from the central flare site \\citep{aschwanden1999,nakariakov1999}; high-frequency, compressible waves traveling along an active region coronal loop \\citep{williams2001,williams2002}; Doppler shift oscillations caused by waves propagating in the upper part of coronal loops \\citep{tian2012}; ubiquitous waves in the solar corona propagating upwards along magnetic field lines \\citep{tomczyk2007,tomczyk2009}; etc. Reviews about waves and oscillations in spicules, prominences, and coronal structures can be found in \\citet{zaqarashvili_erdelyi2009}, \\citet{arregui2012}, \\citet{nakariakov_verwichte2005}, and \\citet{demoortel_nakariakov2012}. Many of these events are examples of magnetic flux tubes being perturbed by an external agent. The transverse loop oscillations described by \\citet{aschwanden1999} and \\citet{nakariakov1999} are a remarkable case because time series of EUV images allow to see the lateral swaying of a coronal loop. This particular phenomenon has been interpreted as a standing fast kink mode oscillation. Fast kink (i.e. transverse) waves propagating along magnetic flux tubes are more frequent in the literature than their standing oscillation counterparts. They have been observed not only in coronal loops \\citep[e.g.][]{tian2012}, but also in spicules \\citep{zaqarashvili2007}, in mottles \\citep{kuridze2012}, in active region fibrils \\citep{pietarila2011}, and in filament threads \\citep{lin2007,okamoto2007}. What is common to all these phenomena is that an external excitation causes the transverse displacement of a magnetic tube and that this perturbation propagates along the tube, where it is usually detected as a time variation of the Doppler velocity or the magnetic tube position. These propagating waves can be excited, e.g., by a periodic driver acting at a fixed position of the magnetic tube. This has been the mechanism invoked by \\citet{VTG2010} and \\citet{TGV2010} to explain the waves propagating along the coronal magnetic field observed by \\citet{tomczyk2007} and \\citet{tomczyk2009}. A fast kink wave propagating along a uniform, cylindrical magnetic tube produces a periodic transverse motion of the tube such as that in Figure~\\ref{fig_cylinder_kink}(a). \\begin{figure}[ht!] \\centerline{(a)\\includegraphics[width=0.3\\textwidth,angle=0]{f1a.eps}\\hspace{1cm}(b)\\includegraphics[width=0.3\\textwidth,angle=0]{f1b.eps}} \\caption{Uniform cylindrical magnetic flux tube subject to (a) a fast kink mode (i.e. transverse, periodic along the tube) perturbation and (b) a localized transverse perturbation. (A color version of this figure is available in the online journal.)} \\label{fig_cylinder_kink} \\end{figure} Propagating transverse waves can also be generated by an impulsive, concentrated transverse displacement of the magnetic tube (see Figure~\\ref{fig_cylinder_kink}(b)). Such a perturbation of the magnetic tube is a superposition of fast kink eigenmodes, each with its own amplitude. In the absence of dispersion, the initial wave form would keep its original shape during its propagation, that is, the bulge of Figure~\\ref{fig_cylinder_kink}(b) would simply propagate unaltered along the cylinder. Fast kink eigenmodes of a magnetic tube are dispersive, however, and so as time evolves the initial hump transforms into an oscillatory train containing several maxima and minima. In other words, away from the excitation point the magnetic tube can suffer not one, but several lateral oscillations about its equilibrium position as the wave train passes by. Hence, transverse oscillations of magnetic tubes do not necessarily require the presence of a continuous driver. The purpose of this work is to study the dispersion of linear fast kink wave trains propagating along uniform, cylindrical magnetic flux tubes. These wave trains are excited by an impulsive, localized initial disturbance (Figure~\\ref{fig_cylinder_kink}(b)). To solve this initial value problem we use a technique based on the method of Fourier integrals \\citep[cf.][Section~11.2]{whitham1974}, that consists of expressing the initial perturbation as a sum of eigenmodes. This means that the properties of the magnetic tube eigenmodes need to be well known in advance, and for this reason their main features are summarized in Sections~\\ref{sect_normal_modes} and \\ref{sect_cont_modes}. Next the method of Fourier integrals is applied to the cylindrical magnetic tube, and an analytical expression describing the spatial and temporal variation of perturbed variables is obtained (Section~\\ref{sect_fourier}). This formula is expressed in terms of an integral that contains contributions from all eigenmodes, with amplitudes that depend exclusively on the initial conditions. Accurate approximations to this integral are computed numerically. The propagation of a concentrated transverse disturbance is considered in Section~\\ref{subsect_numerics}, paying special attention to the dispersion of the wave train during its propagation. An application to coronal loop transverse oscillations is carried out in Section~\\ref{sect_loop_application} and a discussion of the results and our conclusions are presented in Section~\\ref{sect_conclusions}. ", "conclusions": "\\label{sect_conclusions} We have analyzed the propagation of impulsively excited, linear fast kink wave packets in cylindrical magnetic tubes. The method we use is based in decomposing the initial perturbation in a sum of eigenmodes. We have imposed a concentrated transverse disturbance of the magnetic tube and have found that most of its energy is imparted to the global kink mode. The concentrated transverse impulse is the sum of many different global fast kink modes with their own amplitude and that travel at their own group velocity in the packet. For this reason, the wave train dispersion is determined by two key ingredients: on one hand, the dependence of the group velocity on the longitudinal wavenumber, that is represented in Figures~\\ref{fig_dr_over}(c) and \\ref{fig_cg_vae}. On the other hand, the amplitude of the global kink mode as a function of the longitudinal wavenumber (Figure~\\ref{fig_amplr}). We have found that large internal to external density ratios and especially shorter initial perturbations enhance wave dispersion. The signature of a wave train at a fixed point in a coronal loop has been investigated using the radial velocity and the density perturbation as measurable quantities. We have shown that both variables display very fast oscillations with periods varying between 1 and 4~s and with detectable amplitudes. These values have been obtained for a loop radius $a=250$~km and an internal Alfv\\'en speed $\\vai=500$~km~s$^{-1}$, although these are just illustrative values. If the loop radius is doubled or its internal Alfv\\'en speed is halved, then the obtained periods and time scales are doubled. That is, both the horizontal and vertical scales in Figures~\\ref{fig_vae2_delta1_z}(b), (d) and \\ref{fig_vae2_delta1_rho_z}(b) are doubled. In view of the diversity of both $a$ and $\\vai$ in coronal loops, we conclude that the dispersion of a transverse coronal loop perturbation can be detected as periodic variations of the velocity and density with periods ranging from a fraction of a second to tens of seconds. \\citet{murawski1998} performed a numerical simulation of impulsively generated, small-amplitude waves in a slab coronal loop model. Regarding transverse oscillations, these authors concluded that they could be detected as changes in the loop position through difference images of time series taken with a cadence of seconds or better. Moreover, \\citet{selwa2004} solved the same problem for a cylindrical loop geometry and obtained that time scales of the fast kink wave are of the order of a dozen of seconds, in agreement with the results presented here. \\citet{selwa2004} did not describe the dispersive properties of localized transverse loop disturbances and, for a perturbation centered in the loop axis, did not obtain large density variations. This contradicts our conclusions (see Figure~\\ref{fig_vae2_delta1_rho_z}), probably because the initial disturbance of \\citet{selwa2004} has a very small amplitude, of the order of 0.1~km. Impulsively generated transverse waves in coronal loops modeled as straight or curved slabs \\citep{selwa2006,selwa2007,rial2013} or as curved cylinders \\citep{selwa2011a,selwa2011b} have also been investigated, although the main aim of these studies is the analysis of transverse oscillations in line-tied loops, rather than the propagation of localized disturbances. A localized axisymmetric compression (rather than the localized transverse excitation investigated here) is the sum of fast sausage eigenmodes, which in a uniform cylinder are highly dispersive. Studies of propagating fast sausage wave trains have been performed for example by \\citep{roberts1984}. These authors showed that an impulsive axisymmetric excitation of a magnetic cylinder generates a clear time signature when the wave packet arrives at a point along the tube away from the initial disturbance. This time signature consists of a periodic phase, followed by a quasi-periodic phase, and a decaying phase \\citep[see also][]{murawski1993b,murawski1993d,nakariakov2004,nakariakov2009}. This complex temporal behavior is produced by the dispersive character of the fast sausage mode, that results in different frequencies arriving at different times at the detection position. In the study of \\citet{nakariakov2004}, the detected density perturbation contains a continuous variation of the period and its wavelet diagram shows large power distributed between 3 and 10~s, with a total duration of the oscillatory event of about 20~s. In fact, the wavelet spectra of Figure~\\ref{fig_vae2_delta1_z}(d) and Figures~3 to 5 of \\citet{nakariakov2004} are remarkably similar in spite of the different nature of the propagating disturbances. Fast sausage wave trains propagating along coronal loops have been invoked by \\citet{nakariakov2004,nakariakov2005} as the cause of the compressible waves analyzed by \\citet{williams2001,williams2002} and \\citet{katsiyannis2003}. These waves have also been put forward as the cause of the ``wavelet tadpoles'' found in solar decimetric radio bursts \\citep{meszarosova2009,meszarosova2011,karlicky2013}. But in view of the similar density perturbations produced by a transverse excitation and an axisymmetric compression, these conclusions may need additional work to be confirmed or refuted. Our method for solving Equations~(\\ref{eq_mass})--(\\ref{eq_induc}) in a magnetic cylinder has no underlying assumptions, except that the magnetic tube and the environment are uniform and that the internal to external density ratio is larger than one. It has several advantages over the direct numerical solution of Equations~(\\ref{eq_mass})--(\\ref{eq_induc}): the numerical integration of this set of partial differential equations requires using a two-dimensional mesh of points in the $r$- and $z$-directions. Moreover, to obtain the solution at a given time, $t$, the temporal evolution of the five perturbed variables $\\xi_r$, $\\xi_\\varphi$, $b_r$, $b_\\varphi$, and $P$ must be computed from $t=0$ to $t$. To make things worse, to produce the results of Figure~\\ref{fig_vae2_delta1_t}, for example, one needs to consider, at least, the large range of distances $0\\leq z\\leq 160a$, with $a$ the magnetic tube radius. The method we use is a numerical one and so it also suffers from numerical inaccuracy. It is nevertheless optimised to study wave propagation along a uniform magnetic tube. Imposing the initial conditions allows to obtain the amplitudes of the eigenmodes ($A_j^\\pm(k)$ from Equation~(\\ref{A_pm-main}) and $A_\\omega^\\pm(k)$ from Equation~(\\ref{A_pm_om-main})). Once these amplitudes are known, any perturbed variable can be computed at any position and time with Equation~(\\ref{Four_trans_inv-main}). A similar approach was used by \\citet{terradas2007}, based on the work of \\citet{ruderman2006}, to estimate the energy deposited in the normal modes given an external perturbation located in the solar corona. The method can be extended to study the propagation of, for example, an axisymmetric compression of the magnetic tube. Such a disturbance can be expressed as the sum of fast sausage eigenmodes. This study will be almost a complete repetition of the analysis in this paper. The only difference is that, in the case of fast sausage waves, there is a cut-off wavenumber even for the mode fundamental in the radial direction. This implies that we can expect more substantial contribution from improper eigenmodes. In this paper we have considered a magnetic tube with a sharp boundary. If, instead, we consider a tube with the density continuously varying from its value inside the tube to a lower value outside, then there will be wave damping due to resonant absorption. Even in this case the initial value problem can be solved using the same method as the one used in this paper. The only difference is that, in the case of a tube inhomogeneous in the radial direction, it is usually not possible to calculate the proper and improper eigenfunctions analytically, so they have to be calculated numerically." }, "1402/1402.7358_arXiv.txt": { "abstract": "Effective WIMP models are minimal extensions of the standard model that explain the relic density of dark matter by the ``WIMP miracle.'' In this paper we consider the phenomenology of effective WIMPs with trilinear couplings to leptons and a new ``lepton partner'' particle. The observed relic abundance fixes the strength of the cubic coupling, so the parameters of the models are defined by the masses of the WIMP and lepton partner particles. This gives a simple parameter space where collider and direct detection experiments can be compared under well-defined physical minimality assumptions. The most sensitive collider probe is the search for leptons + MET, while the most sensitive direct detection channel is scattering from nuclei arising from loop diagrams. Collider and direct detection searches are highly complementary: colliders give the only meaningful constraint when dark matter is its own antiparticle, while direct detection is generally more sensitive if the dark matter is not its own antiparticle. ", "introduction": "\\label{sec:intro} The observation of dark matter is an unambiguous discovery of physics beyond the standard model. The simplest and most compelling explanation for dark matter is a stable thermal relic, a weakly interacting massive particle (WIMP). Assuming that it freezes out by annihilation to standard model particles via dimensionless order-1 couplings, the correct relic density is obtained for dark matter masses of order 100~GeV to 1~TeV. This ``WIMP miracle'' means that dark matter can potentially be directly produced and studied at high-energy collider experiments such as the LHC. At the same time, this coupling allows for direct detection of astrophysical dark matter through its interactions with ordinary matter. % It is one of the most % ambitious dreams of particle physics to be able to discover dark matter in both types of experiment, and make a direct comparison to properties of dark matter inferred from cosmology and astrophysics. There are many well-motivated models of physics beyond the standard model that have a WIMP candidate (for example, the minimal supersymmetric standard model), but comparing collider and direct detection sensitivity is generally model-dependent due to the large number of % independent parameters in these theories. In a previous paper \\cite{Chang:2013oia}, we proposed a class of minimal models where it is assumed that the only particles beyond the standard model that are important for dark matter phenomenology consist of the WIMP and a standard model ``partner'' particle with the same gauge quantum numbers as one of the standard model particles. This allows a cubic coupling of the form \\beq\\eql{cubic} \\De \\scr{L} \\sim \\la (\\text{SM}) ( \\widetilde{\\text{SM}}) (\\text{DM}), \\eeq where DM denotes the dark matter particle and SM ($\\widetilde{\\text{SM}}$) denote the standard model field and its partner, respectively. The coupling $\\la$ is fixed by requiring the correct relic abundance, so the only parameters in this model are the masses of the WIMP and the standard model partner particle, once the spin and $CP$ properties of the dark matter is fixed. This gives a well-defined % and complete effective theory for dark matter phenomenology, motivating the name ``effective WIMP'' for this class of models. In particular, it allows direct comparison of the collider and direct detection searches for dark matter under well-defined physical minimality assumptions. \\Ref{Chang:2013oia} analyzed the phenomenology in the case where the standard model particle and its partner are colored. In this case collider and direct detection experiments were found to be highly complementary. Similar models were also studied in \\Refs{An:2013xka, Bai:2013iqa,DiFranzo:2013vra,Papucci:2014iwa}. The major difference in our work is that we focus on the parameter space where the WIMP has the correct thermal relic abundance. This has important implications for the global picture. For example, \\Ref{Chang:2013oia} found that the coupling $\\la$ is enhanced in some regions of parameter space, leading to increased sensitivity for collider and direct detection searches. On the other hand, the relic abundance constraint eliminates regions of parameter space where indirect detection limits are important. There has also been other approaches to simplified dark matter such as $s$-channel mediators \\cite{Buchmueller:2013dya,An:2012ue,Shoemaker:2011vi} and dark matter with electroweak interactions \\cite{Cirelli:2005uq, Cheung:2013dua}. In this paper, we consider effective WIMP models where the standard model particle is a lepton ($e, \\mu, \\tau$ or their corresponding neutrinos). In this case, the collider constraints come from the pair production of lepton partners followed by their decay into a WIMP plus a lepton or neutrino, and so the collider constraints are independent of the value of $\\la$. Once again, we find that collider and direct detection experiments are highly complementary. The direct detection constraints depend on $\\la$, with the dominant interactions coming from loop diagrams to a photon. The type of interactions allowed depend on whether the dark matter is its own antiparticle, which strongly affects the scattering rate. If the dark matter is its own antiparticle, the only meaningful constraints come from collider experiments, while in the opposite case direct detection experiments are generally more sensitive. There is a significant region of parameter space where both future collider and direct detection experiments can see a signal. Imposing the relic abundance constraints has an important effect on the direct detection phenomenology. In particular, these constraints extend into the regime of large lepton partner masses because the coupling $\\la$ becomes large in this region to get the correct relic abundance. Imposing the relic abundance constraint also implies that constraints from indirect detection are not important. The models that we consider are listed in Table~1. We consider couplings to left-handed lepton doublets only, since the phenomenology of right-handed leptons is essentially a subset of this case. Note that this fixes the production rate at colliders which leads to slightly stronger constraints than the right-handed lepton partner case. We consider both the cases where the dark matter couples to only the first two generations of leptons, as well as the case where it couples to all generations. We find that the differences between them are small, as one might expect. \\begin{table} \\begin{center} \\begin{tabularx}{.85\\textwidth}{|p{4cm}|p{4cm}|X|} \\hline \\multicolumn{2}{|c|}{Particles} & \\multirow{2}{.2\\textwidth}{\\parbox{4cm}{\\hspace{.5in} $\\mathcal{L}_{\\text{int}}$}}\\\\ \\cline{1-2} Dark matter $\\chi$ & Lepton partner $L$ & \\\\ \\hline \\hline Majorana fermion& Complex scalar& $\\la(\\chi \\ell)L^*+\\text{h.c.} $\\\\ \\hline Dirac fermion & Complex scalar & $\\la(\\chi \\ell)L^*+\\text{h.c.}$ \\\\ \\hline Real scalar & Dirac fermion& $\\la (L^c \\ell)\\chi+\\text{h.c.}$ \\\\ \\hline Complex scalar & Dirac fermion & $\\la (L^c \\ell)\\chi+\\text{h.c.}$ \\\\ \\hline \\end{tabularx} \\begin{minipage}{5.5in} \\caption{\\small Summary of the models considered in this paper. Spinors are written in 2-component notation. Here $\\ell$ is the left-handed lepton doublet of the standard model, $L$ is the lepton partner field, and $\\chi$ is the dark matter field. \\label{table:models}} \\end{minipage} \\end{center} \\end{table} ", "conclusions": "We have considered minimal extensions of the standard model containing electroweak singlet dark matter with renormalizable couplings to leptons and lepton ``partner'' particles. These models naturally explain the observed relic abundance of dark matter by the ``WIMP miracle.'' The strength of the coupling is fixed by requiring the correct relic abundance, so these models are parameterized only by the masses of the dark matter and the lepton partner particles. This gives a simple model where direct detection and dark matter searches at colliders can be compared under well-defined physical minimality assumptions. Previous studies focused on effective WIMPs coupled to colored standard model particles, while this paper focuses on effective WIMPs coupled to leptons. As in the previous case, we find that collider and direct detection experiments are remarkably complementary. The collider limits depend only on the spin and masses of the particles, and not whether dark matter is its own antiparticle. The dominant constraint comes from the dilepton plus MET channel. We find that the mono-lepton channel is less sensitive in these models. Direct detection experiments place interesting bounds only in the case where the dark matter is not its own anti-particle, {\\it i.e.}~Dirac or complex scalar dark matter. In these cases, current direct detection limits are generally stronger than collider limits, although there is an interesting region near $m_\\chi \\sim m_L/2$, where collider limits are competitive or stronger. In this region, it is possible to discover the dark matter in both future collider and direct detection experiments. In particular, the next generation of direct detection experiments (e.g. XENON1T) will be sensitive to the bulk of parameter space. However, if the dark matter is its own antiparticle ({\\it i.e.}~Majorana or real scalar dark matter) the direct detection bounds are well below the reach of current and next generation experiments, while the collider bounds are unaffected. In this case, leptophilic effective WIMPs can be discovered only in collider experiments. {\\it Note: As this work was being completed \\Ref{Bai:2014osa} appeared, which analyzes the same models without the relic abundance constraint and makes projections for lepton partner reaches at the 14~TeV LHC.}" }, "1402/1402.5993_arXiv.txt": { "abstract": "{We study a dark energy model with non-zero anisotropic stress, either linked to the dark energy density or to the dark matter density. We compute approximate solutions that allow to characterise the behaviour of the dark energy model and to assess the stability of the perturbations. We also determine the current limits on such an anisotropic stress from the cosmic microwave background data by the Planck satellite, and derive the corresponding constraints on the modified growth parameters like the growth index, the effective Newton's constant and the gravitational slip.} \\begin{document} ", "introduction": " ", "conclusions": "" }, "1402/1402.0322_arXiv.txt": { "abstract": "{The two objects \\object{1SWASP~J150822.80$-$054236.9} and \\object{1SWASP~J160156.04+202821.6} were initially detected from their SuperWASP archived light curves as candidate eclipsing binaries with periods close to the short-period cut-off of the orbital period distribution of main sequence binaries, at $\\sim$0.2~d. Here, using INT spectroscopic data, we confirm them as double-lined spectroscopic and eclipsing binaries, in contact configuration. Following modelling of their visual light curves and radial velocity curves, we determine their component and system parameters to precisions between $\\sim$2 and 11\\%. The former system contains 1.07 and 0.55~$M_{\\sun}$ components, with radii of 0.90 and 0.68~$R_{\\sun}$ respectively; its primary exhibits pulsations with period 1/6 the orbital period of the system. The latter contains 0.86 and 0.57~$M_{\\sun}$ components, with radii of 0.75 and 0.63~$R_{\\sun}$ respectively.} ", "introduction": "\\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{star10lc.eps}} \\caption{SuperWASP light curve for J150822, folded at period of 22469.219~s, with binned mean curve overplotted. A typical uncertainty range for a single observation is shown. These fluxes correspond to a visual magnitude range of $\\sim$12.4--13.2.} \\label{star10lc} \\end{figure} \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{star28lc.eps}} \\caption{SuperWASP light curve for J160156, folded at period of 19572.136~s, with binned mean curve overplotted. A typical uncertainty range for a single observation is shown. These fluxes correspond to a visual magnitude range of $\\sim$14.1--14.8.} \\label{star28lc} \\end{figure} The orbital period distribution of main sequence binary stars exhibits a fairly sharp lower limit at around 0.2~d \\citep{ruc92, ruc07, szymanski, paczynski}, the cause of which is the subject of ongoing research e.g. \\citet{stepien, stepien12, jiang}. However, despite the inherent interest of this region of parameter space, relatively few eclipsing binaries (EB) have been discovered with periods near the cut-off point. This motivated a search of the archive of the SuperWASP project (Wide Angle Search for Planets: \\citet{pollacco}) for EB candidates with apparent periods $<$20\\,000~s ($\\sim$0.2315~d), reported in \\citet{norton}. 53 candidates were found, 48 of which were new discoveries at the time. A subsequent search of these candidate EBs for evidence of period change \\citep{lohr} corrected the periods of nine to values slightly greater than 20\\,000~s, but still $<$22600~s ($\\sim$0.2616~d). A more rigorous search of the SuperWASP archive \\citep{lohr13} then detected 143 candidate EBs with periods $<$20\\,000~s, including 97 new discoveries since \\citet{norton}, and measured significant period changes in 74 candidates. Here, spectroscopic data allow us to confirm two of these candidates as double-lined EBs in contact configuration (W\\,UMa-type variables): \\object{1SWASP~J150822.80$-$054236.9} (J150822) and \\object{1SWASP~J160156.04+202821.6} (J160156). Both were initially identified in \\citet{norton}; the period of J150822 was revised upwards to 22469.2~s in \\citet{lohr} and so it did not appear in \\citet{lohr13}. We report system and component parameters obtained for these EBs by simultaneous modelling of their SuperWASP light curves and radial velocities. These should be of interest for the study of low-mass dwarfs and W\\,UMa systems in general, and of very short period binaries specifically. ", "conclusions": "J150822 and J160156 are established to be spectroscopic double-lined and eclipsing binary systems in contact configuration, composed of low-mass dwarfs. J150822 has been modelled as consisting of 1.07 and 0.55~$M_{\\sun}$ components (mass ratio 0.51), and J160156 as having 0.86 and 0.57~$M_{\\sun}$ components (mass ratio 0.67). The primary of J150822 appears to be pulsating with a period 1/6 of the orbital period. Both systems are plausibly undergoing mass transfer; this may be related to the primaries' radii being smaller, and the secondaries' radii being larger, than would be typical for single stars with these masses. The parameters obtained here should contribute to our understanding of low-mass stars and contact binary systems, since relatively few binaries are known with such short orbital periods. We hope to follow up further candidate short-period EBs listed in \\citet{lohr13} with multi-colour photometry and spectroscopy, with a view to confirming their binary nature. Many of them should be good prospects for full solution, and capable of significantly extending our knowledge of this aspect of the field." }, "1402/1402.5507_arXiv.txt": { "abstract": "Eclipsing millisecond pulsars in close ($P_b < 1$~day) binary systems provide a different view of pulsar winds and shocks than do isolated pulsars. Since 2009, the numbers of these systems known in the Galactic field has increased enormously. We have been systematically studying many of these newly discovered systems at multiple wavelengths. Typically, the companion is nearly Roche-lobe filling and heated by the pulsar which drives mass loss from the companion. The pulsar wind shocks with this material just above the surface of the companion. We discuss various observational properties of this shock, including radio eclipses, orbitally modulated X-ray emission, and the potential for $\\gamma$-ray emission. Redbacks, whose companions are likely non-degenerate and significantly more massive, generally have more luminous shocks than black widows which have very low mass companions. This is expected since the more massive redback companions intercept a greater fraction of the pulsar wind. We also compare these systems to accreting millisecond pulsars, which may be progenitors of black widows and in some cases can pass back and forth between redback and accretion phases. ", "introduction": "In 1986, the first binary pulsar to exhibit radio eclipses, PSR B1957+20, was discovered at the Arecibo observatory (Fruchter et al. 1988a). This 1.61~ms pulsar is in a 9.2~hr orbit around a very low mass ($M_c \\sim 0.02 M_{\\odot}$) companion. Regular radio eclipses were observed for $\\sim 10\\%$ of the orbit when the companion was at inferior conjunction. During the ingress and egress of the eclipse, delays in the pulse arrival times were observed which, along with the eclipse duration, suggested that the eclipses were due to excess ionized gas within the system (although the exact physical mechanism of the eclipses was and still is unclear eg. Thompson 1995). Optical studies showed large orbital variations in the light from the companion, suggesting the emission was primarily due to illumination by the pulsar (Fruchter et al. 1988b). H$\\alpha$ (Kulkarni et al. 1988) and X-ray (Stappers et al. 2003) imaging showed unambiguous evidence for a pulsar wind nebula around the pulsar, and observations of the non-thermal X-ray point source showed evidence for orbital modulation as well (eg. Huang and Becker 2007). Long term timing has shown that there are orbital period variations which are much larger in magnitude than would be expected from gravitational wave emission, with an orbital period derivative that can even change sign over time (Arzoumanian et al. 1994). These have no easy interpretation, but mass loss from the system and an induced quadrapole moment of the companion could contribute to the measured changes in orbit (Applegate \\& Shaham 1994). The inference drawn from these various observations is that PSR B1957+20 is ablating material off of its companion which shocks with the pulsar wind. Such systems had been proposed as a stage in the formation of isolated millisecond pulsars by Ruderman et al. (1989). Since the pulsar seemed to be in the process of destroying its companion, it was dubbed the ``black widow\" pulsar. Early models of the shock emission expected from the original Black Widow pulsar were developed by Arons and Tavani (1993) and Raubenheimer et al (1995). These predicted electrons could be accelerated to $\\sim 3$ TeV and that orbital modulation could arise from obscuration by the companion of part of the shock, intrinsic beaming of the emission by the magnetic field in the shock, or by doppler boosting. The high energy luminosity could depend on the distance to the shock, the magnetic field of the pulsar, the ion fraction, the optical emission of the companion (in the case of inverse Compton emission), and the magnetization of the wind. However, the overall scaling should be primarily due to spin-down power of the pulsar $\\dot E$, the fraction of the wind involved in the shock $f$, and the distance to the system $d$. \\begin{equation} L_S\\propto f \\dot E/d^2 \\end{equation} No unambiguous detection of $\\gamma$-ray emission arising from the shock in PSR B1957+20 has yet been made (although there has been an intriguing suggestion of GeV emission in $Fermi$ data, see Wu et al. 2012). However, the ability of a pulsar wind to produce emission from radio frequencies up to TeV energies in an intrabinary shock has been demonstrated by observations of the PSR B1259$-$63 system (\\cite{jmm+99,aaa+11,hess13}) which consists of a young pulsar in an highly eccentric 3.4 year orbit around a main sequence Be star. Near periastron the radio pulses of PSR B1259$-$63 are obscured as it enters the region of the dense equatorial outflow of the Be star, and unpulsed radio, X-ray, GeV and TeV emission is observed. While the $\\dot E$ of PSR B1259$-$63 is somewhat higher and a larger fraction of the pulsar wind is shocked by the stellar wind than in PSR B1957+20, the total shock luminosity should not be much more than an order of magnitude greater. However, near periastron, PSR B1259$-$63 is a few hundred times brighter in X-rays than PSR B1957+20, despite being at similar distances. The big difference between the Black Widow shock and those in PSR B1259-63 and the termination shocks of young pulsar wind nebulae is the scale of the systems. In terms of light cylinder radii $R_{lc}$, the companion of the Black Widow pulsar is on the order of $10^4 R_{lc}$ rather than the more typical $10^8 R_{lc}$ to the inner torii of young pulsar wind nebulae. This forces the shock to be in the region where, if the pulsar wind is magnetically dominated at the light cylinder, the magnetization fraction $\\sigma$ should still be relatively high. This may allow direct observational tests of proposed solutions to the ``sigma problem\" of pulsar wind shocks where the inferred magnetization at the termination shock from models of the X-ray emission is very low despite the expectation that at the light cylinder it should be very high (\\cite{a12}). Other potential advantages of studying the shocks in black widow type systems are that pulsar timing and optical observations allow us to determine the geometry of the system very well, and short orbits ($< 1$~d) allow for many repeated observations of the shock as viewed from different orientations. For many years it seemed that eclipsing systems in the Galactic field might be quite rare. In the 20 years following the discovery of PSR B1957+20, only 2 other systems were discovered, both of which had spin-down fluxes $\\dot E/d^2$ much smaller than that of the original Black Widow (see Roberts 2011 and references therein). Observations of globular clusters, however, revealed that a significant fraction of their millisecond pulsars either eclipsed and/or had very low companion masses. Freire (2005) listed 18 such systems. 11 have very low mass companions ($M_c \\la 0.04 M_{\\odot}$)like PSR B1957+20, but 7 have masses more typical of helium white dwarfs. Optical observations of these latter systems tended to indicate the companions were, in fact, non-degenerate. Were these then systems where the pulsar was spun up by one companion, then exchanged the original companion for a non-degenerate one in the very crowded core of the globular cluster? If this were the case, then it would be natural that most such systems would be found in globular clusters. ", "conclusions": "" }, "1402/1402.4737_arXiv.txt": { "abstract": "{ Chameleon, environmentally dependent dilaton, and symmetron gravity are three models of modified gravity in which the effects of the additional scalar degree of freedom are screened in dense environments. They have been extensively studied in laboratory, cosmological, and astrophysical contexts. In this paper, we present a preliminary investigation into whether additional constraints can be provided by studying these scalar fields around black holes. By looking at the properties of a static, spherically symmetric black hole, we find that the presence of a non-uniform matter distribution induces a non-constant scalar profile in chameleon and dilaton, but not necessarily symmetron gravity. An order of magnitude estimate shows that the effects of these profiles on in-falling test particles will be sub-leading compared to gravitational waves and hence observationally challenging to detect. } ", "introduction": "Since it was first published nearly a century ago, general relativity (GR) has earned its place as an incredibly successful and well verified theory of gravity (see e.g.\\ \\cite{Will2005}). There are however, both theoretical and observational reasons to consider alternatives to, and extensions of, GR. On the one hand, string theory provides a framework for describing quantum gravity, but suggests that we live in more than 4 dimensions. The consequences of these extra dimensions have not been definitively predicted, however, a generic feature of dimensional reduction is that extra fields appear in the low energy gravitational sector. On the other hand, observations of high redshift supernovae indicate that the universe is expanding at an accelerating rate~\\cite{Riess1998,Perlmutter1999}, and together with microwave background~\\cite{Hinshaw:2012aka,Ade:2013zuv} and large scale structure~\\cite{Abazajian:2008wr} measurements, suggest that around $70 \\%$ of the energy density of the universe comes in the form of a `dark energy' -- an energy momentum density which has a large negative pressure and is well modelled by a cosmological constant. Since the magnitude of the cosmological constant is extremely small by particle physics standards, explaining its stability under quantum corrections is a challenge. Finding either a natural explanation for its measured value or an alternative to the constant, is a major motivation for developing and studying modified theories of gravity (see \\cite{Clifton2011} for a review of various approaches). Additionally, studying modified gravity theories allows us to better explore where GR has been tested rigorously and to constrain the low energy properties of quantum gravity theories. Scalar-tensor theories, which modify gravity by introducing new, non-minimally coupled scalar fields are an extensively studied alternative to GR. They are theoretically attractive because such light scalar fields are generically predicted in the low energy limit of string theory. For a scalar field to affect cosmological expansion, its mass must be of the order of the Hubble scale, $H_0\\sim 10^{-33}\\,\\text{eV}$. If it interacts with matter however, the presence of a light scalar field will result in a long-range fifth force and would thus be subject to tight constraints from laboratory and solar system tests of gravity, \\cite{Will2005,Bertotti2003,Williams2004}. These constraints are weakened for theories in which the scalar field modifying gravity somehow decouples from matter, or is ``screened'', in dense environments. Models with this property are appealing because they can modify the behaviour of gravity on large scales while recovering the behaviour of GR in environments where local tests have been performed. This paper will focus on three particular scalar-tensor theories which have been found to exhibit screening, known as chameleon, \\cite{Khoury2004}, environmentally dependent dilaton, \\cite{Brax2010}, and symmetron, \\cite{Hinterbichler2011}, modified gravity. These models have been studied and constrained using laboratory \\cite{Mota2008,Upadhye2012b,Upadhye2012,Upadhye2012a}, solar system~\\cite{Bertotti2003}, cosmological~\\cite{Jain2010}, and astrophysical~\\cite{Jain2012} tests. These investigations all have one thing in common: they probe gravity in a regime where gravitational fields and space-time curvatures are relatively weak. In the near future, the direct detection of gravitational waves from compact binary systems will allow us to constrain the behaviour of gravity in the strong field, large curvature regime. Accordingly, attention has increasingly focussed on efforts to test gravity by studying the dynamics of compact objects such as neutron stars and black holes, \\cite{Yunes2013,Psaltis2008}. It is thus natural to ask whether observations of black holes might provide new constraints on screened modified gravity. Studying black holes in the context of a modified gravity theories inevitably leads to the re-examination of the uniqueness of exact solutions, i.e. ,whether the extra physical fields add extra degrees of freedom to black hole solutions, usually referred to as ``black hole hair'', \\cite{Ruffini}. A slightly looser definition of ``hair'', in terms of non-trivial scalar profiles rather than extra measurable charges at infinity known as {\\it dressing} \\cite{Achucarro:1995nu}, has been known to be possible for many years, starting with the (unstable) coloured black holes \\cite{Bartnik:1988am}, or the (stable) black monopoles, \\cite{Lee:1991vy}, quantum hair, \\cite{Coleman:1991jf,Dowker:1991qe}, as well as the explicit vortex hair \\cite{Achucarro:1995nu}. In scalar tensor gravity, a number of no-hair theorems (i.e., demonstrating that the scalar fields takes a constant value around an isolated black hole) have been proven, \\cite{Bekenstein1996,Sotiriou2012,Faraoni2013} (although note the assumptions behind these theorems may not always correspond to desired physical situations, \\cite{Sotiriou:2013qea}). The results of these theorems have been extended to binary black hole systems using perturbative, \\cite{Yunes2012,Mirshekari2013}, and numerical, \\cite{Healy2011}, calculations, demonstrating that dynamical spacetimes with two interacting black holes in scalar-tensor theory will be indistinguishable from general relativity. This might seem to imply that black hole systems will not be useful for constraining screened modified gravity, however, these theorems do not take into account cosmological backgrounds, in which the scalar field will typically be dynamical. A study of cosmologically evolving black holes with a canonical scalar field, \\cite{Chadburn:2013mta}, shows that the scalar field does evolve on the event horizon of the black hole, though there is no evidence for an additional scalar charge. Indeed, there are several solutions in the literature which allow for nontrivial scalar fields around black holes, \\cite{Fonarev:1994xq,Clifton:2004st,Clifton:2006ug,Guariento:2012ri}, although to be fair, many are singular, or ``engineered''. This paper focuses on removing the requirement that the black hole exists in a vacuum. This consideration is clearly relevant for astrophysical black holes that are typically observed in galaxies or galactic centres, and can have energetic accretion discs. In order to explore this question, we consider spherically symmetric distributions of matter around a black hole in screened modified gravity. We perform a `probe' calculation, in that we explore the scalar profile around the black hole without considering the corresponding modification of the black hole geometry. We are therefore not looking at issues of time dependent black holes, such as discussed in \\cite{McVittie:1933zz,Sultana:2005tp,Faraoni:2007es,Carrera:2008pi,Abdalla:2013ara}, as we will see nontrivial profiles even in the static case. Strictly speaking, a non-uniform matter distribution will result in a non-uniform scalar field profile due to the non-minimal coupling of the scalar to gravity. We expect that the complete picture will be a superposition of multipoles, with the dominant features being encapsulated by the monopole, or spherically symmetric, behaviour. Indeed, if we were to discover that the spherically symmetric case precluded the possibility of black hole hair, then it would be indicative that black holes would not carry hair. Conversely, the discovery of non-trivial spherically symmetric scalar profiles would demonstrate that black holes can indeed carry screened scalar hair, although the full solution would be more complicated and involved than the simple picture presented here. The layout of the paper is as follows: we introduce screened modified gravity in \\textsection\\ref{screenedMG}. We then study the effect of matter on the scalar field profile in \\textsection\\ref{profile} under the assumption of spherical symmetry and discuss observational implications for black hole properties in \\textsection\\ref{implications}. ", "conclusions": "The strong field, large curvature limit of modified gravity is a largely unexplored area for constraining various model dependent parameters and differentiating between the numerous models that exist in literature. We have presented an exploratory calculation for theories of modified gravity with screening mechanisms by studying a static, spherically symmetric black hole with an $r$-dependent matter distribution around it. In this construction, we found that chameleon and dilaton fields develop a non-trivial scalar profile, while the symmetron assumes a constant value. An order of magnitude estimate showed that the resulting scalar gradients affect in falling test particles in a way that is sub-leading compared to the quadruple radiation in GR and thus would be observationally challenging to detect. Note that our findings for symmetrons could be qualitatively altered if we used a different matter distribution in our setup. As we discussed in \\textsection\\ref{sym_profile}, if the matter density far from the black hole is small enough to make the symmetron approach $\\phistar=\\mu/\\sqrt{\\lambda}$ as its boundary condition, its field equation will no longer have a constant solution. This could occur, for example, if we used a model of an accretion disk which had finite spatial extent. We would, however, expect any symmetron gradients to be comparable to the ones found above for long range chameleons and dilatons. Similar reasoning can be applied when we consider solutions to the scalar field equation with angular dependence, or solutions on a Kerr background. No hair theorems require that in a stationary system involving matter around a black hole, any gradient in the scalar field will be sourced entirely by non-uniformities in matter density. If we assume that the density contrasts examined above are typical of astrophysical matter distributions near black holes, the magnitude of the scalar gradients produced (and their consequent fifth forces) should likewise be representative. Thus generically we would expect theories of screened modified gravity to produce solutions distinct from GR around stationary black hole systems with sub leading corrections to various observational effects. It would be interesting to explore this strong field limit further. One possible physical effect to study within our setup would be the phase of gravitational waves. Our crude estimates suggest that the rate of GR quadruple radiation will dominate over any radiation in the scalar sector, however, in~\\cite{Berti2012}, the authors showed that in Brans-Dicke theory the presence of a massive scalar field affects the phasing of gravitational radiation from binary systems significantly and they found that for observations of intermediate mass ratio inspirals, this effect could be used to place constraints on model parameters (a lower bound on $\\omega_{BD}$ and an upper bound on the mass of the scalar) which are competitive with Cassini and LLR measurements. Additionally, the requirement that our system and solutions be stationary causes us to disregard the possibility of transient effects associated with superradiance. Superradiance is a property of rotating black holes through which incident waves with at resonant frequency can become amplified by extracting some of the black hole's rotational energy. A number of studies have shown that superradiant effects can give rise to long-lived unstable modes in the presence of a massive scalar \\cite{Cardoso2005,Dolan2007,Dolan2012,Witek2013,Cardoso2013a}. This occurs for isolated black holes, but it also has been shown that when a black hole is surrounded by matter, superradiant modes can be amplified by factors as large as $10^5$~\\cite{Cardoso2013}. If the scalar's mass is very light, the instability timescale for these modes becomes short and would result in gaps in the mass-spin phase space of observed black holes. Because of this, measurements of a black hole's mass and spin can be used to constrain allowed masses for scalar fields. This technique has already been used to place the most stringent upper bound on the mass of the photon~\\cite{Cardoso2013a,Berti2013a}. Yet another potentially observable effect of superradiance occurs in EMRI systems, in what is known as floating orbits \\cite{Yunes2013,Yunes2012,Cardoso2011}. Floating orbits occur when the orbital frequency of the small compact object can excite superradiant modes. This causes the scalar field to transfer rotational energy from the central black hole into the orbit of the small compact object, counteracting the energy lost through quadrupole radiation and slowing the inspiral. Such floating orbits would affect the EMRI's gravitational waveforms significantly, so the detection of an EMRI signal consistent with GR would allow us to place strong constraints on the mass of a light scalar. We note with interest that these effects are typically studied in the context of scalar fields with a mass in the range $10^{-33}\\,\\text{eV} - 10^{-10}\\,\\text{eV}$~\\cite{Berti2013a}, which is relevant for dilatons and symmetrons, as well as chameleons at galactic or cosmological densities. Consideration of gravitational wave phasing, superradiant instabilities, and floating orbits will thus be important if we want to understand how these theories can be constrained by observations of black hole systems. We hope to further explore some of these effects in future works." }, "1402/1402.6442_arXiv.txt": { "abstract": "We present the association rates between solar energetic particles (SEPs) and the radio emission signatures in the corona and IP space during the entire solar cycle 23. We selected SEPs associated with X and M-class flares from the visible solar hemisphere. All SEP events are also accompanied by coronal mass ejections. Here, we focus on the correlation between the SEP events and the appearance of radio type II, III and IV bursts on dynamic spectra. For this we used the available radio data from ground-based stations and the Wind/WAVES spacecraft. The associations are presented separately for SEP events accompanying activity in the eastern and western solar hemisphere. We find the highest association rate of SEP events to be with type III bursts, followed by types II and IV. Whereas for types III and IV no longitudinal dependence is noticed, these is a tendency for a higher SEP-association rate with type II bursts in the eastern hemisphere. A comparison with reports from previous studies is briefly discussed. ", "introduction": "Solar energetic particle (SEP) events are transient flux enhancements of electrons, protons and ions due to acceleration processes in the solar corona and interplanetary (IP) space. The high energy particles can pose a serious risk for the near Earth and ground-based technological devices, may disturb communications and be a health hazard \\citep{2006SSRv..124..303M}. This is why energetic particles are of major space weather interest. The accelerator of these particles, however, is still a subject of debate. Presently, the reconnection processes during flares \\citep{2002JGRA..107.1315C} and the shock-acceleration at coronal mass ejections (CMEs), \\cite{1999SSRv...90..413R}, are recognized as efficient particle accelerators. An important issue is whether both acceleration processes act simultaneously with comparable efficiency or one of them dominates the particle energization process. In the current two-class SEP picture, the flares are thought to dominate the {\\it impulsive} events and the shock waves dominate the large {\\it gradual} events. Observations, however, are not able to relate unambiguously the SEP parameters measured in situ to the parent solar activity (e.g., flares vs. shocks). The large uncertainties usually present when doing different correlation analyses may arise due to the poorly known particle transport in the turbulent IP magnetic field, the magnetic connection between the acceleration site and the Earth, and that SEPs are often just measured at a single point in space. Radio observations can provide additional information and constraints for identifying the particle accelerator in the corona and IP space. When nonthermal electrons propagate through the solar corona they may emit radio waves, if certain conditions apply \\citep{2008SoPh..253....3N}. The different types of radio bursts in the dynamic radio spectrum (see e.g., a review by \\citet{2008A&ARv..16....1P}), are usually interpreted as signatures from electrons accelerated in the vicinity of a shock wave (type IIs), as electron beams (type IIIs) or as electrons confined in closed loop structure during the early evolution of a CME (type IVs). Type III bursts usually extend to low frequencies (reaching as far as the Earth orbit, $\\sim$20 kHz) which is the main indicator that accelerated electrons (and by inference also protons) can efficiently escape from the solar corona \\citep{2002JGRA..107.1315C}. Interplanetary counterparts of type II are also observed, but to lesser extent, whereas type IVs can be seen only to few MHz, but not to lower frequencies. Here we focus on the comparison of the particle (mainly proton) intensities measured in situ with the electromagnetic emission of electrons in the corona and IP space over the entire solar cycle 23. Numerous previous studies reported different association rates of SEPs and emission of radio bursts of type II \\citep{2004ApJ...605..902C,2008AnGeo..26.3033G}, III \\citep{2002JGRA..107.1315C} and IV \\citep{1982ApJ...261..710K}. The aim here is to identify the different types of radio emission from the dynamic spectra, to complement our results with reports from the different radio observatories and to compare them with previous works on the topic. ", "conclusions": "We present the association rates of the SEP events (protons) and their accompanying radio emission (from electrons) in the corona (dm and m wavelength) and IP space (DH-range). Since there are no signatures of protons interacting with the solar atmosphere (with the exception of gamma-ray emission), we use electron signatures as a diagnostic for particle acceleration from the corona up to 1~AU. \\cite{2002JGRA..107.1315C} were the first to identify long-lasting groups of DH type III bursts as a typical radio counterpart of large SEP events. They found the groups to be of significantly longer duration than type~III bursts associated with impulsive flares, see also \\cite{1987SoPh..111..397M}. In the analysis performed here, we did not take into account the burst duration of DH-type III, nor explicitly separate the SEP events into gradual or impulsive. Still we find that SEP events have the highest association rate with type III radio bursts. This implies that the electrons accelerated in the corona (within one solar radius) have a ready access to the IP space, irrespective of whether the SEP event is impulsive or gradual. Shock signatures (type II bursts) were also considered in correlation studies with SEP events. We found a lower association rate (up to 75\\%) of the DH-type II bursts with SEP events, compared to the m- and DH-type III association rates. The number increases when only SEP events with strong intensities are considered, in agreement with \\cite{2002ApJ...572L.103G} and \\cite{2004ApJ...605..902C}. The association of SEP events with types III and IV is comparable in the eastern and western groups (see Figure~\\ref{F-histo}). But for the type II bursts there is a slight trend for a higher association rate in the eastern hemisphere. This result could be understood it terms of different sources contributing to SEP events. Shocks could be the dominant accelerator when the parent activity is poorly connected to Earth (as in the eastern solar hemisphere). In the western hemisphere (where more than twice as many events were detected) both flare and shock acceleration could contribute. When no shock signatures accompany the eastern SEP events, their propagation through the IP space and detection at Earth could be facilitated by a large-scale magnetic structure, e.g., interplanetary coronal mass ejections (ICMEs). \\cite{1991JGR....96.7853R} estimated that for about 15\\% of the eastern events this is likely the case. Recently \\cite{2013SoPh..282..579M} showed similar occurrence rate of ICMEs for the western SEP events (20\\%). The association with the type IIIs seems to be increasing with the decreasing of the radio frequency (from dm to DH-range). At first, the type III identification in the dm-range might be masked due to overlying decimeter continuum often present in the dynamic radio spectra (observational bias). Hence, the obtained association rates of dm-IIIs are to be considered as lower limits only. In addition, high frequencies and dense plasma impede the production and growth of Langmuir waves. Also, the electron beam must travel an `instability distance' from the acceleration site before Langmuir waves are generated \\citep{2011A&A...529A..66R}. This can cause a lack of high frequency type III emission in even the most energetic of electron beams. Moreover, collisions of both waves and electrons can suppress Langmuir wave growth at high frequencies, e.g., \\citet{1982ApJ...263..423K,2012SoPh..tmp..109R}. Without the intention to give a complete overview on the results from previous work, we selected few statistical studies that are (partially) co\\-vering solar cycle 23. The different association rates are given in Table~\\ref{T-Rates_Sum} and are mostly consistent, although in many earlier studies only large SEP events were considered." }, "1402/1402.3267_arXiv.txt": { "abstract": "{ Scaling properties of galaxy cluster observables with cluster mass provide central insights into the processes shaping clusters. Calibrating proxies for cluster mass that are relatively cheap to observe will moreover be crucial to harvest the cosmological information available from the number and growth of clusters with upcoming surveys like \\textit{eROSITA} and \\textit{Euclid}. The recent \\textit{Planck} results led to suggestions that X-ray masses might be biased low by $\\sim\\!40$~\\%, more than previously considered. }{ We aim to extend knowledge of the weak lensing -- X-ray mass scaling towards lower masses (as low as $1\\!\\times\\!10^{14}\\,\\mathrm{M}_{\\odot}$) in a sample representative of the $z\\!\\sim\\!0.4$--$0.5$ population. Thus, we extend the direct calibration of cluster mass estimates to higher redshifts. }{ We investigate the scaling behaviour of MMT/Megacam weak lensing (WL) masses for $8$ clusters at $0.39\\!\\leq\\!z\\!\\leq\\!0.80$ as part of the \\emph{400d} WL programme with hydrostatic \\textit{Chandra} X-ray masses as well as those based on the proxies, e.g. $Y_{\\mathrm{X}}\\!=\\!T_{\\mathrm{X}}M_{\\mathrm{gas}}$. }{ Overall, we find good agreement between WL and X-ray masses, with different mass bias estimators all consistent with zero. When subdividing the sample into a low-mass and a high-mass subsample, we find the high-mass subsample to show no significant mass bias while for the low-mass subsample, there is a bias towards overestimated X-ray masses at the $\\sim\\!2\\sigma$ level for some mass proxies. The overall scatter in the mass-mass scaling relations is surprisingly low. Investigating possible causes, we find that neither the greater range in WL than in X-ray masses nor the small scatter can be traced back to the parameter settings in the WL analysis. }{ We do not find evidence for a strong ($\\sim\\!40$~\\%) underestimate in the X-ray masses, as suggested to reconcile recent \\textit{Planck} cluster counts and cosmological constraints. For high-mass clusters, our measurements are consistent with other studies in the literature. The mass dependent bias, significant at $\\sim\\!2\\sigma$, may hint at a physically different cluster population (less relaxed clusters with more substructure and mergers); or it may be due to small number statistics. Further studies of low-mass high-$z$ lensing clusters will elucidate their mass scaling behaviour. } ", "introduction": "Galaxy cluster masses hold a crucial role in cosmology. In the paradigm of hierarchical structure formation from tiny fluctuations in the highly homogeneous early cosmos after inflation, clusters emerge via the continuous matter accretion onto local minima of the gravitational potential. Depending sensitively on cosmological parameters, the cluster mass function, i.e.\\ their abundance as function of mass and redshift $z$, provides observational constraints to cosmology \\citep[e.g.,][]{2009ApJ...692.1060V,2011ARA&A..49..409A,2013arXiv1303.5080P}. Observers use several avenues to determine cluster masses: properties of the X-ray--emitting intracluster medium (ICM), its imprint on the cosmic microwave background via the Sunyaev-Zel'dovich (SZ) effect in the sub-mm regime, galaxy richness estimates and dynamical masses via optical imaging and spectroscopy, and gravitational lensing. Across all wavelengths, cluster cosmology surveys are under preparation, aiming at a complete cluster census out to ever higher redshifts, e.g. \\textit{eROSITA} \\citep{2010SPIE.7732E..23P,2012arXiv1209.3114M,2012MNRAS.422...44P} and \\textit{Athena} \\citep{2013arXiv1306.2307N,2013arXiv1306.2319P} in X-rays, \\textit{Euclid} \\citep{2011arXiv1110.3193L,2012arXiv1206.1225A}, DES and LSST in the optical/near-infrared, CCAT \\citep{2012SPIE.8444E..2MW} and SKA at sub-mm and radio frequencies. Careful X-ray studies of clusters at low and intermediate redshifts yield highly precise cluster masses, but assume hydrostatic equilibrium, and in most cases spherical symmetry \\citep[e.g.][]{2008A&A...487..431C,2013SSRv..177..119E}. Observational evidence and numerical modelling challenge these assumptions for all but the most relaxed systems \\citep[e.g.][]{2008MNRAS.384.1567M,2012NJPh...14e5018R,2013SSRv..177..155L,2013ApJ...765...24N}. While simulations find X-ray masses to only slightly underestimate the true mass of clusters that exhibit no indications of recent mergers and can be considered virialised, non-thermal pressure support can lead to a $>\\!20$~\\% bias in unrelaxed clusters \\citep{2010A&A...510A..76L,2012NJPh...14e5018R}. \\citet{2014ArXiv1401.7657S} modelled the pressure due to ICM turbulence analytically and found a $\\sim\\!10$~\\% underestimate of cluster masses compared to the hydrostatic case. Weak lensing (WL), in contrast, is subject to larger stochastic uncertainties, but can in principle yield unbiased masses, as no equilibrium assumptions are required. Details of the mass modelling however can introduce biases, in particular concerning projection effects, the source redshift distribution and the departures from an axisymmetric mass profile \\citep{2009MNRAS.396..315C,2011ApJ...740...25B,2012MNRAS.421.1073B,2013SSRv..177...75H}. For individual clusters, stochastic uncertainties dominate the budget; however, larger cluster samples benefit from improved corrections for lensing systematics, driven by cosmic shear projects \\citep[e.g.][]{2013MNRAS.429..661M}. Most of the leverage on cosmology and structure formation from future cluster surveys will be due to clusters at higher $z$ than have been previously investigated. Hence, the average cluster masses and signal/noise ratios for all observables are going to be smaller. Even and especially for the deepest surveys, most objects will lie close to the detection limit. Thus the scaling of inexpensive proxies (e.g. X-ray luminosity $L_{\\mathrm{X}}$) with total mass needs to be calibrated against representative cluster samples at low and high $z$. Weak lensing and SZ mass estimates are both good candidates as they exhibit independent systematics from X-rays and a weaker $z$-dependence in their signal/noise ratios. Theoretically, cluster scaling relations arise from their description as self-similar objects forming through gravitational collapse \\citep{1986MNRAS.222..323K}, and deviations from simple scaling laws provide crucial insights into cluster physics. For the current state of scaling relation science, we point to the recent review by \\citet{2013SSRv..177..247G}. As we are interested in the cluster population to be seen by upcoming surveys, we focus here on results obtained at high redshifts. Self-similar modelling includes evolution of the scaling relation \\emph{normalisations} with the Hubble expansion, which is routinely measured \\citep[e.g.][]{2011A&A...535A...4R, 2013AN....334..354E}. Evolution effects beyond self-similarity, e.g.\\ due to declining AGN feedback at low $z$, have been claimed and discussed \\citep[e.g.][]{2007MNRAS.382.1289P,2010MNRAS.408.2213S,2010ApJ...715.1508S,2012MNRAS.421.1583M}, but current observations are insufficient to constrain possible evolution in slopes \\citep{2013SSRv..177..247G}. Evidence for different scaling behaviour in groups and low-mass clusters was found by, e.g.,\\ \\citet{2011A&A...535A.105E,2012MNRAS.422.2213S,2014arXiv1402.0868B}. \\citet{2011A&A...535A...4R} and \\citet{2012MNRAS.421.1583M} investigated X-ray scaling relationships including clusters at $z\\!>\\!1$, and both stressed the increasing influence of selection effects at higher $z$. Larger weak lensing samples of distant clusters are just in the process of being compiled \\citep{2011ApJ...737...59J,2012A&A...546A.106F,2011ApJ...726...48H,2012MNRAS.427.1298H,2012A&A...546A..79I,2012arXiv1208.0597V,2012ApJS..199...25P}. Thus most WL scaling studies are currently limited to $z\\!\\!\\lesssim\\!\\!0.6$, and also include nearby clusters \\citep[e.g.][M13]{2012MNRAS.427.1298H,2013ApJ...767..116M}. The latter authors find projected WL masses follow the expected correlation with the SZ signal $Y_{\\mathrm{SZ}}$, corroborating similar results for more local clusters by \\citet{2009ApJ...701L.114M,2012ApJ...754..119M}. \\citet{2013MNRAS.429.3627M} performed a detailed WL analysis of a $z\\!=\\!0.81$ cluster discovered in the SZ using the Atacama Cosmology Telescope, and compared the resulting lensing mass against the \\citet{2012ApJ...751...12R} $Y_{\\mathrm{SZ}}$--$M$ scaling relation, in what they describe as a first step towards a high-$z$ SZ-WL scaling study. By compiling \\textit{Hubble Space Telescope} data for $27$ massive clusters at $0.83\\!<\\!z\\!<\\!1.46$, \\citet{2011ApJ...737...59J} not only derive the relation between WL masses $M^{\\mathrm{wl}}$ and ICM temperature $T_{\\mathrm{X}}$, but also notice a good correspondence between WL and hydrostatic X-ray masses $M^{\\mathrm{hyd}}$. As they focus on directly testing cosmology with the most massive clusters , these authors however stop short of deriving the WL--X-ray scaling. Also using \\textit{HST} observations, \\citet{2011ApJ...726...48H} investigated the WL mass scaling of the optical cluster richness (i.e.\\ galaxy counts) and $L_{\\mathrm{X}}$ of $25$ moderate-$L_{\\mathrm{X}}$ clusters at $0.3\\!<\\!z\\!<\\!0.6$, thus initiating the study of WL scaling relations off the top of the mass function. Comparisons between weak lensing and X-ray masses for larger cluster samples were pioneered by \\citet{2008MNRAS.384.1567M} and \\citet{2008A&A...482..451Z}, collecting evidence for the ratio of weak lensing to X-ray masses $M^{\\mathrm{wl}}/M^{\\mathrm{hyd}}\\!>\\!1$, indicating non-thermal pressure. \\citet{2010ApJ...711.1033Z}, analysing $12$ clusters from the \\textit{Local Cluster Substructure Survey} (LoCuSS), find this ratio to depend on the radius . Likewise, a difference between relaxed and unrelaxed clusters is found \\citep{2010ApJ...711.1033Z,2013ApJ...767..116M}. \\citet{2012NJPh...14e5018R} show that the gap between X-ray and lensing masses is more pronounced in simulations than in observations, pointing to either an underestimate of the true mass also by WL masses \\citep[cf.][]{2012MNRAS.421.1073B} or to simulations overestimating the X-ray mass bias. The current disagreement between the cosmological constraints derived from \\textit{Planck} primary cosmic microwave background (CMB) data with \\textit{Wilkinson Microwave Anisotropy Probe} data, supernova data, and cluster data \\citep{2013arXiv1303.5080P} may well be alleviated by, e.g.\\ sliding up a bit along the \\textit{Planck} degeneracy curve between the Hubble factor $H_{0}$ and the matter density parameter $\\Omega_{\\mathrm{m}}$. Nevertheless, as stronger cluster mass biases than currently favoured $(\\sim\\!40$~\\%) have also been invoked as a possible explanation, it is very important to test the cluster mass calibration with independent methods out to high $z$, as we do in this work. This article aims to test the agreement of the weak lensing and X-ray masses measured by \\citet{2012A&A...546A..79I} for $8$ relatively low-mass clusters at $z\\!\\gtrsim\\!0.4$ with scaling relations from the recent literature. The \\emph{400d} X-ray sample from which our clusters are drawn has been constructed to contain typical objects at intermediate redshifts, similar in mass and redshift to upcoming surveys. Hence, it does not include extremely massive low-$z$ clusters. We describe the observations and WL and X-ray mass measurements for the $8$ clusters in Sect.~\\ref{sec:obsdat}, before presenting the central scaling relations in Sect.~\\ref{sec:res}. Possible explanations for the steep slopes our scaling relations exhibit are discussed in Sect.~\\ref{sec:puzzle}, and we compare to literature results in Sect.~\\ref{sec:m13disc}, leading to the conclusions in Sect.~\\ref{sec:conclu}. Throughout this article, $E(z)\\!=\\!H(z)/H_{0}\\!=\\!\\sqrt{\\Omega_{\\mathrm{m}}(z+1)^{3}+\\Omega_{\\Lambda}}$ denotes the self-similar evolution factor (Hubble factor $H(z)$ normalised to its present-day value of $H_{0}\\!=\\!72\\,\\mbox{km}\\,\\mbox{s}^{-1}\\,\\mbox{Mpc}^{-1}$), computed for a flat universe with matter and Dark Energy densities of $\\Omega_{\\mathrm{m}}\\!=\\!0.3$ and $\\Omega_{\\Lambda}\\!=\\!0.7$ in units of the critical density. ", "conclusions": "\\label{sec:conclu} In this article, we analysed the scaling relation between WL and X-ray masses for $8$ galaxy clusters drawn from the \\emph{400d} sample of X-ray--luminous $0.35\\!\\leq\\!z\\!\\leq\\!0.89$ clusters. WL masses were measured from the \\citet{2012A&A...546A..79I} MMT/Megacam data, and X-ray masses were based on the V09a \\textit{Chandra} analysis. We summarise our main results as follows: \\begin{description} \\item[1.] We probe the WL--X-ray mass scaling relation, in an unexplored region of the parameter space for the first time: the $z\\!\\sim\\!0.4$--$0.5$ redshift range, down to $1\\!\\times\\!10^{14}\\,\\mathrm{M}_{\\odot}$. \\item[2.] Using several X-ray mass estimates, we find the WL and X-ray masses to be consistent with each other. Most of our clusters are compatible with the $M^{\\mathrm{X}}\\!=\\!M^{\\mathrm{wl}}$ line. \\item[3.] Assuming the $M^{\\mathrm{wl}}$ not to be significantly biased, we do not find evidence for a systematic underestimation of the X-ray masses by $\\sim\\!40$~\\%, as suggested as a possible solution to the discrepancy between the \\textit{Planck} CMB constraints on $\\Omega_{\\mathrm{m}}$ and $\\sigma_{8}$ (the normalisation of the matter power spectrum) and the \\textit{Planck} SZ cluster counts \\citep{2013arXiv1303.5080P}. While our results favour a small WL--X-ray mass bias, they are consistent with both vanishing bias and the $\\sim\\!20$~\\% favoured by studies of non-thermal pressure support. \\item[4.] For the mass-mass scaling relations involving $M^{\\mathrm{wl}}$, we observe a surprisingly low scatter $0.5\\!\\!<\\!\\!\\chi^{2}_{\\mathrm{red}}\\!\\!<\\!\\!0.6$, although we use only stochastic uncertainties and allow for correlated errors via a Monte Carlo method. Because the errors in $M^{\\mathrm{wl}}$ are largely determined by the intrinsic WL shape noise $\\sigma_{\\varepsilon}$, we however deem a drastic overestimation unlikely (Sect.~\\ref{sec:chisqdisc}). For the scaling relations involving $M_{\\mathrm{G}}$, however, we observe a large scatter, contrary to \\citet{2010ApJ...721..875O} and M13. \\item[5.] Looking in detail, there are intriguing indications for a mass-dependence of the WL--X-ray mass ratios of our relatively low-mass $z\\!\\sim\\!0.4$--$0.5$ clusters. We observe a mass bias in the low--$M^{\\mathrm{wl}}$ mass bin at the $\\sim\\!2\\sigma$ level when splitting the sample at $\\log{(M_{\\mathrm{piv}}/\\mathrm{M}_{\\odot})}\\!=\\!14.5$ This holds for the masses V09a report based on the $Y_{\\mathrm{X}}$, $T_{\\mathrm{X}}$, and $M_{\\mathrm{G}}$ proxies. \\end{description} We thoroughly investigate possible causes for the mass-dependent bias and tight scaling relations. First (Sect.~\\ref{sec:c200}), we confirm that by using a mass-concentration relation instead of directly fitting $c_{200}$ from WL, we already significantly reduced the bias due to conversion from $r_{200}$ to $r_{500}$. We emphasise that, on average, the NFW shear profile represents a suitable fit for the cluster population \\citep[cf.][]{2013ApJ...769L..35O}. Measuring $M^{\\mathrm{hyd}}$ within $r_{500}^{\\mathrm{wl}}$ induces correlation between the data points in Fig.~\\ref{fig:mxml}. Removing this correlation by plotting both masses within a fixed physical radius, we still find small scatter (Sect.~\\ref{sec:chisqdisc}). We notice that the mass range occupied by the $M^{\\mathrm{wl}}$ exceeds the X-ray mass ranges, Partially, this higher WL mass range can be explained by the correction for dilution by member galaxies, which could be applied only where colour information was available (Paper~II). Coincidentally, this is the case for the more massive half of the MMT sample in terms of $M^{\\mathrm{wl}}$, thus boosting the range of measured WL masses (Sect.~\\ref{sec:ndc}). This result underscores the importance of correcting for the unavoidable inhomogeneities in WL data due to the demanding nature of WL observations \\citep[cf.][]{2012arXiv1208.0605A}. We find no further indications for biases via the WL analysis. Furthermore the tight scaling precludes strong redshift effects, and we find that our small MMT subsample is largely representative of the complete sample of $36$ clusters, judging from the $M^{\\mathrm{Y}}$--$M^{\\mathrm{T}}$ relation (Sect.~\\ref{sec:fluke}). For the $M^{\\mathrm{Y}}$--$M^{\\mathrm{G}}$ and $M^{\\mathrm{T}}$--$M^{\\mathrm{G}}$ relations, significant scatter ($\\chi^{2}_{\\mathrm{red}}\\!>\\!2$) is present in the larger sample. The former relation also shows indications for a significant bias of $M^{\\mathrm{Y}}\\!\\approx\\!1.15 M^{\\mathrm{G}}$. Weak lensing and hydrostatic masses for the \\emph{400d} MMT clusters are in good agreement with the $z\\!>\\!0.35$ part of the \\citet{2013ApJ...767..116M} sample and the $M^{\\mathrm{wl}}_{500}$--$M^{\\mathrm{hyd}}_{500}$ relation derived from it (Sect.~\\ref{sec:hydlit}). The M13 and \\citet{2012A&A...546A.106F} samples include three \\emph{400d} clusters with CFHT WL masses. These clusters neither point to significantly higher scatter nor to a less mass-dependent bias (Fig.~\\ref{fig:lit}). We are planning a re-analysis of the CFHT data, having demonstrated in Paper~II that lensing catalogues from MMT and CFHT are nicely compatible. Such reanalysis is going to be helpful to identify more subtle WL analysis effects potentially responsible for the steep slopes and tight correlation of WL and X-ray masses. An alternative explanation are intrinsic differences in the low-mass cluster population. That the \\emph{400d} MMT sample probes to slightly lower masses ($1\\!\\times\\!10^{14}\\,\\mathrm{M}_{\\odot}$) than M13 or F12 becomes especially obvious from the $M^{\\mathrm{wl}}$--$Y^{\\mathrm{X}}$ relation (Fig.~\\ref{fig:lit}, Sect.~\\ref{sec:yxlit}). Because the \\emph{400d} sample is more representative of the $z\\!\\sim\\!0.4$--$0.5$ cluster population, it is likely to contain more significant mergers relative to the cluster mass, skewing mass estimates (Sect.~\\ref{sec:physics}). Hence, the \\emph{400d} survey might be the first to see the onset of a mass regime in which cluster physics and substructure lead the WL--X-ray scaling to deviate from what is known at higher masses. Remarkably, Giles et al.\\ (in prep.) are finding a different steep slopes in their low-mass WL--X-ray scaling analysis. Detailed investigations of how their environment shapes clusters like CL\\,1416+4446 might be necessary to improve our understanding of the cluster population to be seen by future cosmology surveys. Analysis systematics might also behave differently at lower masses. A turn for WL cluster science towards lower mass objects, e.g.\\ through the completion of the \\emph{400d} WL sample, will help addressing the question of evolution in lensing mass scaling relations." }, "1402/1402.3857.txt": { "abstract": "Two formation scenarios have been proposed to explain the tight orbits of hot Jupiters. They could be formed in orbits with a small inclination (with respect to the stellar spin) via disk migration, or in more highly inclined orbits via high-eccentricity migration, where gravitational interactions with a companion and tidal dissipation are at play. Here we target hot Jupiter systems where the misalignment $\\lambda$ has been inferred observationally and we investigate whether their properties are consistent with high-eccentricity migration. Specifically, we study whether stellar tides can be responsible for the observed distribution of $\\lambda$ and orbital separations. Improving on previous studies, we use detailed models for each star, thus accounting for how convection (and tidal dissipation) depends on stellar properties. In line with observations suggesting that hotter stars have higher $\\lambda$, we find that $\\lambda$ increases as the amount of stellar surface convection decreases. This trend supports the hypothesis that tides are the mechanism shaping the observed distribution of $\\lambda$. Furthermore, we study the past orbital evolution of five representative systems, chosen to cover a variety of temperatures and misalignments. We consider various initial orbital configurations and integrate the equations describing the coupled evolution of the orbital separation, stellar spin, and misalignment. We account for stellar tides and wind mass loss, stellar evolution, and magnetic braking. We show that the current properties of these five representative systems can be explained naturally, given our current understanding of tidal dissipation and with physically motivated assumptions for the effects driving the orbital evolution. ", "introduction": "\\label{Intro} The plethora of exoplanets discovered in recent years has revealed that planetary systems exist in a much greater variety than we had ever imagined. To date, more than 1000 exoplanets have been confirmed using different observational techniques. Almost 200 of these planets are similar in mass to Jupiter, but revolve around their parent star every 10\\,days or less (NASA Exoplanet Archive), thus challenging our understanding of planet formation and evolution. Different scenarios have been proposed to explain how these so-called {\\it hot Jupiters} formed in such tight orbits. % and an additional 3,600 candidates have been provided by NASA's {\\it Kepler}'s satellite. These discoveries have shown that planetary systems exist in a much greater variety than we had ever imagined. In this paper we consider hot Jupiters, giant planets which challenge our understanding of planet formation and evolution by revolving around their parent stars in very tight orbits. Different formation scenarios have been proposed to explain these planets. One way to distinguish between these models is to investigate the current properties of the many discovered systems. Our focus here lies on systems where the obliquity $\\lambda$ (the sky-projected angle between the stellar spin and orbital angular momentum vectors) has been constrained observationally. Two migration models have been invoked to bring gas giants from their birthplace at several AU into the tight orbits we observe today: {\\it disk migration} and {\\it high-eccentricity migration} (however, see also \\citealt{TutukovFedorova2012, Thies+11}). These models predict different orientations of the planet's orbit at present. In the disk migration scenario, planets could migrate inwards through their interactions with the protoplanetary gas disk \\citep{GoldreichTremaine80,Ward97, MurrayHHT98, Lin+96,Guillochon+11}. As the disks tend to damp the orbital inclination \\citep{Cresswell+07,XiangPapaloizou14}, this model would naturally lead to nearly circular orbits and small obliquities (e.g., \\citealt{GoldreichTremaine80,PapaloizouLarwood00}). In the high-eccentricity migration scenario, gravitational interactions either between several planets or with companion stars could lead to highly eccentric orbits and high obliquities (\\citealt{Kozai62, Lidov62, WuLithwick11,Naoz+11,Nagasawa08,FabryckyTremaine07,WuMurray03,RasioFord96, ChatterjeeFMR08}; see also \\citealt{PlavchanBilinski13} for empirical evidence). As tidal dissipation tends to damp the eccentricity while decreasing the orbital separation, close-in planets could then result from tidal circularization. \\cite{WinnFAJ10} used a sample of 19 systems in which the projected spin-orbit angle $\\lambda$ was measured %with a $1\\sigma$ precision of $10^{o}$ or better via the Rossiter-McLaughlin (RM) effect. The RM effect occurs when a transiting planet blocks the blue- or red-shifted part (or both, depending on the orbital inclination with respect to the stellar spin) of the spinning star as it passes across the stellar disk, thus distorting the star's spectral line profile. \\cite{WinnFAJ10} investigated the behavior of the sky-projected misalignment as a function of the host star's effective temperature ($T_{\\rm eff *}$). Their results suggested that the degree of misalignment increases for hotter stars. In particular, a sharp increase in $\\lambda$ seems to occur at $T_{\\rm eff *}\\,\\simeq\\,$6250~K. These findings were later confirmed by \\citeauthor{Albrecht+12} (\\citeyear{Albrecht+12}, hereafter A12) with a sample of RM measurements twice as large as the one available to \\cite{WinnFAJ10}; see also e.g., \\cite{MortonJohnson11}. Since $\\simeq\\,$6250~K is the temperature at which the outer convective zone in main sequence stars starts becoming negligible, \\cite{WinnFAJ10} proposed that the mechanism responsible for the trend observed in the data is convective dissipation of tides in the star. Hot Jupiters could then be produced via a single formation process yielding a broad distribution of obliquities. Later on, tidal dissipation in cool stars would damp the obliquity within a few Gyr. Correspondingly, the high degree of sky-projected misalignment observed in hot stars would result from the inefficiency of tidal dissipation. While its simplicity is appealing, this scenario presents a major weakness. In fact, tidal dissipation in the star acts both on the misalignment and on the orbital separation, causing orbital decay whenever the stellar spin frequency is lower than the orbital frequency, and it thus fails in explaining the currently observed aligned hot Jupiters (see, e.g., \\citealt{RogersLin13}, hereafter R13). %\\cite{RogersLin13} (hereafter R13) integrated the equations describing the tidal evolution of the orbital separation, stellar spin and obliquity in the weak friction approximation considering both slowly and rapidly rotating stars. For slow rotators the authors find that, while tides tend to damp the obliquity, they also transfer angular momentum from the orbit to the star at an increasingly faster rate. As a result, the planet would fall into the star before the obliquity is completely damped. On the other hand, for fast rotators the authors show that tidal damping of the misalignment occurs together with a significant orbital expansion. A significant degree of obliquity evolution can occur only if the planet starts its orbital evolution from inside the star. Possible solutions to this evolutionary conundrum were presented by \\cite{WinnFAJ10} and \\cite{Lai12}. \\cite{WinnFAJ10} suggested that, if the star's radiative interior is weakly coupled to the outer convective region and to the planet, then tides would act on the obliquity faster than on the orbital separation (see also A12). However, a large amount of differential rotation inside the star would potentially lead to fluid instabilities, which would tend to quickly re-couple the star's convective and radiative regions. To overcome this problem, \\cite{Lai12} presented a different scenario, following the idea that different physical processes dissipate tides with different efficiencies. His model invokes the excitation and damping of inertial waves in a stellar convective zone (see \\S~\\ref{Tidal Contribution due to inertial wave dissipation} for a summary). These waves are driven by the Coriolis force and are excited only in misaligned systems. In this configuration, the tidal potential to the leading quadrupole order has several terms. Each of these terms generates tidal disturbances with its own dissipation efficiency. Among these terms, \\cite{Lai12} identified a component of the tidal torque which acts only on the misalignment without affecting the orbital separation, thus providing a more efficient mechanism to modify the misalignment. The validity of this prescription was recently questioned by R13, who considered a random distribution of initial obliquities for 50 objects (nearly the number of observed systems considered by A12) and integrated the equations derived by \\cite{Lai12} forward in time. The authors computed the evolution of the misalignment {\\it alone} while keeping the orbit and stellar spin fixed and found that tides would lead to a nearly equal amount of prograde ($\\lambda \\textless 90^{o}$), retrograde ($\\lambda \\textgreater 90^{o}$), and 90$^{o}$ orbits. This appears inconsistent with the observations, as the majority of observed obliquities are smaller than $90^{o}$. However, we note that R13's investigation has two major limitations. First, it neglects the simultaneous evolution of the orbital separation, stellar spin, and misalignment, while previous investigations have shown that it is essential to consider the coupled evolution of the orbital elements and spins (e.g., \\citealt{Jackson+08,BarkerOgilvie2009,MatsumuraPR2010,Xue+14}). Furthermore, it does not account for the various physical effects that might compete in the evolution of the system (e.g., magnetic braking and the radial expansion of the star as a result of stellar evolution). This last simplification was recently adopted also by \\cite{Xue+14}, who integrated the full set of equations presented by \\cite{Lai12} and showed that all intermediate states found by R13 eventually evolve towards alignment. While this appears to be consistent with the majority of observed obliquities ($\\textless\\,90^{o}$), it cannot explain the currently observed intermediate misalignments. %We show in this paper that the orbital configurations of misaligned hot Jupiters can indeed be explained, once all physical effects are properly taken into account. In this paper, we reconsider tidal dissipation in the star as a possible mechanism responsible for the observed distribution of misalignments and orbital separations. We carefully examine the observed relation between $\\lambda$, $T_{\\rm eff *}$, and the amount of convection inside the host star. In contrast to previous studies, we use detailed stellar evolution models, thus accounting for the dependence of convection on stellar properties, such as mass, metallicity, effective temperature, and age. Furthermore, we integrate the full set of equations describing the evolution of the orbital separation, stellar spin, and misalignment. We take into account the effects of tidal dissipation in the star, stellar wind mass loss, changes in the star's internal structure as a result of stellar evolution, and magnetic braking. The tidal prescription adopted follows \\cite{Lai12}, and includes both tides in the weak friction approximation \\citep{Zahn1977, Zahn1989} and convective damping of inertial waves. In the weak-friction regime, a body's response to tides is generally measured via a tidal quality factor $Q$ \\citep{GoldreichSoter66}, which parametrizes the efficiency of tidal dissipation. This term measures how a tidally-deformed body undergoing a forced oscillation dissipates part of the associated energy during each oscillation period. It is formally defined as the ratio of the maximum energy stored in the tidal distortion over the energy lost during each cycle. %how a tidally deformed body undergoing a forced oscillation dissipates part of the associated energy during each oscillation period (similarly to the quality factor of a forced, damped harmonic oscillator). %, it is defined as %\\begin{align} %Q = 2\\pi E_{0}\\left(\\oint -\\dot{E}dt\\right)^{-1}. %\\end{align} %Here $E_{0}$ is the maximum energy stored in an oscillation and the integral is the energy dissipated over one cycle. The value of $Q$ is the result of complex dissipative processes occurring within a body and it thus varies for bodies of different masses and types. Furthermore, $Q$ depends on the tidal forcing frequency and thus on the spins and orbital configuration. As a result, $Q$ is expected to vary by orders of magnitudes \\citep{PenevSasselov11} and it is clear that different $Q$ values are needed to explain different systems (e.g., \\citealt{MatsumuraPR2010}, hereafter M10). In this work, we prefer not to introduce additional model parameters and instead use a parametrization for tidal dissipation calibrated from observations of binary stars (e.g., \\citealt{VerbuntPhinney1995,RasioTLL1996,HurleyTP02,BelczynskiStartrack2008}). For quick reference, the notations adopted in this work for the components and orbital parameters are summarized in Table~\\ref{Tab:ParamsDefinition}. The paper is organized as follows. In \\S~\\ref{The Sample} we present the sample of hot Jupiters considered in this work. In \\S~\\ref{Detailed Stellar modeling with MESA} we present the procedure adopted to model in detail the host stars in our sample. In \\S~\\ref{Orbital Evolution Model}, we summarize the equations that we integrate to study the orbital evolution of misaligned hot Jupiters (tests on the orbital evolution code developed for this work are presented in Appendix~\\ref{Tests on the Orbital Evolution code}). In \\S~\\ref{Detailed orbital evolution for four representative systems} we present possible evolutionary sequences of five representative systems: \\mbox{HAT-P-6}, WASP-7, 15, 16, and 71 (the results are summarized in Table~\\ref{Tab:OrbitalEvolResultsPandTheta}, where we also include a few additional examples, without describing their evolution in detail). We discuss the assumptions adopted in this work in \\S~\\ref{Discussion}. We summarize and conclude in \\S~\\ref{Conclusions}. As mentioned above, we consider tides in the weak friction approximation. Specifically, we account for both convective damping of the equilibrium tide and radiative damping of the dynamical tide. For the latter, we use results of detailed calculations presented by \\cite{Zahn1975}, which are valid in the limit of small tidal forcing frequencies (in the weak-friction regime). In Appendix~\\ref{Dynamic Tides in WASP-71} we solve the full set of equations describing non-adiabatic non-radial forced stellar oscillations, and we discuss the significance of dynamic tides in the most massive system among those studied in detail in \\S~\\ref{Detailed orbital evolution for four representative systems} for a wide spectrum of tidal forcing frequencies. %%%%%%%%%%%%%%%%%%%%%%%%%% \\begin{table}[!h] \\centering \\caption{Definition of the various parameters used in this work.} \\begin{tabular}{lr} \\hline {\\bf Parameter} & {\\bf Definition} \\\\ \\\\[-1.0em] \\hline \\\\[-1.0em] \\smallskip $M_{*}$, $M_{\\rm pl}$, $\\Delta\\,M_{\\rm CZ}$ & Mass\\\\ \\smallskip $R_{*}$, $\\Delta\\,R_{\\rm CZ}$ & Radius\\\\ \\smallskip $L_{*}$ & Bolometric luminosity\\\\ \\smallskip $T_{\\rm eff *}$ & Effective temperature\\\\ \\smallskip Fe/H (or $Z$) & Metallicity\\\\ \\smallskip $I_{*}$ & Moment of inertia\\\\ \\smallskip $\\Omega_{*}$ ($\\Omega_{o}$) & Spin (orbital) frequency\\\\ \\smallskip $i_{*}$ ($i_{o}$) & Stellar (orbital) inclination\\\\ \\smallskip $v_{{\\rm rot}}{\\rm sin}~i_{*}$ & Rotational velocity\\\\ \\smallskip $a$ & Semimajor axis\\\\ \\smallskip $P_{\\rm orb}$ & Orbital period\\\\ \\smallskip $e$ & Eccentricity\\\\ \\smallskip $S$ & Spin angular momentum\\\\ \\smallskip $L$ & Orbital angular momentum\\\\ $\\lambda$ ($\\Theta_{*}$) & Sky-projected (true) misalignment\\\\ \\\\[-1.0em] \\hline \\\\[-1.0em] \\end{tabular} \\label{Tab:ParamsDefinition} \\tablecomments{The subscripts ``*'' and ``pl'' refer to the star and planet, respectively, while ``CZ'' refers to the stellar surface convection zone (see \\S~\\ref{Detailed Stellar modeling with MESA} for details) . We denote with $i_{*}$ the angle between the stellar spin axis and the line of sight, while $i_{o}$ denotes the angle between the orbital angular momentum and the line of sight.} \\end{table} %%%%%%%%%%%%%%%%%%%%%%%%%% %In what follows we use $M_{*}$, $R_{*}$, $T_{\\rm eff *}$, Fe/H (or $Z$), $\\lambda$ ($\\Theta_{*}$), $I_{*}$, and $v_{{\\rm rot}}{\\rm sin}~i$ to indicate the star's mass, radius, effective temperature, metallicity, sky-projected (true) misalignment, moment of inertia, and rotational velocity, respectively. The planet's mass is denoted with $M_{\\rm pl}$. The stellar spin frequency and planet's orbital frequency are $\\Omega_{*}$ and $\\Omega_{o}$, respectively. The orbital period, semi major axis, and eccentricity are denoted with $P_{\\rm orb}$, $a$ and $e$, respectively. The angle between the stellar spin axis and the line of sight is $i_{*}$, while the angle between the orbital angular momentum and the line of sight is $i_{o}$. %%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "Two formation models have been proposed to explain the tight orbits of hot Jupiters. These giant planets could migrate inward in a disk (in the so-called {\\it disk migration} scenario), or they could be formed via tidal circularization of a highly eccentric orbit following gravitational interactions with a companion (in the so-called {\\it high-eccentricity migration} scenario). Disk migration yields orbits where the stellar spin and orbital angular momentum vectors are nearly aligned, while high-eccentricity migration results in a broad range of misalignments. Here we targeted the known hot Jupiters where the obliquity has been inferred observationally (following and updating the sample considered by \\citealt{Albrecht+12}) and investigated whether their properties are consistent with high-eccentricity migration. In contrast to previous studies, we modeled in detail each host star and showed that the observed increase in misalignment $\\lambda$ with the star's effective temperature $T_{\\rm eff}$ is shaped by the amount of convection inside the host star, as originally suggested by \\cite{WinnFAJ10}. Specifically, higher degrees of misalignment are found in stars with less surface convection, especially if one considers the radial extent of the surface convective region. This result supports the hypothesis that giant planets are formed with a broad initial distribution of misalignments. During the subsequent orbital evolution, convective dissipation of tides in the star is the mechanism that shapes the observed distribution of misalignments. To further test this hypothesis, we computed the coupled evolution of the orbital elements and stellar spin of five representative systems: one aligned, two prograde, and two retrograde systems. We studied in detail \\mbox{HAT-P-6}, WASP-7, 15, 16, and 71, and provided results for few more systems, as additional examples. Within the regime of validity of the tidal prescription adopted (\\S~\\ref{Tidal Contribution due to inertial wave dissipation}), WASP-16 is an aligned star among the coolest ones. Instead, \\mbox{WASP-71} and \\mbox{WASP-7} (\\mbox{WASP-15} and \\mbox{HAT-P-6}) are among the coolest and hottest prograde (retrograde) stellar hosts, respectively. In contrast to previous studies on the observed $\\lambda\\,-\\,T_{\\rm eff *}$ correlation (e.g., \\citealt{Albrecht+12,RogersLin13}), we took into account the combined effects of tides, stellar wind mass loss, magnetic braking, and stellar evolution. The tidal prescription adopted combines tides in the weak friction approximation and inertial wave dissipation, and it was recently proposed by \\cite{Lai12} to explain the currently observed aligned hot Jupiters. For the efficiency of inertial wave dissipation, we followed numerical results by \\cite{OgilvieLin2007} and \\cite{BarkerOgilvie2009} and considered tidal quality factors $Q'_{\\rm 10}$ ranging between $10^{6}\\,-\\,10^{10}$. Furthermore, we scanned a broad parameter space in initial orbital periods, misalignments, and degrees of asynchronism between the stellar spin and the planet's orbital frequency ($\\Omega_{*}/\\Omega_{o}$). Our results show that, accounting for all the relevant physical mechanisms and considering the simultaneous evolution of the orbital separation, stellar spin, and misalignment, the current properties of the variety of systems considered here can be naturally explained. For HAT-P-6, WASP-7, and 15, with F-dwarfs, we found that both orbital decay and obliquity damping are small even on Gyr timescales. This supports the notion that many of the F stars are presently not capable of either destroying the planet or damping the obliquity, and that the high observed obliquities are the result of the hot-Jupiters formation process. For WASP-71, we found that both the orbit and obliquity are actively damping, but, for the smallest $Q'_{\\rm 10}$ value considered, the obliquity evolves on a timescale that can be almost 1 order of magnitude shorter at present (and up to 3 orders of magnitude shorter in the past, Fig.~\\ref{fig:DetailedOrbitalEv_WASP-71_timescales}). The same is true for WASP-16, where the obliquity damping timescale is more than 1 order of magnitude shorter than the orbital decay timescale at present (Fig.~\\ref{fig:DetailedOrbitalEv_WASP-16_timescales}). This system in particular, together with WASP-4 described in \\cite{VR14}, supports the idea that obliquity damping in cool and less massive G-dwarfs can occur more rapidly than orbital decay, provided that inertial wave dissipation is actively driving the misalignment evolution. This mechanism does indeed provide an explanation for the population of currently known aligned hot Jupiters. Finally, our results show that the physical effects included in this work can all play a significant role in the orbital evolution of misaligned hot Jupiters systems, depending on the properties of the star and planet, as well as the orbital configuration, as summarized below. By using detailed stellar models we took into account how changes in the star's moment of inertia affect its spin. Stellar evolution efficiently decreases $\\Omega_{*}$ in systems hosting F-dwarfs, where it becomes increasingly significant as the star expands during its main sequence evolution. Another efficient driver of stellar spin-down is magnetic braking, included here according to \\citeauthor{Skumanich72}'s (\\citeyear{Skumanich72}) law. We varied its strength by changing the parameter $\\gamma_{\\rm MB}$ entering the magnetic braking prescription, but the values of $\\gamma_{\\rm MB}$ which yield \\mbox{HAT-P-6-}, \\mbox{WASP-15-}, and \\mbox{71-}type systems agree with those adopted in the literature. WASP-7, and 16 are exceptions, as their present properties can be matched either {\\it without} magnetic braking ($\\gamma_{\\rm MB}\\,=\\,0$, for WASP-7, with an F-dwarf) or with $\\gamma_{\\rm MB}\\,\\leq\\,0.5$ (for WASP-16, with a G-dwarf). However, we argued that more systems resembling WASP-7 with $\\gamma_{\\rm MB}\\,>\\,0$ could be found considering higher initial levels of asynchronism between the stellar spin and the planet's orbital frequency. For tides, the \\citeauthor{Lai12}'s (\\citeyear{Lai12}) prescription is valid for sub-synchronous systems ($\\Omega_{*}\\,\\textless\\,\\Omega_{o}$) and inertial wave dissipation acts only on the stellar spin and the misalignment. Instead, in super-synchronous systems ($\\Omega_{*}\\,\\textgreater\\,\\Omega_{o}$), this dissipation channel might also affect the orbital separation, but we did not explore this possibility and we targeted only configurations where $\\Omega_{*}\\,\\leq\\,\\Omega_{o}$ (and discussed cases in which $\\Omega_{*}\\,=\\,\\Omega_{o}$). %The results do not change significantly for $Q'_{\\rm 10}\\textgreater 10^{6}$, for most system configurations. %In fact, explaining the currently observed misaligned systems {\\it solely} via tides in the weak-friction approximation presents major problems (e.g., \\citealt{RogersLin13}). %This tidal prescription was recently questioned by R13, however, their investigation present major weaknesses. In fact, while it is clear from the literature (e.g., this paper, \\citealt{Jackson+08,BarkerOgilvie2009,MatsumuraPR2010}) that it is essential to consider the coupled evolution of the orbital and rotational elements, R12 computed the evolution of the misalignment {\\it alone, only} accounting for inertial wave dissipation, thus neglecting all the various and potentially significant physical effects included here. As mentioned above, we showed that inertial wave dissipation can be significant, mainly for the evolution of the misalignment. This dissipation channel can increase the misalignment in systems like \\mbox{HAT-P-6}, while it can actively damp it in WASP-16- and 71-type systems. It is unimportant for WASP-15- and 7-type systems and the associated timescales can be orders of magnitude longer than those related to the main drivers of spin and misalignment evolution. For systems like \\mbox{WASP-7}, in particular, this finding suggests that the evolution of the orbital separation would not be significantly affected by inertial wave dissipation, if super-synchronous configurations were considered. Expanding the initial parameter space to include $\\Omega_{*}/\\Omega_{o}\\,\\textgreater\\,1$ would likely yield \\mbox{WASP-7-type} systems where the magnetic braking coefficient $\\gamma_{\\rm MB}$ is consistent with values adopted in the literature. Tides in the weak friction approximation are the main driver of orbital decay (when $\\Omega_{*}\\,\\textless\\,\\Omega_{o}$) and they can also affect the misalignment. As weak-friction tides always tend to decrease the misalignment, they can either counteract the effect of inertial wave dissipation (e.g., in \\mbox{HAT-P-6-type} systems) or strengthen it (e.g., in \\mbox{WASP-71-type} systems), depending on $Q'_{\\rm 10}$. Tides in the weak friction approximation are less important for the evolution of the stellar spin, except for \\mbox{WASP-71}: for this system, this mechanism becomes relevant towards the end of the evolution and it counteracts the spin-down driven by stellar evolution and magnetic braking. Finally, stellar wind mass loss is negligible for all systems considered here. In a \\mbox{WASP-71-type} system it could have been the main mechanism driving the evolution of the orbital separation early in its past, but the associated timescales were too long to significantly affect it. To conclude, our detailed stellar modeling, the orbital evolution calculations, and the few more examples summarized in Table~\\ref{Tab:OrbitalEvolResultsPandTheta}, provide support to the high-eccentricity migration scenario for the formation of hot Jupiters. A detailed calculation similar to the one presented here on {\\it all} misaligned systems is needed to provide a definite answer (this will be the subject of future work, see also \\S~\\ref{Discussion}). However, here we demonstrated that tidal dissipation in the host star, together with stellar wind mass loss, magnetic braking, and stellar evolution, can account for the observed distribution of misalignments (and alignments) and orbital separations found around stars of different temperatures. In particular, the orbital configuration of five representative hot Jupiter systems covering a variety of misalignments and stellar temperatures can be explained with reasonable assumptions on the physical effects driving the orbital evolution in circular and misaligned systems, despite the limited parameter space considered." }, "1402/1402.2342_arXiv.txt": { "abstract": "We report resolved near-infrared spectroscopic monitoring of the nearby L dwarf/T dwarf binary WISE~J104915.57$-$531906.1AB (Luhman~16AB), as part of a broader campaign to characterize the spectral energy distribution and temporal variability of this system. A continuous 45-minute sequence of low-resolution IRTF/SpeX data spanning 0.8--2.4~$\\micron$ were obtained, concurrent with combined-light optical photometry with ESO/TRAPPIST. Our spectral observations confirm the flux reversal of this binary, and we detect a wavelength-dependent decline in the relative spectral fluxes of the two components coincident with a decline in the combined-light optical brightness of the system over the course of the observation. These data { are successfully modeled as a combination of achromatic (brightness) and chromatic (color) variability in the T0.5 Luhman~16B, consistent with variations in overall cloud opacity; and no significant variability in L7.5 Luhman~16A, consistent with recent resolved photometric monitoring.} We estimate a peak-to-peak amplitude of 13.5\\% at 1.25~$\\micron$ over the full lightcurve. Using a simple { two-spot} brightness temperature model for Luhman~16B, we infer { an average cold} covering fraction of { $\\approx$30--55\\%, varying by 15--30\\% over a rotation period assuming a $\\approx$200--400~K difference between hot and cold regions. We interpret these variations as changes in the covering fraction of a high cloud deck and corresponding ``holes'' which expose deeper, hotter cloud layers, although other physical interpretations are possible. A Rhines scale interpretation for the size of the variable features explains an apparent correlation between period and amplitude for Luhman~16B and the variable T dwarfs SIMP 0136+0933 and 2MASS~J2139+0220, and predicts relatively fast winds (1--3~{\\kms}) for Luhman~16B consistent with lightcurve evolution on an advective time scale (1--3 rotation periods). The strong variability observed in this flux reversal brown dwarf pair supports the model of a patchy disruption of the mineral cloud layer as a universal feature of the L dwarf/T dwarf transition.} ", "introduction": "The driving mechanism for the transition between the L dwarf and T dwarf spectral classes has emerged as one of the outstanding problems in brown dwarf astrophysics. Spectroscopically, this transition is defined by the { appearance} of {\\meth} absorption features at near-infrared wavelengths \\citep{1999ApJ...519..802K,2006ApJ...637.1067B}, accompanied by a substantial reduction of condensate cloud opacity \\citep{1996Sci...272.1919M,2000ApJ...531..438B,2001ApJ...556..357A}. Both effects drive near-infrared spectral energy distributions (SEDs) to transition from red ($J-K \\approx 1.5-2.5$) to blue ($J-K \\approx 0-0.5$; \\citealt{2000ApJ...536L..35L,2002ApJ...568..335M,2006ApJ...637.1067B}), with strengthening molecular gas bands delineating the spectral subclasses. What is remarkable about the L dwarf/T dwarf transition is that it appears to take place over a relatively narrow range of effective temperatures ({\\teff}s) and luminosities, based on absolute magnitude trends (e.g., \\citealt{2002AJ....124.1170D,2004AJ....127.2948V}), broad-band SED measurements (e.g., \\citealt{2004AJ....127.3516G}) and spectral model fits (e.g., \\citealt{2008ApJ...678.1372C,2009ApJ...702..154S}). The L/T transition also exhibits an apparent excess of binaries \\citep{2007ApJ...659..655B}, gaps in color distributions \\citep{2012ApJS..201...19D} and a decline in number densities as a function of spectral type \\citep{2008ApJ...676.1281M}, trends that suggest the transition is rapid in time as well as temperature. The important role of photospheric cloud evolution for this transition is seen in the observation that early-type T dwarfs with minimal cloud opacity are often significantly brighter at 1~$\\micron$ than their hotter, cloudier L dwarf counterparts. This is true in both color-magnitude diagrams of local populations \\citep{2003AJ....126..975T,2012ApJ...752...56F,2012ApJS..201...19D} and among components of ``flux-reversal'' binaries that straddle the L/T transition \\citep{2006ApJS..166..585B,2006ApJ...647.1393L,2008ApJ...685.1183L}. The 1~$\\micron$ region is a minimum of molecular gas opacity---the local pseudocontinuum---so condensate grain scattering can dominate the overall opacity at these wavelengths \\citep{2001ApJ...556..872A}. The 1~$\\micron$ brightening has thus been interpreted as a depletion of photospheric condensate clouds over a narrow range of {\\teff} and/or time. The geometry of the depletion has been modeled as both global changes in photospheric chemistry (e.g., \\citealt{2004AJ....127.3553K,2005ApJ...621.1033T,2006ApJ...640.1063B,2008ApJ...689.1327S}) and hole formation that allows light to emerge from hotter regions \\citep{2001ApJ...556..872A,2002ApJ...571L.151B,2010ApJ...723L.117M}. The latter hypothesis predicts an enhancement of rotationally-modulated photometric variability at the L/T transition, particularly in the 1~$\\micron$ region, { depending on the sizes and distribution of the cloud gaps}. Recent brown dwarf monitoring observations support this prediction, as the two most prominent variables identified to date, SIMP~J013656.5+093347 (\\citealt{2006ApJ...651L..57A,2009ApJ...701.1534A}; hereafter SIMP~J0136+0933) and 2MASS J21392676+0220226 (\\citealt{2008AJ....136.1290R,2012ApJ...750..105R}; hereafter 2MASS~J2139+0220) are both early-type T dwarfs. Their variability can be reproduced with spot models assuming regions with thick and thin clouds at different temperatures { assumed to probe different layers in the atmosphere} \\citep{2012ApJ...750..105R,2013ApJ...768..121A}. The spectral character of the observed variability is nevertheless complex. Rather than variability being limited to pseudocontinuum regions where gas opacity is a minimum, broad-band chromatic and achromatic variations are seen across the infrared \\citep{2009ApJ...701.1534A,2012ApJ...750..105R}. The light curve shapes themselves are also seen to change over several rotation periods, suggesting dynamic evolution of features at rates considerably faster than the Solar gas giants \\citep{2013ApJ...776...85S}. Finally, variability measurements over widely-separated spectral regions have recently revealed evidence of pressure-dependent phase variations, indicating vertical structure in { the features driving the variability} \\citep{2012ApJ...760L..31B,2013ApJ...778L..10B}. The considerable level of detail on brown dwarf cloud structure and atmospheric dynamics garnered from these monitoring studies is of relevance to exoplanet atmospheres, where clouds are { now seen} as a key opacity source \\citep{2011ApJ...733...65B,2011ApJ...737...34M,2012ApJ...754..135M,2013MNRAS.432.2917P,2013A&A...559A..33C}. The recently-discovered, nearby binary brown dwarf system {\\name}AB (hereafter {\\namesh}AB; \\citealt{2013ApJ...767L...1L}) has emerged as a potential benchmark for studying the L/T transition. With spectral types of L7.5 and T0.5 \\citep{2013ApJ...770..124K,2013ApJ...772..129B}, its components straddle the transition. Its T dwarf secondary is brighter than the primary in the 0.95--1.3~$\\micron$ range, making it a flux-reversal system \\citep{2013ApJ...772..129B}. {\\namesh}AB is also a significant variable. Combined-light red optical photometry by \\citet{2013A&A...555L...5G} revealed peak-to-peak variability of $\\sim$10\\% with a period of 4.87$\\pm$0.01~hr, with large changes in the light curve structure over daily timescales. The variability was attributed primarily to the T dwarf component. Resolved photometry by \\citet{2013ApJ...778L..10B} extended the observed variability into the near-infrared, confirmed {\\namesh}B as the dominant variable, and revealed pressure-dependent phase variations. As such, this system embodies nearly all of the remarkable characteristics of the L/T transition---multiplicity, variability, and flux reversal---while residing only 2.020$\\pm$0.019~pc from the Sun \\citep{2013arXiv1312.1303B}. In April 2013, our consortium organized a week-long monitoring campaign of {\\namesh}AB using telescopes in Chile, Australia and Hawaii, with the aim of characterizing its variability panchromatically (radio, optical and infrared) and spectroscopically, while simultaneously obtaining kinematic data (radial and rotational velocities) to constrain its orbit and viewing geometry. This article reports low-resolution near-infrared spectroscopic monitoring observations obtained over 45 minutes with the SpeX spectrometer \\citep{2003PASP..115..362R} on the 3.0m NASA Infrared Telescope Facility (IRTF), coincident with combined-light optical photometry obtained with the TRAnsiting Planets and PlanetesImals Small Telescope (TRAPPIST; \\citealt{2011Msngr.145....2J}). In Section~2 we describe our observation and data reduction procedures, including period analysis of the TRAPPIST lightcurve around this epoch. In Section~3 we describe our spectral extraction and variability analysis of the SpeX data, and create an empirical model to replicate both the SpeX and TRAPPIST observations. In Section~4 we discuss our results, examining the nature of {\\namesh}B's inferred variability in the context of a simple { two-spot} brightness temperature model, and compare this source to other significantly variable L/T transition objects. We summarize our results in Section~5. ", "conclusions": "\\subsection{{ Interpreting} the Nature of {\\namesh}B's Spectral Variability} Spectral trends in variability have been examined in several L and T dwarfs to date, through pure spectroscopy (e.g., \\citealt{2008MNRAS.384.1145B,2008A&A...487..277G,2013ApJ...768..121A}) and simultaneous or near-simultaneous broad-band imaging (e.g., \\citealt{2004MNRAS.354..466K,2009ApJ...701.1534A,2012ApJ...760L..31B,2012ApJ...750..105R,2013AJ....145...71K,2013ApJ...767..173H,2013ApJ...778L..10B}). The most significant variables up until now, SIMP~J0136+0933 and 2MASS~J2139$-$0220, both exhibit color trends in { near-infrared} photometric variability, with larger amplitude changes at $J$ as compared to $K$, again consistent with variable condensate cloud opacity. However, spectroscopic variability measurements of these same two sources over 1.0--1.7~$\\micron$ by \\citet{2013ApJ...768..121A} indicate that achromatic or near-achromatic variations dominate the psuedocontinuum. These authors propose { a two-layer cloud model with a thick shallow cloud and thin deep cloud as a means of reproducing} both achromatic psuedocontinuum and chromatic broadband variability. { Matched to atmosphere models, this framework can replicate observed trends in the colors and spectral shapes of SIMP~J0136+0933 and 2MASS~J2139$-$0220 over 1-3 rotation periods, although detailed fits to the data remain poor (see also \\citealt{2012ApJ...750..105R}).} For {\\namesh}B, we also find that both achromatic and chromatic variations must be present in the psuedocontinuum to properly model the observations. Achromatic variation yields a decline in the overall flux, amounting to roughly 0.03~mag in broad-band $J$ over the observing period. The concurrent chromatic variation simultaneously reddens the spectrum of this source by $\\Delta(J-K)$ = 0.02~mag, resulting in a relative flux variation amplitude of $\\Delta{F_{K_s}}/\\Delta{F_J}$ = 0.41$\\pm$0.18, similar to values reported for SIMP~J0136+0933 \\citep{2009ApJ...701.1534A} and 2MASS~J2139$-$0220 \\citep{2012ApJ...750..105R}. Combined, the achromatic and chromatic terms nearly cancel in the $K$-band, a region that is gas opacity dominated (H$_2$O, CH$_4$ and H$_2$). Hence, { our linear spectral model} is functionally consistent with condensate clouds being the primary driver of variability in {\\namesh}B. \\subsection{ A Brightness Temperature Spot Model for Luhman~16B} { Given the known shortcomings in reproducing the near-infrared spectra of L/T transition brown dwarfs (e.g., \\citealt{2008ApJ...682.1256L,2008ApJ...678.1372C,2009ApJ...702..154S}), we forgo detailed modeling of the spectra in lieu of a simply brightness temperature variation model, focusing at 1.25~$\\micron$ where gas opacity is a minimum and cloud structure variations are expected to have the greatest influence \\citep{2001ApJ...556..872A}. The simplest model for replicating the surface flux $\\langle{F}\\rangle$ of a patchy brown dwarf is two sets of regions with differing brightness temperatures covering the surface:} \\begin{equation} \\langle{F}\\rangle \\propto \\langle{T_{br}}^4\\rangle \\equiv {A}T_{cold}^4 + (1-A)T_{hot}^4. \\label{eqn:tsurf} \\end{equation} { Here, $\\langle{T_{br}}^4\\rangle$ is the disk-averaged brightness temperature,} $T_{cold}$ and $T_{hot}$ are the brightness temperatures of cold and hot { regions}, respectively, $A \\equiv F_{cold}/\\langle{F}\\rangle \\leq 1$ is the areal covering fraction of the cold { regions}, and we ignore limb darkening. Our interpretation of this model is that the cold { regions} correspond to the { highest cloud layer in the brown dwarf atmosphere}, while the hot { regions} correspond to gaps in these clouds that probe to some as-yet undetermined deeper layer with brightness temperature $T_{hot}$; { a cartoon perspective of this is shown in Figure 6 of \\citet{2013ApJ...768..121A}. We note that this is not the only interpretation of a two-spot model, which could also arise from magnetic interaction at the photosphere (i.e., starspots) or updrafts of warm air pockets driven by convective flows. Nevertheless, we will occasionally refer to the cold { region} as ``clouds'' and hot { regions} as ``holes'' in the following discussion.} \\begin{figure}[h] \\center \\epsscale{1.0} \\plotone{f5.eps} \\caption{{ Cold spot} covering fraction $A$ and change in covering fraction $\\Delta{A}$ (solid lines) on {\\namesh}B { over a full cycle}, as a function of the hot spot brightness temperature at 1.25~$\\micron$. Values for $A$ are derived from Eqn.~\\ref{eqn:tsurf} and assume $\\langle{T_{br}}\\rangle_{L16B}$ = { 1560~K} and $T_{cold}$ = ${\\langle}T_{br}\\rangle_{L16A}$ = { 1510~K}, based on Figure~\\ref{fig:variability} (vertical triple-dot dash lines). Values for $\\Delta{A}$ are derived from Eqns.~\\ref{eqn:da1} and~\\ref{eqn:da2} and assume $\\Delta{F}/F$ = 13.5\\%. The point where $\\Delta{A}$ = 100\\% indicates the minimum $T_{hot}$; { the temperature at which $A = 0.5$ is also labeled, separating cold surfaces with hot spots from a hot surface with cold spots.} Also labeled are the range of $T_{cr}$ values from \\citet{2005ApJ...621.1033T} and \\citet{2012ApJ...760..151S} that estimate the height of cloud tops, and the temperature at which all condensates are assumed to be evaporated ($T_{evap}$). Estimates for $A$ and $T_{hot}$ for SIMP~J0136+0933 and 2MASS~J2139$-$0220 from \\citet{2012ApJ...750..105R} and \\citet{2013ApJ...768..121A} are indicated, assuming the same cloud top temperature. Finally, we label estimates of jet size scales for wind velocities of { $U$ = 1.6~{\\kms} and 3.4~{\\kms} based on a Rhines length scale (Eqn.~\\ref{eqn:rhines}); these intersect the $\\Delta{A}$ curve at $T_{hot}$ = 1700~K and 1900~K}. } \\label{fig:patchy} \\end{figure} For {\\namesh}B, ${\\langle}T_{br}\\rangle$ = { 1560~K} at the 1.25~$\\micron$ $J$-band peak continuum (Figure~\\ref{fig:variability}; see also \\citealt{faherty14}). If we take the brightness temperature of {\\namesh}A at this wavelength, { 1510}~K, as an estimate for $T_{cold}$ for both sources,\\footnote{The assumption can be justified in part by the nearly identical brightness temperatures of the two sources in the 1.15~$\\micron$ region, where {\\wat} and {\\meth} opacity play a larger role than clouds.} then we can jointly constrain $A$ and $T_{hot}$, as illustrated in Figure~\\ref{fig:patchy}. { Coverage of cold regions} is essentially negligible for $T_{hot} <$ 1570~K, then climbs to over 50\\% at $T_{hot} \\approx$ 1860~K. At hotter temperatures, { our model predicts that the atmosphere of {\\namesh}B would be overall similar to that of {\\namesh}A with occasional hot spots, which we assume to be less than the evaporation temperature of mineral condensate species} ($T_{evap} \\approx$ 2000~K; \\citealt{1999ApJ...519..793L}). We note that { equal hot-cold spot coverage for {\\namesh}B occurs in} the 1700--1900~K range that \\citet{2005ApJ...621.1033T} estimate as the effective top of a brown dwarf cloud layer ($T_{cr}$; see \\citealt{2012ApJ...760..151S}). \\citet{2012ApJ...750..105R} and \\citet{2013ApJ...768..121A} also provide estimates for $A$ and $\\Delta{T}_{hc} = T_{hot}-T_{cold}$ for SIMP~J0136+0933 and 2MASS~J2139$-$0220 based on their own { two-spot modeling}. While temperature values reported in these studies are based on model effective temperatures, if we assume that the brightness temperature offsets are the same as their preferred $\\Delta{T}_{hc} \\approx$ 300~K, this would also place the hot { regions of {\\namesh}B in the same temperature range as} the cloud tops of the \\citet{2005ApJ...621.1033T} models. { Thus, if our spot model is interpreted as probing different layers of {\\namesh}B's atmosphere, the best estimates of the temperature differential is in line with the conjecture of \\citet{2013ApJ...768..121A} that gaps in the highest cloud deck still probe regions influenced by condensate opacity. However, we stress that our data cannot independently determine $A$ or $T_{hot}$, and other interpretations of these temperature differences are conceivable.} The fractional peak-to-peak variation in observed flux that occurs as { hot and cold regions} rotate in and out of view is \\begin{equation} \\frac{\\Delta{F}}{\\langle{F}\\rangle} = \\frac{\\Delta{A}}{A-\\epsilon} \\label{eqn:da1} \\end{equation} where $\\langle{F}\\rangle$ is the average flux, $\\Delta{F}$ the change in total flux, $\\Delta{A}$ the change in cloud coverage (increasing $A$ decreases the total flux), and $\\epsilon \\equiv F_{hot}/(F_{hot}-F_{cold}) = T_{hot}^4/(T_{hot}^4-T_{cold}^4) \\geq 1$; see also \\citet{2012ApJ...750..105R}.\\footnote{Our expression differs slightly from \\citet{2012ApJ...750..105R} because we assign $A$ to be the cloud-covering fraction, whereas they define the equivalent parameter $a$ as the cloud-cleared fraction.} In terms of brightness temperatures: \\begin{equation} \\Delta{T_{br}} = \\frac{\\Delta{F}}{4\\langle{F}\\rangle}\\langle{T_{br}}\\rangle = \\frac{\\Delta{A}}{4(A-\\epsilon)}\\langle{T_{br}}\\rangle. \\label{eqn:da2} \\end{equation} During our observations, we observed a 7.5\\% variation in the $J$-band peak continuum that was coincident with a 2.5\\% variation in TRAPPIST red-optical photometry. We therefore assume that the full 4.5\\% peak-to-peak variation in TRAPPIST photometry around our spectral observations (Figure~\\ref{fig:trappist}) corresponds to a 13.5\\% variation at $J$, or a peak-to-peak temperature fluctuation of $\\Delta{T_{br}} \\approx$ 50~K\\footnote{The maximum variation observed by \\citet{2013A&A...555L...5G} over weeks of monitoring is 10\\%, which would correspond to 30\\% variations in $J$, exceeding those observed in 2MASS~J2139$-$0220. However, we restrict our analysis here to the period around the spectral observations { since the spectral response of larger fluctuations may differ}.} following Eqn~\\ref{eqn:da2}. This temperature offset is notably similar to the temperature difference between {\\namesh}A and B at these wavelengths (Figure~\\ref{fig:variability}). Using the { relationship between} $A$ and $T_{hot}$ above, we computed $\\Delta{A}$ as a function of $T_{hot}$, also shown in Figure~\\ref{fig:patchy}. Not surprisingly, the { areal} variation required to reproduce the observed brightness variations declines with higher $T_{hot}$; i.e., with greater contrast between cold and hot { regions}. An important reference point is the temperature at which { areal} variations become smaller than the total { cold region} coverage, which occurs for $T_{hot} >$ 1710~K and $A >$ 30\\%. { The corresponding $\\Delta{T}_{hc}$ = 150~K is on the low end of estimates} for SIMP~J0136+0933 and 2MASS~J2139$-$0220, and just above the minimum $T_{cr}$ from \\citet{2005ApJ...621.1033T}. { For the range 1700~K $< T_{hot} <$ 1900~K, which we again take as a reasonable estimates of the hot spot temperature, the inferred cold covering fraction is roughly 30--55\\%, intermediate between similar values inferred for SIMP~J0136+0933 (25-30\\%) and 2MASS~J2139$-$0220 (50-65\\%) by \\citet{2013ApJ...768..121A}. \\subsection{ Interpretation: Rhines Length Scale and Advective Time Scale} For 1700~K $< T_{hot} <$ 1900~K, { cold spot coverage} must vary by 15--30\\% over a single period to replicate the observed variability amplitude, { implying a $\\sim$ 30-100\\% variation between hemispheres if the spot patterns are static. Organized jet features in the atmospheres of the giant Solar planets generally scale in size with the Rhines length \\citep{1970GApFD...1..273R,2008ASPC..398..419S}, $L_{Rh} \\sim \\left({U}/{2\\Omega{R}\\cos{\\phi}}\\right)^{1/2}$, where $U$ the characteristic wind speed, $R$ is the radius, $\\Omega= 2\\pi/P$, $P$ is the rotation period and $\\phi$ is the latitude of the feature. If we assume that the same scaling occurs for features in brown dwarf atmospheres (e.g., \\citealt{2013ApJ...768..121A,2013ApJ...776...85S}), then their maximum fractional size scale is: \\begin{equation} \\alpha_{Rh} \\sim \\left(\\frac{L_{Rh}}{R}\\right)^2 \\approx 2\\%\\left(\\frac{U}{\\rm km/s}\\right)\\left(\\frac{P}{\\rm hr}\\right)\\left(\\frac{R_{Jup}}{R}\\right). \\label{eqn:rhines} \\end{equation} where we have assumed mid-latitude features. If we now relate this maximum scale to the areal spot variation inferred here ($\\alpha_{Rh} \\sim \\Delta{A}$), the known rotational period and assumed radius of {\\namesh}B implies characteristic wind speeds of 1.6~{\\kms} $< U <$ 3.4~{\\kms} for 1700~K $< T_{hot} <$ 1900~K (Figure~\\ref{fig:patchy}). These speeds are somewhat higher than the range favored by the circulation models of \\citet{2013ApJ...776...85S}, assuming winds are driven by inefficient conversion of convective heat (10--300~{\\ms})}. However, the speeds do give advection timescales, $\\tau_{adv} \\sim R/U \\sim (2-5)\\times10^4~s \\sim 1-3$ rotation periods, that are consistent with the timescale of lightcurve evolution observed in {\\namesh}B \\citep{2013A&A...555L...5G}. The convergence between the inferred variation and Rhines length scales, and the advective and evolutionary time scales, suggest that our gross estimates for $T_{cold}$, $T_{hot}$, $A$ and $\\Delta{A}$ are not too far off the mark. However, we have made a number of major assumptions that require confirmation through more detailed spectroscopic monitoring and modeling, in particular to ascertain whether the spot regions have spectral characteristics (features and line profile shapes) consistent with the inferred brightness temperatures. Nonetheless, our basic model of a cold cloud deck disrupted by warm dynamic features shows promising agreement with planetary analogs and current brown dwarf circulation models.} \\subsection{Trends in L/T Transition Variability} \\begin{deluxetable*}{lcccccl} \\tablecaption{Comparison of Highly Variable L/T Transition Dwarfs. \\label{tab:comp}} \\tabletypesize{\\small} \\tablewidth{0pt} \\tablehead{ \\colhead{Source} & \\colhead{SpT} & \\colhead{$P_{rot}$} & \\colhead{$\\Delta{F}/F$} & \\colhead{$A$} & \\colhead{$\\Delta{K}/\\Delta{J}$} & \\colhead{Ref} \\\\ & & \\colhead{(hr)} & \\colhead{at 1.25~$\\micron$} \\\\ } \\startdata SIMP~J0136+0933\t& T2.5 & 2.3895$\\pm$0.0005 & 5.5\\% & 25-30\\% & 0.48$\\pm$0.06 & 1,2 \\\\ {\\namesh}B & T0.5 & 4.87$\\pm$0.01 & 13.5\\%\\tablenotemark{a} & 30-55\\%\\tablenotemark{b} & 0.41$\\pm$0.18 & 3,4 \\\\ 2MASS~J2139$-$0220 & T1.5 & 7.721$\\pm$0.005 & 30\\% & 50-65\\% & 0.45--0.83 & 2,5 \\\\ \\enddata \\tablenotetext{a}{Based on the maximum peak-to-peak TRAPPIST variability amplitude during the current observing period.} \\tablenotetext{b}{Assuming $T_{hot}-T_{cold}$ = 300$\\pm$100~K; see \\citet{2013ApJ...768..121A}.} \\tablerefs{ (1) \\citet{2009ApJ...701.1534A}; (2) \\citet{2013ApJ...768..121A}; (3) \\citet{2013A&A...555L...5G}; (4) This paper; (5) \\citet{2012ApJ...750..105R}.} \\end{deluxetable*} {\\namesh}B joins SIMP~J0136+0933 and 2MASS~J2139$-$0220 as the three most variable L/T transition objects detected to date, so it is worth comparing the variability properties of these sources, summarized in Table~\\ref{tab:comp}. The variability period, $J$-band variability amplitude and inferred cloud covering fraction of {\\namesh}B are all intermediate between those of SIMP~J0136+0933 and 2MASS~J2139$-$0220, although epoch-to-epoch changes in these values are considerable. As the Rhines scale scales linearly with the rotation period,\\footnote{\\citet{2013ApJ...768..121A} incorrectly state a spot scaling law of ${P}^{-2}$ in the text, but infer a spot scaling between 2MASS~J2139$-$0220 and SIMP~J0136+0933 that is consistent with $A \\propto P$; the former is likely a typographical error.}, { its interpretation as an estimate of surface feature size is consistent with {\\namesh}B's intermediate period and intermediate variability amplitude, as a few large features are more likely to give rise to stronger disk-integrated variations than many small features \\citep{2013ApJ...768..121A}.} There also appears to be a correlation between rotation period and cloud covering fraction, although temperature effects may play a role in this statistic. The source with the smallest cloud coverage, SIMP~J0136+0933, is also the latest-type and presumably coldest brown dwarf in the sample. Finally, we find essentially no difference in color variability among these sources. As noted above, our estimate of $\\Delta{F_{K_s}}/\\Delta{F_{J}}$ for {\\namesh}B is consistent with similar measures for SIMP~J0136+0933 and 2MASS~J2139$-$0220 { (although the latter can exhibit more extreme color terms; \\citealt{2012ApJ...750..105R}), suggesting that the condensate clouds responsible for the variations in these sources are likely to} have similar opacities and physical properties (i.e., composition, grain size distribution, vertical structure, etc.). { However, confirmation of this agreement will again require more careful spectral modeling to accurately determine cloud properties.}" }, "1402/1402.5081_arXiv.txt": { "abstract": "{% In Monte-Carlo simulations of gamma-ray or cosmic-ray detector arrays on the ground (here mainly arrays of imaging atmospheric Cherenkov telescopes), the atmosphere enters in several ways: in the development of the particle showers, in the emission of light by shower particles, and in the propagation of Cherenkov light (or fluorescence light or of particles) down to ground level. Relevant parameters and their typical impact on energy scale and so on are discussed here. } ", "introduction": "Ground-based gamma-ray and cosmic-ray detectors make use of the atmosphere as a part of their instruments. Incoming energetic particles (neglecting neutrinos or exotic particles for the purpose of this paper) will interact with the atmosphere, produce secondary particles and initiate a particle cascade or extensive air shower. The level of inclusion of the atmosphere into the instrument differs between different detection techniques. Particle detector arrays are affected by the development of extensive air showers. Beyond the atmospheric overburden on ground, the atmospheric profile will also matter, due to the competition between interactions and decays of secondary unstable particles produced in interactions. Since most of the development of air showers happens at altitudes of up to a few ten kilometers of altitude, the change of atmospheric composition beyond about 80 km altitude has, fortunately, no significant impact onto the air shower development and detection. For atmospheric Cherenkov detectors as well as for air fluorescence detectors, the atmosphere enters at two additional levels. The first is the emission process and the second is the propagation of the emitted light down to the photo-detectors. The Cherenkov and fluorescence processes depend in different ways on changes in density profiles. For the fluorescence technique, there is a competition between fluorescence light emission and collisional de-excitation, after the nitrogen molecules get excited by the charged particles in the air shower. The higher efficiency of light emission is balanced basically by lower excitation in lower density air. As a consequence, the fluorescence process depends relatively weakly on the air density profile, except for the shower development itself. The Cherenkov emission process is more strongly impacted by the density profile. Higher density, resulting in an almost proportionally enhanced refractivity (index of refraction minus one, $n-1$), results in enhanced emission as well as lowering the minimum particle velocity necessary for emission. It also increases the Cherenkov emission angle and spreading out the light on the ground over a larger area. Both the Cherenkov and fluorescence techniques are affected by the light propagation, by absorption as well as by scattering. The absorption processes are not just affected by the main composition components but also affected by trace gases, like ozone or water vapour. At that stage the detailed composition profile thus enters the game. Scattering includes Rayleigh scattering - thus directly linked to the density profile - and Mie scattering on aerosols (also resulting in light absorption). As a consequence, these detection techniques are affected by the aerosol density, chemical composition, size distributions -- and all of these depend on altitude and change with time. Simulations of air showers and detector response usually need some simplifications, in order to run in an efficient way. This paper will also try to point out some often-used simplifications which should be applied with care. Since this paper is focusing on the impact of atmospheric parameters, a number of other site-related topics will be omitted, including the impact of the night-sky background (NSB) on Cherenkov or fluorescence detectors and the impact of the geomagnetic field. ", "conclusions": "Knowledge of the atmospheric density profile is important for proper simulation of any type of ground-based air shower instrument, for particle detector arrays as well as for Cherenkov or fluorescence detectors. For the latter two the profile enters also in the light emission processes. It is particularly relevant for the Cherenkov emission where the related index of refraction determines the minimum velocity for Cherenkov emission, the number of photons emitted per unit path length and also the Cherenkov cone opening angle. Most simulations assume a wavelength-independent index of refraction but the wavelength dependence can be turned on, at the expense of the simulation efficiency. The extinction of Cherenkov and fluorescence light depends in addition to Rayleigh scattering and absorption on O$_2$ also on aerosols and on trace gases like ozone and water vapour. The tropospheric ozone absorption is of particular concern for Cherenkov detectors with extended UV response using muon rings for calibration. Hardly any simulations make use of measured ozone profiles. Except for this ozone part, the atmospheric transmission tables used in the simulations -- in particular the total aerosol optical depth -- can be best checked with star light extinction. For further improving the accuracy of the simulations, the aerosol profile along the line of sight should be monitored. Nevertheless tricky are aerosol layers intersecting the shower development, requiring dedicated and well adapted simulations for correcting their effect. Scattered light, in particular scattered Cherenkov light, is highly relevant to the fluorescence technique but only has a small effect for IACTs. For IACT simulations, it is generally ignored." }, "1402/1402.0989_arXiv.txt": { "abstract": "Spectral and Temporal properties of black hole candidates can be explained reasonably well using Chakrabarti-Titarchuk solution of two component advective flow (TCAF). This model requires two accretion rates, namely, the Keplerian disk accretion rate and the halo accretion rate, the latter being composed of a sub-Keplerian, low angular momentum flow which may or may not develop a shock. In this solution, the relevant parameter is the relative importance of the halo (which creates the Compton cloud region) rate with respect to the Keplerian disk rate (soft photon source). Though this model has been used earlier to manually fit data of several black hole candidates quite satisfactorily, for the first time, we made it user friendly by implementing it into XSPEC software of GSFC/NASA. This enables any user to extract physical parameters of the accretion flows, such as two accretion rates, the shock location, the shock strength etc. for any black hole candidate. We provide some examples of fitting a few cases using this model. Most importantly, unlike any other model, we show that TCAF is capable of predicting timing properties from the spectral fits, since in TCAF, a shock is responsible for deciding spectral slopes as well as QPO frequencies. ", "introduction": "Compact objects, such as black holes, neutron stars, etc. are identified by electromagnetic radiations emitted from the accreting matter. Understanding the spectral and timing properties of this radiation is essential for model builders and theorists alike. In a binary system, matter from the companion star accrets into the black hole through the Roche lobe, and/or through capturing its winds. This matter produces a disk like structure around the primary compact object. There are a large number of theoretical models in the literature which explain the physics of accretion around a black hole. Evidences of the standard disk proposed by \\citet{SS73} and \\citet{NT73} % are present in most of the binary systems. However, the emitted spectrum of the radiation is multi-color in nature and contains both thermal and non-thermal components and a standard disk cannot explain the entire X-ray spectral features. Moreover, the inner region of the standard disk may be unstable due to the viscous and thermal effects \\citep{Lightman74,Kobayashi03}. % Simply put, one of the components of the spectrum is a multi-color blackbody radiation from the standard Keplerian disk and the other is a power-law component formed due to repeated Compton scatterings of the low energy (soft) photons of this blackbody by the hot electrons of the `Compton' cloud \\citep{ST80,ST85}. % There are many speculations regarding the nature of this Compton cloud ranging from a magnetic corona \\citep{Galeev79}, to hot gas corona over the disk \\citep{Haardt93,Zdziarski03}. Since the formation process of a static corona around an accretion disk is totally unknown, and since a low angular dynamic flow may naturally act as a corona, \\citet[][hereafter CT95]{CT95} % proposed that a disk having two distinct components, a Keplerian disk submerged inside a sub-Keplerian halo is enough to explain all the spectral properties very satisfactory. Observational evidences also started to support this so-called two-component advective flow (TCAF) \\citep[e.g.,][]{Soria01,Smith02,Wu02,Cambier13}. % While creating a self-consistent TCAF solution, the properties of a viscous transonic flow was made use of, in which a flow having viscosity above a critical value naturally forms a Keplerian disk and the region with a lower viscosity, due to centrifugal barrier, forms a shock wave, typically, at a few tens of Schwarzschild radii. The post-shock region (from the shock and the inner sonic point) basically evaporates the Keplerian component and together acts as a Compton cloud which produces a power-law component (hard photons) with exponential cut-off in the spectrum through thermal Comptonization. From the inner sonic point to the horizon of the black hole (bulk motion dominated advective flow or BDAF) the matter is advected rapidly to the black hole. The bulk motion in this region also up-scatters the soft photons and produces a second power-law component even when the temperature of the region is zero. If the centrifugal barrier is not strong enough, the shock may not form, but the flow still slows down. The spectral properties in this case are discussed in \\citet[][hereafter C97]{C97}. The CENtrifugal pressure supported BOundary Layer or CENBOL, referred to the post-shock region or centrifugal force dominated region, which is also the base of the outflows where the pre-Jet is launched, plays the most important role in the black hole physics. As usual, this CENBOL, pre-Jet and BDAF intercept soft photons from the Keplerian disk and reprocess them to high energies via inverse Compton scattering. In this {\\it letter}, we will implement the TCAF solution to study spectral properties of black hole candidates (BHCs) using widely used user-friendly spectral analysis software package, developed by GSFC/NASA, called XSPEC. For the sake of concreteness, we focus only the cases where only the CENBOL is present. We ignore the effects of BDAF and pre-Jet. In our next version of analysis, these components and spin of the black hole would be included. The Galactic transient black hole candidates are very interesting objects to study in X-rays because these sources generally show rapid evolutions in their temporal and spectral properties during their outburst phases, which are strongly correlated to each other \\citep[see for a review,][]{RM06}. In general, four basic states - {\\it hard, hard-intermediate, soft-intermediate}, and {\\it soft} states are observed during an outburst of the BHCs \\citep[see,][and references therein]{Nandi12}. The evolutions of these spectral states are observed, which indeed form a hysteresis loop during the outburst with hard states are found to be at the beginning and end time of the outbursts, whereas soft and intermediate spectral states are observed in between. The evolution of spectral states are strongly dependent on the variation of the accretion rates. According to the TCAF solution, accretion flow rates may be controlled by a physical parameter, such as the magnetic viscosity, perhaps owing to the enhanced magnetic activity of the companion \\citep{Wu02,Nandi12,DD13}. During the rising phase of the outburst, viscosity may cause an increase in the accretion rate of the Keplerian matter. As the viscosity is reduced, the Keplerian rate is reduced and declining phase starts. The Keplerian disk itself recedes away leaving behind only the low-angular sub-Keplerian flow causing a hard state. Thus, a rigorous fit with TCAF model is expected to throw light on how the accretion rates and the flow geometry evolve with time. In general, low and intermediate frequency quasi-periodic oscillations (LFQPOs) are observed in hard and intermediate (hard-intermediate and soft-intermediate) spectral states of transient black hole candidates. These QPOs are reported extensively in literature, although still there are debates on the origin of these QPOs. However, according to the shock oscillation model (SOM) by Chakrabarti and his collaborators, LFQPOs are originated due to the oscillation of the post-shock region (\\citealt{MSC96}, hereafter MSC96; \\citealt{CAM04}, hereafter CAM04; \\citealt{GGC14}, hereafter GGC14) when the resonance occurs between the infall time scale and the cooling time scale in CENBOL. During oscillation, the shape of the Compton cloud and the degree of interception change periodically. Since from our spectral fit, we can directly extract values of physical parameters related to this shock wave, we can also predict what should be the frequency of the observed low frequency QPO (if present; see \\S 4.1 for details). The {\\it paper} is organized in the following way: in the next Section, we briefly describe properties of the TCAF model. In \\S 3, we discuss the method of the implementation of the TCAF model in XSPEC for spectral fittings. In \\S 4, TCAF model fitted results obtained from the spectral fit of three different BHCs. Finally, in \\S 5, we make concluding remarks and our future work plans. ", "conclusions": "" }, "1402/1402.1177_arXiv.txt": { "abstract": "{ Polaris the Cepheid has been observed for centuries, presenting surprises and changing our view of Cepheids and stellar astrophysics, in general. Specifically, understanding Polaris helps anchor the Cepheid Leavitt law, but the distance must be measured precisely. The recent debate regarding the distance to Polaris has raised questions about its role in calibrating the Leavitt law and even its evolutionary status. In this work, I present new stellar evolution models of Cepheids to compare with previously measured CNO abundances, period change and angular diameter. Based on the comparison, I show that Polaris cannot be evolving along the first crossing of the Cepheid instability strip and cannot have evolved from a rapidly-rotating main sequence star. As such, Polaris must also be at least 118~pc away and pulsates in the first overtone, disagreeing with the recent results of Turner et al. (2013). } ", "introduction": "The North Star, Polaris (HD~8890), has fascinated people for centuries, guided explorers, and appears in the mythologies of numerous cultures. Polaris is also the nearest classical Cepheid, making it a powerful laboratory for understanding stellar evolution as well as anchoring the calibration of the Cepheid Leavitt Law (period-luminosity relation). However, one of the greatest hindrances to understanding Polaris is its distance. \\cite{vanLeeuwen2007} measured a distance $d = 129 \\pm 2~$pc from revised Hipparcos parallaxes, whereas \\cite{Turner2013} found a distance $d = 99\\pm 2$~pc based on spectroscopic line ratio measurements. The two distances are both measured precisely, yet disagree significantly. Furthermore, these distances are particularly interesting because they suggest different pulsation properties and evolutionary histories for Polaris. Analysis of interferometric observations suggest an angular diameter for Polaris of $\\theta = 3.123 \\pm 0.008$~mas \\citep{Merand2007}, hence a mean radius of either $33.4 \\pm 0.6~R_\\odot$ or $43.5 \\pm 0.8~R_\\odot$ depending on which distance one considers. If the radius is the former value then the period-radius relation \\citep[e.g.][]{Gieren1997, Neilson2010, Storm2011} implies that Polaris is pulsating as a fundamental mode Cepheid, whereas if the radius is the larger value then Polaris must be pulsating as a first overtone Cepheid. Understanding the pulsation mode is necessary for calibrating the Cepheid Leavitt Law \\citep{vanLeeuwen2007} and constraining the structure of the Cepheid instability strip on the Hertzsprung-Russell diagram. The distance debate also presents challenges for understanding the evolution of Polaris. Assuming Polaris is at the closer distance, \\cite{Turner2013} argued that Polaris is evolving along the first crossing of the Cepheid instability strip because of its large rate of period change \\citep{Turner2006, Neilson2012a}. In that case, Polaris has evolved from a main sequence star but has not yet become a red giant star. \\cite{Turner2013} also suggested that Polaris must have rotated rapidly as a main sequence star to explain the measured nitrogen and carbon abundances \\citep{Usenko2005}. On the other hand, \\cite{Neilson2012a} suggested that Polaris cannot be evolving along the first crossing if it is at the farther distance because the measured rate of period change is too small relative to a first-crossing Cepheid with a similar luminosity. Hence, Polaris is evolving along the third crossing of the instability strip on the Cepheid blue loop and its observed CNO abundances are consistent with dredge up during the previous red giant stage. \\cite{Neilson2012a} further suggested that to account for discrepancy between measured and predicted rates of period change, Polaris must be losing mass in an enhanced stellar wind with $\\dot{M} = 10^{-7}$ -- $10^{-6}~M_\\odot$~yr$^{-1}$. Our understanding of Polaris is biased by the assumed distance, therefore measuring a precise and consistent distance is crucial. However, there is no a priori reason to prefer one measured distance over another. \\cite{Turner2013} argued that Hipparcos parallax must be wrong because it is inconsistent with other measurements such as space velocities of nearby stars, and main sequence fitting of Polaris's F3~V binary companion. \\cite{vanLeeuwen2013} countered that the first argument could not be reproduced, suggesting that Polaris is not a member of the local cluster. Furthermore, the F3~V stars have a large enough color distribution that would be consistent with both distances. \\begin{figure*}[th] \\begin{center} \\includegraphics[width=0.48\\textwidth]{plot1-9.eps}\\includegraphics[width=0.48\\textwidth]{plot1-6.eps} \\end{center} \\caption{Predicted rates of period change for stars crossing the Cepheid instability strip assuming different prescriptions for Cepheid mass loss: $\\dot{M} = 10^{-9}~M_\\odot~$yr$^{-1}$ (left) and $10^{-6}~M_\\odot~$yr$^{-1}$ (right), the results for the other two cases are nearly indistinguishable. The grey band refers to the measured rate of period change for Polaris \\citep{Neilson2012a}.} \\label{plot_pdot} \\end{figure*} There are also questions regarding the spectroscopic method employed by \\cite{Turner2013}. That method is based on the FGK supergiant calibration of \\cite{Kovtyukh2007} and \\cite{Kovtyukh2010}, hence, stellar effective temperatures and luminosities are determined by measuring various iron line ratios from optical spectra. However, Cepheid variable stars are not static yellow supergiants. For instance, \\cite{Kervella2006} and \\cite{Merand2006, Merand2007} detected infrared excesses about a sample of Galactic Cepheids, but not for the yellow supergiant star $\\alpha$ Persei. There are also observations of dynamic motions in the photosphere driven by pulsation that do not occur in static yellow supergiant stars \\citep[e.g.][]{Nardetto2008}. \\cite{Engle2012} reported X-ray and UV observations of the nearest Cepheids including Polaris, suggesting the presence of hot ($\\approx 100,000$~K) plasma in the photosphere, which may affect the ionization rates of various species. These differences could affect the systematic uncertainties of the effective temperature and luminosity measured for Polaris. However, in spite of these differences, \\cite{Turner2013} reported a mean effective temperature $\\langle T_{\\rm{eff}} \\rangle = 6025\\pm 1$~K and $\\langle M_V \\rangle = 3.07\\pm 0.01$, where the presented error is statistical only. There are no systemic uncertainties presented. Convection alone would lead to a greater variation of temperature. However, to be consistent with the Hipparcos parallax the uncertainty of the luminosity must be $\\ge 0.6$ dex. The current distance debate to Polaris is confusing our understanding of the star from the context of stellar evolution and pulsation. In this article, I compute new state-of-the-art stellar evolution models to compare with various observations of Polaris, such as the rate of period change \\citep{Turner2005, Neilson2012a}, angular diameter \\citep{Merand2006}, and CNO abundances \\citep{Usenko2005}. We will also compare these models to the estimates of the distance and effective temperature, hence radius and luminosity. Comparing stellar evolution models to the $[N/H]$ and $[C/H]$ abundances test the hypothesis that Polaris is evolving along the first crossing of the instability strip, i.e., Polaris was once a rapidly rotating during main sequence star. Comparing models with the observed rate of period change provides insight into the mass-loss rate and which crossing of the instability strip Polaris is evolving, while comparing the models to the angular diameter constrains the radius and luminosity of the star, hence its distance. In Sect.~2, I briefly describe the stellar evolution models computed. In Sect.~3, I describe how Cepheid rates of period change are computed from stellar evolution models. I also compute rates of period change as a function of stellar mass-loss rate to compare to the observed rate of period change. I compare stellar evolution models with the observed nitrogen and carbon abundance to constrain the hypothesis that Polaris is a first-crossing Cepheid in Sect.~4. In Sect.~5, I discuss the implications of these tests for the distance and pulsation mode. ", "conclusions": "The debate over the distance to Polaris is important for the calibration of the Cepheid Leavitt Law, as the nearest Cepheids will calibrate the zero-point \\citep[e.g.][]{Freedman2012}. Resolving the debate is also important for constraining stellar evolution physics, and constraining the properties of other Cepheids, which will become particularly important as GAIA observations will measure parallaxes for thousands of Cepheids \\citep{Windmark2011}. In this work, I presented new stellar evolution models to reanalyze the fundamental properties of Polaris, this time without assuming a distance. The observed rate of period change is found to be inconsistent with stellar evolution models for Cepheids, for both the first and third crossings of the instability strip, regardless of whether the \\cite{vanLeeuwen2013} or \\cite{Turner2013} distance is assumed. There is no obvious theory where Polaris is consistent with stellar evolution on the first crossing of the instability strip, since that time scale is determined by dynamical time scales. On the other hand, mass loss could explain the rate of period change for a Cepheid on the third crossing, even though mass loss also reduces the effective temperature width of the Cepheid blue loop making it difficult to compare with the observed properties of Polaris. However, it can be noted that the measured period change is inconsistent with a first-crossing Cepheid. I further computed new rotating stellar evolution models to explore the hypothesis that Polaris is a first-crossing Cepheid with nitrogen and carbon abundances changed by rotational mixing. Rotational mixing models predict abundances roughly consistent with measurements by \\cite{Usenko2005} during the first-crossing of the instability strip, however, the predicted rotation rates are too large to be consistent with the measured turbulent velocity for Polaris \\citep{Usenko2005} and rotational velocities for a sample of Galactic Cepheids \\citep{Bersier1996}. As such, the properties of Polaris are inconsistent with being a first-crossing Cepheid and must be a third-crossing Cepheid. Because Polaris is a third-crossing Cepheid, then it must also be at a distance greater than $118$~pc, significantly greater than that measured by \\cite{Turner2013} and must also be pulsating as a first-overtone Cepheid. While the results of \\cite{Turner2013} are inconsistent with stellar evolution calculations, there is still much to explore about the nearest Cepheid, Polaris to constrain the physics behind stellar multiplicity \\citep[e.g.][]{Evans2013}, mass loss \\citep{Neilson2012a, Neilson2012b}, ultraviolet and x-ray radiation \\citep{Engle2012}, the apparent brightening of Polaris \\citep{Engle2014} and late-stage evolution \\citep[e.g.][]{Langer2012} as we approach an era of precision stellar astrophysics." }, "1402/1402.4108_arXiv.txt": { "abstract": "In this follow-up work to the High Energy Physics Community Summer Study 2013 (HEP CSS 2013, a.k.a. \\textsc{Snowmass}), we explore the scientific capabilities of a future Stage-IV Cosmic Microwave Background polarization experiment (CMB-S4) under various assumptions on detector count, resolution, and sky coverage. We use the Fisher matrix technique to calculate the expected uncertainties in cosmological parameters in $\\nu\\Lambda$CDM that are especially relevant to the physics of fundamental interactions, including neutrino masses, effective number of relativistic species, dark-energy equation of state, dark-matter annihilation, and inflationary parameters. To further chart the landscape of future cosmology probes, we include forecasted results from the Baryon Acoustic Oscillation (BAO) signal as measured by DESI to constrain parameters that would benefit from low redshift information. We find the following best 1-$\\sigma$ constraints: $\\sigma(M_{\\nu})$ $= 15$ meV, $\\sigma(N_{\\rm eff})$ $= 0.0156$, Dark energy Figure of Merit = 303, $\\sigma(p_{ann})$ $= 0.00588\\times3\\times10^{-26}$ cm$^3$/s/GeV, $\\sigma(\\Omega_K)$ $= 0.00074$, $\\sigma(n_s)$ $= 0.00110$, $\\sigma(\\alpha_s)$ $= 0.00145$, and $\\sigma(r)$ $= 0.00009$. We also detail the dependences of the parameter constraints on detector count, resolution, and sky coverage. ", "introduction": "In the past two decades, CMB experiments have made great strides in sensitivity improvement. For satellite experiments, there was a factor of 10 better in sensitivity from \\textit{COBE} to \\textit{WMAP} and from \\textit{WMAP} to \\textit{Planck}~\\cite{2013arXiv1309.5383A}. These satellites, along with numerous ground-based and balloon-borne experiments, as listed in Ref. \\cite{Lambda}, collectively moved us from setting upper limits on the temperature anisotropy ($C_\\ell^{TT}$), the E- and B-mode polarization ($C_\\ell^{EE}$ and $C_\\ell^{BB}$) spectra to precision measurements of the temperature power spectrum (e.g.~\\cite{2013ApJ...779...86S, 2013arXiv1303.5075P, 2013arXiv1301.1037D}), shrinking error bars on the E-mode polarization spectrum (e.g.~\\cite{2012ApJ...760..145Q, 2013arXiv1310.1422B}), and detecting gravitational lensing in the B-mode polarization power spectrum~\\cite{2013arXiv1307.5830H, 2013arXiv1312.6645P,2013arXiv1312.6646P}. Tremendous opportunities still lie in CMB polarization. CMB polarization maps can be projected into a curl-free component (E-modes), and a divergence-free component (B-modes). Scalar and tensor perturbations seeded the temperature anisotropies in the CMB. While E-modes, like temperature anisotropies, could be sourced by both scalar and tensor perturbations through Thomson scattering, the only primordial source for B-modes is from tensor perturbations i.e. gravitational waves. Detecting this B-mode polarization signal, which peaks at the degree angular scale, would be a powerful probe of inflation. B-modes at smaller angular scales are generated by gravitational lensing of E-modes by Large-Scale Structures (LSS). The conversion from E- to B-modes provides an exceptionally clean way of reconstructing the lensing potential $\\phi$ and its power spectrum $C_\\ell^{\\phi\\phi}$ \\cite{2003PhRvD..67h3002O}. The latter is sensitive to the properties of the LSS, making it a probe of the late universe ($z < 10$ \\cite{2006PhR...429....1L}). As a result, among other astrophysical properties, it is central to the measurements of the total neutrino mass and dark-energy equation of state. Different techniques have been developed to reconstruct lensing potential from CMB temperature and polarization maps \\cite{2003PhRvD..67h3002O, Hirata:2003ka}. Recently, experiments have detected gravitational lensing in CMB temperature maps with high enough significance to measure the lensing potential power spectrum ~\\cite{2011PhRvL.107b1301D,2012ApJ...756..142V,2013arXiv1303.5077P}. Moreover, the first reconstruction through E-to-B channel has recently been demonstrated \\cite{2013arXiv1312.6646P}. The lensing B-mode signal also acts as a foreground to the inflationary B-mode signal. Specifically, if the tensor-to-scalar ratio $r$, which parametrizes the amplitude of the tensor perturbation, is smaller than 0.02, the level of inflationary B-modes would be lower than the lensing B-modes \\cite{Kesden:2002ku,Knox:2002pe}. Algorithms have been developed to remove the lensing-induced B-modes \\cite{2012JCAP...06..014S} to study the primordial signal. Therefore, measuring the lensing potential to high precision is fundamental to advancing our understanding of both the early and the late universe. We envision the Stage-IV\\footnotemark \\footnotetext{Each advance in stage refers roughly to an order of magnitude increase in number of detectors \\cite{2013arXiv1309.5383A}.} Cosmic Microwave Background polarization experiment (CMB-S4) to be a high-resolution, high-sensitivity experiment for this very purpose. This paper explores how much CMB-S4 could help answer the pressing questions in High Energy Physics (HEP) --- more broadly, physics of fundamental interactions. In what follows, we outline the main HEP topics that can be addressed by CMB polarization.\\\\ \\textbf{Cosmic Neutrino background --} The relic neutrino number density for each species is 112 cm$^{-3}$ \\cite{Fukugita}. This makes neutrinos the second most abundant particle in the universe, after photons. Although neutrinos are massless in the Standard Model, solar and atmospheric neutrino oscillation experiments have shown that they are massive. Assuming normal hierarchy, a lower limit on the total neutrino mass is $\\sim58$ meV. Though minuscule in mass, their large cosmic abundance enables us to observe their cumulative gravitational effect on structure formation: neutrinos suppress structure growth below scales defined by their free streaming distances. As a result, we can measure the total neutrino mass by observing the matter power spectrum -- which CMB lensing is particularly sensitive to~\\cite{2003PhRvL..91x1301K}. Another way to study the cosmic neutrino background is to measure the extra number of relativistic species $N_{\\text{eff}}$ aside from photons at recombination. In the Standard Model, no other particles except neutrinos contribute as extra relativistic species. In this scenario, $N_{\\text{eff}}$ is predicted to be 3.046. Deviation from this value hints at new physics. The CMB is uniquely sensitive to the photon-baryon-dark matter interaction at recombination. The current constraints on $N_{\\text{eff}}$ are $3.30^{+0.54}_{-0.51}$ (at the $95\\%$ C.L.) from \\textit{Planck} + \\textit{WMAP} polarization + small scale CMB data + BAO measurements, and the current limit on total neutrino mass is $< 0.23\\, \\text{eV}$ (at the $95\\%$ C.L.) from \\textit{Planck} data + BAO data \\cite{2013arXiv1303.5076P}. \\\\ \\textbf{Dark Energy --} Since the discovery of dark energy \\cite{1998AJ....116.1009R,1999ApJ...517..565P}, its contribution to the energy density of the universe $\\Omega_{\\Lambda}$ has been determined to percent level over the past 15 years. However, its fundamental nature still remains mysterious. The three broad possibilities are: (1) vacuum energy manifesting as a cosmological constant; (2) a spatially homogeneous dynamical field; (3) modified gravity on cosmological scales. These hypotheses can be tested experimentally by measuring the dark-energy equation of state $w$ through the clustering of LSS as a function of redshift. So far, measurements are consistent with a cosmological constant. Over the next decade, photometric and spectroscopic surveys such as LSST, Euclid, and DESI will take these tests to the next level of precision, opening up opportunities for new discoveries \\cite{Albrecht2012}. CMB lensing is highly complementary to these surveys, because it provides high-redshift information in the linear regime, cf.~\\cite{2001PhR...340..291B}. It is especially valuable for searches of early evolution of $w$ when $z>2$, which would not be possible with galaxy surveys. In addition, CMB lensing can help optical lensing surveys calibrate their shear bias and boost constraints on cosmological parameters, \\cite{2013arXiv1311.0905P, 2013arXiv1311.2338D, 2013arXiv1311.6200H}. The current constraints of $w_0$ and $w_a$ in the $w = w_0+w_a(1-a)$ model are: $w_0 = -1.04^{+0.72}_{-0.69}$ and $w_a < 1.32$ (\\textit{Planck}+\\textit{WMAP} polarization + BAO data), cf.~\\citet{2013arXiv1303.5076P}. \\\\ \\textbf{Dark-matter annihilation--} The gravitational properties of dark matter (DM) have been well constrained by data from a plethora of measurements, e.g. weak~\\cite{2003ARA&A..41..645R} and strong~\\cite{1998ApJ...498L.107T} lensing, multi-wavelength studies of the Bullet Cluster \\cite{2006ApJ...648L.109C}, distant supernovae~\\cite{1998AJ....116.1009R,1999ApJ...517..565P}, and the CMB~\\cite{2013arXiv1303.5076P}. However, the particle nature of dark matter remains unknown. There are several types of experiments aiming to understand dark matter interactions: (1) direct detection experiments, which rely on nuclear recoil signature of DM-Standard Model particle scattering, (2) indirect detection experiments, which are sensitive to the products of DM decay or annihilation, and (3) collider experiments. For a DM theory and detection review, see Ref. \\cite{2010ARA&A..48..495F}. An alternative observable is the CMB two-point correlation function. When DM annihilates, heat is transferred to the photon-baryon fluid, atoms are ionized and/or excited~\\cite{2004PhRvD..70d3502C}. \\citet{2013PhRvD..87l3513S} performed a detailed study regarding the energy deposition history into the photon-baryon fluid. Dark-matter annihilation also leads to growing ionization fraction perturbations and amplified small-scale cosmological perturbations, leaving an imprint on the CMB bispectrum \\cite{2013PhRvD..87j3522D}. CMB temperature and polarization spectra can constrain the parameter $p_{ann} \\equiv f \\langle \\sigma v \\rangle\\,/\\,m_{DM}$, where $f$ is the fraction of energy deposited into the plasma, $\\langle \\sigma v \\rangle$ is velocity-weighted cross section, and $m_{DM}$ is the mass of the DM particle. Current constraints coming from \\textit{WMAP} $9$-year data, \\textit{Planck}, ACT, SPT, BAO, HST and SN data exclude DM masses below $26$ GeV at the 2$\\sigma$ level, assuming that all the energy is deposited in the plasma \\cite{2013arXiv1310.3815M}. We show in this paper that CMB-S4 will tighten these constraints by a factor of $10$.\\\\ \\textbf{Inflation --} A detection of degree-scale B-mode polarization would help us learn about the physics of inflation. In particular, the amplitude of tensor perturbations is directly related to the energy scale of inflation. In addition, whether $r$ is above or below $\\sim$0.01 determines whether the inflaton field range is sub- or super-Planckian in a broad class of inflationary models. Beyond B-modes, small angular scale E-mode polarization, which is less contaminated by foregrounds than the temperature maps, can test the following predictions from slow-roll inflation: (1) the nearly scalar invariant primordial power spectrum where $n_s$ is close to but not exactly 1, (2) the small running of $n_s$, and (3) the almost flat mean spatial curvature, $\\Omega_K \\approx 10^{-4}$. The current tightest constraint for $n_s$ is $0.9603 \\pm 0.0073$ at the $68\\%$ C.L., for $dn_s/d\\log k$ is $-0.014^{+0.016}_{-0.017}$ at the $95\\%$ C.L., and for $\\Omega_K$ is $-0.0005^{+0.0065}_{-0.0066}$ at the $95\\%$ C.L. \\cite{2013arXiv1303.5076P}. \\\\ We show in this paper, given the range of possible experimental configurations (survey coverage, depth, resolution), how sensitive future ground-based experiments like CMB-S4 will be in probing these areas of new physics. We deliberately extend the experimental configuration space stated in the Snowmass study for CMB-S4 so that we understand the steepness of the improvement in parameter constraints as a function of experimental setups. Furthermore, we detail the forecast methods which are lacking in the Snowmass report. The paper is organized as follows. First, in Section~\\ref{sec:methods}, we describe the methodology and the fiducial cosmology. Sections ~\\ref{sec:neutrinos} to \\ref{sec:Inflation} are assigned to the aforementioned HEP topics and parameters of interests: \\begin{itemize} \\item Section~\\ref{sec:neutrinos}: total neutrino mass and extra relativistic species -- $M_{\\nu}$ and $N_{\\rm{eff}}$ \\item Section~\\ref{sec:DarkEnergy}: dark-energy equation of state -- $w_0$ and $w_a$ \\item Section~\\ref{sec:DarkMatter}: dark-matter annihilation -- $p_{ann}$ \\item Section~\\ref{sec:Inflation}: inflation -- $\\Omega_K$, $n_s$, $\\alpha_s$ (running of $n_s$), $r$, and $n_t$ \\end{itemize} For each section above, we present a brief overview of the underlying physical phenomena and their effects on the CMB power spectra, and present and discuss the results of the forecast. For the tensor-to-scalar ratio $r$, we have a more thorough method section outlined in Section~\\ref{sssec:r-method}. Finally, we conclude and summarize the main results in Section~\\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} In this work, we forecasted how well a highly capable next-generation ground-based CMB experiment can constrain cosmological parameters of fundamental physics relevant to both high energy physics and cosmology. We forecast for a range of experimental inputs -- $10^4$ to $10^6$ detectors, $1'-4'$ beams ($6'$, $8'$ in some cases), $1-75\\%$ sky fraction in order to see how the constraints for each parameters vary with these inputs. We detailed in section~\\ref{sec:methods} the methods used to estimate the performance of a given CMB experimental design. We presented our results in sections~\\ref{sec:neutrinos},~\\ref{sec:DarkEnergy}, ~\\ref{sec:DarkMatter}, and ~\\ref{sec:Inflation}. Here we quote the range of constraints each parameter falls in for experiments with $10^4-10^6$ detectors, $1'$ to $4'$ beam, and 25\\% to 75\\% $f_{sky}$ as illustrations and summarize the CMB only parameter improvement dependence on $N_{\\rm det}$, beam size, and $f_{sky}$: \\begin{itemize} \\item $ 0.0156 \\leq \\sigma(N_{\\rm eff}) \\leq 0.0690$ (CMB): \\\\ increasing $N_{\\rm det}$ from $10^4$ to $10^5$ improves $\\sigma(N_{\\rm eff})$ by about 30\\%, same for going from $10^5$ to $10^6$; decreasing beam size improves $\\sigma(N_{\\rm eff})$ by $\\sim$10\\% per arc-minute; $\\sigma(N_{\\rm eff})$ is not sample variance limited. \\item $ 15 \\leq \\sigma(M_{\\nu}) \\leq 24 $ [meV] (CMB+BAO): \\\\ increasing $N_{\\rm det}$ from $10^4$ to $10^5$ improves $ \\sigma(M_{\\nu})$ by 10-20\\% (smaller beam gives better improvement with $N_{\\rm det}$), while increasing $N_{\\rm det}$ from $10^5$ to $10^6$ improves $\\sigma(M_{\\nu})$ by 5-15\\%; decreasing beam size improves $\\sigma(M_{\\nu})$ at percent-levels per arc-minute; $\\sigma(M_{\\nu})$ is sample variance limited. \\item $164 \\leq$ DETF-FoM $\\leq 303$ (CMB+BAO+$H_0$):\\\\ increasing $N_{\\rm det}$ from $10^4$ to $10^5$ improves FoM by more than factor of 2, same for going from $10^5$ to $10^6$; decreasing beam size improves the FoM by a few \\% to 10s of \\% depending on configuration; FoM is not sample variance limited. \\item $0.00588 \\leq \\sigma(p_{ann}) \\leq 0.0110\\, [3\\times10^{-26}$ cm$^3$/s/GeV] (CMB):\\\\ increasing $N_{\\rm det}$ from $10^4$ to $10^5$ improves $\\sigma(p_{ann})$ by about 4\\%, same for going from $10^5$ to $10^6$; decreasing beam size improves $\\sigma(p_{ann})$ by $\\lesssim$ 1\\%; $\\sigma(p_{ann})$ is sample variance limited. \\item $0.00074 \\leq \\sigma(\\Omega_K) \\leq 0.0014 $ (CMB+BAO+$H_0$):\\\\ increasing $N_{\\rm det}$ from $10^4$ to $10^5$ improves $\\sigma(\\Omega_K)$ by about 20\\%, while increasing $N_{\\rm det}$ from $10^5$ to $10^6$ improves $\\sigma(\\Omega_K)$ by 10-20\\% (smaller beam gives a better improvement with $N_{\\rm det}$); decreasing beam size improves $\\sigma(\\Omega_K)$ at percent levels per arc-minute; $\\sigma(\\Omega_K)$ is sample variance limited. \\item $0.00110 \\leq \\sigma(n_s) \\leq 0.00236$ (CMB+BAO): \\\\ increasing $N_{\\rm det}$ from $10^4$ to $10^5$ improves $\\sigma(n_s)$ by $\\sim\\, 5-10\\%$ (smaller beam gives a better improvement with $N_{\\rm det}$), same for going from $10^5$ to $10^6$; decreasing beam size improves $\\sigma(n_s)$ at percent levels per arc-minute; $\\sigma(n_s)$ is sample variance limited. \\item $0.00145 \\leq \\sigma(\\alpha_s) \\leq 0.00330$ (CMB+BAO): \\\\ increasing $N_{\\rm det}$ from $10^4$ to $10^5$ improves $\\sigma(\\alpha_s)$ by 16-20\\%, and by 13-17\\% going from $10^5$ to $10^6$ (smaller beam gives a better improvement with $N_{\\rm det}$); decreasing beam size improves $\\sigma(\\alpha_s)$ at percent levels per arc-minute; $\\sigma(\\alpha_s)$ is sample variance limited. \\item $0.00009 \\leq \\sigma(r) \\leq 0.00203$ for 1\\% and 10\\% foreground residual: \\\\ $\\sigma(r)$ is foreground limited; when foreground is high, the optimal $f_{sky}$ shifts higher; decreasing beam size improves $\\sigma(r)$ at percent levels per arc-minute (the slope of this trend with beam increases with lower sky coverage). \\end{itemize} Detailed constraints in specific cases can be read off from tables listed in each section.\\\\ Besides learning the approximate ranges of how well these parameters can be constrained with CMB-S4, we also learn how the constraints improve as functions of $N_{\\rm det}$, beam size, and $f_{sky}$: \\begin{itemize} \\item For all parameters, except $r$, increasing $f_{sky}$ always improve the constraints even though the overall sensitivity of the experiment decreases. \\item For all parameters, except those related to dark-energy equation of state, going from $10^5$ to $10^6$ detectors yields the same or less percentage improvement on the constraints than going from $10^4$ to $10^5$ detectors. The improvements range from a few to tens of percent. For $M_{\\nu}$, the improvement beyond $10^5$ detectors is marginal when BAO signal is added. \\item Dependence on beam size is quite mild -- constraints on $N_{\\rm eff}$ improves by about 10\\% per arc-minute decrease while for all other parameters it is around a few \\% improvement per arc-minute decrease. \\end{itemize} We envision CMB-S4 to be a powerful next-generation ground-based CMB polarization experiment with high-resolution and high-sensitivity. With CMB-S4, we showed that most constraints on cosmological parameters are sample variance limited. Combining these data sets with space-borne observations will allow access to larger sky fraction, thus further improving the constraints on sample variance limited parameters." }, "1402/1402.1207_arXiv.txt": { "abstract": "A high performance Space-Time Reference in orbit could be realized using a stable atomic clock in a precisely defined orbit and linking that to high accuracy atomic clocks on the ground using a laser based time-transfer link. This would enhance performance of existing systems and provide unique capabilities in navigation, precise timing, earth sciences, geodesy and the same approach could provide a platform for testing fundamental physics in space. Precise laser time- and frequency-transfer from the ground to an orbiting satellite would make it possible to improve upon the current state of the art in timing (about 1 to 30\\,ns achieved with GPS) by roughly a factor of 1000 to the 1\\,ps level. ", "introduction": "There have been tremendous advances in the performance of atomic frequency standards (clocks) over the past 40 years, and, for compelling reasons, there are growing efforts to put more advanced atomic clocks into space. Prominent examples are the PHARAO cold-cesium atomic clock that is part of the European ACES mission\\cite{ESA_site} scheduled to fly on the International Space Station about 2016, the compact Hg$^+$ ion standard of JPL designed for space applications, and other promising systems under development for the future (the ESA Space Optical Clock,\\cite{Schiller_SOC_project} the DARPA Slow Beam Optical Clock\\cite{DARPA_AOSense}). Advanced laser systems have already vastly improved the performance of atomic clocks and optical frequency synthesis and division. Lasers can do the same for time transfer to space. ", "conclusions": "" }, "1402/1402.0751_arXiv.txt": { "abstract": "In order to build more realistic single stellar population (SSP) models with variable $\\alpha$-enhancement, we have recently determined [Mg/Fe] in a uniform scale with a precision of about 0.1 dex for 752 stars in the MILES empirical library. The [$\\alpha$/Fe] abundance ratio is commonly used as a good temporal scale indicator of star formation, taking Mg as a template for $\\alpha$ elements. Calcium is another element whose abundance is currently being investigated for the MILES stars. The MILES library is also being expanded by around 20\\% by including stars with known {\\it T}$_{\\rm eff}$, log~$g$, [Fe/H] and [Mg/Fe]. The transformation of their photospheric parameters to the MILES system has been carried out, but the calibration of their [Mg/Fe] is still in progress. In parallel, C, N and O abundances are also being compiled from literature for the library stars because they play an important role in the photospheric opacity, particularly influencing the blue spectral region. The Galactic kinematic classification of MILES stars with compiled [Mg/Fe] has been just computed such that this information can be considered in the SSP modelling. Comparisons of theoretical stellar predictions of the Lick line-strength indices against the MILES data have revealed the good behaviour of Fe-sensitive indices predictions, while highlighting areas for improvement in some models for the higher order H-Balmer features. ", "introduction": "\\label{s:intro} The main limitation of current empirical stellar spectral libraries is the use of stars whose individual elemental abundances are not adequately considered. Moreover, the stars in these libraries are basically selected from the solar neighbourhood, in which the chemical evolution of Milk Way is imprinted. That limitation is also propagated to the SSP isochrone-based models that have been built on these libraries. Typically, only iron is taken as a metallicity tracer, but the spectral energy distributions of stars and stellar systems considerably depend on abundance ratios of other metals relative to iron (e.g. CNO group and alpha elements). For instance, [$\\alpha$/Fe] is known as a good temporal scale indicator for the star formation since the alpha and iron-group elements have distinct nucleosynthetic origins with different time scales, respectively type II and type Ia supernovae. The higher this ratio is, the shorter the time scale for stellar formation. Magnesium is usually assumed as a template for $\\alpha$ elements. Therefore we have been taking steps to improve one of the most widely used empirical stellar spectral libraries - MILES \\citep{2006MNRAS.371..703S} \\& \\citep{2007MNRAS.374..664C}, and to use it to test theoretical models. Our ultimate aim is to utilise a combination of well characterised stars and theoretical model spectra to predict SSP spectra for a wide range of abundance patterns. Furthermore, building a new generation of semi-empirical SSP models with variable [$\\alpha$/Fe] will open new possibilities to minimise the effect of extrapolations of [$\\alpha$/Fe] in the models parametric space as well as to break the age-metallicity degeneracy over the model line-strength predictions, improving the recovery of star formation history in galaxies. ", "conclusions": "\\label{s:maths} We have determined [Mg/Fe] with a precision of $\\sim$0.1 dex for about 80\\% of MILES stars that are placed on a uniform scale (available on request or in Milone et al. 2011). A robust spectroscopic analysis was carried out using the MILES spectra and LTE spectral synthesis of two Mg features. This same semi-automated approach can be applied to recover [Ca/Fe] and perhaps [C/Fe]. MILES is currently being expanded by around 20\\% through the inclusion of stars with known {\\it T}$_{\\rm eff}$, log~$g$, [Fe/H] and [Mg/Fe] that fills in some gaps of the library 4-D parameter space and increase the star density in other regions. The transformation of their photospheric parameters to the MILES homogeneous system has been carried out, and the calibration of their [Mg/Fe] will be completed soon. The abundances of calcium, carbon, nitrogen and oxygen are currently being investigated for the MILES stars (compilation from literature for all of them and Ca atomic features spectroscopically analysed at mid-resolution). Specifically carbon and nitrogen most likely play an important role in the blue spectral region as we have concluded from our empirical analysis based on MILES spectral ratios between pairs of similar stars in {\\it T}$_{\\rm eff}$ and log~$g$. The Galactic kinematic classification of MILES stars with compiled [Mg/Fe] has been just computed such that this information can be included in the SSP modelling. Careful comparisons of the theoretical stellar predictions of the Lick indices against the MILES data have revealed the good behaviour of Fe-sensitive indices predictions, while highlighting areas for improvement in some models for the higher order H-Balmer features. We intend to compute a set of self-consistent semi-empirical SSP models with variable $\\alpha$-enhancement for a range in ages and metallicities around solar." }, "1402/1402.4458_arXiv.txt": { "abstract": "name}{Abstract} \\begin{abstract} In this work we focus on the evolution of the linear perturbations in the novel hybrid metric-Palatini theory achieved by adding a $f(\\mathcal{R})$ function to the gravitational action. Working in the Jordan frame, we derive the full set of linearized evolution equations for the perturbed potentials and present them in the Newtonian and synchronous gauges. We also derive the Poisson equation, and perform the evolution of the lensing potential, $\\Phi_{+}$, for a model with a background evolution indistinguishable from $\\Lambda$CDM. In order to do so, we introduce a designer approach that allows to retrieve a family of functions $f(\\mathcal{R})$ for which the effective equation of state is exactly $w_{\\textrm{eff}} = -1$. We conclude, for this particular model, that the main deviations from standard General Relativity and the Cosmological Constant model arise in the distant past, with an oscillatory signature in the ratio between the Newtonian potentials, $\\Phi$ and $\\Psi$. ", "introduction": "Introduction} Einstein's General Relativity (GR) is modern cosmology's main framework, providing a set of equations that dictate the dynamics of our Universe according to its material constituents. By them, our Universe could be expanding, static, or even collapsing. However, it is now well established that our Universe is currently undergoing an accelerated expansion which was preceded by phases of matter and radiation domination where gravitational attraction resulted in a decelerated expansion. And, at the beginning, it should have experienced a period of quasi-exponential inflation, so that any primordial spatial curvature would have been wiped out, leading to the spatially-flat and homogeneous Universe we observe. The simplest explanation for the Universe's accelerated expansion is a cosmological constant, $\\Lambda$, with a constant ratio of pressure to density (usually defined as the equation of state, $w$) equal to $-1$. Despite being in agreement with supernovae observations \\cite{accel1,accel2,accel3,accel4}, data from the cosmic microwave background (CMB) \\cite{cmb1,cmb2} including the recent {\\it Planck} data \\cite{planck1}, and large-scale structure (LSS) data \\cite{lss1}, cosmologists still struggle to account for the difference between the theoretically-expected value for its energy density and the observed one. If it exists, observationally, it should account for approximately $70 \\%$ of the Universe's total energy density, a value $121$ orders of magnitude smaller than that obtained from quantum field theory (for a review on $\\Lambda$, see Ref.~\\cite{lambdareview}). In light of these issues, new physics may be in order to account for that major component of our Universe, usually labeled dark energy (DE). Some theories, such as quintessence, k-essence, and so on, propose scalar fields rolling in a potential (see Ref.~\\cite{quintessence} and references therein for a comprehensive review). Other theories consider higher dimensions, as in braneworld models such as the DGP model \\cite{DGP,branes}, or assume that GR fails on cosmological scales and propose corrections to Einstein's action. The latter are grouped as the so-called Modified Gravity Theories (MGT), such as the Brans--Dicke scalar--tensor theory \\cite{bd}, Galileon models \\cite{gall}, the Fab Four \\cite{fabfour}, $f(R)$ theories \\cite{frreview}, and many others. For an extensive review on MGT, see Ref.~\\cite{reviewall}. In this paper, we focus on a novel model, the hybrid metric-Palatini gravity \\cite{main1,main2}. In this type of theories, the usual Einstein-Hilbert Lagrangian is supplemented with an $f(\\mathcal{R})$ Palatini term. This type of hybrid theory arises when perturbative quantization methods are considered on Palatini backgrounds \\cite{quantization}, which are connected with non-perturbative quantum geometries \\cite{quantumgeo}. Like the pure metric and Palatini cases, the hybrid theory has a dynamically equivalent scalar-tensor representation \\cite{main1,main2}. Those authors have also shown that the scalar field need not be massive in order to pass the stringent Solar System constraints \\cite{main1}, in contrast to the metric $f(R)$ theories, while possibly modifying the cosmological \\cite{hybridcosmo} and Galactic \\cite{hybridgala} dynamics due to its light, long-range interacting nature. In Ref. \\cite{hybridcosmo}, the criteria for obtaining cosmic acceleration was discussed. Alongside that, several cosmological solutions were derived, depending on the form of the effective scalar field potential, describing both accelerating and decelerating Universes. In this work, we focus on cosmological perturbations in the hybrid metric-Palatini formalism in the Jordan frame. Therefore, in Section \\ref{I} we briefly introduce the hybrid metric-Palatini model and, in Section \\ref{designer}, we present the designer approach which allows us to retrieve a family of solutions for $f(\\mathcal{R})$ whose effective equation of state is $w_{\\textrm{eff}} = -1$. In Section \\ref{II} we derive the full set of perturbed cosmological equations and present them in the Newtonian and Synchronous gauges. Then, in Section \\ref{III} we derive the Poisson equation and re-express the perturbed potentials in terms of the lensing potential, $\\Phi_{+}$, and the slip, $\\chi$, which we numerically evolve. We finish in Section \\ref{conclusion} with the conclusions of this work. ", "conclusions": "Conclusion} In this paper, we have derived the full set of perturbed Einstein equations for the novel hybrid metric-Palatini theory of gravity and presented them in the Newtonian and synchronous gauges. The latter, in particular, open the possibility of implementing this model in CAMB \\cite{camb} and give an in-depth analysis of the effects it can have at early times and even constrain its parameter space, which we leave for future work. We have introduced a designer approach to obtain a family of functions $f(\\mathcal{R})$ that reproduce a cosmology indistinguishable from $\\Lambda$CDM, with an effective equation of state exactly equal to $w_{\\textrm{eff}} = -1$. This particular approach leads to models where the modifications from standard General Relativity are more significant in the distant past. And, even though one can tweak the free parameters to control such modifications in order that they are quite negligible at a redshift of $z_{i} \\approx 1000$, this can prove problematic at even earlier times if the departure from GR gets increasingly larger, as it seems to do. Potentially, one could observe an inversion in the sign of $G_{\\textrm{eff}}$, leading to an inversion of the effect of gravity. This was, however, avoided in our analysis. We would like to point out that, other background solutions were neglected, for now, due to the inability to consistently set the initial conditions for $F$ when $w_{\\textrm{eff}} \\neq -1$. This is being explored for future work. We also derived the Poisson equation, which is substantially different from standard GR, given the inclusion of several extra dynamical elements. We then introduced the lensing potential, $\\Phi_{+}$, and the slip between the Newtonian potentials, $\\chi$, which we numerically evolved using the designer approach. We note that the departure from GR is more noticeable at the beginning of evolution. More specifically, $\\chi$ oscillates with a frequency proportional to the mode's wave number, and the oscillations' amplitude is the largest at this point. It is then gradually damped due to the evolution, tending to a GR value of $0$. Nevertheless, these oscillations never end up reflecting upon the lensing potential, $\\Phi_{+}$, which remains practically indistinguishable from $\\Lambda$CDM apart from a negligible enhancement at the start of evolution. However, they do reflect upon the Newtonian potentials, translating into a signature on the ratio between them, which oscillates rather quickly, signaling a clear departure of this model from standard GR. We also note that the evolution of $\\chi$ and $\\Phi_{+}$ in the hybrid metric-Palatini theory is related, like in metric $f(R)$ models, to the effective mass of the additional scalar degree of freedom, $m^{2}_{\\phi}$. Since the latter is smaller at early times, the range of the action of the additional fifth force will be larger. Hence, the enhancement in the perturbations, specially $\\chi$, will be greater then. And it will be greater the smaller the scale under consideration is, since these scales start their evolution deep within the range of the additional force." }, "1402/1402.6341_arXiv.txt": { "abstract": "The far-ultraviolet (FUV; 912~--~1700~\\AA) radiation field from accreting central stars in Classical T Tauri systems influences the disk chemistry during the period of giant planet formation. The FUV field may also play a critical role in determining the evolution of the inner disk ($r$~$<$~10 AU), from a gas- and dust-rich primordial disk to a transitional system where the optically thick warm dust distribution has been depleted. Previous efforts to measure the true stellar+accretion-generated FUV luminosity (both hot gas emission lines and continua) have been complicated by a combination of low-sensitivity and/or low-spectral resolution and did not include the contribution from the bright Ly$\\alpha$ emission line. In this work, we present a high-resolution spectroscopic study of the FUV radiation fields of 16 T Tauri stars whose dust disks display a range of evolutionary states. We include reconstructed Ly$\\alpha$ line profiles and remove atomic and molecular disk emission (from H$_{2}$ and CO fluorescence) to provide robust measurements of both the FUV continuum and hot gas lines (e.g., Ly$\\alpha$, \\ion{N}{5}, \\ion{C}{4}, \\ion{He}{2}) for an appreciable sample of T Tauri stars for the first time. We find that the flux of the typical Classical T Tauri Star FUV radiation field at 1 AU from the central star is $\\sim$~10$^{7}$ times the average interstellar radiation field. The Ly$\\alpha$ emission line contributes an average of 88\\% of the total FUV flux, with the FUV continuum accounting for an average of 8\\%. Both the FUV continuum and Ly$\\alpha$ flux are strongly correlated with \\ion{C}{4} flux, suggesting that accretion processes dominate the production of both of these components. On average, only $\\sim$~0.5\\% of the total FUV flux is emitted between the Lyman limit (912~\\AA) and the H$_{2}$ (0~--~0) absorption band at 1110~\\AA. The total and component-level high-resolution radiation fields are made publicly available in machine-readable format. ", "introduction": "Considerable observational effort has been invested to characterize the high-energy spectra of Classical T Tauri stars (CTTS), Class-II protostars with gas-rich circumstellar environments and active accretion. Near-ultraviolet (NUV; $\\lambda$~=~1700~--~3200~\\AA; 7~$>$~$h\\nu$~$>$~4~eV), far-ultraviolet (FUV; $\\lambda$~=~912~--~1700~\\AA; 13.6~$>$~$h\\nu$~$>$~7~eV), extreme-ultraviolet (EUV; $\\lambda$~=~120~--~912~\\AA; 100~$>$~$h\\nu$~$>$~13.6~eV), and X-ray ($\\lambda$~$<$ 100~\\AA; $h\\nu$~$>$~0.1~keV) irradiances are important ingredients for a complete picture of the chemistry and evolution of protoplanetary environments around young stars during the epoch of giant planet formation, migration, and the growth of rocky planet cores~\\citep{ward97,armitage02,ida04}. FUV molecular spectra provide unique insight into the physical conditions and composition of the warm inner-disk ($r$~$<$~10 AU) surface layer~\\citep{herczeg04,france11b,schindhelm12b,france12b} and molecular outflows~\\citep{walter03,herczeg06,krull07}. NUV spectra of the Balmer continuum excess provide perhaps the most direct measure of the protostellar mass accretion rate~(e.g.; Ingleby et al. 2011b, 2013 and references therein), while detailed studies of the fluxes and profiles of FUV spectral lines~\\citep{giovannelli95,lamzin96,ardila02,gunther08,ardila13} and X-ray spectra~\\citep{schneider08,brickhouse10} provide constraints on the magnetospheric accretion paradigm.\\nocite{ingleby11b,ingleby13} Further, the intrinsic energetic radiation field, generated from a combination of accretion processes and atmospheric activity on the protostar, provide an essential input to models of the chemistry and evolution of CTTS disks. % {\\it Disk Chemistry~--~} The strength and shape of the FUV radiation field has a strong influence on the chemical abundances of the disk, both at planet-forming radii ($r$~$<$ 10 AU; Walsh et al. 2012) and at larger radii ($r$~$>$~50 AU) where the majority of the disk mass resides (see e.g., Bergin et al. 2007 and references therein).\\nocite{bergin07} The stellar FUV continuum controls the dissociation of the most abundant disk molecules (H$_{2}$ and CO; Shull \\& Beckwith 1982; van Dishoeck \\& Black 1988)). The propagation of the FUV continuum is mainly regulated by dust grains~\\citep{zadelhoff03}; the processes of grain-growth and settling likely allow these photons to penetrate deeper into the disk as the protoplanetary environment evolves~\\citep{aikawa06,vasyunin11}. \\nocite{walsh12,shull82,vdb88} \\citet{bergin03} first emphasized the importance of accretion-generated \\ion{H}{1} Ly$\\alpha$ to the disk chemistry, and more recently it has been shown that the FUV spectral energy distribution of $all$ CTTSs is overwhelmingly dominated ($\\gtrsim$~80\\%) by Ly$\\alpha$ emission~\\citep{schindhelm12b}. Unlike the FUV continuum emission, the radiative transfer of Ly$\\alpha$ photons is controlled mainly by resonant scattering in the upper, atomic disk atmosphere~\\citep{bethell11}. Subsequent detailed disk modeling has demonstrated the importance of properly accounting for Ly$\\alpha$ radiation from the central star, finding significant ($\\gtrsim$ 1 order of magnitude) depletions in the abundances of C$_{2}$H$_{4}$, CH$_{4}$, HCN, NH$_{3}$, and SO$_{2}$ when Ly$\\alpha$ is included~\\citep{fogel11}. Interestingly, some species with large photo-absorption cross-sections at Ly$\\alpha$ ($\\lambda$~=~1216~\\AA), such as H$_{2}$O, do not show significant depletion because the enhanced dissociation rate is balanced by Ly$\\alpha$-driven photodesorption of water molecules from dust grains. It is clear now that Ly$\\alpha$ is a mandatory component of FUV radiation fields used for chemical modeling. However most large CTTS spectral atlases in the literature do not provide spectral coverage at 1216~\\AA\\ (e.g., Yang et al. 2012), or are dominated by geocoronal emission, such as in archival data from the {\\it International Ultraviolet Explorer}. When Ly$\\alpha$ spectral coverage is included, scattering in the interstellar and circumstellar environment prevents a direct measurement of the local Ly$\\alpha$ environment (as with measurements from the {\\it Hubble Space Telescope}-Space Telescope Imaging Spectrograph; $HST$-STIS).\\nocite{yang12} {\\it Disk Evolution~--~}Primordial gas disks are known to dissipate on timescales of~$\\leq$~10~Myr, at which point mass accretion onto the central star halts~\\citep{fedele10}. While there is a growing body of evidence that inner gas disks can survive longer than the typical 2~--~4 Myr lifetime of inner dust disks (e.g., Salyk et al. 2009; France et al. 2012b)\\nocite{salyk09,france12b}, the physical process by which the inner disk is cleared is not yet established. Various mechanisms including photoevaporation~\\citep{alexander06,gorti09} and dynamical clearing by exoplanetary systems~\\citep{rice03,dodson11}, possibly aided by a magnetorotational instability~\\citep{chiang07}, can reproduce certain transitional disk observations.~\\nocite{hernandez07,chiang07} Photoevaporation was initially considered for EUV photons from the central star~\\citep{clarke01,alexander06}, and more recent work has demonstrated that X-rays~\\citep{owen10} and FUV photons~\\citep{gorti09} can also play an important role in disk dispersal. Models that simultaneously treat FUV, EUV, and X-ray irradiation from the central star have shown that the FUV illumination can control the total evaporation rate (and hence the disk lifetime) by driving the heating at intermediate ($r$~$\\sim$~3~--~30 AU) and large radii ($r$~$\\geq$~100 AU; Gorti \\& Hollenbach 2009). FUV radiation also controls the gas temperature at the base of the evaporative flows through the generation of photoelectrons released by FUV-illuminated dust grains. Grain-growth and dust settling in the disk, part of the first stages of the planet formation process, enable deeper penetration of the FUV radiation. Therefore, planet formation itself can lead to an alteration of the temperature and chemical structure of the planet-forming region. \\citet{ingleby11} presented a comprehensive study of the evolution of FUV (excluding Ly$\\alpha$) and X-ray radiation fields over the 10 Myr lifetimes of gas disks. However, the low-resolution data used in most previous studies suffers from molecular disk contamination (line-blending with photo-excited H$_{2}$ and CO emissions) and no previous surveys of FUV radiation fields have included a proper treatment of Ly$\\alpha$ or isolation of the FUV continuum. Therefore, most data in the literature or readily available in the archive may not be representative of the strength $or$ the shape of the true disk-dispersing radiation fields. {\\it This Work~--~} In order to provide a more accurate and complete (including local Ly$\\alpha$ emission profiles) observational basis for models of disk chemistry and evolution, we present new measurements of the FUV line and continuum spectra generated by accretion and magnetic processes near the protostellar photosphere. These FUV radiation fields are available to the community in a machine-readable format\\footnote{ {\\tt http://cos.colorado.edu/$\\sim$kevinf/ctts\\_fuvfield.html} }. These spectra were obtained with the Cosmic Origins Spectrograph (COS) and the STIS aboard $HST$. While {\\it Far-Ultraviolet Spectroscopic Explorer} ($FUSE$) observations of a small number of bright CTTSs are available in the literature~\\citep{wilkinson02,herczeg04,herczeg05,herczeg06,gunther08}, $FUSE$ did not have the sensitivity to study ``typical'' Taurus-Auriga CTTSs in detail. We therefore use $FUSE$ observations to constrain our 912~--~1150~\\AA\\ radiation field creation (in particular the shapes and strengths hot gas emission lines and continua), but we present an approach where the short-wavelength radiation fields are inferred from longer-wavelength $HST$ data. \\begin{figure} \\figurenum{1} \\begin{center} \\epsfig{figure=f1.eps,width=3.4in,angle=00} \\vspace{+0.25in} \\caption{Complete FUV spectra of the 16 CTTSs studied in this work, including reconstructed Ly$\\alpha$ emission lines. These spectra are coadditions of $HST$-COS observations in the G130M and G160M modes (except for TW Hya, which was observed with STIS E140M; Herczeg et al. 2002) at several central wavelengths and focal-plane split positions. Almost all of the structures seen in these data are real atomic and molecular emission and absorption features. The data have been corrected for interstellar reddening (Table 1), scaled to the flux at 1 AU from the central star for comparison, and binned by three spectral resolution elements (21 pixels) for display. \\label{cosovly}} \\end{center} \\end{figure} \\begin{figure} \\figurenum{2} \\begin{center} \\epsfig{figure=f2.eps,width=3.4in,angle=00} \\vspace{+0.25in} \\caption{The binned FUV continuum spectra are shown as gray filled circles. A second order polynomial fit is extrapolated down to the Lyman Limit (912~\\AA) and is shown as the red dashed line. The ``1600~\\AA\\ Bump'' (spanning $\\sim$~1520~--~1660~\\AA) is prominent (detected at $>$ 3$\\sigma$ significance) in 10/16 targets. \\label{cosovly}} \\end{center} \\end{figure} In Section 2, we describe the targets and the $HST$ observations. In Section 3, we briefly describe the spectral deconvolution performed to separate the continuum, hot gas atomic line, and molecular line emission. The details of this deconvolution can be found in the Appendix. We detect the quasi-continuous emission feature near 1600~\\AA, the ``1600~\\AA\\ Bump'' in $\\sim$~70\\% of the sources; the analysis of this emission~\\citep{bergin04,ingleby09,france11a} will be the subject of a future work. In Section 4, we discuss the stellar+accretion spectra and present correlations suggesting that the majority of the Ly$\\alpha$ and FUV continuum are generated by accretion processes. % We present a summary of this work in Section 5. ", "conclusions": "There are numerous indicators of the gas and dust content of a young protoplanetary system. Three important observables are the warm dust content of the inner disk, the presence of circumstellar gas, and signs of active accretion. Our $HST$ observations provide measurements of the last two, while the first has been extensively studied in the IR. % The large transition probabilities of the H$_{2}$ electronic band systems and the lack of photospheric emission at $\\lambda$~$<$~1700~\\AA\\ in low-mass stars make fluorescent H$_{2}$ one of the most sensitive indicators for the presence of molecular gas in the inner ~$\\sim$~10~AU of young circumstellar disks~\\citep{france12b}. H$_{2}$ emission line spectroscopy can directly probe gas surface densities as small as $\\Sigma_{H2}$~$\\lesssim$~10$^{-6}$ g cm$^{-2}$. The FUV bandpass also contains resonance lines of hydrogen, carbon, nitrogen, and oxygen which are powerful diagnostics of warm/hot gas ($T_{form}$~$\\sim$~10$^{4}$~--~3~$\\times$~10$^{5}$ K; \\ion{H}{1} Ly$\\alpha$, \\ion{C}{3} $\\lambda$977, \\ion{C}{4} $\\lambda$$\\lambda$1548,1550, \\ion{N}{5} $\\lambda$$\\lambda$1239,1243, and \\ion{O}{6} $\\lambda$$\\lambda$1032,1038; as well as the H$\\alpha$ line of ionized helium, \\ion{He}{2} $\\lambda$1640) formed near the accretion shock region\\footnote{including the heated photosphere at the base of the accretion column, the pre-shock region, shock surface, and post-shock region} by collisional- and photo- excitation and ionization processes. Our observations enable the first measurements of the FUV continuum in many of these targets, a potentially useful diagnostic of the accretion environment that has only recently been made accessible by the high sensitivity and spectral resolution of $HST$-COS (e.g., France et al. 2011a).\\nocite{france11b} \\subsection{Ly$\\alpha$, Hot Gas Lines, and the FUV Continuum: Constraints on Physical Origins % } By bringing together the reconstructed Ly$\\alpha$ emission profiles, the ``H$_{2}$ cleaned'' hot gas emission lines, and measurements of the uncontaminated FUV continuum, we can explore the relations between these different components and empirically constrain the origin of the strong Ly$\\alpha$ and FUV continuum emission in these sources. The prevailing picture for the formation of the UV continuum excess is that it arises from ionized gas ($T_{cont}$~$\\sim$~1.5~--~3~$\\times$~10$^{4}$ K) in an optically thin pre-shock region at the base of the accretion column~\\citep{calvet98,costa00}. In this picture, the FUV continuum may be an extension of the Balmer continuum used for accretion rate determination in the NUV (see e.g., Ingleby et al. 2013); and the connection between the NUV and FUV continua has been explored previously for DF Tau~\\citep{france11a}. In the following subsection, we show that for a subsample of six of our targets with nearly simultaneous FUV and NUV observations, the FUV and NUV continua are characterized by a significantly different slope. This demonstrates that the connection proposed by~\\citet{france11a} was most likely an artifact of comparing non-simultaneous observations, although this cannot be conclusively demonstrated for all of our targets in our sample without contemporaneous FUV and NUV data. We find that while some of our sources display a monotonically decreasing FUV continuum towards shorter wavelengths (BP Tau, DF Tau, RU Lup, see also Herczeg et al. 2005), many of the FUV continua are flat across the FUV band (GM Aur, RECX-15, UX Tau, V4046 Sgr). This suggests that multiple physical processes likely contribute in the complex emitting region near the stellar photosphere and the inner accretion disk. Our aim here is to provide the data for use in detailed models of disk chemistry and evolution, and therefore we do not attempt to present a comprehensive model for the continuum emission region. The hot gas emissions from \\ion{C}{4} and \\ion{N}{5} have been long considered to be related to accretion processes~\\citep{krull00}. A detailed spectral line analysis of a larger sample of COS and STIS observations has shown that these lines are likely produced by hot gas in accretion spots near the stellar photosphere, the edges of accretion columns, or multiple accretion columns of varying densities~\\citep{ardila13}. The \\ion{He}{2} lines however appear to be dominated by a combination of emission from the magnetically active stellar atmosphere and the pre-shock region. Ly$\\alpha$ is presumably generated at several places in the near-star environment, however we are not aware of any previous work that explores the observational connection between Ly$\\alpha$ luminosity and accretion processes (see Muzerolle et al. 2001 and Kurosawa et al. 2006 for a detailed description of Balmer lines).\\nocite{muzerolle01,kurosawa06} \\begin{figure*} \\figurenum{6} \\begin{center} \\epsfig{figure=f6.eps,width=5.0in,angle=90} \\vspace{+0.25in} \\caption{ A comparison between the reddening corrected FUV (gray filled circles) and NUV (black line) continua for the six stars in our sample with contemporaneous observations. The solid orange line and dashed pink lines are linear fits to the FUV and NUV continua, respectively. The linear fit to the FUV data is created using fluxes from 1150~--~1174~\\AA\\ and 1680~--~1750~\\AA, excluding contributions from the molecular continuum and the ``1600~\\AA\\ Bump''. The linear fits show that the FUV continuum is brighter than suggested by an extrapolation of the NUV continuum for most targets, and that the FUV slopes are shallower, possibly suggesting multiple emitting regions. \\label{cosovly}} \\end{center} \\end{figure*} In order to empirically constrain the relationship between Ly$\\alpha$ and the FUV continuum and accretion processes in the following subsections, we assume that the integrated \\ion{C}{4} flux is a proxy for the mass accretion rate~\\citep{krull00,ardila13}. It is critical to have this quasi-simultaneous accretion diagnostic because accretion rates and UV fluxes from CTTSs are known to vary by factors of $\\sim$~2 on timescales of months-to-years~\\citep{valenti00}. Therefore, accretion rates in the literature will not necessarily be well correlated with the UV emission at the time of our observations. In Figure 5 {\\it top left}, we demonstrate the tight correlation between the accretion-dominated \\ion{C}{4} and \\ion{N}{5} flux for our stars (see also the larger sample from Ardila et al. 2013). For all of the correlations shown in Figure 5, we present the Spearman rank correlation coefficient ($\\rho$) and the likelihood that this population is drawn from a random sample (i.e., not correlated; $n$). In general, $\\vert$$\\rho$$\\vert$~$>$~0.5 and $n$~$<$~0.05 indicates that a real correlation exists between the two quantities. For instance, [$\\rho$,$n$] = [0.81, 1.3~$\\times$~10$^{-4}$] for the plot of \\ion{N}{5} and \\ion{C}{4}, indicating a strong correlation, as expected. \\subsubsection{FUV Continuum} The FUV continuum flux (in units of erg cm$^{-2}$ s$^{-1}$), evaluated at 1 AU from the central star, is computed by integrating the polynomial fit to the binned intra-emission-line $HST$ spectra (\\S A.3) from 912~--~1650~\\AA\\ (Figures 2 and A.4~--~A.6). The FUV continuum is shown to be tightly correlated with the \\ion{C}{4} emission (Figure 5, {\\it top right}, [$\\rho$,$n$] = [0.93, 1.5~$\\times$~10$^{-7}$]). We interpret this as strong evidence that the CTTS FUV continuum is generated by accretion processes, either directly or powered by accretion luminosity. However, it may originate in a spatially separate region from the NUV Balmer continuum, as mentioned above. In Figure 6, we compare data for six of our targets (DM Tau, DR Tau, GM Aur, HN Tau, RECX-11, and RECX-15) with FUV and NUV observations separated by less than a few hours. Linear fits to the FUV (excluding the molecular quasi-continuum) and NUV regimes are shown overplotted as solid orange and dashed pink lines, respectively. The observed FUV continuum is shown to be brighter than predicted from a simple extrapolation of the NUV continuum, and has a significantly shallower slope (see also Herczeg et al. 2004, 2005).\\nocite{herczeg04,herczeg05} We note that the S/N ratios of individual binned data points are relatively high, demonstrating that the excess FUV continuum flux is statistically significant in all cases. The average S/N per binned continuum point ranges from 2.5~--~14.8 from 1140~--~1340~\\AA\\ and 0.9~--~17.0 from 1660~--~1740~\\AA\\ (Table A.1). While it seems clear that the FUV continuum is related to accretion processes, these differences argue that the FUV and NUV continua may be formed in spatially distinct regions near the interaction of the accretion streams and the stellar photosphere. Recent models of the NUV Balmer excess have incorporated multiple accretion components with a range of energy fluxes and densities (e.g., Ingleby et al. 2013), and future work may be able to connect the FUV and NUV continua by incorporating these different components. Our analysis finds that the FUV continuum extends to $\\lambda$~$<$~1216~\\AA\\ in all cases, ruling out the possibility that this flux originates from \\ion{H}{1} 2-photon (2$s$~$^{2}S_{1/2}$~--~1$s$~$^{2}S_{1/2}$) emission. Furthermore, \\ion{H}{1} 2-photon emission is suppressed in high-density regions ($n_{H}$~$>$~1.5~$\\times$~10$^{4}$ cm$^{-3}$), consistent with the scenario in which this emission is the high-frequency extension of the Balmer continuum generated in the relatively dense pre-shock region at the base of the accretion column~\\citep{calvet98}. A future space-borne instrument capable of simultaneous FUV and NUV spectroscopy or photometry could provide more insight on the potential link between the FUV continuum and the NUV Balmer continuum in CTTSs. \\begin{deluxetable*}{lccccccc} \\tabletypesize{\\scriptsize} \\tablecaption{Relative Contributions to the Stellar + Accretion FUV Radiation Field\\label{lya_lines}} \\tablewidth{0pt} \\tablehead{ \\colhead{Target} & \\colhead{$F_{tot}$\\tablenotemark{a}} & \\colhead{log$_{10}$($F_{tot}$/G$_{o}$)\\tablenotemark{b} } & \\colhead{FUV Continuum} & \\colhead{Ly$\\alpha$ \\tablenotemark{c}} & \\colhead{\\ion{C}{4}} & \\colhead{Other lines\\tablenotemark{d}} & \\colhead{$\\lambda$~$\\leq$~1110~\\AA} \\\\ & (erg cm$^{-2}$ s$^{-1}$) & & (\\%) & (\\%) & (\\%) & (\\% ) & (\\% ) } \\startdata AATAU & 3.1~$\\times$~10$^{4}$ & 7.3 & 2.8 & 95.9 & 0.8 & 0.6 & 0.1 \\\\ BPTAU & 3.5~$\\times$~10$^{4}$ & 7.3 & 16.8 & 75.7 & 4.5 & 3.1 & 1.8 \\\\ DETAU & 1.5~$\\times$~10$^{4}$ & 7.0 & 9.2 & 88.0 & 2.0 & 0.9 & 0.3 \\\\ DFTAU & 2.3~$\\times$~10$^{5}$ & 8.2 & 2.1 & 97.2 & 0.4 & 0.2 & 0.1\\\\ DMTAU & 4.3~$\\times$~10$^{3}$ & 6.4 & 8.4 & 88.0 & 1.7 & 1.9 & 0.3 \\\\ DRTAU & 4.0~$\\times$~10$^{4}$ & 7.4 & 46.2 & 49.4 & 3.3 & 1.0 & 0.7 \\\\ GMAUR & 1.2~$\\times$~10$^{4}$ & 6.9 & 9.9 & 86.2 & 2.4 & 1.5 & 0.4 \\\\ HNTAU & 1.2~$\\times$~10$^{4}$ & 6.9 & 10.1 & 88.4 & 0.9 & 0.5 & 0.1 \\\\ LKCA15 & 1.6~$\\times$~10$^{4}$ & 7.0 & 6.5 & 90.4 & 2.1 & 1.0 & 0.3 \\\\ RECX11 & 2.3~$\\times$~10$^{3}$ & 6.2 & 4.6 & 90.9 & 2.9 & 1.7 & 0.6 \\\\ RECX15 & 4.8~$\\times$~10$^{3}$ & 6.5 & 2.8 & 96.7 & 0.3 & 0.3 & 0.1 \\\\ RULUPI & 2.8~$\\times$~10$^{4}$ & 7.2 & 17.6 & 79.8 & 2.0 & 0.7 & 0.3 \\\\ SUAUR & 3.3~$\\times$~10$^{4}$ & 7.3 & 8.4 & 88.3 & 2.1 & 1.2 & 0.4 \\\\ TWHYA & 9.7~$\\times$~10$^{3}$ & 6.8 & 16.7 & 71.9 & 6.5 & 4.9 & 0.9 \\\\ UXTAU & 5.6~$\\times$~10$^{3}$ & 6.5 & 4.8 & 92.2 & 1.7 & 1.3 & 0.3\\\\ V4046SGR & 1.5~$\\times$~10$^{4}$ & 7.0 & 5.3 & 91.8 & 1.2 & 1.8 & 0.9 \\\\ \\tableline Average\\tablenotemark{e} & & 7.0~$\\pm$~0.5 & 8.4~$\\pm$~5.2 & 88.1~$\\pm$~7.3 & 2.1$~\\pm$~1.6 & 1.4~$\\pm$~1.2 & 0.5~$\\pm$~0.4 \\enddata \\tablenotetext{a}{Integrated 912~--~1650~\\AA\\ stellar+accretion FUV radiation field (not including molecular fluorescence lines, low-ionization atomic emission, or the ``1600~\\AA\\ Bump''), evaluated at 1 AU from the central pre-main sequence star. } \\tablenotetext{b}{Ratio of the integrated FUV radiation field at 1 AU to the average interstellar radiation field (1.6~$\\times$~10$^{-3}$ erg cm$^{-2}$ s$^{-1}$; Habing 1968). } \\tablenotetext{c}{Intrinsic Ly$\\alpha$ emission, integrated over 1211~--~1221~\\AA .} \\tablenotetext{d}{Other stellar+accretion hot gas emission lines, \\ion{C}{3} $\\lambda$977~\\AA\\ + \\ion{O}{6} $\\lambda$$\\lambda$1032,1038~\\AA\\ + \\ion{N}{5} $\\lambda$$\\lambda$1239, 1243~\\AA\\ + \\ion{He}{2} $\\lambda$1640~\\AA. } \\tablenotetext{e}{Average quantities are calculated excluding DR Tau, whose Ly$\\alpha$ profile reconstruction is compromised by multiple Ly$\\alpha$ emission sources (\\S A.1).} \\end{deluxetable*} \\subsubsection{Ly$\\alpha$ Line Emission} The intrinsic Ly$\\alpha$ flux (in units of erg cm$^{-2}$ s$^{-1}$), evaluated at 1 AU from the central star, is computed by integrating the reconstructed Ly$\\alpha$ line profile from 1211~--~1221~\\AA. We note that we take the intrinsic reconstructed line profile as opposed to the line profile seen by the H$_{2}$ molecules at the disk surface after absorption by atomic hydrogen in the outflow (see Figure A.1). A discussion of the Ly$\\alpha$ profiles including the neutral outflow absorption component is presented by~\\citet{schindhelm12b}, and the uncertainty on the intrinsic Ly$\\alpha$ flux used here is $\\sim$~10~--~30\\%, depending on the detectability and S/N of the H$_{2}$ fluorescence lines used for the reconstruction~\\citep{france12b}. The middle panels of Figure 5 show the correlation between Ly$\\alpha$ and the FUV continuum ([$\\rho$,$n$] = [0.71, 1.9~$\\times$~10$^{-3}$]) and the \\ion{C}{4} emission ([$\\rho$,$n$] = [0.77, 4.8~$\\times$~10$^{-4}$]). The Ly$\\alpha$ flux is observed to correlate with both quantities, but with less significance than the correlations between \\ion{C}{4} and \\ion{N}{5}, and \\ion{C}{4} and the FUV continuum. This is possibly confirmed by another representation on the bottom panels of Figure 5, where the ratio of Ly$\\alpha$/FUV continuum is plotted against the published mass accretion rates ($left$, [$\\rho$,$n$] = [$-$0.31, 0.24]) and the \\ion{C}{4} fluxes ($right$, [$\\rho$,$n$] = [$-$0.50, 0.047]). The plots are qualitatively similar; Figure 5 ({\\it bottom left}) does not show a significant anti-correlation between Ly$\\alpha$/FUV continuum and the mass-accretion rate, Ly$\\alpha$/FUV continuum and \\ion{C}{4} show a weak anti-correlation suggesting that the FUV continuum is more tightly correlated with the accretion than the Ly$\\alpha$ emission. \\begin{figure} \\figurenum{7} \\begin{center} \\epsfig{figure=f7.eps,width=3.5in,angle=00} \\vspace{+0.0in} \\caption{ A comparison between observed CTTS and WTTS Ly$\\alpha$ profiles for a few example objects. The CTTS Ly$\\alpha$ profiles (GM Aur and BP Tau) uniformly display higher fluxes and line widths ($\\Delta$$v$~$>$~$\\pm$~500 km s$^{-1}$) than their WTTS counterparts (TWA 7 and LkCa19), suggesting that accretion processes contribute (and probably dominate) to the total Ly$\\alpha$ luminosity. The narrow central emission is from Earth's geocorona and the broad depression seen at the core of the CTTS line profiles is due to resonant scattering in the interstellar medium along the line-of-sight. \\label{cosovly}} \\end{center} \\end{figure} The production of Ly$\\alpha$ is still likely dominated by accretion processes. Figure 7 shows the observed Ly$\\alpha$ emission lines from two representative CTTSs from this work (GM Aur and BP Tau) and two WTTSs (TWA 7 and LkCa19). In all cases, the CTTS Ly$\\alpha$ lines are much stronger and broader than WTTS Ly$\\alpha$ profiles for a similar age and stellar mass. All CTTSs display broad Ly$\\alpha$ wings to $v_{Ly\\alpha}$~$>$~500 km s$^{-1}$, and many show Ly$\\alpha$ emission wings in excess of $\\pm$~1000 km s$^{-1}$. There may be several physical reasons that the correlation between Ly$\\alpha$ and \\ion{C}{4} is weaker than between the FUV continuum and \\ion{C}{4}, including uncertainties in the Ly$\\alpha$ reconstruction process, particularly when there are multiple sources of Ly$\\alpha$ photons in the systems (stellar atmosphere, outflows and/or disk surface) and resonant scattering of Ly$\\alpha$ photons in the gas-rich CTTS circumstellar environment. \\subsection{Absolute Fluxes at Planet-forming Radii and Relative Contributions} In order to understand the UV-driven photochemistry of the protoplanetary environment, one needs constraints on both the absolute flux and the shape of the illuminating radiation field. The absolute flux of the radiation environment determines the photoexcitation and photodissociation rates for disk molecules and the SED determines the primary scattering and absorption mechanisms for diffusing FUV photons through the disk~\\citep{bergin03,bethell11,fogel11}. In Table 2, we give the total, reddening-corrected flux from our targets (evaluated at 1 AU to enable inter-comparison) integrated over the 912~--~1650~\\AA\\ bandpass. In the second column, this quantity is given in units of the average interstellar radiation field from~\\citet{habing68}. The average incident flux from the accreting central star is~$\\sim$~10$^{7}$ times the contribution from scattered OB starlight at 1 AU from the central star, although the shapes of the CTTS radiation field and the interstellar field are very different. By spectrally resolving the various components of the FUV field and including the reconstructed Ly$\\alpha$ emission, we can provide an inventory of the FUV luminosity sources in $\\sim$~1~--~10 Myr CTTSs. Table 2 shows the relative contribution of each component to the total FUV irradiance from each target. As argued by previous authors~\\citep{bergin03,herczeg04,schindhelm12b}, Ly$\\alpha$ dominates the FUV radiation output from these sources, with an average fractional luminosity of 88.1~$\\pm$~7.3~\\%\\footnote{The only object for which Ly$\\alpha$ may not be the largest contributor to the FUV radiation field is DR Tau. However, the Ly$\\alpha$ reconstruction for this star is rather uncertain due to the presence of multiple sources of Ly$\\alpha$ emission in the system, most likely a second, narrow component associated with an outflow. }. The FUV continuum is the second largest energy source, with 8.4~$\\pm$~5.2~\\%. The contribution from the \\ion{C}{4} $\\lambda$$\\lambda$1548,1550~\\AA\\ doublet is 2.1~$\\pm$~1.6~\\%, and the remaining hot gas emission lines make up $<$~5\\% in all cases, although the exact contribution is rather uncertain because we assumed the TW Hya \\ion{C}{3} and \\ion{O}{6} fluxes, scaled by the relative \\ion{C}{4} fluxes, for the remaining targets. Lower ionization species likely contribute $<$ 1\\% to the total field strength (see A.2). The primary transitions leading to dissociation of H$_{2}$ and CO, through the H$_{2}$ $B$~--~$X$ and $C$~--~$X$ band systems (radiatively dissociative) and the CO $E$~--~$X$ band system (predissociative), are located at $\\lambda$~$\\leq$~1110~\\AA. The $\\lambda$~$\\leq$~1110~\\AA\\ radiation field also plays an important role in the production of photo-electrons that heat the gas in the disk surface layers~\\citep{pedersen11}. The 912~--~1110~\\AA\\ bandpass contains between 0.1~\\% and 1.8~\\% of the total FUV flux, with an average (excluding DR Tau) of 0.5~$\\pm$~0.4~\\%. On average, the 912~--~1110~\\AA\\ CTTS radiation field has a flux of $\\sim$~150 erg cm$^{-2}$ s$^{-1}$ at 1 AU from the central star. \\subsection{Comparison with Low-resolution Spectra} In this subsection, we quantify the impact of low-resolution data on the determination of the total FUV flux from CTTSs. Previous large FUV spectroscopic samples of CTTSs were acquired at low spectral resolution ($R$~$\\sim$~80~--~1000; Johns-Krull et al. 2000; Valenti et al. 2000; Ingleby et al. 2009, 2011; Yang et al. 2012). Most of these observations did not have access to Ly$\\alpha$ directly due to instrumental bandpass limitations and/or contamination by geocoronal emission. Furthermore, low-resolution prevents the necessary measurements of individual H$_{2}$ or CO fluorescent emission lines necessary for reconstructing the Ly$\\alpha$ emission profile. Consequently, most quoted measurements of the FUV radiation field strength are integrated over $\\approx$ 1230~--~1700~\\AA~\\citep{ingleby11,yang12}, $excluding$ the dominant FUV flux source, \\ion{H}{1} Ly$\\alpha$. Low resolution data are also insufficient for separating the accretion-powered FUV continuum from the wealth of narrow molecular emission lines and the molecular continuum (the ``1600~\\AA\\ Bump''). In order to quantify the differences between high- and low-resolution CTTS FUV radiation fields, data for 10 objects with previous low-resolution FUV observations from $HST$ were retrieved from the~\\citet{yang12} atlas of low-resolution T Tauri star spectra hosted at MAST. The average flux ratio between the observed low-resolution 1235~--~1700~\\AA\\ flux from the~\\citet{yang12} atlas ($F(1235-1700)_{LR}$) and the 1235~--~1700~\\AA\\ fluxes from this work ($F(1235-1700)_{fit}$), $F(1235-1700)_{LR}$/$F(1235-1700)_{fit}$, is 1.41~$\\pm$~0.52 (Table 3). The average being greater than unity most likely reflects the inclusion of Ly$\\alpha$-pumped H$_{2}$ and CO fluorescence in low-resolution data and the scatter is most likely the time-variability that is characteristic of these sources (recall these data were acquired over a multi-year baseline; and see e.g., Giovannelli et al. 1995; G{\\'o}mez de Castro \\& Fern{\\'a}ndez 1996). The two sources which display the largest FUV variability are DM Tau and SU Aur, showing a factor of $\\sim$~3 variability between the two epochs. If we screen sources farther than 3--$\\sigma$ from the mean of the distribution, the average drops to $F(1235-1700)_{LR}$/$F(1235-1700)_{fit}$ = 1.06~$\\pm$~0.34. \\begin{deluxetable}{lccc} \\tabletypesize{\\normalsize} \\tablecaption{Comparison to Low-Resolution CTTS Spectra \\label{lya_lines}} \\tablewidth{0pt} \\tablehead{ \\colhead{Target} & \\colhead{$HST$ Mode\\tablenotemark{a}} & \\colhead{ $\\frac{ F(1235-1700)_{LR} }{ F(1235-1700)_{fit} }$ \\tablenotemark{b} } & \\colhead{ $\\frac{ F(1235-1700)_{LR} }{ F(1150-1700)_{fit} }$ \\tablenotemark{c} } } \\startdata AATAU & ACS & 1.18 & 0.07 \\\\ BPTAU & STIS & 0.70 & 0.21 \\\\ DETAU & ACS & 1.68 & 0.27 \\\\ DMTAU & STIS & 2.73 & 0.31 \\\\ DRTAU & ACS & 0.77 & 0.55 \\\\ GMAUR & ACS & 0.99 & 0.14 \\\\ GMAUR & STIS & 0.96 & 0.14 \\\\ HNTAU & ACS & 1.04 & 0.17 \\\\ LKCA15 & STIS & 1.43 & 0.19 \\\\ SUAUR & STIS & 3.24 & 0.60 \\\\ V4046SGR & ACS & 0.76 & 0.05 \\\\ \\tableline Average & & 1.41~$\\pm$~0.52 & 0.25~$\\pm$~0.17 \\enddata \\tablenotetext{a}{$HST$ Low-resolution mode used for comparison. Low-resolution data taken from~\\citet{yang12}. ACS refers to ACS/SBC PR130L mode, STIS refers to STIS G140L. } \\tablenotetext{b}{Ratio of the observed low-resolution spectrum from 1235~--~1700~\\AA\\ to the extracted, intrinsic CTTS spectrum from 1235~--~1700~\\AA. } \\tablenotetext{c}{Ratio of the observed low-resolution spectrum from 1235~--~1700~\\AA\\ to the extracted, intrinsic CTTS spectrum from 1150~--~1700~\\AA, including the reconstructed Ly$\\alpha$ emission line. % } \\end{deluxetable} Comparing the ratio of the observed low-resolution 1235~--~1700~\\AA\\ flux from the~\\citet{yang12} atlas and our 1150~--~1700~\\AA\\ fluxes that {\\it include the reconstructed Ly$\\alpha$ profiles} ($F(1150-1700)_{fit}$), we find $F(1235-1700)_{LR}$/$F(1150-1700)_{fit}$~=~0.25~$\\pm$~0.17. Applying the 3--$\\sigma$ screen, $F(1235-1700)_{LR}$/$F(1150-1700)_{fit}$~=~0.17~$\\pm$~0.08. The general result is that even with the additional contribution from H$_{2}$ emission, and the uncertainty associated with the time-variability of the sources, low-resolution FUV spectra that do not include a reconstructed Ly$\\alpha$ emission line underestimate the total FUV flux by a factor of $\\approx$~6. While a factor of $\\approx$~6 is the average flux underestimate, Table 3 shows that in the extreme cases (e.g., AA Tau and V4046 Sgr), the low-resolution 1235~--~1700~\\AA\\ data underestimate the true intrinsic stellar+accretion FUV radiation field strength by factors of ~$\\geq$~15. Our conclusion from this comparison with low-resolution data in the literature is that ACS/SBC PR130L and STIS G140L-based flux estimates underestimate the true FUV radiation field strength from CTTSs by approximately an order-of-magnitude. Our findings quantitatively support the assertions of previous authors~\\citep{bergin03,herczeg04,schindhelm12b} that Ly$\\alpha$ must be included to make an accurate of estimate of $both$ the strength and the shape of the FUV radiation field in protoplanetary environments around low-mass stars. It should also be emphasized that the largest uncertainty on the inferred local FUV radiation field strength is likely the uncertainty on the line-of-sight reddening, both interstellar and circumstellar. Typical dispersions on the derived values of $A_{V}$ in the literature are 0.3~--~1.0 magnitudes, the relative contributions of interstellar and circumstellar grains are unclear~(see e.g., McJunkin et al. 2014 and references therein), and the scattering and extinction properties of grains in regions like Taurus may be quite different than those found in the diffuse interstellar medium~\\citep{calvet04}.\\nocite{mcjunkin14} With empirically-derived UV radiation fields now available, the uncertainty on the extinction is likely the limiting factor on our knowledge of the absolute flux of the energetic radiation environments around young stars." }, "1402/1402.6031_arXiv.txt": { "abstract": "We perform $N$-body simulations on a multiple massive black hole (MBH) system in a host galaxy to derive the criteria for successive MBH merger. The calculations incorporate the dynamical friction by stars and general relativistic effects as pericentre shift and gravitational wave recoil. The orbits of MBHs are pursed down to ten Schwarzschild radii ($\\sim 1$AU). As a result, it is shown that about a half of MBHs merge during 1~Gyr in a galaxy with mass $10^{11}M_{\\odot}$ and stellar velocity dispersion $240$~km~s$^{-1}$, even if the recoil velocity is two times as high as the stellar velocity dispersion. The dynamical friction allows a binary MBH to interact frequently with other MBHs, and then the decay of the binary orbits leads to the merger through gravitational wave radiation, as shown by \\citet{Tanikawa11}. We derive the MBH merger criteria for the masses, sizes, and luminosities of host galaxies. It is found that the successive MBH mergers are expected in bright galaxies, depending on redshifts. Furthermore, we find that the central stellar density is reduced by the sling-shot mechanism and that high-velocity stars with $\\sim 1000$~km~s$^{-1}$ are generated intermittently in extremely radial orbits. ", "introduction": "Massive black holes (hereafter MBHs) with the mass of more than $10^6$ solar mass ($M_{\\odot}$) have been found in the centres of galaxies. The mass of MBHs is correlated with the properties of the spheroidal components of their host galaxies, with respect to the mass \\citep{Kormendy95,Magorrian98,Marconi03}, the velocity dispersions \\citep{Ferrarese00,Tremaine02,Gultekin09}, and the number of globular clusters \\citep{Burkert10,Harris11}. The origin of MBHs is an open issue of great significance. In the last decade, quasars (QSOs) that possess $\\sim 10^9~M_{\\odot}$ MBHs have been found at high redshifts of $z \\gtrsim 6$ \\citep[e.g.,][]{Fan01}, that is, at the cosmic age of $\\lesssim 1$~Gyr. Conservatively speaking, the seeds of the MBHs could be stellar mass black holes as massive star remnants. In particular, the remnants of first stars are one of plausible candidates, since first stars are likely to be as massive as a few hundred solar mass \\citep{Abel00,Nakamura01,Bromm02,Yoshida06}, several tens solar mass \\citep{Clark11}, or about $50M_{\\odot}$ \\citep{Hosokawa11}. However, in order for first star remnants to grow up to $\\sim 10^9~M_{\\odot}$ in $1$~Gyr, the Eddington ratio of mass accretion rate should be larger than unity \\citep[e.g.,][and references therein]{Umemura01,Greene12}. Super-Eddington accretion is one of possible solutions for the MBH growth \\citep[e.g.][]{Abramowicz88,Kawaguchi03,Ohsuga05}. On the other hand, the integration of the QSO luminosity function is concordant with the integrated mass function of MBHs in the local universe, as long as the Eddington ratios are between 0.1 and 1.7 \\citep{Soltan82,Yu02,Marconi04}. This implies that supermassive black holes acquire the bulk of mass through gas accretion in the late evolutionary stages and the mass accretion rates are not highly super-Eddington. Also, the gas accretion onto the seeds should be intermittent, and on average could be lower than the Eddington accretion rate \\citep{Milosavljevic09a,Milosavljevic09b}. If the merger of multiple black holes precedes the growth via gas accretion, the merged MBH can be a seed of a supermassive black hole, and therefore the constraint for the BH growth can be weaker. In the cold dark matter cosmology, larger galaxies form hierarchically through mergers of smaller galaxies. Hence, many MBHs are assembled in a larger galaxy, if smaller galaxies already possess MBHs. Furthermore, MBHs could be born in hyper-massive star clusters formed by galaxy collisions \\citep{Matsui11}. Thus, galaxy merger remnant can contain many MBHs, even if precursory galaxies have no MBHs. Observationally, multiple active galactic nucleus (AGN) systems have been discovered recently. They include a triple AGN in the galaxy SDSS J1027+1749 at $z = 0.066$ \\citep{Liu11}, three rapidly growing MBHs of $10^6-10^7 M_\\odot$ in a clumpy galaxy at $z = 1.35$ \\citep{Schawinski11}, a firstly-discovered physical quasar triplet QQQ J1432--0106 within the projected separation of $30-50$kpc at $z = 2.076$ \\citep{Djorgovski07}, and a second quasar triplet QQQ J1519+0627 within the projected separation of $200$kpc, which is likely to be harboured in a yet-to-be-formed massive system at $z=1.51$ \\citep{Farina13}. According to the hierarchical merger history, galaxies with many MBHs are likely to form at higher redshifts. Although the galaxy merger proceeds through the violent relaxation, the merger of MBHs has difficulty. As pointed out in \\citet{Begelman80}, two MBHs in a galaxy are likely to form a binary, but unlikely to merge directly due to the so-called loss cone depletion (the depletion of stars on orbits that intersect the binary MBH). A binary MBH cannot reach sub-parsec separation due to the loss cone depletion, which is called the final parsec problem \\citep[e.g.][]{Merritt04}. A possible way to evade the loss cone depletion is the nonaxisymmetric potential of the host galaxy \\citep{Merritt04,Berczik06,Khan11,Khan12}, which is the natural consequence of the galaxy merger. A binary MBHs can merge in the nonaxisymmetric potential in $10$~Gyr or $0.3$~Gyr, when the galaxy contains stars with $10^9M_{\\odot}$ or $10^{11}M_{\\odot}$, respectively \\citep{Khan11}. However, since this timescale is comparable to or longer than the galactic dynamical timescale, other galaxies harbouring MBHs can intrude before two MBHs merge. This is likely to occur at higher redshifts of $z \\gtrsim 6$, at which the universe age is less than $1$~Gyr. If there are more than two MBHs in a galaxy, the dynamical relaxation of MBHs is significantly controlled by the gravity of MBHs themselves, especially by three-body interaction. When a third MBH intrudes into a binary MBH, one of the three MBHs carries away angular momentum from the rest two MBHs, reducing the binary separation, and eventually the binary merges \\citep[e.g.][]{Iwasawa06}. So far, galaxy structures have not been investigated when the galaxies contain more than three MBHs, although they have been investigated in the cases of two MBHs \\citep[e.g.][]{Khan11,Khan12} and three MBHs \\citep[e.g.][]{Iwasawa08}. \\cite{Tanikawa11} (hereafter, paper~I) scrutinised a system of multiple MBHs in a galaxy by high-resolution $N$-body simulations, and found that multiple MBHs produce one dominant MBH through successive mergers. Binary MBHs lose their angular momentum owing to sling-shot mechanism, which induces the decay of the binary orbits through gravitational wave (GW) radiation. In paper~I, we investigated one model of a galaxy containing multiple MBHs. In this paper, we explore the evolution of multiple MBHs in galaxies with different 3-dimensional stellar velocity dispersions to derive the criteria of the MBH merger. We also consider the effect of the recoil by anisotropic GW radiation at the MBH merger. Since the recoil velocity typically reaches several hundred km~s$^{-1}$ \\citep{Kesden10}, it could suppress the MBH growth. Furthermore, we investigate the impact by the MBH merger on the galaxy structure. The paper is organised as follows. In section \\ref{sec:model}, we describe the simulation model. In section \\ref{sec:results}, we show numerical results. In section \\ref{sec:criteria}, the results are translated to derive the criteria for MBH merger, which are applied for high and low redshift galaxies. In section \\ref{sec:back-reaction}, the back-reaction to a host galaxy is discussed with respect to the galactic structure and the production of high velocity stars. In section \\ref{sec:summary}, we summarise this paper. ", "conclusions": "\\label{sec:summary} We have performed $N$-body simulations to investigate successive mergers of MBHs in galaxies with different masses and radii. We have found that about a half of multiple MBHs successively merge to one bigger MBH within $140t_{\\rm dy,g}$ in galaxies with the velocity dispersion larger than $\\sim 180$~km~s$^{-1}$. The merger of MBHs is promoted, such that the loss cone of binary MBHs is refilled by MBHs losing their angular momenta due to dynamical friction. GW recoil does not affect the merger process, if the recoil velocity is of the order of the stellar velocity dispersion. Galaxies which allow multiple MBHs to merge should reside in dark matter haloes with the mass more than $4 \\times 10^{10}M_{\\odot}$, if these dark matter haloes form at high redshifts. These galaxies could correspond to LAEs or LBGs brighter than the UV magnitude $M_{\\rm UV} \\simeq -19$ at high redshifts. On the other hand, an MBH which has experienced the successive merger can inhabit low-redshift galaxies brighter than $M_{\\rm UV} \\simeq -18$. We have also investigated the evolution of the galactic structure and the generation of high-velocity stars as the back-reaction by the successive merger of MBHs. We have found that the dynamics of MBHs affects the central regions of galaxy that contain about ten times the total mass of MBHs. The mass density profile is transformed to $\\rho \\propto r^{-0.5}$, which is the same as the mass density profile in the case of a galaxy with two and three MBHs. The mass density in the central regions is $1.5$ times smaller than in the case of the galaxy with two MBHs. In a galaxy with ten MBHs, high-velocity stars are generated intermittently, while they are generated at a constant rate in the case of a galaxy with two MBHs. Such features should enable us to constrain the merger mechanism of MBHs in a galaxy." }, "1402/1402.3614_arXiv.txt": { "abstract": "Recently the number of main-sequence and subgiant stars exhibiting solar-like oscillations that are resolved into individual mode frequencies has increased dramatically. While only a few such data sets were available for detailed modeling just a decade ago, the {\\it Kepler} mission has produced suitable observations for hundreds of new targets. This rapid expansion in observational capacity has been accompanied by a shift in analysis and modeling strategies to yield uniform sets of derived stellar properties more quickly and easily. We use previously published asteroseismic and spectroscopic data sets to provide a uniform analysis of 42 solar-type {\\it Kepler} targets from the Asteroseismic Modeling Portal (AMP). We find that fitting the individual frequencies typically doubles the precision of the asteroseismic radius, mass and age compared to grid-based modeling of the global oscillation properties, and improves the precision of the radius and mass by about a factor of three over empirical scaling relations. We demonstrate the utility of the derived properties with several applications. ", "introduction": "\\label{SEC1} It is difficult to overstate the impact of the {\\it Kepler} mission on the observation and analysis of solar-like oscillations in main-sequence and subgiant stars. In a review from just a decade ago, \\cite{bk03} highlighted the tentative detections of individual oscillation frequencies in just a few such stars from ground-based observations, and {\\it Kepler} was not even mentioned. Despite funding issues that delayed the mission from an original deployment date in 2006, {\\it Kepler} finally launched in March 2009 and operated almost flawlessly for more than 4 years, slightly exceeding its design lifetime \\citep{bor10}. The archive of public data now includes nearly uninterrupted observations for many thousands of solar-type stars, including short-cadence data \\citep[58.85~s sampling,][]{gil10} for hundreds of these targets. In the span of a decade, the study of solar-like oscillations has been transformed dramatically \\citep{cm13}. During the first 10 months of science operations, {\\it Kepler} performed a survey for solar-like oscillations in more than 2000 main-sequence and subgiant stars, yielding detections in more than 500 targets from the 1-month data sets. The initial analysis of this ensemble, using empirical scaling relations \\citep{kb95} to generate estimates of radius and mass, suggested a significant departure from the mass distribution expected from Galactic population synthesis models \\citep{cha11}. Subsequent analysis of the sample, using updated effective temperatures \\citep{pin12} and a substantial grid-based modeling effort, led to more precise estimates of the radii and masses as well as information about the stellar ages \\citep{cha14}. Some of the brightest stars from the survey were subjected to a more detailed analysis, including spectroscopic follow-up to determine more precise atmospheric properties \\citep{bru12} plus the identification and detailed modeling of dozens of oscillation frequencies in each star \\citep{mat12}. These studies gave us a preview of what to expect from the subsequent phase of the mission. Starting with Quarter 5 (Q5), {\\it Kepler's} short-cadence study of solar-like oscillations transitioned to a specific target phase, where extended observations began for a fixed sample of stars identified during the survey. The target list during this phase gave priority to stars showing oscillations with the highest signal-to-noise ratio (S/N), but it also retained the brightest main-sequence stars cooler than the Sun, where the lower intrinsic oscillation amplitudes yielded relatively weak detections from the survey. From Q5 through the end of the mission (Q17), about 200 of the 512 available short-cadence targets were typically specified by the {\\it Kepler} Asteroseismic Science Consortium \\citep[KASC,][]{kje10} and about half of those were intended for the study of solar-like oscillations. Just like the exoplanet side of the mission, the KASC team gradually improved the data reduction and analysis methods while additional data swelled the archive. Never in the history of the field had such extended monitoring been possible at all, let alone for such a large sample of stars. As a consequence, the availability of reliable sets of input data for stellar modeling lagged well behind the continually expanding time-series for each star in the archive. This delay was primarily due to the challenge of coordinating the efforts of multiple teams, first to produce optimized light curves from the raw {\\it Kepler} data \\citep{gar11}, then to fit the global oscillation properties and remove the stellar granulation background from the power spectra \\citep{ver11,mat11}, and finally to extract and identify the individual oscillation frequencies using so-called ``peak-bagging'' techniques \\citep{app12}. Also like the exoplanet program, ground-based follow up observations were difficult to obtain for the faintest targets, further limiting the number of stars for which detailed modeling was feasible. Even after reliable sets of observational constraints became available, an analogous effort was required to consolidate the results from many stellar modeling teams. Initially this effort sought to define objective metrics of model quality, and to use the ensemble of results from different codes and methods to estimate the systematic uncertainties for a few specific targets \\citep{met10,met12,cre12,dog13,sil13}. The first large sample to emerge from the survey made this ``boutique'' modeling approach impractical, and motivated the initial large-scale application of the Asteroseismic Modeling Portal \\citep[AMP,][]{met09,woi09}. \\cite{mat12} presented a uniform analysis of 22 {\\it Kepler} targets observed for 1 month each during the survey phase, and compared detailed modeling from AMP with empirical scaling relations and with results from several grid-based modeling methods that matched the global oscillation properties ($\\Delta\\nu$ and $\\nu_{\\rm max}$, see below) rather than the individual frequencies from peak-bagging. The results clearly demonstrated the improved level of precision that was possible from detailed modeling of the individual oscillation frequencies, particularly for stellar ages. In this paper we present stellar modeling results from AMP for the first large sample of {\\it Kepler} targets with extended observations during the specific target phase of the mission. In section~\\ref{SEC2} we describe the sample, which was drawn from the most recently published observations. We outline our stellar modeling approach in section~\\ref{SEC3}, including several improvements to the previous version of AMP and using slightly customized procedures for different types of stars. We present the main results and initial applications in section~\\ref{SEC4}, and we discuss conclusions and future prospects in section~\\ref{SEC5}. ", "conclusions": "\\label{SEC5} Our uniform asteroseismic analysis for a large sample of {\\it Kepler} main-sequence and subgiant stars yielded precise determinations of many stellar properties (see Table~\\ref{tab1}), and some important lessons for future work. Most of the simple stars have values of $f\\le10$, indicating a reasonable match between the models from AMP and the observational constraints. The exceptions (KIC\\,6603624, 8694723, 8760414 and 11244118) are stars with the most extreme metallicities in our sample, which may require modifications to the solar-composition mixture in the opacity tables employed by AMP for more accurate modeling. More of the F-like stars have $f>10$, but these are generally targets where the $\\ell=0$ and $\\ell=2$ modes are most difficult to separate. The {\\'e}chelle diagrams in Figure~\\ref{fig3} show that the extracted frequencies for F-like stars from \\cite{app12} include variations in the small separations $d_{0,2}(n)$ that cannot be reproduced by models---in which the $\\ell=0,2$ ridges always curve together with only monotonic variations in $d_{0,2}(n)$. This suggests that the smoothness criteria in peak-bagging pipelines may need to be modified for F-like stars. Automated fitting of the mixed-mode stars is limited to those with fewer than $\\sim$3 avoided crossings and the match to the observed frequencies is generally less precise, with most stars showing $f>10$. The greatest difficulties are again encountered for extreme metallicity targets (KIC\\,7976303 and 10018963), but the qualitative agreement with the locations of avoided crossings for most of the mixed-mode stars in Figure~\\ref{fig3} is remarkable. Ultra-precise constraints on the properties of these stars must rely on future dense grid-modeling to refine the estimates from Table~\\ref{tab1}. Our modifications to the fitting procedures to try and avoid a bias towards low-helium solutions have helped, but the problem was not eliminated entirely. As noted in section~\\ref{SEC3.1}, of the 22 stars considered by \\cite{mat12} six of them (27\\%) yielded an initial helium mass fraction significantly below the primordial value, while four additional targets (18\\%) were marginally below $Y_{\\rm P}$. We attempted to address this issue by including the frequency ratios as additional constraints, and by adopting larger uncertainties at higher frequencies where the surface correction is larger. A smaller fraction of our sample of 42 stars appears to be affected by the low-helium bias, but Table~\\ref{tab1} still includes one star (2.4\\%) with $Y_{\\rm i}$ significantly below $Y_{\\rm P}$ and three additional targets (7.1\\%) that fall marginally below $Y_{\\rm P}$. Further improvement may require an alternative surface correction \\citep[e.g.,][]{jcd12}. The precision of asteroseismic data sets from extended observations with {\\it Kepler} now demands that we address the dominant sources of systematic error in the stellar models. The largest source arises from incomplete modeling of the near-surface layers. Although something like the empirical correction of \\cite{kje08} will continue to be useful, we are now in a position to capture more of the relevant physics. For example, \\cite{gru13} performed a Bayesian analysis of the 22 {\\it Kepler} targets from \\cite{mat12}, including a simplified non-adiabatic treatment of the pulsations. They found that the Bayesian probabilities were higher when non-adiabatic rather than adiabatic frequencies were fit to the observations, and that for most stars the surface effect was minimized and in some cases even eliminated. Their non-adiabatic model accounted for radiative losses and gains but neglected perturbations to the convective flux and turbulent pressure \\citep{gue94}. In the case of the Sun, the stability and frequency of the oscillation modes depends substantially on turbulent pressure and the inclusion of non-local effects in the treatment of convection \\citep{bal92a,bal92b,hou10}, and the resulting frequency shift is uncertain. The sensitivity of the results to the model of convection and the temperature profile in the super-adiabatic layer was also emphasized by \\cite{ros99} and \\cite{li02}. Regardless, the Bayesian approach has the advantage of incorporating the unknown sources of systematic error directly into the uncertainties on the derived stellar properties, and can reveal which approach to the pulsation calculations generally improves the model fits. For future analyses, we intend to augment ADIPLS with the non-adiabatic stellar oscillation code GYRE \\citep{tt13}, which currently includes a limited treatment of non-adiabatic effects but is flexible enough to incorporate additional contributions. We would also like to take advantage of the numerical stability and modular architecture of the open-source MESA code \\citep{pax13} to explore different chemical mixtures and to include heavy element diffusion and settling in the evolutionary models, which is not currently stable for all types of stars with ASTEC. With these new modules for stellar evolution and pulsation calculations, we can embed a Bayesian formalism into the parallel genetic algorithm to complement the simple $\\chi^2$ approach. The complete sample of asteroseismic targets from extended {\\it Kepler} observations spanning up to four years will provide a rich data set to validate these new ingredients for the next generation of AMP. The golden age of asteroseismology for main-sequence and subgiant stars owes a great debt to the {\\it Kepler} mission, but it promises to continue with the anticipated launch of NASA's Transiting Exoplanet Survey Satellite \\citep[TESS,][]{ric14} in 2017. While {\\it Kepler} was able to provide asteroseismic data for hundreds of targets and could simultaneously monitor 512 stars with 1-minute sampling, TESS plans to observe $\\sim$500,000 of the brightest G- and K-type stars in the sky at a cadence sufficient to detect solar-like oscillations. The data sets will be nearly continuous for at least 27 days, but in two regions near the ecliptic poles the fields will overlap for durations up to a full year. These brighter stars will generally be much better characterized than the {\\it Kepler} targets---with parallaxes from {\\it Hipparcos} and ultimately {\\it Gaia} \\citep{per01}, and reliable atmospheric constraints from ground-based spectroscopy---making asteroseismic characterization more precise and accurate. With several years of development time available, AMP promises to be ready to convert this avalanche of data into reliable inferences on the properties of our solar system's nearest neighbors." }, "1402/1402.4811_arXiv.txt": { "abstract": "We present high-energy (3--30~keV) {\\it NuSTAR} observations of the nearest quasar, the ultraluminous infrared galaxy (ULIRG) Markarian~231 (Mrk 231), supplemented with new and simultaneous low-energy (0.5--8~keV) data from {\\it Chandra}. The source was detected, though at much fainter levels than previously reported, likely due to contamination in the large apertures of previous non-focusing hard X-ray telescopes. The full band (0.5--30 keV) X-ray spectrum suggests the active galactic nucleus (AGN) in Mrk~231 is absorbed by a patchy and Compton-thin (N$_{\\rm H} \\sim1.2^{+0.3}_{-0.3}\\times10^{23}$~cm$^{-2}$) column. The intrinsic X-ray luminosity (L$_{\\rm 0.5-30~keV}\\sim1.0\\times10^{43}$~erg~s$^{-1}$) is extremely weak relative to the bolometric luminosity where the 2--10~keV to bolometric luminosity ratio is $\\sim$0.03\\% compared to the typical values of 2--15\\%. Additionally, Mrk~231 has a low X-ray-to-optical power law slope ($\\alpha_{\\rm OX}\\sim-1.7$). It is a local example of a low-ionization broad absorption line (LoBAL) quasar that is intrinsically X-ray weak. The weak ionizing continuum may explain the lack of mid-infrared [O IV], [Ne V], and [Ne VI] fine-structure emission lines which are present in sources with otherwise similar AGN properties. We argue that the intrinsic X-ray weakness may be a result of the super-Eddington accretion occurring in the nucleus of this ULIRG, and may also be naturally related to the powerful wind event seen in Mrk 231, a merger remnant escaping from its dusty cocoon. ", "introduction": "\\label{sec:intro} The presence of outflows in luminous quasars and active galactic nuclei (AGN) is most evident in broad absorption line (BAL) systems \\citep{lynds67}. Approximately 20\\% of quasars show BAL features at some level \\citep[e.g.,][]{gibson09}, understood to be the result of an observational sightline through AGN-driven wind material \\citep[e.g.,][]{weymann91, ogle99, dipompeo13}. In a common model for BAL quasars \\citep[e.g.,][]{murray95, proga00}, the wind is launched from the accretion disk approximately 0.01--0.1 pc from the black hole and radiatively driven by ultraviolet (UV) line pressure. However, X-ray emission from typical quasars would over-ionize the gas, quenching the line-driving mechanism and thus preventing the appearance of BAL features. Therefore, this so-called ``disk-wind'' model requires that BAL quasars have lower levels of X-ray emission, which could either be intrinsic or be due to shielding very close to the black hole, near the base of the wind. Indeed, BAL systems are very often observed to be X-ray faint \\citep[e.g.,][]{g06, gibson09, wu10}, though observations have yet to provide a clear picture of the origin of this X-ray faintness. Absorption of a nominal X-ray continuum has been established in several cases \\citep[e.g.,][]{g02, grupe03, shemmer05, giustini08}, but there may be diversity among the population. Following the terminology of other studies of X-ray emission in BAL quasars \\citep[e.g.,][and references therein]{luo13}, the term ``X-ray weakness or faintness'' refers to the observed X-ray emission being significantly lower than expected given the optical-UV emission. This may be caused by a normal X-ray continuum modified by absorption from gas and dust. The term ``intrinsic X-ray weakness'' refers to an innate property of the AGN, which is one possible cause of the observed X-ray faintness. In order to investigate whether intrinsic X-ray weakness or shielding gas is the most likely explanation for the observed X-ray faintness of BAL quasars, sensitive observations above 10~keV are highly desirable as they can distinguish between the presence of Compton-thick ($N_{\\rm H} \\geq 1.5 \\times 10^{24}$~cm$^{-2}$) shielding material and a less-obscured, intrinsically X-ray weak nucleus. The {\\it Nuclear Spectroscopic Telescope Array} ({\\it NuSTAR}; \\citealp{nustar}) is the first focusing high-energy X-ray telescope in orbit, and provides unprecedented sensitivity and angular resolution for high-energy, or hard (3--79 keV) X-ray photons. Recently, {\\it NuSTAR} observed two of the optically brightest BAL quasars known, PG~1004+130 and PG~1700+518, with the goal of investigating the origin of their X-ray faintness \\citep{luo13}. Unfortunately, low count rates prevented a definitive answer, but they conclusively showed that any absorption present must be Compton-thick. \\citet{luo13} were also able to demonstrate that about 17--40\\% of BAL quasars are intrinsically X-ray weak by stacking {\\it Chandra} observations of $z\\approx1.5-3$ BAL quasars in the Large Bright Quasar Survey. Confirmation of intrinsically X-ray weak AGN in some BAL quasars would have important consequences regarding the origin of BAL features. Instead of invoking shielding gas with an arbitrarily thick column density, intrinsic X-ray weakness would favor a scenario in which the BAL features are coupled to abnormally faint coronal X-ray emission, perhaps due to some mechanism that (periodically) quenches the X-ray corona (see \\S~4.2 of \\citealp{luo13} for discussion). To this end, we have obtained an extremely sensitive, broadband (0.5--30~keV) X-ray spectrum of Markarian 231 (Mrk 231), the nearest BAL quasar, using {\\it Chandra} and {\\it NuSTAR}. This paper is organized as follows: \\S2 details the multi-wavelength properties of Mrk~231; \\S3 presents the new X-ray observations obtained for this study; \\S4 discusses the previously reported hard X-ray spectra of Mrk~231, which are inconsistent with our new data; \\S5 presents our modeling of the new broadband X-ray spectrum, \\S6 discusses our favored scenario, that Mrk~231 is indeed intrinsically X-ray weak; and \\S7 summarizes our results. Throughout this paper, we adopt $H_0$ = 71 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_M = 0.27$, and $\\Lambda = 0.73$ \\citep{cosmo}. Luminosities taken from the literature have been recalculated for our assumed cosmology. \\begin{figure*}[ht] \\centering \\includegraphics[width=5.5in, angle=270]{mrk231_orientation_pretty.ps} \\caption{Schematic showing the complexity of Mrk~231's structure and our assumed viewing geometry based on multiwavelength data. The host galaxy is face-on on the plane of the sky with optical emission extending out to a radius of 8~kpc ($9\\farcs8$) \\citep[e.g.,][]{rv11}. The thin (23~pc) molecular disk is tilted by 10$^\\circ$ with the denser inner region extending from $\\sim$75~pc ($0\\farcs1$) to 1.1~kpc ($1\\farcs3$) and the outer molecular disk extending out to 2.5~kpc ($3\\farcs0$) \\citep{downes98}. The accretion disk is assumed to be orthogonal to the pc-scale jet, which is assumed to be inclined by 65$^{\\circ}$ (see \\S\\ref{sec:torus}). In this figure, north is up.} \\label{fig:geometry} \\end{figure*} ", "conclusions": "\\label{sec:summary} We have obtained the most sensitive broadband (0.5--30~keV) spectrum of Mrk~231, the closest BAL quasar and an ULIRG, to date. {\\it NuSTAR} detects Mrk~231 at a fainter level above 10~keV than previous observations with poorer angular resolution by {\\it BeppoSAX} PDS and {\\it Suzaku} PIN. Two epochs of {\\it NuSTAR} data separated by seven months show no evidence of hard X-ray variability so contamination is the likely cause of the discrepancy between {\\it NuSTAR} and previous, non-focusing hard X-ray observations. Analysis of contemporaneous {\\it Chandra} and {\\it NuSTAR} broadband spectra find significant X-ray emission from a powerful circumnuclear starburst. The direct AGN emission is absorbed and scattered by a patchy torus that is Compton-thin with N$_{\\rm H}\\sim1.2^{+0.3}_{-0.3}\\times10^{23}$~cm$^{-2}$. The Compton-thin absorption is compatible with the apparent weak Fe~K emission (EW$\\sim$0.2~keV).% Mrk~231 appears intrinsically X-ray weak. The absorption-corrected 2--10~keV luminosity is $\\lesssim0.1$\\% of the AGN bolometric luminosity with $\\alpha_{\\rm OX}\\sim-1.7$. A reason for the low X-ray luminosity could be that the AGN is in a rare ultra soft state similar to those seen in X-ray binaries, though this is inconsistent with the measured $\\Gamma_{\\rm AGN}$. It is unclear why the AGN is intrinsically X-ray weak, but there are not very many known examples of such sources as points of comparison. The weak ionizing continuum explains the lack of mid-infrared [O IV], [Ne V], and [Ne VI] fine structure emission lines which are present in sources with otherwise similar AGN properties. % The intrinsic X-ray weakness may be a result of the super-Eddington accretion occurring in the nucleus of the ULIRG. These results on Mrk~231 are consistent with its being a merger remnant emerging from its dusty cocoon through a powerful wind event." }, "1402/1402.4350_arXiv.txt": { "abstract": "Relativistic contributions to the dynamics of structure formation come in a variety of forms, and can potentially give corrections to the standard picture on typical scales of 100\\,Mpc. These corrections cannot be obtained by Newtonian numerical simulations, so it is important to accurately estimate the magnitude of these relativistic effects. Density fluctuations couple to produce a background of gravitational waves, which is larger than any primordial background. A similar interaction produces a much larger spectrum of vector modes which represent the frame-dragging rotation of spacetime. These can change the metric at the percent level in the concordance model at scales below the equality scale. Vector modes modify the lensing of background galaxies by large-scale structure. This gives in principle the exciting possibility of measuring relativistic frame dragging effects on cosmological scales. The effects of the non-linear tensor and vector modes on the cosmic convergence are computed and compared to first-order lensing contributions from density fluctuations, Doppler lensing, and smaller Sachs-Wolfe effects. The lensing from gravitational waves is negligible so we concentrate on the vector modes. We show the relative importance of this for future surveys such as Euclid and SKA. We find that these non-linear effects only marginally affect the overall weak lensing signal so they can safely be neglected in most analyses, though are still much larger than the linear Sachs-Wolfe terms. The second-order vector contribution can dominate the first-order Doppler lensing term at moderate redshifts and are actually more important for survey geometries like the SKA. ", "introduction": "\\label{sec1} Relativistic corrections to the standard model of cosmology come in a variety of forms, from the altering the dynamics of structure formation to the various effects associated to the interpretation of observations, in particular modifying the propagation of light. There has been considerable debate as to the importance and amplitude of these effects on the dynamics of the expansion of the universe and the growth of large scale structure (see, e.g., Ref.~\\cite{Clarkson:2011zq} for an overview), and the amplitude and importance of these dynamical effects are still actively debated~\\cite{buchertetal,nobackreac,backsig}. Though subdominant for linear structure formation, relativistic corrections are a generic prediction of General Relativity and are inevitable at a non-linear level through mode-mode coupling. The scalar gravitational potential induces rotational frame-dragging modes in spacetime (so-called vector modes) as well as gravitational waves (tensor modes). Neither of these have counterparts in Newtonian gravity as they both induce a non-zero magnetic Weyl curvature which is absent in Newtonian gravity and difficult to take into account in N-body numerical simulations~\\cite{Bruni:2013mua,Adamek:2013wja}. They therefore serve as an important tool in understanding purely relativistic aspects of structure formation and its observational consequences, as they set a lower limit on the amplitude of relativistic corrections. On top of dynamical corrections, relativistic effects also induce corrections to the propagation of light since it probes the complete spacetime geometry. This can alter the interpretation of cosmological observations at a level that cannot be neglected in an era of ``precision cosmology''. Provided one works within perturbation theory, the amplitude of these effects is computable and completely fixed once the normalisation of the scalar power spectrum, at the linear level, is determined. For instance, some relativistic effects have been taken into account on the cosmic microwave background~\\cite{pub} and shown to be below the constraints on non-Gaussianity derived by \\textit{Planck}~\\cite{planckXXIX}, but nevertheless in principle detectable on small angular scales, in particular through spectral distortions~\\cite{spectral}. This article focuses on the effect of relativistic corrections on weak lensing observations, focusing mainly on the induced vector mode background. Weak gravitational lensing by the large-scale structure of the Universe has now become a major tool of cosmology~\\cite{revuesWL}, used to study questions ranging from the distribution of dark matter to tests of general relativity~\\cite{testGR}. The propagation of light in an inhomogeneous universe gives rise to both distortion and magnification induced by gravitational lensing. The effect of non-linear corrections on the Hubble diagram have been considered~\\cite{BenDayan:2012wi,cemuu,Umeh:2012pn,Umeh:2014ana} and shown to be non-negligible given the accuracy of contemporary observations~\\cite{Fleury:2013sna,Fleury:2013uqa,voidslensing,BenDayan:2012ct,BenDayan:2013gc}. In this article we consider the effect on the weak lensing convergence of non-linear effects that induce the existence of a vector and tensor modes background. We compare this to the various contributions to the convergence at first-order~-- the usual integral of the density contrast along the line of sight~\\cite{revuesWL}, the contribution from the Doppler effect which is dominant at low redshifts and large scales~\\cite{Bonvin:2008ni,voidslensing,Bacon:2014uja}, the Integrated Sachs-Wolfe (ISW) and Sachs-Wolfe (SW) terms which are relatively small and mainly neglected when computing cosmic convergence. The induced background of gravitational waves from scalar-scalar coupling was presented in Ref.~\\cite{Ananda:2006af} during the radiation era, and its present-day spectrum calculated in Ref.~\\cite{Baumann:2007zm}, with shear lensing effects studied in Ref.~\\cite{Sarkar:2008ii}, all following the pioneering analysis of Ref.~\\cite{Mollerach:2003nq}. Surprisingly it was found that the induced gravitational wave background is significantly larger than any primordial background (even for a tensor-scalar ratio $r\\sim0.1$) on intermediate scales of $\\sim$100\\,Mpc, which is around the equality scale, though of course it is much smaller on small scales. Similarly, the induced vector mode background was presented in Refs.~\\cite{Lu:2007cj,Lu:2008ju}, and again a spectrum was found that peaks on 100\\,Mpc scales. Remarkably, however, it was found that the amplitude of the background of vector modes for the metric potential behaves on small scales with the same scaling as the gravitational potential, with nearly 1\\% of its amplitude. While both of these induced degrees of freedom have little effect on the dynamics of structure formation (they cannot directly source the density fluctuation as it is a scalar degree of freedom) they can influence the gravitational lensing produced by large-scale structure. Is it significant, and could it be a new way to detect relativistic aspects of structure formation? The effects of these contributions on weak lensing convergence predictions are computed in order to understand if they can either be detected or, in the worst case, bias the analysis of future weak lensing experiments, such as Euclid or SKA; {\\em i.e.}, if the interpretation of the observation by assuming that the observed convergence corresponds to the convergence sourced by scalar modes only is an accurate enough assumption or whether some of these effects have to be included in the analysis. This article addresses this question and computes the effect of these two non-linear effects on weak lensing observations by considering second order vector and tensor background. We restrict our analysis to the hypothesis that the Born approximation still holds. In principle, one needs also to take into account second order effects on the geodesic deviation equation~\\cite{Seitz:1994xf, Cooray:2002mj,Dodelson:2005zj,Schaefer:2005up,Shapiro:2006em}, as fully described in Refs.~\\cite{Bernardeau:2009bm,Bernardeau:2011tc}. This effect is however small. In Section~\\ref{sec2} we describe the vector and gravity waves background induced by the non-linear dynamics and then, in Section~\\ref{sec3}, the computation of the weak lensing power spectra, splitting the effects of the scalar, vector, tensor, Doppler, ISW and SW contributions in order to compare their magnitude. Since the contribution of the tensor modes remains negligible and both ISW and SW being relatively small, we focus in Section~\\ref{sec4} on the vector and Doppler contribution, estimating their magnitude in surveys such as Euclid and SKA. Technical details are gathered in Appendices~\\ref{app0}-\\ref{appC}. ", "conclusions": "\\label{sec5} This article has evaluated the amplitude of relativistic contributions to the weak lensing power spectra. We have considered the gravitational wave and vector mode backgrounds which are sourced at second-order by density perturbations. The amplitude of these backgrounds are completely fixed once the normalisation of the scalar power spectrum in the linear regime is determined. As these are purely relativistic degrees of freedom they set the lower limit for all relativistic effects on cosmological modelling. While the gravitational wave background is very small in relation to the scalars, the vectors, which represent frame dragging in the metric, give corrections to the metric at nearly the percent level. The effect of these contributions on weak lensing convergence predictions have been computed in order to understand if they can either be detected, or bias the analysis of future weak lensing experiments, such as Euclid or SKA. We have compared them to the usual gravitational lensing contribution, the two Sachs-Wolfe contributions as well as the Doppler lensing contribution~\\cite{Bacon:2014uja}. First, we have shown that even though the non-linear tensor mode background dominates over any possible primordial gravitational wave contribution, its effect on weak lensing is completely negligible, by 10 to 12 order of magnitudes (see Figs.~\\ref{p12} and~\\ref{points}). Then, we have shown that the vector contribution to the convergence, while small, can dominate over the Doppler lensing at high redshift~-- but there it is swamped by gravitational lensing by density perturbations. We have shown this both for point sources and for two survey geometries. The vectors are actually more important than the Doppler term for SKA-like source distributions on small scales, but not for a Euclid like survey. For both of these surveys the vectors only reach about $10^{-3}$\\% that of the normal gravitational lensing contribution, and so can be safely neglected. Nevertheless, it is interesting that the vector contribution can be as important as some linear terms. We have also recovered that although the frame dragging effect is small, it becomes more important than both ISW and SW above $\\ell \\ge 50$. This comes to corroborate the fact that for observations, neglecting the 2 first order Sachs-Wolfe terms is a good approximation. In this analysis, the non-linear effects of the metric perturbations have been described at second order while weak lensing was described assuming that the Born approximation still holds. In principle, one needs also to take into account second order effects on the geodesic deviation equation~\\cite{Seitz:1994xf, Cooray:2002mj,Dodelson:2005zj,Schaefer:2005up,Shapiro:2006em}, as fully described in Refs.~\\cite{Bernardeau:2009bm,Bernardeau:2011tc}. There are a huge variety of second-order effects which come into the convergence. We have only considered two contributions which arise from non-linear dynamical effects which happen as structure forms. Many contributions appear when calculating the lensing convergence itself~\\cite{Umeh:2012pn,Umeh:2014ana,BenDayan:2012wi}, and these also need to be analysed in a similar manner to that presented here to determine whether relativistic effects are important for future observations of magnification." }, "1402/1402.4485_arXiv.txt": { "abstract": "{We present the most extensive analysis of Fourier-based \\mbox{X-ray} timing properties of the black hole binary Cygnus X-1 to date, based on 12 years of bi-weekly monitoring with RXTE from 1999 to 2011. Our aim is a comprehensive study of timing behavior across all spectral states, including the elusive transitions and extreme hard and soft states. We discuss the dependence of the timing properties on spectral shape and photon energy, and study correlations between Fourier-frequency dependent coherence and time lags with features in the power spectra. Our main results are: (a) The fractional rms in the 0.125--256\\,Hz range in different spectral states shows complex behavior that depends on the energy range considered. It reaches its maximum not in the hard state, but in the soft state in the Comptonized tail above 10\\,keV. (b) The shape of power spectra in hard and intermediate states and the normalization in the soft state are strongly energy dependent in the 2.1--15\\,keV range. This emphasizes the need for an energy-dependent treatment of power spectra and a careful consideration of energy- and mass-scaling when comparing the variability of different source types, e.g., black hole binaries and AGN. PSDs during extremely hard and extremely soft states can be easily confused for energies above $\\sim$5\\,keV in the 0.125--256\\,Hz range. (c) The coherence between energy bands drops during transitions from the intermediate into the soft state but recovers in the soft state. (d) The time lag spectra in soft and intermediate states show distinct features at frequencies related to the frequencies of the main variability components seen in the power spectra and show the same shift to higher frequencies as the source softens. Our results constitute a template for other sources and for physical models for the origin of the \\mbox{X-ray} variability. In particular, we discuss how the timing properties of \\mbox{Cyg~X-1} can be used to assess the evolution of variability with spectral shape in other black hole binaries. Our results suggest that none of the available theoretical models can explain the full complexity of \\mbox{X-ray} timing behavior of \\mbox{Cyg~X-1}, although several ansatzes with different physical assumptions are promising.} ", "introduction": "The canonical states of accreting black hole binaries have first been observed and defined in the spectral domain: a hard state with a power law spectrum with a photon spectral index of $\\sim$1.7 and a soft state with a spectrum dominated by thermal emission from an accretion disk. They are joined by usually short-lived transitional or intermediate states. The whole sequence of states -- from hard state over the intermediate into soft and then again into intermediate and finally into hard state -- can best be depicted on a hardness-intensity-diagram (HID), where transient sources that undergo a full outburst follow a \\textsf{q}-shaped track \\citep[][see \\citeauthor*{McClintock_Remillard_2006a}, \\citeyear{McClintock_Remillard_2006a}, for a different nomenclature]{Fender_2004a}. Radio emission is detected in the hard state, with jets imaged for \\object{\\mbox{Cyg~X-1}} \\citep{Stirling_2001a} and \\object{GRS~1915$+$105} \\citep{Dhawan_2000a,Fuchs_2003a}. In the soft state, radio emission is strongly quenched. Evidence for similar spectral behavior also exists in several other classes of accreting objects such as neutron star \\mbox{X-ray} binaries \\citep[e.g.,][\\object{Aql~X-1}]{Maitra_2004a}, active galactic nuclei \\citep[AGN;][]{Koerding_2006a}, and dwarf novae \\citep[][\\object{SS~Cyg}]{Koerding_2008a}. The spectral states, including the different flavors of the intermediate state, show distinct \\mbox{X-ray} timing characteristics, such as shapes of power spectra or time lags between emission at different energies. Timing parameters seem to be a remarkably sensitive tool for defining state transitions \\citep[e.g.,][]{Pottschmidt_2003b,Fender_2009a,Belloni_2010a}. While the radio emission, and possibly also the gamma-ray emission above 400\\,keV \\citep{Laurent_2011a,Jourdain_2012a} originate in the jet, the origin of the \\mbox{X-ray}s is still unclear. As shown, e.g., by \\citet{Nowak_2011a}, the combination of the best resolution and the most broadband \\mbox{X-ray} spectra available today fails to enable us to statistically distinguish between jet models \\citep{Markoff_2005a, Maitra_2009a} and thermal and/or hybrid Comptonization in a corona \\citep[e.g.,][]{Coppi_1999a,Coppi_2004a}. The fluorescent Fe K$\\alpha$ line and a reflection hump point towards a contribution by reflection, independent of the origin of the continuum \\citep[see][for a review and \\citealt{Duro_2011a}, \\citealt{Tomsick_2014a}, and references therein for \\mbox{Cyg~X-1}]{Reynolds_2003a}. Spectro-timing analysis holds the promise of solving this ambiguity, as a truly physical model has to consistently describe both spectral and timing behavior. While no self-consistent models that would encompass all parameters exist yet, some studies address, e.g., simultaneous modeling of photon spectra and root mean square variability (rms) spectra \\citep{Gierlinski_2005a} or photon spectra and Fourier-dependent time lags \\citep{Cassatella_2012a}. Theoretical ansatzes to describe the \\mbox{X-ray} variability include propagating mass accretion rate fluctuations \\citep[usually based on][see, e.g., \\citealt{Ingram_2013a} for an analytical model]{Lyubarskii_1997a}, up-scattering in a jet \\citep{Reig_2003a,Kylafis_2008a}, and full magnetohydrodynamic simulations \\citep[][who concentrate on spectra but also address timing properties]{Schnittman_2013a}. There is mounting evidence for similarities in timing behavior between \\mbox{X-ray} binaries, AGN \\citep[see][for a review]{McHardy_2010a}, and recently also cataclysmic variables \\citep[see][for discovery of Fourier-dependent time lags]{Scaringi_2013a}. Because of their brightness and short variability time scales, \\mbox{X-ray} binaries remain the best laboratories to investigate these phenomena and are key to deciphering the complex and currently highly disputed interplay of accretion and ejection processes. We note especially that AGN are usually viewed on a single state due to the very much longer variation timescale, and thus \\mbox{X-ray} binaries are needed to study the variety of states and their inter-relationships. The first step to understanding the \\mbox{X-ray} timing characteristics is a fundamental overview of their evolution with the spectral shape that can only be achieved with a high number of high quality observations densely covering all states, including the elusive transitions. Most previous works concentrate on energy-independent evolution of rms and power spectral distributions (PSDs) with spectral state \\citep[e.g.,][]{Pottschmidt_2003b,Belloni_2005a,Axelsson_2006a,Klein-Wolt_2008a}, although further spectral shape-, energy-, and Fourier frequency-dependent correlations have been noted in individual observations and smaller samples \\citep[e.g.,][]{Homan_2001a,Rodriguez_2002b,Kalemci_2004a,Rodriguez_2004a, Boeck_2011a,Cassatella_2012b,Stiele_2013a}, often with a focus on the behavior of narrow quasi-periodic oscillations. Missing are consistent analyses of the energy-resolved evolution of rms and PSDs, Fourier-frequency dependent evolution of cross-spectral quantities (coherence function and lags), and correlations of features in PSDs and Fourier-frequency dependent cross-spectral quantities over the full range of spectral states and over multiple transitions. In this paper, we address these questions with an extraordinarily long and well sampled RXTE campaign on the high mass black hole binary \\mbox{Cyg~X-1} that enables us to conduct the most comprehensive spectro-timing analysis of a black hole binary to date. Located at a distance of $1.86^{+0.12}_{-0.11}$\\,kpc \\citep[][consistent with \\citealt{Xiang_2011a}]{Reid_2011a}, \\mbox{Cyg~X-1} is bright (in the hard state $\\sim$$7\\times10^{-9}$\\,erg\\,cm$^{-2}$\\,s$^{-1}$ in the 1.5--12\\,keV band of RXTE-ASM) and persistent and therefore a prime target for both spectral and timing studies. Often considered a prototypical black hole binary, it is frequently used for comparisons with other black hole binaries \\citep[e.g,][]{Munoz-Darias_2010a} and AGN \\citep{Markowitz_2003a,McHardy_2004a,Papadakis_2009a}. Although the spectrum of the source is never fully dominated by the disk and the bolometric luminosity only changes by a factor of $\\sim$4 \\citep[][and references therein]{Cui_1997a,Shaposhnikov_2006a,Wilms_2006a}, \\mbox{Cyg~X-1} often undergoes state transitions \\citep[e.g.,][]{Pottschmidt_2003b,Wilms_2006a,Grinberg_2013a} and is therefore also well suited for studies of the intermediate states. Here, we analyze data from the full set of RXTE observations of \\mbox{Cyg~X-1}. This paper is a part of a series where we previously analyzed spectro-timing correlation in the hard state 1998 to 2001 \\citep{Pottschmidt_2003b}, the rms-flux relation \\citep{Gleissner_2004a}, the radio-\\mbox{X-ray} correlations \\citep{Gleissner_2004b}, the spectral evolution 1999--2004 \\citep{Wilms_2006a}, and states and state transitions 1996--2012 with all sky monitors \\citep{Grinberg_2013a}. We start Sect.~\\ref{sect:data} by introducing the data and the general behavior of the source during the time period covered by this analysis and follow with a description of the spectral analysis and employed \\mbox{X-ray} timing techniques. In Sect.~\\ref{sect:rms_and_power}, we discuss the energy independent and dependent evolution of the rms and the PSDs with spectral shape. In Sect.~\\ref{sect:cross}, we discuss the evolution of cross-spectral quantities with spectral shape. We address the implication of our results for the analysis of other sources in Sect.~\\ref{sect:other} and for theories explaining the origin of \\mbox{X-ray} variability in Sect.~\\ref{sect:theory}. Sect.~\\ref{sect:summary} summarizes our finding with a focus on the implied directions for further investigations. ", "conclusions": "\\label{sect:summary} In the following we briefly summarize the highlight results of this paper that we expect to have the strongest constraints on future theoretical models and to offer the strongest stimuli for future investigations of the variability of \\mbox{X-ray} binaries: \\begin{itemize} \\item The fractional rms shows a complex behavior that strongly depends on the spectral state of the source and on the considered energy range. In particular, the fractional variability in the soft state is low at energies below $\\sim$4.5\\,keV ($\\sim$15\\%) but increases with energy up to $\\sim$40\\% in the 9.4--15\\,keV range. The ratio of fractional rms at different energies shows a tight dependence on spectral shape. \\item The shape of the power spectra is highly dependent on the energy band used, with a striking change of behavior at $\\Gamma_1 \\approx 2.6$--2.7. This reiterates the importance of an energy-dependent approach to PSDs and of carefully addressing the scaling of typical energies with mass when comparing variability from different source types. \\item The power spectra in hard and intermediate states are dominated by two main variability components that shift to higher frequencies as the source spectrum softens. In the $\\Gamma_1 \\approx 2.4$--2.7 range, the 9.4--15\\,keV band is dominated by the higher frequency component, while in the 2.1--4.5\\,keV and the 4.5--5.7\\,keV bands the lower frequency component is more prominent. \\item The power spectra of the hardest states show further variability components at higher Fourier frequencies. In the 0.125--256\\,Hz range, these PSDs can easily be confused with the softest observations. \\item The coherence is close to unity at frequencies below $\\sim$10\\,Hz in the hard and soft states. At the transition from intermediate to soft state, the coherence drops strongly, with hints of structure similar to the two dominant components of the power spectra. It recovers in the soft state. Above $\\sim$10\\,Hz, the coherence cannot be constrained in our orbit-wise approach. \\item The time lag spectra in the hard and intermediate states show features with a double-humped structure similar to the two dominant components of the PSDs and with the same evolution to higher frequencies as the photon spectra soften. No structure is visible in the time lag spectra with soft photon spectra ($\\Gamma_1 \\gtrsim 2.7$). \\item The average time lags in the 3.2--10\\,Hz band show a non-linear increase with $\\Gamma_1$ until $\\Gamma_1 \\approx 2.65$. For $\\Gamma_1 \\gtrsim 2.65$, the average lags are low and show no dependence on spectral shape. \\end{itemize} We have shown that taking into account spectro-timing correlation does not only provide a more holistic description of source properties, but can also help to assess the quality of spectral models. The spectro-timing analysis presented here for the exceptional data set of \\mbox{Cyg~X-1} can be used as a template for other sources with a worse coverage of different spectral shapes. The behavior it describes will allow to test future theoretical models for accretion and ejection processes against observations. Better constraints of some elusive timing parameters, such as the detailed structure of time lag spectra in individual observations, require larger area instruments and further monitoring of the long-term evolution as may be provided with future satellites such as ASTROSAT \\citep{Paul_2013a} and NICER \\citep{Gendreau_2012a}." }, "1402/1402.4166_arXiv.txt": { "abstract": "In this paper I present dynamic models of the radio source Centaurus A, and critique possible models of in situ particle reacceleration (ISR) within the radio lobes. The radio and $\\g$-ray data require neither homogeneous plasma nor quasi-equipartition between plasma and magnetic field; inhomogeneous models containing both high-field and low-field regions are equally likely. Cen A cannot be as young as the radiative lifetimes of its relativistic electrons, which range from a few to several tens of Myr. Two classes of dynamic models -- flow driven and magnetically driven -- are consistent with current observations; each requires Cen A to be on the order of a Gyr old. Thus, ongoing ISR must be occuring within the radio source. Alfven-wave ISR is probably occuring throughout the source, and may be responsible for maintaining the $\\g$-ray-loud electrons. It is likely to be supplemented by shock or reconnection ISR which maintains the radio-loud electrons in high-field regions. ", "introduction": "The full story of the radio galaxy Centaurus A is still not known, despite much careful study over the years. Cen A is of course worth study for its own sake, but it also provides a nearby, well-documented example from which we can understand the dynamic evolution of other, similar radio galaxies. In this paper I focus on two unresolved questions related to Cen A's large-scale radio lobes. The first question is the age of Cen A. To answer this we need to understand the physical state of its radio lobes. So-called Fanaroff-Riley Type II (``FRII'') radio galaxies are well understood: a collimated, supersonic jet develops a terminal shock where it impacts the ambient medium. The shock is apparent as an outer hot spot, and jet plasma which has been through the shock is left behind as a cocoon. Ram pressure drives the growth of the source, as the end of the jet advances into the ambient medium. However, Cen A is not an FRII, but a Fanaroff-Riley Type I (``FRI''; implicitly defined these days as any radio galaxy that is not an FRII). No robust models exist yet for the dynamics and development of FRI sources. The inner jets in some FRI's have been modelled as turbulent, entraining flows, but that model does not address the variety of FRI morphologies, nor does it describe the history and growth of such a source. More work is needed before we can understand the dynamics of FRI sources. The second question is how Cen A keeps shining long past the radiative lifetimes of its radio-loud and $\\g$-ray loud electrons. In this paper I discuss three alternative dynamical models which may describe Cen A. All of these models require the source to be quite old, with an age between several hundred Myr to more than a Gyr. This age range is much longer than the electrons' radiative lifetimes, which range from a few to a few tens of Myr. It follows that the radiative lifetimes are not a good indicator of the age of the source; in situ reacceleration (ISR) must be ongoing within the large radio lobes. Thus, Cen A provides a nearby case study of ISR in radio galaxies. To address these questions, I begin in Section \\ref{BasicStuff} by describing the setting and collecting important facts to be used in the analysis. In Section \\ref{InternalConditions} I review observational constraints on plasma and magnetic field within the outer lobes. I argue that inhomogeneous models, containing both strong and weak fields, are physically likely and consistent with the data. In Section \\ref{ModelsThatDontWork} I show that the short radiative lifetimes require ISR; models fail which try to grow the source so quickly. In Section \\ref{ModelsThatWork} I present three alternative dynamic models -- buoyant, magnetic tower and flow-driven -- each of which requires the source age to be on the order of a Gyr. The buoyant model does not fare well when compared to the data but both the flow-driven and magnetically-driven models seem consistent with current observations. In Section \\ref{InSituReacceleration} I critique competing ISR models. I speculate that Alfven-wave ISR is operating throughout the lobes and probably maintains the $\\g$-ray-loud electrons, while the radio-loud electrons are energized by shock or reconnection ISR in high-field regions within the source. Finally in Section \\ref{Conclusions} I summarize my arguments and discuss what this exercise reveals about the astrophysics of Cen A. I relegate necessary but tedious details to the Appendices. \\begin{figure*}[htb] {\\center \\includegraphics[width=0.95\\columnwidth]{EilekFigure1.eps}} \\caption{The large-scale structure of Centaurus A, displayed with north to the right. Neither lobe shows any sign of limb brightening; the volume of each lobe appears to be filled with localized, radio-bright filaments and loops. Elliptical rings in the center of this image are not physical, but are part of the interferometer response (to the very strong central source) which has not been fully removed from the image. The inner lobes sit within and are obscured by these artifacts. Mosaiced 1.4-GHz image using ATCA and Parkes data, $60'' \\times 40''$ resolution (data are publicly available on the NASA Extragalactic Database; see also Feain \\etal 2011).} \\label{Fig:OuterLobes} \\end{figure*} ", "conclusions": "\\label{Conclusions} In this section I summarize my previous arguments in this paper, discuss which of the models I have proposed might best describe Cen A, and speculate on how these models connect to other evidence we have on the recent history of the galaxy. \\subsection{Summary I: tentative models for Cen A} In this paper I have shown that the radio galaxy Cen A is on the order of 1 Gyr old; substantially younger models violate observations. Because this is much older than the radiative lifetimes of the relativistic electrons seen in radio and $\\g$-rays, in situ reacceration (ISR) must be ongoing throughout the 300-kpc-scale radio lobes. I developed the argument in three parts, as follows. $\\bullet$ I showed (in Section \\ref{InternalConditions}, also Table \\ref{OL_Bfields}) that homogeneous, quasi-equipartion ($u_B \\sim u_e$) models are {\\it not} required by the data. Inhomogeneous models, in which high-B and low-B regions coexist, are also consistent with the data and probably closer to the real physics. The high-field regions could be shocks within the outer lobes or force-free regions within a large-scale magnetic structure. $\\bullet$ I introduced three toy models to describe the dynamical growth of Cen A (in Section \\ref{ModelsThatWork}, also Tables \\ref{Model_summary} and \\ref{Model_critique}). (1) The slowest-growing model assumes the outer ends of the radio lobes are driven by {\\it buoyancy}. (2) A second model assumes the radio lobes are magnetically dominated, with large-scale organized structures akin to {\\it magnetic towers}. (3) A third, faster-growing model assumes radio lobe growth is driven by {\\it internal plasma flow}. These models require Cen A to be at least several hundred Myr old, and possibly older than a Gyr. By comparison, the radiative lifetimes range from a few Myr for $\\g$-ray-loud electrons, to at most $\\sim 80$ Myr for radio-loud electrons; thus ISR is needed. $\\bullet$ I explored (in Section \\ref{InSituReacceleration}, also Table \\ref{Model_critique}) three competing ISR models relevant to Cen A. (1) {\\it Alfven wave turbulence} can explain the data if the radio lobe plasma is tenuous and highly variable in time and/or space. (2) {\\it Weak shocks} are an attractive ISR mechanism in the transonic, flow-driven model; they may account for most or all of the radio emission in that model. (3) {\\it Reconnection} is another attractive ISR mechanism, especially in the magnetically driven model, {\\it if} reconnection site physics is consistent with current thinking. \\subsection{Summary II: which models might work for Cen A?} Having introduced three dynamic models, and explored what ISR mechanism(s) can work within each one, I then compared the models to the data (in Section \\ref{Compare_dynamic_models} and Section \\ref{Compare_ISR_models}; also Table \\ref{Model_critique}). While the simplest form of the buoyant model had severe challenges, two models remain viable: flow driven and magnetically driven. \\begin{itemize} \\item {\\em Flow-driven models} are the closest to traditional models of FRII sources, and may provide a good description of many tailed FRI sources. Their two-part structure (channel flow plus surrounding cocoon) gives a natural explanation for the transverse spectral structure seen in Cen A. If the channel flow is transonic, the internal shocks it drives are an attractive way to explain the radio filaments and radio spectrum of Cen A. \\item {\\em Magnetically driven models}, while less popular in the radio galaxy literature, are an intriguing alternative to flow-driven models. The large-scale, self-organized structures inherent in these models are a tempting explanation for the large structures and gradients seen in total and polarized radio emission. If these models describe Cen A, large-scale reconnection sites probably supplement Alfvenic ISR throughout the lobes. \\end{itemize} However, neither of these models has been developed to the point where it can confidently be said to describe Cen A. Until my suggested scenarios in this paper are supported by dedicated modelling and/or simulations, they remain no more than speculations. While the internal physics of competing ISR mechanisms clearly needs further work, the most immediate -- and addressible -- test of each model is under what conditions it can explain the large-scale structure of Cen A. Flow-driven models are perhaps most challenged by the lack of brightening at the outer ends of the lobes, where the channel flow should be decelerating and/or developing terminal shocks. I have speculated that lower Mach number flows will not suffer such problems, but -- because most work on flow-driven models has considered high Mach number flows -- it is not yet clear under what conditions, if at all, models driven by slower flows can avoid disruptive instabilities and create large-scale radio lobes resembling those in Cen A. Magnetically-driven models are perhaps most challenged by the lack of large, synchrotron-faint internal cavities as predicted by existing models. I have suggested that the radio lobes may contain more complex, self-organized magnetic structures, but this remains to be demonstrated. Work is especially needed to learn whether AGN-driven plasmas can access the necessary high-order magnetic modes in a soft-boundary system (such as a radio lobe in the IGM), and whether the plasma/field mix inside such systems is sufficiently radio-loud to create radio lobes resembling those in Cen A. \\subsection{Connect to the recent history of the galaxy} \\label{RecentHistory} To close, some possible connections between Cen A and recent events in the history of its parent galaxy deserve mention. \\paragraph{Very recent history: disrupted jets?} Although active jets from the AGN are feeding the kpc-scale inner radio lobes, no detectable jets connect the inner lobes to the outer lobes on either side of the AGN (Neff \\etal 2014). If ongoing energy transport requires a detectable jet, then both outer lobes are disconnected from their engine. Are they dying? If so, how quickly will they fade from view? Because the radio- and $\\g$-ray-loud electrons require frequent ISR in order to keep shining, we must ask, on what timescale the ISR will cease if energy flow to the radio lobes is disrupted. Turbulence is the common denominator in all three ISR mechanisms. It is clearly required for Alfven-wave ISR, and it is implicated in our two alternative ISR models. Thus, the lobes will fade when the turbulence decays. We know that nondriven fluid turbulence will decay in no more than several turnover times, $\\tau_t = \\lambda_t/v_t$ (where $\\lambda_t$ and $v_t$ are the characteristic scale and velocity of the turbulence. We can estimate $\\lambda_t \\sim \\lambda_{max} \\sim 30$ kpc from the images, but we have no direct probe of $v_t$. The flow-driven model allows a possible {\\it ansatz}: if the field is maintained by a turbulent dynamo, $B^2 / 4 \\pi \\ltw \\rho v_t^2$, and thus $v_t \\gtw v_A$. Alternatively, with the magnetically driven model we can directly take $v_A$ as the turbulent turnover speed. Thus, our limit $v_A \\gtw 4000$ km/s (from Section \\ref{AlfvenISR}) suggests $\\tau_t \\ltw 8$ Myr. We therefore expect turbulence in the outer lobes to decay in no more than $\\sim 30$ Myr. This is longer than the lifetime of the $\\g$-ray-loud electrons, and comparable to the lifetime of the radio-loud electrons; the lobes will fade quickly if the turbulence stops being driven. It follows that, if the outer lobes are disconnected from the AGN right now, this situation cannot have lasted long. Whatever disrupted the energy flow from the AGN must have happened no more than a few tens of Myr ago. Did the AGN go through a quiescent period, for the past few tens of Myr, and then revive itself only a few Myr ago? If so, was this part of a regular duty cycle, or is it evidence of an unusual disruptive event in the galaxy's recent history? \\paragraph{Previous history: did a merger launch the radio galaxy?} We have seen that Cen A is on the order of a Gyr old. It is interesting to compare this age range to the dynamic history of NGC 5128. The evidence suggests that at least two significant events -- encounters or megers -- have happened to this galaxy in the past few Gyr. One line of evidence comes from the dramatic gas/dust disk in the core of NGC5128. Because it is warped and twisted we must be seeing it fairly soon after its formation, as it settles into a stable state within the galaxy's gravitational potential. The exact rate at which the disk settles depends on details of the galaxy's structure. If the galaxy is prolate, models suggest the disk is only $\\sim 200$ Myr old (Quillen \\etal 1993); if the galaxy is oblate, the disk is $\\sim 700$ Myr old (Sparke 1996). If, as generally assumed, the disk has been left behind by a recent merger, that event happened less than a Gyr ago. Another line of evidence comes from the galaxy's stars; NGC5128 has a complex star formation history. While the bulk of the stars are at least $10$ Gyr old, recent work has found a younger population, which may be as young as $\\sim 2\\!-\\!3$ Gyr (\\eg, integrated light spectroscopy of globular clusters, Woodley \\etal 2010; resolved studies of halo stars, Rejkuba \\etal 2011). If this recent burst of star formation was triggered by an encounter or merger, that event happened $\\sim 2\\!-\\!3$ Gyr ago. While we do not know what initiates an activity cycle in an AGN, the ages of these two recent events are intriguing. The (relatively young) age of the disk and the (somewhat older) age of the young stellar population in NGC5128 bracket the ages that the dynamic models predict for the outer lobes. Did one of these recent encounters also trigger AGN activity and create the radio galaxy Cen A? \\ack It is a pleasure to thank Susan Neff, Frazer Owen, Sarka Wykes, Hui Li and Martin Hardcastle for many illuminating conversations about radio galaxy physics in general, and Cen A in particular. Insightful comments from the referees contributed significantly to this paper. Norbert Junkes kindly provided his Parkes image of Cen A. This research made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities Inc. \\begin{appendix}" }, "1402/1402.6449_arXiv.txt": { "abstract": "We construct a self-interacting scalar dark matter (DM) model with local discrete $Z_{3}$ symmetry that stabilizes a weak scale scalar dark matter $X$. The model assumes a hidden sector with a local $U(1)_X$ dark gauge symmetry, which is broken spontaneously into $Z_3$ subgroup by nonzero VEV of dark Higgs field $\\phi_X$ ($ \\langle \\phi_X \\rangle \\neq 0$). Compared with global $Z_3$ DM models, the local $Z_3$ model has two new extra fields: a dark gauge field $Z^{'}$ and a dark Higgs field $\\phi$ (a remnant of the $U(1)_X$ breaking). After imposing various constraints including the upper bounds on the spin-independent direct detection cross section and thermal relic density, we find that the scalar DM with mass less than $125$ GeV is allowed in the local $Z_3$ model, in contrary to the global $Z_3$ model. This is due to new channels in the DM pair annihilations open into $Z^{'}$ and $\\phi$ in the local $Z_3$ model. Most parts of the newly open DM mass region can be probed by XENON1T and other similar future experiments. Also if $\\phi$ is light enough (a few MeV $\\lesssim m_\\phi \\lesssim$ O(100) MeV), it can generate a right size of DM self-interaction and explain the astrophysical small scale structure anomalies. This would lead to exotic decays of Higgs boson into a pair of dark Higgs bosons, which could be tested at LHC and ILC. ", "introduction": "Although Planck \\cite{Ade:2013zuv} has already given the dark matter(DM) relic density $\\Omega h^2=0.1199\\pm 0.0027$ with a high precision, we still do not know particle physics nature of DM at all. So far all the compelling evidences for the existence of DM come from astrophysics and cosmology, due to its gravitational interaction. Still, many particle physics models for DMs have been proposed, and most of them have a stable collisionless cold DM(CCDM) candidate whose self-interaction can be ignored. The collisionless cold DM has been very successful when explaining the large scale structure of our Universe. However, anomalies from the small scale astrophysical observations~\\cite{Oh:2010ea,BoylanKolchin:2011de,BoylanKolchin:2011dk} indicate that DM may have strong interactions between themselves. Such self-interaction~\\cite{Spergel:1999mh} would make DM have a flat core density profile rather than a cusp one predicted by CCDM. Recent simulations show that in order to flatten the cores of galaxies the cross section for DM scattering should be around $\\sigma\\sim M_X \\times \\textrm{ barn }\\GeV^{-1}$~\\cite{Vogelsberger:2012ku, Rocha:2012jg, Zavala:2012us}, which is in fact a huge cross section compared with typical weak-scale cross sections $\\sigma\\sim 10^{-12}$ barn or $1$ pb. Some light particle mediator in the dark sector could be an origin of such strong self-interaction between DMs. In this paper, we propose a scalar DM model with a local $Z_3$ symmetry. Unlike models based on global symmetries, local discrete symmetries can protect symmetry-breaking from quantum gravity effects and guarantee the longevity or absolute stability of DM particles. Also a light mediator can exist in the models with local symmetry, and generate the correct self-interaction for DM in explaining the anomalies mentioned in the previous paragraph. The outline of this paper is as follows. In Sec.~II, we introduce the model with a local $Z_3$ symmetry, establish the convention for parameters and give the physical mass spectra. Then we discuss both theoretical and experimental constraints on the parameters in Sec.~III. Then in Sec.~IV, we discuss the relic density and DM direct searches, paying attentions to the semi-annihilation feature, and compare with the global $Z_3$ mode. In Sec.~V, we show that a light scalar mediator in our model can induce strong interaction for DM. Finally we summarize the results in Sec.~VI. ", "conclusions": "In this paper, we have proposed a self-interacting scalar DM model with a local dark $Z_3$ symmetry. Unlike global dark symmetries, local ones can guarantee that DM is absolutely stable even in the presence of higher dimensional nonrenormalizable operators due to the underlying local gauge symmetry. Then we discussed perturbativity constraints on the scalar potential and the experimental limit on the kinetic mixing. Compared with a global $Z_3$ model, our scenario has two new particles, $Z^{'}$ and $H_2$, and there are new channels in the DM pair annihilations for thermalizing DMs. Therefore much ampler parameter space is allowed including a light DM with $M_X<125$ GeV, most region of which can be probed with future DM direct searches. Also, motivated by the small scale astrophysical anomalies, we investigated the phenomenology of a MeV scalar $H_2$ in our model which has no counterpart in the minimal global $Z_3$ model. Thanks to the velocity dependence of DM self-interaction cross section, such a light $H_2$ can mediate strong interaction for DM scattering at Dwarf galaxy scale while satisfying Milky Way and cluster scale constraints. Similar arguments go for the light $Z^{'}$ as well. For such a light $H_2$ or $Z^{'}$, there could be exotic decays of the 126 GeV Higgs boson, which could be studied in the upcoming LHC running and at future lepton colliders." }, "1402/1402.0868_arXiv.txt": { "abstract": "{} { We aim to investigate cool-core and non-cool-core properties of galaxy groups through X-ray data and compare them to the AGN radio output to understand the network of intracluster medium (ICM) cooling and feedback by supermassive black holes. We also aim to investigate the brightest cluster galaxies (BCGs) to see how they are affected by cooling and heating processes, and compare the properties of groups to those of clusters.} {Using Chandra data for a sample of 26 galaxy groups, we constrained the central cooling times (CCTs) of the ICM and classified the groups as strong cool-core (SCC), weak cool-core (WCC), and non-cool-core (NCC) based on their CCTs. The total radio luminosity of the BCG was obtained using radio catalogue data and/or literature, which in turn was compared to the cooling time of the ICM to understand the link between gas cooling and radio output. We determined K-band luminosities of the BCG with 2MASS data, and used a scaling relation to constrain the masses of the supermassive black holes, which were then compared to the radio output. We also tested for correlations between the BCG luminosity and the overall X-ray luminosity and mass of the group. The results obtained for the group sample were also compared to previous results for clusters.} {The observed cool-core/non-cool-core fractions for groups are comparable to those of clusters. However, notable differences are seen: 1)~for clusters, all SCCs have a central temperature drop, but for groups this is not the case as some have centrally rising temperature profiles despite very short cooling times; 2)~while for the cluster sample, all SCC clusters have a central radio source as opposed to only 45\\% of the NCCs, for the group sample, \\textit{all} NCC groups have a central radio source as opposed to 77\\% of the SCC groups; 3)~for clusters, there are indications of an anticorrelation trend between radio luminosity and CCT. However, for groups this trend is absent; 4)~the Indication of a trend of radio luminosity with black hole mass observed in SCC clusters is absent for groups; and 5)~similarly, the strong correlation observed between the BCG luminosity and the cluster X-ray luminosity/cluster mass weakens significantly for groups.} {We conclude that there are important differences between clusters and groups within the ICM cooling/AGN feedback paradigm and speculate that more gas is fueling star formation in groups, than in clusters where much of the gas is thought to feed the central AGN.} \\titlerunning{ICM Cooling, AGN Feedback and BCG properties of Galaxy Groups} \\authorrunning{V.~Bharadwaj et al.} ", "introduction": "The central entropy is another parameter often used to classify CC/NCC clusters (e.g.~\\citealt{ 2008ApJ...687..899R}) and displays a tight correlation with the CCT (\\citealt{ 2010A&A...513A..37H}). This is not surprising since the cooling time, $t_{\\mathrm{cool}}$, is related to the gas entropy, $K$, through the relation $t_{\\mathrm{cool}} \\propto K^{3/2}/T$ for pure Bremsstrahlung. We also show a histogram of the distribution of the central entropy in Fig.~\\ref{fig:Khist}. The entropy values are given in Table~\\ref{CCT}. The plot of the CCT and $K_{0}$ values is shown below in Fig.~\\ref{fig:entropy}. As expected, we see an excellent correlation between the two quantities with a Spearman rank correlation coefficient of 0.99, 0.99, and 0.96 for all, SCC, and non-SCC groups respectively. A CCT of 1 Gyr corresponds to $\\approx~20~\\mathrm{keV~cm}^{2}$. \\begin{figure} \\includegraphics[trim = 30mm 0mm 22mm 8mm, clip, scale=0.33]{./Entropyhistgraylogspace_3107.eps} \\caption{Histogram of the central entropy. Light grey represents SCC groups, medium grey represents WCC groups, dark grey represents NCC groups.} \\label{fig:Khist} \\end{figure} \\begin{figure*} \\centering \\includegraphics[scale=0.50,angle=-90]{Ktcool_3007.eps} \\caption{$K_{0}$ vs. CCT. Magenta is the best fit for all groups, blue for only SCC groups and red for WCC+NCC groups.} \\label{fig:entropy} \\end{figure*} \\subsection{Radio properties} \\subsubsection{CRS fractions-CC/NCC dichotomy}\\label{CRSfract} Table~\\ref{CCT} also lists whether or not there is a central radio source present in a group. We see that while \\textit{all} NCC groups have a CRS, the fractions of SCCs and WCCs containing a CRS are 77\\% and 57\\%, respectively. The overall CRS fraction for CC groups is 70\\%. Figure~\\ref{fig:fraction} shows the CRS fractions in the group sample. \\subsubsection{Total radio luminosity vs.~central cooling time} \\cite{ 2009A&A...501..835M} show that there is an anti-correlation trend between the CCT and the total radio luminosity for CC clusters (correlation coefficient of $-0.63$), which breaks down for cooling times shorter than 1~Gyr. Figure~\\ref{fig:Rtot} shows the same plot for groups. We do not find indications of a trend between the two quantities. Here we show the best fit obtained for the CC clusters from \\cite{ 2009A&A...501..835M} to highlight the difference between clusters and groups. The power-law fit for the CC clusters from \\cite{ 2009A&A...501..835M} is given by: \\begin{equation} L_{\\mathrm{tot}} = (0.041\\pm0.016)\\times (t_{\\mathrm{cool}})^{-3.16\\pm0.38}~. \\end{equation} It is interesting to note that \\textit{all} SCC groups and most of the WCC groups show a much lower radio output than the best fit for clusters. This was first alluded to in \\cite{ 2009A&A...501..835M}, where the groups in that sample (all SCC) were clear outliers and we confirm this for the first time with a large sample of groups. We discuss this in detail in Sect.~\\ref{AGNact}. \\begin{figure*} \\includegraphics[scale=0.35,angle=-90]{tcoolRtot_3007.eps} \\includegraphics[scale=0.35,angle=-90]{SMBHradfinal_3007.eps} \\caption{Left: Total radio luminosity vs. CCT with best-fit line for \\textit{clusters} from \\cite{ 2009A&A...501..835M}. Right: Total radio luminosity vs. mass of SMBH. The blue asterisks represent SCC groups, the green triangles represent WCC groups and the red circles represent NCC groups.} \\label{fig:Rtot} \\end{figure*} \\subsubsection{Total radio luminosity vs.~SMBH mass} The total radio luminosity shows no trend with the SMBH mass (Fig.~\\ref{fig:Rtot}). Classifying the sample as SCC, WCC, and NCC also does not yield any discernible correlations. This is in contrast with the HIFLUGCS sample, which shows a weak correlation for the SCC clusters (correlation coefficient of 0.46). We calculate a correlation coefficient of 0.29 for all groups and 0.20 for only SCC groups. In Table~\\ref{meanSMBHmassLrad} we present the mean SMBH masses and radio luminosities for the different CC types along with their standard errors to investigate whether different CC groups have systematically different masses and/or luminosities. We observe that the NCC groups have a systematically higher SMBH mass and radio luminosity than the SCC and WCC groups. \\begin{table*} \\begin{center} \\caption{Mean SMBH masses and radio luminosities for different CC groups.}\\label{meanSMBHmassLrad} \\begin{tabular}{|c|c|c|} \\hline Group & Mean SMBH mass ($10^{9}$ solar masses) & Mean radio luminosity ($10^{42}$ erg/s) \\\\ \\hline SCC &1.17$\\pm$0.18&0.0050$\\pm$0.0023\\\\ WCC &1.39$\\pm$0.24&0.0034$\\pm$0.0028\\\\ NCC &1.98$\\pm$0.27&0.24$\\pm$0.15\\\\ \\hline \\end{tabular} \\end{center} \\end{table*} \\subsection{BCG properties} We present the scaling relation between the BCG and the cluster/group X-ray luminosity and mass in Figs.~\\ref{fig:LxLbcg} and \\ref{fig:M500Lbcg}. The fits shown here are for a combined relation for groups and clusters. The details are summarized in Table~\\ref{scat}. The derivation of the scatter is explained in Appendix~\\ref{Scatcalc}. We observe that most group BCGs lie above the best-fit relations. Additionally, extending the scaling relations from clusters to groups leads to a higher intrinsic scatter in most cases. \\subsubsection{BCG luminosity vs.~X-ray luminosity} Figure~\\ref{fig:LxLbcg} shows the relation between the two quantities for both the HIFLUGCS clusters and the groups, with fits for all the members, SCC members, and non-SCC members. The correlation coefficients are 0.65, 0.79, and 0.46 for all, SCC and non-SCC systems, respectively. It is seen that most group BCGs have a higher luminosity than expected from the derived scaling relation. The best-fit power-law relation is given by \\begin{equation} \\left ( \\frac{L_{\\mathrm{BCG}}}{10^{11}L_{\\odot}} \\right ) = c\\times \\left ( \\frac{L_{\\mathrm{X}}}{10^{44} \\mathrm{ergs/s}} \\right )^{m}~, \\end{equation} where $m = 0.34\\pm0.03$ and $c = 6.98\\pm0.16$ for all systems, $m = 0.32\\pm0.03$ and $c = 6.89\\pm0.20$ for SCC systems, and $m = 0.43\\pm0.09$ and $c = 7.01\\pm0.30$ for non-SCC systems. We observe that when the relations are extended to the group regime, there is no significant difference in slopes and normalisations for the subsets. \\begin{figure*} \\centering \\includegraphics[scale=0.50, angle=270]{LxLbcg_3007.eps} \\caption{BCG luminosity vs.~X-ray luminosity. The black line is the best fit for all systems, the blue line for only SCC systems, and the red line for non-SCC systems.} \\label{fig:LxLbcg} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[scale=0.50, angle=270]{M500Lbcg_3007.eps} \\caption{BCG luminosity vs.~$ M_{500}$. The black line is the best fit for all systems, the blue line for only SCC systems, and the red line for non-SCC systems.} \\label{fig:M500Lbcg} \\end{figure*} \\subsubsection{BCG luminosity vs.~\\texorpdfstring{$ M_{500}$}{M500}} Figure~\\ref{fig:M500Lbcg} shows that the luminosity of the BCG grows with the cluster/group mass. The correlation coefficients are 0.65, 0.86 and 0.44 for all, SCC and non-SCC systems respectively. The best fit powerlaw relation is given by: \\begin{equation} \\left ( \\frac{L_{\\mathrm{BCG}}}{10^{11}L_{\\odot}} \\right ) = c \\times \\left (\\frac{M_{500}}{10^{14} M_{\\odot}} \\right)^{m} \\end{equation} where $m = 0.49\\pm0.03$ and $c = 4.74\\pm0.12$ for all systems, $m = 0.52\\pm0.04$ and $c = 5.15\\pm0.14$ for SCC systems, and $m = 0.56\\pm0.08$ and $c = 4.17\\pm0.03$ for non-SCC systems. We see that the normalisations for the SCC systems are around 23\\% higher than those for non-SCC systems. The total mass $M_{500}$ was calculated from the virial temperature (taken from \\citealt{ 2011A&A...532A.133M} and \\citealt{ Eckmiller} for the clusters and groups, respectively) through the scaling relation given in \\cite{2001A&A...368..749F}: \\begin{equation} \\left ( \\frac{M_{500}}{10^{13} h^{-1}_{71} M_{\\odot}} \\right ) = (2.5\\pm0.2)\\left ( \\frac{kT_{\\mathrm{vir}}}{1 \\mathrm{keV}} \\right )^{1.676\\pm0.054}~. \\label{M500} \\end{equation} \\begin{table*} \\centering \\caption{Best-fit results and scatter for BCG-cluster scaling relations: ``stat'' refers to statistical scatter and ``int'' refers to intrinsic scatter.} \\begin{tabular}[]{| c c c c c c c |} \\hline \\textbf{Category} & \\textbf{Slope} & \\textbf{Normalisation} & $\\sigma_{\\mathrm{int},L_{\\mathrm{X}}}$ & $\\sigma_{\\mathrm{stat},L_{\\mathrm{X}}}$ & $\\sigma_{\\mathrm{int},L_{\\mathrm{BCG}}}$ & $\\sigma_{\\mathrm{stat},L_{\\mathrm{BCG}}}$ \\\\ \\hline $L_{\\mathrm{X}}-L_{\\mathrm{BCG}}$ Clusters & 0.36$\\pm$0.03 &4.54$\\pm$0.34 & 0.67 & 0.07 & 0.24 & 0.03 \\\\ $L_{\\mathrm{X}}-L_{\\mathrm{BCG}}$ Clusters+Groups &0.34$\\pm$0.03&6.98$\\pm$0.16& 0.85 & 0.08 & 0.21 & 0.03 \\\\ $L_{\\mathrm{X}}-L_{\\mathrm{BCG}}$ Clusters (SCC) &0.32$\\pm$0.03&5.15$\\pm$0.38& 0.64 & 0.08 & 0.21 & 0.02 \\\\ $L_{\\mathrm{X}}-L_{\\mathrm{BCG}}$ Clusters+Groups (SCC) &0.32$\\pm$0.03&6.89$\\pm$0.20& 0.57 & 0.08 & 0.18 & 0.02 \\\\ $L_{\\mathrm{X}}-L_{\\mathrm{BCG}}$ Clusters (NSCC) &0.50$\\pm$0.07&3.49$\\pm$5.09&0.61 & 0.06 & 0.21 & 0.03 \\\\ $L_{\\mathrm{X}}-L_{\\mathrm{BCG}}$ Clusters+Groups (NSCC) &0.43$\\pm$0.09&7.01$\\pm$0.30& 0.64 & 0.07 & 0.29 & 0.03 \\\\ \\hline &&&$\\sigma_{\\mathrm{int},M_{500}}$ & $\\sigma_{\\mathrm{stat},M_{500}}$ & $\\sigma_{\\mathrm{int},L_{\\mathrm{BCG}}}$ & $\\sigma_{\\mathrm{stat},L_{\\mathrm{BCG}}}$ \\\\ \\hline $M_{500}-L_{\\mathrm{BCG}}$ Clusters &0.62$\\pm$0.05&3.52$\\pm$0.28& 0.30 & 0.07 & 0.19 & 0.04 \\\\ $M_{500}-L_{\\mathrm{BCG}}$ Clusters+Groups &0.49$\\pm$0.03&4.74$\\pm$0.12& 0.40 & 0.07 & 0.20 & 0.03 \\\\ $M_{500}-L_{\\mathrm{BCG}}$ Clusters (SCC) &0.62$\\pm$0.01&4.30$\\pm$0.29& 0.21 & 0.07 & 0.13 & 0.04 \\\\ $M_{500}-L_{\\mathrm{BCG}}$ Clusters+Groups (SCC) &0.52$\\pm$0.04&5.15$\\pm$0.14& 0.31 & 0.06 & 0.16 & 0.03 \\\\ $M_{500}-L_{\\mathrm{BCG}}$ Clusters (NSCC) &0.75$\\pm$0.09&2.55$\\pm$0.36& 0.31 & 0.07 & 0.23 & 0.05 \\\\ $M_{500}-L_{\\mathrm{BCG}}$ Clusters+Groups (NSCC) &0.56$\\pm$0.08&4.17$\\pm$0.03& 0.53 & 0.07 & 0.29 & 0.03 \\\\ \\hline \\end{tabular} \\label{scat} \\end{table*} ", "conclusions": "\\label{Summary} With a sample of 26 Chandra galaxy groups we have peformed a study of the ICM cooling, AGN feedback and the BCG properties on the galaxy group scale. The major results of our study are as follows. \\begin{itemize} \\item The group sample has similar SCC, WCC, and NCC fractions to the HIFLUGCS cluster sample. \\item We find that 23\\% of the groups that have CCT $\\leq$ 1 Gyr do not show a central temperature drop. We speculate that this could be due to a partial reheating of the cool core in the past. Additionally, we also speculate that this might be a characteristic feature of fossil groups. \\item An increase in the CRS fraction with decreasing CCT is \\textit{not} seen, unlike for the HIFLUGCS sample. This is the first indication of differences between clusters and groups in the AGN heating/ICM cooling paradigm. \\item There is no correlation seen between the CCT and the integrated radio luminosity of the CRS. We notice that CRSs for the SCC groups, in particular, have a much lower radio luminosity than clusters. \\item We extend the scaling relations between $L_{\\mathrm{BCG}} $ and global cluster properties ($L_{\\mathrm{X}} $ and $M_{500}$) into the group regime. Most group BCGs have a BCG luminosity above the best fit and we think this may be due to a higher stellar mass content in group BCGs than in cluster BCGs, for a given $L_{\\mathrm{X}}$ and $M_{500}$. \\item We have speculated that star formation is a possible, effective answer to the fate of the cool gas in groups where, because of fueling star formation, not enough gas (low as it is) is being fed to the SMBH and hence the radio output of the CRSs is not as high. This could also explain why some SCC groups do not show a CRS. \\end{itemize} In conclusion, we have demonstrated differences between clusters and groups vis-a-vis ICM cooling, AGN feedback, and BCG properties. These results lend support to the idea that groups are not simply scaled-down versions of clusters. In the future, it would be interesting to study the impact of these processes on scaling relations for galaxy groups." }, "1402/1402.4269.txt": { "abstract": "{We extend the resummation method of Anselmi \\& Pietroni (2012) to compute the total density power spectrum in models of quintessence characterized by a vanishing speed of sound. For standard $\\Lambda$CDM cosmologies, this resummation scheme allows predictions with an accuracy at the few percent level beyond the range of scales where acoustic oscillations are present, therefore comparable to other, common numerical tools. In addition, our theoretical approach indicates an approximate but valuable and simple relation between the power spectra for standard quintessence models and models where scalar field perturbations appear at all scales. This, in turn, provides an educated guess for the prediction of nonlinear growth in models with generic speed of sound, particularly valuable since no numerical results are yet available.} ", "introduction": "Ever since the discovery of the cosmic acceleration found by the observations of type Ia supernovae in 1998 \\cite{RiessEtal1998, PerlmutterEtal1999}, a tremendous amount of observational and theoretical effort has been put in place in order to understand the cause of this phenomenon. A plethora of alternative models attribute the observed acceleration to either the presence of a mysterious energy component dubbed dark energy (DE) which dominates at the current epoch, or to a modification of the gravitational laws on cosmological distances (see, e.g. \\cite{CopelandSamiTsujikawa2006} for an overview of proposed theoretical models). While the standard $\\Lambda$CDM model, where the acceleration is driven by a cosmological constant $\\Lambda$, provides a good fit to observations, several models cannot, so far, be ruled-out. The next generation of large-scale structure (LSS) surveys such as EUCLID \\cite{LaureijsEtal2011} and DESI\\footnote{\\href{http://desi.lbl.gov/}{http://desi.lbl.gov/}} will provide an unprecedented determination of observables such as the weak lensing shear and galaxy power spectra over a very large range of scales, and will be therefore sensitive to the details of structure formation, representing a unique opportunity to discriminate between competing models. However, the improvement in the quality and quantity of observational data represents a challenge to the accuracy of theoretical predictions, particularly for non-standard models. A fundamental complication, in this respect, is given by the nonlinear nature of structure formation. Indeed many cosmological probes of matter density correlations will present small statistical errors on scales smaller than $\\sim 100\\Mpc$, where gravitational instability is responsible for relevant nonlinear corrections to linear theory predictions. The solution to the problem entails two goals: to understand and predict these corrections on one hand and, on the other, to achive a fast way (seconds) to compute numerically such predictions, in order to efficiently sample the theory parameter space in the data analysis. Our current understanding of the nonlinear regime, rather limited even for the standard paradigm, is very poor for non--standard cosmological models. In the following we will present an important step in this direction for the case of models with perturbations in the quintessence field. Quintessence is one of the most popular models of DE, where the acceleration of the Universe is attributed to a scalar field with negative pressure \\cite{ZlatevWangSteinhardt1999}. Its standard formulation involves a minimally coupled canonical scalar field, whose fluctuations propagate at the speed of light $c_s=1$. In this case, sound waves maintain the quintessence homogeneous on scales smaller than the Hubble radius $H^{-1}$ \\cite{FerreiraJoyce1997}. The quintessence contribution to the total energy density affects the expansion of the Universe and the growth of dark matter perturbations, allowing observational constraints on the DE equation of state $w(t)=p_Q(t)/\\rho_Q(t)$, where $p_Q$ and $\\rho_Q$ are the pressure and energy density of the scalar field. In this set-up, quintessence clustering effects take place only on scales of order $H^{-1}$, and their detection is therefore severely limited by cosmic variance. However, this may not be the case for more general quintessence models where the scalar field kinetic term is non-canonical\\footnote{ The name ``quintessence'' is usually reserved for theories where the dark energy is given by a scalar field with a canonical kinetic term, however, following previous works on the subject, we will refer as quintessence to a generic minimally coupled scalar field, with either canonical or non-canonical kinetic term.}. In fact, in \\cite{CreminelliEtal2006B,CreminelliEtal2009}, following the approach of \\cite{CheungEtal2008}, the authors constructed the most general effective field theory describing quintessence perturbations around a given Friedman-Robertson-Walker background space-time, showing the existence of theoretically consistent models where the fluctuations propagate at a speed $c_s\\neq 1$. In particular, by a careful analysis of the theoretical consistency of the models, namely, the classical stability of the perturbations and the absence of ghosts, ref.~\\cite{CreminelliEtal2009} concluded that the region of parameter space with $w<-1$ (the so- called ``phantom regime'') is in general excluded, unless the fluctuations are characterized by a practically vanishing, imaginary speed of sound, e.g. $|c_s|\\lesssim 10^{-15}$ (note that observational limits do not exclude the case $w<-1$ \\cite{Planck2013parameters,RestEtal2013}). In the latter case, the stability of the perturbations would be guaranteed by the presence of higher derivative operators with negligible effects on cosmological scales. As was emphasized in \\cite{CreminelliEtal2006B,CreminelliEtal2009}, such a tiny speed of sound should not be considered as a fine tuning, since in the limit $c_s\\to 0$ one recovers the ghost condensate model of ref.~\\cite{ArkaniHamedEtal2004A}, which is invariant under a shift symmetry of the field, and hence a value of $c_s\\simeq 0$ can be interpreted as a deformation of the limit where such symmetry is recovered. What makes these models with $c_s\\simeq 0$ of particular interest is that quintessence perturbations can cluster on all observable scales, opening up to a rich phenomenology. In fact, for such models the quintessence field is expected to affect not only the growth history of dark matter through the background evolution but also by actively participating to structure formation, both at the linear and nonlinear level. In the linear regime, the observational consequences of clustering quintessence have been investigated by several authors: see for instance \\cite{WellerLewis2003, BeanDore2004, DeDeoCaldwellSteinhardt2003, Hannestad2005, EricksonEtal2002, DePutterHutererLinder2010, SaponeKunz2009} for large scale CMB anisotropies, \\cite{Takada2006} for galaxy redshift surveys, \\cite{TorresRodriguezCress2007} for neutral hydrogen observations or \\cite{HuScranton2004, CorasanitiGiannantonioMelchiorri2005} for the cross-correlation of the Integrated Sachs-Wolfe effect in the CMB and the LSS. However, the possibility to detect quintessence perturbations effects focusing on observations in the linear regime alone is limited by large degeneracies with cosmological parameters, particularly when quintessence clusters on all relevant scales. In this case, in fact, such effects consist essentially in a change of the fluctuations amplitude. On the other hand the relation between the linear {\\em and} nonlinear growth of density perturbations can be affected by clustering quintessence in a peculiar way, leading to specific signatures that could in principle allow to distinguish these models from $\\Lambda$CDM and homogeneous quintessence cosmologies \\cite{SefusattiVernizzi2011, DAmicoSefusatti2011, AnselmiBallesterosPietroni2011}. Robust predictions of the nonlinear dynamics of the structure formation process ultimately require numerical simulations. The case of homogeneous quintessence (i.e., $c_s=1$), as it affects only the background evolution has been investigated by means of N-body simulations in several articles (see for instance \\cite{Baldi2012, KuhlenVogelsbergerAngulo2012} and references therein). The case of clustering quintessence, however, represents a much more difficult problem and numerical results are still lacking. The extension of common fitting formulas such as \\texttt{halofit} \\cite{SmithEtal2003, TakahashiEtal2012, Zhao2013}, based in turn on numerical simulations, is also, {\\em a priori} not possible. Precisely for this reason, the study of analytical approximations in cosmological perturbation theory (PT) to investigate the mildly nonlinear regime is particularly relevant. On one hand it can give insights in the physical processes at play. One of the main results of this work, in fact, consists in showing how fitting functions valid for tested, standard cosmologies can be used in clustering quintessence models. Such extensions have been already considered in the literature without a proper theoretical justification \\cite{SaponeKunz2009, AmendolaKunzSapone2008, SaponeKunzAmendola2010}. On the other hand, PT predictions can as well provide useful checks for future numerical results, which are likely to comprise significant approximations in their description of the dynamics of quintessence perturbations. The nonlinear regime of structure formation in quintessence models with $c_s=0$ has been studied in previous literature. Several works focussed on the spherical collapse of structures in the presence of quintessence perturbations \\cite{MotaVanDeBruck2004, NunesMota2006, AbramoEtal2007}. In addition, ref.~\\cite{CreminelliEtal2010} makes use of the spherical collapse model \\cite{GunnGott1972} (see \\cite{BasseEggersBjaeldeWong2011, BatistaPace2013} for the case of arbitrary sound speed) along with the Press-Schechter formalism \\cite{PressSchechter1974}, to provide predictions for the halo mass function. We will consider here, instead, analytical predictions in cosmological perturbation theory for the mildly nonlinear regime of the density power spectrum (see \\cite{BernardeauEtal2002} for a review of standard PT). In the context of standard, $\\Lambda$CDM cosmologies, perturbative techniques have witnessed a significant progress, motivated by the need for accurate predictions for Baryon Acoustic Oscillations measurements, following the seminal papers of Crocce \\& Scoccimarro \\cite{CrocceScoccimarro2006A, CrocceScoccimarro2006B, CrocceScoccimarro2008}. Several different \u00d2resummed\u00d3 perturbative schemes and approaches have been proposed in the last few years \\cite{MatarresePietroni2007, TaruyaHiramatsu2008, Pietroni2008, BernardeauCrocceScoccimarro2008, BernardeauVanDeRijtVernizzi2012, JurgensBartelmann2012, BlasGarnyKonstandin2014}, along with public codes implementing some of the methods \\cite{CrocceScoccimarroBernardeau2012,TaruyaEtal2012}. The key improvement over standard PT consisted in reorganizing the perturbative series in a more efficient way by means of a new building block, the nonlinear propagator, a quantity measuring the response of the density and velocity perturbations to their initial conditions. The possibility to analytically compute this new object revolutionized indeed the cosmological perturbation theory approach. A first extension of standard PT to the specific case of clustering quintessence can be found in \\cite{SefusattiVernizzi2011}, where the pressure-less perfect fluid equations have been worked out, while in \\cite{DAmicoSefusatti2011, AnselmiBallesterosPietroni2011} it was shown how to compute the nonlinear power spectrum (PS) by means of the Time-Renormalization Group (TRG) method of \\cite{DAmicoSefusatti2011, AnselmiBallesterosPietroni2011}. However, the computation of the nonlinear propagator for these cosmologies was still lacking. This is the first achievement of the present work, allowing the extension of a large set of powerful resummed techniques to clustering Dark Energy models. We compute the nonlinear PS exploiting the resummation scheme proposed by Anselmi \\& Pietroni \\cite{AnselmiPietroni2012} (hereafter AP). Such scheme takes into account the relevant diagrams in the perturbative expansion and proposes a new interpolation procedure between small and large scales. What makes AP so appealing is the possibility to extend the predictions, for the first time, to scales beyond the Baryonic Acoustic Oscillation (BAO) range, together with a fast and easy code implementation. It is important to stress that the AP approach, as other perturbative techniques, is intrinsic limited by the emergence of multi-streaming effects at small scales, i.e. the generation of velocity dispersion due to orbit crossing. Clearly, the breaking-down of the pressureless perfect fluid (PPF) approximation employed as a starting point for the fluid equations has nothing to do with the perturbative scheme used. The AP prediction agrees with N-body simulations at the 2\\% level accuracy on BAO scales while at smaller scales shows discrepancies only of a few percents up to $k \\simeq 1\\kMpc$ at $z \\gtrsim 0.5$. As we expect, the level of the agreement is consistent with the findings of \\cite{PueblasScoccimarro2009} where the authors estimated the departures from the PPF assumption by means of high resolution N-body simulations. Clearly, it would be desirable to extend our results beyond the PPF description. However, so far, even in the standard scenario this is still subject of intense investigations (see \\cite{PueblasScoccimarro2009, ValageasNishimichi2011A, PietroniEtal2011, CarrascoHertzbergSenatore2012, PajerZaldarriaga2013, ValageasNishimichiTaruya2013} for an incomplete list of contributions). Taking advantage of the AP approach, we will show that additional nonlinear corrections due DE clustering become larger than $1\\%$ for scales smaller than BAO for 10\\% variation from $w=-1$. Thus it allows us to finally test the approximation employed in the TRG predictions of \\cite{AnselmiBallesterosPietroni2011} and the assumptions made in several works forecasting the detectability of dark energy models in presence of quintessence perturbations \\cite{SaponeKunz2009, AmendolaKunzSapone2008, SaponeKunzAmendola2010}. Furthermore, as a by-product, we provide a theoretically motivated mapping from the nonlinear matter power spectrum in {\\em smooth} quintessence models to the nonlinear, total density power spectrum in {\\em clustering} quintessence models that could possibly serve as a consistency check for numerical results once these will become available. This is not of small significance, given the challenge posed by accurate simulations of structure formation in this kind of models. This paper is organized as follows. In Section \\ref{EqMotionLinear} we review the equations of motion for the quintessence--dark matter fluid. In Section \\ref{dynamics} we recast the starting fluid equations in a compact form and make explicit the nonlinear behavior of perturbations. In Section \\ref{StatisticalObs}, in order to compute the nonlinear density power spectrum, we apply the selected resummation scheme to the clustering quintessence fluid. In Section \\ref{results} we present our results and a well motivated mapping from the smooth to the clustering quintessence PS. Section \\ref{discussion} is devoted to our conclusions and possible future developments. Some details about the resummation scheme and a comparison with recent numerical simulations for a $\\Lambda$CDM cosmology are included as Appendices. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "\\label{discussion} Regardless any proper theoretical motivation or degree of naturalness, quintessence models characterized by a vanishing speed of sound are extremely interesting under a phenomenological point of view. In this case, in fact, quintessence perturbations are present at all scales and, more importantly, in the single-stream approximation, they are comoving with perturbations in the matter component, i.e.~they share the same velocity field. This allows to easily extend sophisticated techniques in cosmological perturbation theory already developed for standard, $\\Lambda$CDM cosmologies, \\cite{SefusattiVernizzi2011}. At the same time, $c_s=0$ represents a limiting case of a much broader class of quintessence models, parametrized by the equation of state parameter $w$ (with allowed values close to $-1$) and the sound speed parameter $c_s$ taking values, in principle, in the relatively wide interval $0\\le c_s\\le1$. Needless to say, the possible detection of any departure from the standard values $w=-1$ and $c_s=1$ would be of great significance toward an understanding of the mechanism driving the observed acceleration of the Universe. Still, the theoretical description of the nonlinear evolution of quintessence perturbation poses a remarkable challenge. The difficulty of the task is highlighted by the fact that no numerical simulation for these quintessence models has been performed to date. Furthermore, it is easy to imagine that when they {\\em will} be performed they will likely involve relevant approximations. Precisely for this reason any analytical insight into the nonlinear behavior of a central quantity as the density power spectrum is greatly valuable. Approximate solutions for the power spectrum of (standard) models with $c_s=1$ and (extreme) models with a vanishing speed of sound will therefore mark the limits of the broader set of solutions for models with intermediate values of $c_s$ and provide boundaries for useful interpolations, both under a practical as under a theoretical point of view. With this goal in mind, we applied, in this paper, the resummation scheme proposed in \\cite{AnselmiPietroni2012} to quintessence models with $c_s=0$, providing predictions for the total (i.e. matter and quintessence) density power spectrum, in the standard single-fluid approximation. This approach extends the range of validity of perturbative solutions beyond the acoustic oscillations scales up to $k\\sim 0.6\\kMpc$ with an accuracy below the 2\\% level at redshifts $z\\ge 0.5$, therefore comparable to the accuracy delivered by the \\texttt{halofit} fitting formula \\cite{SmithEtal2003, TakahashiEtal2012} or the emulator approach of \\cite{HeitmannEtal2014}. In addition, our analytical investigation suggests an approximate but useful and simple mapping between the nonlinear power spectra of models with $c_s=1$ and models with $c_s=0$ when all other cosmological parameters are held at the same values. We show, in fact, that the ratio between these two quantities is very close to the ratio of the square of their respective linear growth factors, eq.~(\\ref{nlApp}). This is the likely consequence of the fact, already suggested in earlier works \\cite{SefusattiVernizzi2011, AnselmiBallesterosPietroni2011, DAmicoSefusatti2011}, that nonlinear growth is driven predominantly by the matter component. This is quite an interesting result as it provides, for the first time, a theoretical motivation for simple predictions of the density power spectrum that can make use of tools as \\texttt{halofit} or more in general results from existing, numerical simulations of more standard models. Waiting for a proper N-body description of clustering quintessence models, our findings will enable more robust assessments of the ability of future cosmological probes to constrain this class of models. We will address these issues in future work." }, "1402/1402.0603_arXiv.txt": { "abstract": "We report on spatially resolved X-ray spectroscopy of the north-eastern part of the mixed morphology supernova remnant (SNR) W28 with {\\it XMM-Newton}. The observed field of view includes a prominent and twisted shell emission forming the edge of this SNR as well as part of the center-filled X-ray emission brightening toward the south-west edge of the field of view. The shell region spectra are in general represented by an optically thin thermal plasma emission in collisional ionization equilibrium with a temperature of $\\sim$0.3~keV and a density of $\\sim$10~cm$^{-3}$, which is much higher than the density obtained for inner parts. In contrast, we detected no significant X-ray flux from one of the TeV $\\gamma$-ray peaks with an upper-limit flux of 2.1$\\times$10$^{-14}$ erg cm$^{-2}$ s$^{-1}$ in the 2--10~keV band. The large flux ratio of TeV to X-ray, larger than 16, and the spatial coincidence of the molecular cloud and the TeV $\\gamma$-ray emission site indicate that the TeV $\\gamma$-ray of W28 is $\\pi^{0}$-decay emission originating from collisions between accelerated protons and molecular cloud protons. Comparing the spectrum in the TeV band and the X-ray upper limit, we obtained a weak upper limit on the magnetic field strength $B\\lesssim$ 1500~$\\mu$G. ", "introduction": "Supernova remnants (SNRs) are one of the most promising acceleration sites of cosmic rays up to $\\sim10^{15.5}$~eV~(the knee energy). \\citet{1995Natur.378..255K} discovered synchrotron X--rays from the shell of SN 1006, indicating the existence of extremely high-energy electrons up to $\\sim$~TeV produced by the first-order Fermi acceleration. Following this discovery, the synchrotron X-ray emission has been discovered from a few more young shell-type SNRs, such as RX~J1713.7--3946 \\citep{1997PASJ...49L...7K}, RCW~86 \\citep{2000PASJ...52.1157B,2001ApJ...550..334B}, and RX~J0852.0$-$4622 \\citep{2000PASJ...52..887T}. On the other hand, TeV $\\gamma$--rays have also been detected from some SNRs \\citep{2004Natur.432...75A,2006A&A...448L..43A}. The radiation of TeV $\\gamma$-ray is explained by (1) Inverse-Compton scattering (IC) of cosmic microwave background photons by the same high energy electron giving rise to the X-ray synchrotron emission, (2) non-thermal bremsstrahlung by high energy elections or (3) the decay of neutral pions that are generated by collisions between high energy protons and dense interstellar matter. Most of the SNRs with such a X-ray and TeV $\\gamma$-ray evidence have been young with an age of less than several thousands of years. Recently, Fermi discovered GeV $\\gamma$-rays from several SNRs, such as W44 \\citep{2010Sci...327.1103A}, IC443 \\citep{2010ApJ...712..459A}, W51C \\citep{2009ApJ...706L...1A}, and so on. Interesting fact is that they are not only young but rather old SNRs compared with the TeV $\\gamma$-ray emitting SNRs. It can be due to the escape of high energy particles from the shocks \\citep{2011MNRAS.410.1577O,2011ApJ...731...87E,% 2012MNRAS.421..935L,2012A&A...541A.153T,2013MNRAS.429.1643N}. Most SNRs with GeV $\\gamma$-rays are interacting with molecular clouds, so it may be related with the escape. However, it is still unknown the detailed picture how particles escapes from the shocks. The SNR W28, locating at ($l$, $b$) = ($\\timeform{6D.4}$, $-\\timeform{0D.1}$), is an interesting target as a cosmic-ray accelerator from which both GeV and TeV $\\gamma$-rays were detected from the eastern edge of the radio shell~(\\cite{2009ApJS..183...46A,2008A&A...481..401A}). The diameter and the distance to W28 are 48 arcmin and 1.9~kpc, respectively~\\citep{2002AJ....124.2145V}. The age seems to be several times 10$^{4}$ year, which means the remnant is middle-aged~\\citep{2002ApJ...575..201R}. W28 is classified as ``mixed--morphology'' SNR, showing center-filled X-rays and a shell-like radio emission. The shell-like radio emission peaks at the northern and northeastern boundaries where interaction of the SNR matter with the molecular cloud is established~\\citep{1981ApJ...245..105W}. This interaction was revealed by a lot of OH maser spots \\citep{1994ApJ...424L.111F,1997ApJ...489..143C,2005ApJ...620..257H}, which are signposts of molecular interactions and the location of high density shocked gas ($n \\ge$10$^{4}$ cm$^{-3}$; \\cite{1999PASJ...51L...7A}). TeV $\\gamma$-ray emission was detected by H.E.S.S.~\\citep{2008A&A...481..401A} from the eastern edge of the radio shell. Recently, {\\it{Fermi}} and {\\it AGILE} also detected $\\gamma$-rays at 100~MeV $\\textless E \\textless$ 100~GeV ~\\citep{2010ApJ...718..348A,2010A&A...516L..11G} from the same region of the TeV emission. In contrast, the emission in the X-ray band had been believed to comprise of thermal radiation. \\citet{2002ApJ...575..201R} evaluated temperatures of 1.5~keV in the southwest, 0.56~keV in the northeast, and the central region requires two-temperature plasma with 0.6~keV and 1.8~keV. The long ionization timescales in the northeast and central region imply that the gas is close to the ionization equilibrium. Recently, Suzaku discovered that the thermal X-ray emission from the inner region is over-ionized, implying that the plasma may underwent the sudden rarefaction \\citep{2012PASJ...64...81S}. These facts imply that W28 is an ideal target to study the particle escape from the shock. We analyzed XMM-Newton archival data of the north--eastern part of W28, where the molecular clouds, OH maser spots, bright X-ray shells, GeV and TeV emission were detected. XMM-Newton has large effective area and high angular resolution. These characteristics enable us to carry out high quality spatially resolved spectroscopy. In \\S~2, we present the observation log and data reduction method. Imaging and spectral analyses are shown in \\S~3 and \\S~4, respectively. From one of the TeV $\\gamma$-ray emission region, we obtained only the upper limit on the X-ray flux. Discussions are made on the basis of these results in \\S~5 on the nature of the thermal and non-thermal components in multi-wavelength. Finally we summarize our results in \\S~6. ", "conclusions": "We analyzed the XMM-Newton data of the north-eastern part of the supernova remnant W28. The observed X-ray image showed the bright and twisted north-eastern shell and the inner emission region which is part of the center-filled emission brightening toward the south-west end of the field of view. The X-ray emission from north-eastern shell is found to reach collisional ionization equilibrium state and can be fitted well with a single temperature optically thin thermal emission model with $kT$ of $\\simeq$0.3~keV. From the emission measure and the apparent volume, the electron density is found to be as high as $\\simeq$10~cm$^{-3}$. Since a bunch of molecular cloud spatially coincides with the outer edge of a part of the shell, this high density is due to the collision of the plasma with the molecular cloud. In contrast, there is no significant X-ray emission from one of the TeV $\\gamma$-ray peaks. We only obtained the 90\\% upper limit flux of 2.1$\\times$ 10$^{-14}$ erg cm$^{-2}$ s$^{-1}$ in the 2--10~keV band, assuming a power-law spectrum with the same photon index as in the TeV $\\gamma$-ray (= 2.66). The spatial coincidence of the molecular cloud and the TeV $\\gamma$-ray emission site suggests that TeV $\\gamma$-ray is hadronic origin. We calculate the spectra of hadrons including pions, kaons, nucleons and so on produced through proton-proton scatterings and their daughter particles (gamma-rays, electrons, neutrinos and so on), and found that $\\pi^{0}$-decay emission is dominant in the TeV $\\gamma$-ray band. A weak upper limit on the magnetic field strength is obtained $B\\lesssim$ 1500$\\mu$G from the X-ray flux upper limit." }, "1402/1402.4436_arXiv.txt": { "abstract": "{We report the discovery of large--scale diffuse radio emission South-East of the galaxy cluster MACS\\,J0520.7-1328, detected through high sensitivity Giant Metrewave Radio Telescope 323~MHz observations. This emission is dominated by an elongated diffuse radio source and surrounded by other features of lower surface brightness. Patches of these faint sources are marginally detected in a 1.4 GHz image obtained through a re-analysis of archival NVSS data. Interestingly, the elongated radio source coincides with a previously unclassified extended X-ray source. We perform a multi--wavelength analysis based on archival infrared, optical and X-ray \\chandra data. We find that this source is a low--temperature ($\\sim$3.6 keV) cluster of galaxies, with indications of a disturbed dynamical state, located at a redshift that is consistent with the one of the main galaxy cluster MACS\\,J0520.7-132 (z=0.336). We suggest that the diffuse radio emission is associated with the non-thermal components in the intracluster and intergalactic medium in and around the newly detected cluster. We are planning deeper multi--wavelength and multi-frequency radio observations to accurately investigate the dynamical scenario of the two clusters and to address more precisely the nature of the complex radio emission. } ", "introduction": "\\begin{figure*} \\centering \\includegraphics[width=0.95\\textwidth]{fig1.pdf} \\caption{ {\\em Left panel:} 323~MHz GMRT full--resolution (11.8\\secpoint $\\times$ 7.8\\secpoint, p.a. 16.9\\degpoint) image of the 19\\minpoint$\\times$19\\minpoint area around the cluster M0520. Contours are spaced by a factor of two and start at $\\pm$3$\\sigma$=0.33 mJy/b (negative contours are dashed). The dashed circle indicates the central region of 1 Mpc radius around the cluster M\\,0520 . {\\em Right panel:} zoom in of the area around the diffuse elongated source; radio contours (same as in {\\em Left panel}) are overlaid on the DSS2 optical red image. The black circles show position of the three galaxies from DSS catalog, which are likely optical counterparts of radio sources S1 and S2. } \\label{fig:fig1} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=1.\\textwidth]{fig2.pdf} \\caption { {\\em Left panel}: 323~MHz GMRT low--resolution (21.1\\secpoint $\\times$ 17.2\\secpoint, p.a. 28.8\\degpoint) image of the same 19\\minpoint$\\times$19\\minpoint region around the cluster MACS\\,J0520 (same as in Fig. \\ref{fig:fig1}); black contours are spaced by a factor of two and start at $\\pm$3$\\sigma_{323\\rm MHz}$=0.82 mJy/b (negative contours are dashed). To highlight the position of discrete compact radio sources in the field, contours from the full--resolution image (see Fig. \\ref{fig:fig1}) are over--plotted in red, starting from $\\pm6\\sigma$=0.66 mJy/b. Dotted ellipses mark the location of diffuse emission, labelled D1 to D4. {\\em Right panel:} \\vla 1.4 GHz contours (in blue) of an image obtained from re--processed NVSS data (45.8\\secpoint $\\times$ 29.9\\secpoint, p.a. -70.0\\degpoint); contours start at $\\pm2\\sigma_{1400\\rm MHz}=0.5$ mJy/b and are spaced by a factor of 2 (blue dashed are negative). The contours are overlaid on a 323~MHz GMRT image of similar resolution (41.5\\secpoint$\\times$33.5\\secpoint, p.a. 21.3\\degpoint), shown in grey scale (the grey dashed contour corresponds to 3$\\sigma_{323\\rm MHz}$=1.65 mJy/b). } \\label{fig:fig_2} \\end{figure*} In the currently supported hierarchical scenario of structure formation, galaxy clusters assemble and evolve through mergers and accretion of smaller units of matter, at the intersection of cosmic filaments connecting clusters into the large--scale structure of the Universe. In addition to the most common methods for clusters detections -- based on the study of galaxy over-densities, lensing, extended X-ray emission due to thermal bremmstrahlung from the intracluster medium (ICM), and the sub-mm signal due to Inverse Compton scattering of CMB photons by hot intracluster electrons (SZ effect) -- observations in the radio band can also probe the existence of clusters and/or large--scale structures. Powerful high-redshift radio galaxies are known to be efficient tools for detecting galaxy clusters at high redshifts (z$>$1.5; \\eg \\citealt{galametz13}, and reference therein). At lower redshift, tailed radio galaxies are being extensively and successfully used for cluster searches. Since they are known to be associated with galaxy clusters, they can be used to trace the high density environments in the Universe (\\eg \\citealt{gv09}). Several previously unobserved clusters have been actually identified thanks to the detection of tailed sources (\\eg \\citealt{blanton03}; \\citealt{smolcic07}; \\citealt{gv09}; \\citealt{kantharia09}; \\citealt{mao2010}). \\\\ Beyond tailed radio galaxies, it is nowadays known that a fraction of merging clusters host diffuse Mpc--scale radio emission (such as radio halos and relics, that are named according with their location and observational properties; \\eg see \\citealt{ferrari08}). This synchrotron emission is unrelated to individual cluster radio galaxies and directly probes the existence of non-thermal components (relativistic particles and magnetic fields) mixed with the thermal ICM. To date, all these sources are found to be associated with dynamically disturbed / merging systems (\\eg \\citealt{cassano10a}; \\citealt{cassano13}; \\citealt{feretti12}, for a recent review). The detection of such sources can become -- in principle -- a powerful tool to trace previously unidentified clusters. \\\\ Upcoming deep and wide-field radio surveys with new radio interferometers like LOFAR (Low Frequency Array; \\eg \\citealt{vanhaarlem13}), ASKAP (Australian Square Kilometer Array Pathfinder; \\eg \\citealt{johnston08}), MeerKAT (\\eg \\citealt{booth12}) and MWA (Murchison Widefield Array; \\eg \\citealt{tingay2013}) will allow systematic searches for new galaxy clusters based on the detection of tailed radio galaxies and diffuse extended emission (see \\eg \\citealt{ensslin02}; \\citealt{cassano10b}). For these studies, LOFAR and MWA are particularly suitable, thanks to the very large field of view covered at low radio frequencies that makes them a powerful survey telescope (see \\ie the ongoing all-sky surveys LOFAR MSSS, \\citealt{heald13}, and MWA GLEAM, \\citealt{lss}, in preparation).\\\\ In this paper we present an example of a new cluster detection made possible thanks to deep low--frequency radio observations. We report the serendipitous discovery with the Giant Metrewave Radio Telescope (\\gmrt) at 323 MHz of new large--scale diffuse radio emission in the proximity of the cluster MACS\\,J0520.7-1328 (M\\,0520 hereinafter). \\\\ M\\,0520 is one of the X-ray brightest clusters of galaxies from the MAssive Cluster Survey (MACS, \\citealt{Ebeling10}). It is located at $z$=0.336 and has an X-ray luminosity of L$_{\\rm r500, [0.1-2.4 keV]} = 7.8 \\times 10^{44}$ erg/s \\citep{Mantz10}\\footnote{Corrected for our slightly different choice of cosmological parameters.}. M0520 is classified as a relatively relaxed system on the basis of its morphological properties, \\ie a good alignment between the galaxy and ICM distributions, and quite concentric X-ray surface brightness contours in its central 1.5 Mpc region \\citep[morphology code 2 in][]{Ebeling10}. This cluster is also found to be a lensing system, though no mass estimate has been derived from lensing studies \\citep{Horesh10}. M\\,0520 is identified in the Planck SZ cluster catalogue as PSZ1 G215.29-26.09, with a mass estimated from the SZ of M$_{500}\\sim$6.15$\\times$10$^{14}$ M$_{\\odot}$ \\citep{plk2013}. \\\\ For our analysis, we made use of our new \\gmrt\\ radio observations and of multi--wavelength data available from archives and the literature. The paper is presented as follows: in Sect. 2 we describe the new \\gmrt observations and data reduction, along with the analysis of the detected emission; in Sect. 3, archival \\chandra X-ray data analysis of the cluster is presented; in Sect. 4, we report optical/IR analysis based on available public catalogues. A multi--wavelength comparison is reported in Sect. 5, and the results are discussed in Sect. 6. \\\\ The adopted cosmology is $\\Lambda$CDM, with ${\\rm H}_0$=71 km ${\\rm s}^{-1} {\\rm Mpc}^{-1}$, $\\Omega_{\\rm m} = 0.27$, $\\Omega_{\\Lambda} = 0.73$. At the redshift of the cluster 1$^{\\prime}$ corresponds to $\\sim$290 kpc. ", "conclusions": "We have reported high sensitivity \\gmrt\\ radio observations at 323 MHz of the galaxy cluster M\\,0520, that allowed the detection of complex diffuse radio emission. The main feature of this emission is located at a projected distance of $\\sim$8' SE of the cluster centre and was found to coincide with an extended X-ray source seen in the archival \\chandra image. We have performed a multi wavelength analysis based on archival X-ray and optical/IR data, that allowed us to classify it as a galaxy cluster, namely 1WGA\\,0521. Our X-ray analysis allowed us to characterise this newly discovered cluster as a low--temperature, relatively small system. From the spectral fitting we derived a redshift that is consistent with the optical redshift of M\\,0520, z=0.336. However the uncertainties on our X-ray estimate (Sec. \\ref{sec:Xspec}, Table \\ref{tab:x}) do not allow us to firmly conclude that the two clusters lie at the same distance \\footnote{ Note that, at z$\\sim$0.336, the difference of 0.004 between the X-ray redshift estimates (see Table \\ref{tab:x}) corresponds to 25 Mpc in luminosity distance. If we also consider their uncertainties from spectral fitting, the distance between the two clusters could be much larger. }. Despite the fact that at z=0.336 their projected separation of $\\sim$2.3 Mpc would easily include the clusters virial radii, the \\chandra\\ image does not show evidence of a bridge of X-ray emission between M\\,0520 and 1WGA0521, that could be expected in case of interaction (though this may be also due to the low sensitivity of the exposure). \\\\ On the other hand, our optical/IR analysis supports the hypothesis that the two clusters lie at similar redshifts, since their galaxy populations lie in statistically similar regions of the colour magnitude diagram. Optical spectroscopic observations and deeper X-ray observations are needed to accurately determine the 1WGA\\,0521 redshift and to investigate the dynamical scenario of these two systems and possible ongoing physical interaction. \\\\ The optical iso-density map derived from candidate cluster members shows that 1WGA\\,0521 has an elongated morphology, suggesting a disturbed dynamical state. Conversely, M\\,0520 seems more relaxed (as found by \\citealt{Ebeling10}), with a more regular shape, except for its inner region where the galaxy distribution is slightly elongated. Our radio images at low resolution reveal a complex distribution of diffuse emission, that we have analysed by identifying four structures (namely D1, D2, D3 and D4, see Sect. \\ref{sec:radio}). The diffuse elongated source D1 is the brightest of these features. It is characterised by large--scale and low--surface--brightness, and it is coincident with the newly detected, probably unrelaxed cluster of galaxy 1WGA\\,0521. These properties are typical for cluster--scale diffuse radio sources. Due to its projected location in the central region of the cluster, the emission could be a cluster radio halo with an elongated morphology. Such shape may also suggest the source is a radio relic (though located closer to the cluster centre with respect to what typically observed). The emission in this case would be associated with the passage of a shock wave, that may have either accelerated particles from the thermal ICM, or have revived fossil radio plasma from a previous activity of a radio--loud AGN (that could be one of the two nearby radio galaxies S1 and S2). Following the definition by \\cite{kempner2004}, the source would be classified as \\textit{radio gischt} in the former scenario, or a \\textit{radio phoenix} in the latter. The relic could be associated with the cluster M\\,520, rather than to 1WGA\\,0521. This hypothesis seems less likely, due to its large distance from the cluster centre ($\\sim$ 2.3 Mpc), much larger than what is typically observed, and the fact that M\\,520 seems to be a relaxed system, while relics are always found to be associated to merging clusters. Another possibility that cannot be firmly excluded is that D1 is actually the tail of a tailed radio galaxy, whose nucleus may be identified with S1 or S2. Our present radio data, however, do not allow us to disentangle between these possibilities; spectral and polarisation information is needed to clearly classify this source. \\\\ Beyond the prominent source D1, additional diffuse emission on large--scale is located around it (D2, D3, D4). The two elongated structures D2 and D3 are difficult to classify. They have an even lower brightness, their shape is filamentary, and they are located in a region in between 1WGA0521 and M\\,0520. D4 is even harder to interpret, due to its roundish shape and its location in a region where our present data does not show any significant optical or X-ray emission. We suggest that these features might be related to non-thermal components in the faint inter-galactic medium surrounding 1WGA0521, that is not detected by the current shallow X-ray data due to sensitivity limits, but may be revealed through more extensive observations (\\eg \\citealt{plck2013} and references therein). \\\\ Multi--frequency high sensitivity radio observations are essential to properly classify all the features of the complex extended radio emission, and to understand their origin. Follow up observations are planned. \\\\ Regardless the nature of the diffuse radio emission and its possible relation to the cluster's dynamical state, the most important result of our analysis is that our detection of large--scale diffuse radio emission has led us to the discovery of a new massive galaxy cluster, which was not previously known. Indeed, despite the fact that M0520 is a system in well studied cluster samples \\citep{Ebeling10,Mantz10,Horesh10} and that our X-ray and optical/IR analyses are based on archival data, the existence of the neighbouring cluster 1WGA0521 has been ignored up to now. Although it is not uncommon that X-ray surveys miss clusters (see \\eg \\citealt{cagnoni01}), in this case the radio observations have been crucial to spot its existence. \\\\ Thanks to on-going and future deep all-sky radio surveys (e.g. LOFAR--Surveys, EMU--ASKAP, MeerKAT--MighTEE, MWA GLEAM), diffuse radio sources will become a powerful tool for the discovery of dynamically disturbed galaxy clusters and large--scale structure merging/accretion events." }, "1402/1402.6433_arXiv.txt": { "abstract": "Many dark energy models fail to pass the cosmic age test. In this paper, we investigate the cosmic age problem associated with nine extremely old Global Clusters (GCs) and the old quasar APM $08279+5255$ in the $R_h=ct$ Universe. The age data of these oldest GCs in M31 is acquired from the Beijing-Arizona-Taiwan-Connecticut system with up-to-date theoretical synthesis models. They have not been used to test the cosmic age problem in the $R_h=ct$ Universe in previous literature. By evaluating the age of the $R_h=ct$ Universe with the observational constraints from the type Ia supernovae and Hubble parameter, we find that the $R_h=ct$ Universe can accommodate five GCs and the quasar APM 08279+5255 at redshift $z=3.91$. But for other models, such as $\\Lambda$CDM, interacting dark energy model, generalized Chaplygin gas model and holographic dark energy model, can not accommodate all GCs and the quasar APM 08279+5255. It is worthwhile to note that the age estimates of some GCs are controversial. So, unlike other cosmological models, the $R_h=ct$ Universe can marginally solve the cosmic age problem, especially at high redshift. ", "introduction": "Many astronomical observations, such as type Ia supernovae (SNe Ia) \\cite{Riess,Perlmutter,Schmidt}, the cosmic microwave background (CMB) \\cite{Spergel,Komatsu,Planck}, gamma-ray bursts \\cite{Wang11a,Wang11b} and large-scale structure (LSS) \\cite{Tegmark}, indicate that the Universe is undergoing an accelerated expansion, which suggests that our universe may have an extra component like dark energy. The nature of dark energy is still unknown, but the simplest and most interesting candidate is the cosmological constant \\cite{Carroll}. This model can consist with most of astronomical observations. The latest observation gives that the present cosmic age is about $t_0=13.82$ Gyr in $\\Lambda$CDM model \\cite{Planck}, but it still suffer from the cosmic age problem \\cite{Yang,Wang}. The cosmic age problem is that some objects are older than the age of the universe at its redshift $z$. In previous literatures, many cosmological models have been tested by the old quasar APM $08279+5255$ with age $2.1\\pm0.3$ Gyr at $z=3.91$ \\cite{Friacas,Komossa}, such as the $\\Lambda$CDM \\cite{Friacas,Alcaniz2003a}, $\\Lambda(t)$ model \\cite{Cunha}, the interacting dark energy models \\cite{Wang}, Generalized Chaplygin gas model \\cite{Alcaniz2003b,Wang2009}, holographic dark energy model \\cite{Wei}, braneworld models \\cite{Movahed,Alam,Pires} and conformal gravity model \\cite{Yang2013}. But all of these models have a serious age problem except the conformal gravity model, which can accommodate this quasar at $3\\sigma$ confidence level \\cite{Yang2013}. In this paper, we will use the old quasar APM 08279+5255 at redshift $z=3.91$ and the 9 extremely old Global Clusters \\cite{Ma2009,WangAj2010} to investigate the cosmic age problem in the $R_h=ct$ Universe. The data of these 9 extremely old GCs listed in Table 1 is acquired from the Beijing-Arizona-Taiwan-Connecticut system with up-to-date theoretical synthesis models. The evolutionary population synthesis modeling has become a powerful tool for the age determination \\cite{Tinsley68,Searle73}. In \\cite{Ma2009,WangAj2010}, they get the ages of those GCs by using multi-color photometric CCD data and comparing them with up-to-date theoretical synthesis models. But the ages of GCs derived by different authors based on different measurements using same method are not always consistent \\cite{Ma2009}. We find that the 9 of those GCs can give stronger constraints on the age of universe than the old quasar APM $08279+5255$. Those 9 extremely old GCs have been used to test the cosmic age problem in dark energy models in previous work and many dark energy models have a serious age problem \\cite{Wang}. The $R_h=ct$ Universe is a cosmic model which is closely restricted by the cosmological principle and Weyl's postulate \\cite{Melia2007}. In the $R_h=ct$ Universe, the gravitational horizon $R_h$ is always equal to $ct$. The $R_h=ct$ Universe can fit the SNe Ia data well \\cite{Melia2012} , explain the growth of high-$z$ quasars \\cite{Melia2013}, account for the apparent absence in CMB angular correlation \\cite{Melia2014}. As we discuss above, many cosmological models can not pass the age test. But whether the $R_h=ct$ Universe suffers the cosmic age problem in still unknown. The structure of this paper is as follows. In section 2, we introduce the $R_h=ct$ Universe. In section 3, we give the constraints on the $R_h=ct$ Universe from SNe Ia and H(z) data. Then we will test the $R_h=ct$ Universe with the 9 extremely oldest GCs and the old quasar APM $08279+5255$. The age test in other cosmological models is presented in section 4. Conclusions will be given in section 5. ", "conclusions": "\\label{sec:Conclusions} In this paper, we test the cosmic age problem in several cosmological models by using nine extremely old GCs in M31 and the old quasar APM $08279+5255$. We find that the best-fit value of Hubble constant in the $R_h=ct$ Universe is $H_0 = 70.01\\pm0.40 {\\rm~km~s^{-1}~Mpc^{-1}}$ at $1\\sigma$ confidence level by using SNe Ia data. In this case, the age of local $R_h=ct$ universe $t_0 = 13.97{\\pm}0.08 \\rm Gyr$. If we fit the $R_h=ct$ Universe with the SNe Ia and H(z) data, the Hubble constant is $H_0 = 69.83\\pm0.40{\\rm~ km~s^{-1}~Mpc^{-1}}$ at the $1\\sigma$ confidence level. The age of local universe is $t_0 = 14.01{\\pm}0.08$ Gyr. From Fig. \\ref{fig:Age_Ia_Hz_Q_fig}, we find that the $R_h=ct$ Universe can accommodate the old quasar APM $08279+5255$ at more than 3$\\sigma$ confidence level. From Fig. \\ref{fig:Age_Ia_Hz_G_fig}, we find that there are five GCs (B239, B144D, B260, B383, B495) can be accommodated by the $R_h=ct$ Universe at $1\\sigma$ confidence level. But the age estimates of some GCs are controversial. For example, the metallicities of B129, B024, B297D and B050 measured by \\cite{Ma2009,WangAj2010} and \\cite{Galleti} are significantly different. So the derived ages are different. Due to the uncertainty of age determination, we can claim that the $R_h=ct$ Universe can marginally solve the cosmic age problem. Using the same method, we also test some other cosmological models, such as $\\Lambda$CDM, interacting dark energy model, generalized Chaplygin gas model and holographic dark energy model. In Sec.(\\ref{sec:OtherModels}), we show that these models can not accommodate all nine old GCs in M31. Meanwhile, for the old quasar APM $08279+5255$ at $z=3.91$, the $R_h=ct$ model can accommodate it at more than $3\\sigma$ confidence level. But these models can not accommodate it. The generalized Chaplygin gas model is in tension (over $2\\sigma$ confidence level) with the age of APM $08279+5255$. So the $R_h=ct$ Universe can marginally solve the cosmic age problem, especially at high redshift." }, "1402/1402.6743_arXiv.txt": { "abstract": "We report the discovery of 24 Lyman-break candidates at $7\\lesssim z \\lesssim 10.5 $, in the Hubble Frontier Fields (HFF) imaging data of Abell 2744 ($z=0.308$), plus \\emph{Spiter}/IRAC data and archival ACS data. The sample includes a triple image system with a photometric redshift of $z\\simeq 7.4$. This high redshift is geometrically confirmed by our lens model corresponding to deflection angles that are 12\\% larger than the lower-redshift systems used to calibrate the lens model at $z= 2.019$. The majority of our high-redshift candidates are not expected to be multiply lensed given their locations in the image plane and the brightness of foreground galaxies, but are magnified by factors of $\\sim 1.3 - 15$, so that we are seeing further down the luminosity function than comparable deep field imaging. It is apparent that the redshift distribution of these sources does not smoothly extend over the full redshift range accessible at $z<12$, but appears to break above $z=9$. Nine candidates are clustered within a small region of $20\\arcsec$ across, representing a potentially unprecedented concentration. Given the poor statistics, however, we must await similar constraints from the additional HFF clusters to properly examine this trend. The physical properties of our candidates are examined using the range of lens models developed for the HFF program by various groups including our own, for a better estimate of underlying systematics. Our spectral-energy-distribution fits for the brightest objects suggest stellar masses of $\\simeq 10^{9}$~\\Msun, star-formation rates of $\\simeq 4$~\\Msun~yr$^{-1}$, and a typical formation redshift of $z\\lesssim 19$. ", "introduction": "Our understanding of the first few billion years of cosmic time has increased significantly in recent years, thanks to the {\\em Hubble Space Telescope's} (\\HST) Wide-Field Camera~3/Infrared Channel \\citep[WFC3/IR,][]{wfc3} as well as the {\\em Spitzer Space Telescope's} Infrared Array Camera \\citep[IRAC,][]{irac}. Until recently, the Hubble Ultra Deep Field \\citep{beckwith,garth} has provided our deepest view of the Universe, revealing a considerable number of galaxy candidates at $z > 7$, including one candidate at $z\\simeq 10$ \\citep{bouwens, bouwens1, ellis, oesch, garth}. The cosmic epoch of $z\\simeq 10$ is important to study as it marks the dawn of galaxy formation and the beginning of reionization of the intergalactic medium. However, galaxies at that redshift are extremely faint, making it difficult to discover and study the abundant population of galaxies below $L^{*}$, the knee of the luminosity function. Fortunately, the magnification boost afforded by gravitational lensing combined with \\HST's exquisite imaging capabilities in the near-infrared (NIR), provides an avenue for both discovering and characterizing the intrinsic properties of galaxies around $z\\simeq 12$, when the Universe was less than half a billion years old. The Cluster Lensing And Supernova survey with Hubble \\citep[CLASH,][]{postman} carried out \\HST{} imaging of 25 galaxy clusters in 16 broad bands between $0.2-1.7$ $\\mu$m to a depth of AB magnitude $\\sim 27$ with a total of 20 orbits per cluster. The CLASH program has led to many interesting discoveries of magnified, intrinsically faint galaxies. Several hundred dropout galaxies have been uncovered in the range $z\\simeq 6-8$, with a few notable examples at higher redshifts of $z\\simeq 9-11$ \\citep[see][]{zheng, bouwens2, coe, bradley}, helping to motivate dedicated deeper lensing surveys. The Hubble Frontier Fields (HFF) is a new initiative now being carried out to observe the distant Universe to an unprecedented depth, combining the power of deep \\HST\\ imaging and gravitational lensing. In \\HST's Cycles 21 and 22, 560 orbits of Director's Discretionary Time have been allocated to observe four clusters. The observations are carried out with four WFC3/IR filters (F160W, F140W, F125W, F105W) and three ACS filters \\cite[Advanced Camera for Surveys,][F814W, F606W, F435W]{acs}. It is anticipated that 280 orbits will be allocated in Cycle 23 to observe two additional clusters. In addition, deep \\emph{Spitzer} and \\emph{Chandra} observations are planned for the six HFF fields. These coordinated observations will enable us to probe the star formation rate density at $z\\gtrsim 9$, study the faint end of the galaxy population at $z \\simeq 3-8 $, and map the dark matter in these clusters in unprecedented detail via many multiple images of background sources (Hubble Deep Fields Initiative 2012 Science Working Group Report).\\footnote{http://www.stsci.edu/hst/campaigns/frontier-fields/HDFI\\_SWGReport2012.pdf} We report the discovery of 24 candidate Lyman-break galaxies (LBGs) at $z\\gtrsim 7$ in the field of \\cl, based on the HFF observations and archival data. The faintest sources detected are around AB magnitude 29. These objects, listed in Tables~\\ref{tbl-z9}-\\ref{tbl-z7}, have ``secure'' photometric redshifts greater than 7 and a negligible probability ($<1\\%$) of being at lower redshift. We adopt a concordance cosmology with $\\Omega_M=0.3$, $\\Omega_\\Lambda=0.7$ and $h=H_0/100\\,{\\rm km\\,s^{-1}\\,Mpc^{-1}}=0.7$, and the AB magnitude system throughout. ", "conclusions": "We find 24 LBG candidates at $7 \\lesssim z \\lesssim 10.5$ in the HFF imaging of \\cl, reaching an intrinsic magnitude of $\\sim 32$. One source at $z\\simeq 7.4$ is lensed into three images. Significant clustering is observed on the intrinsic scale of $10-100$~kpc. Thanks to gravitational lensing, we are able to carry out Spitzer/IRAC photometry for 16 of the sources. SED fitting to the brightest candidates suggests stellar masses of $\\simeq 10^9$~\\Msun, star-formation rates of $\\simeq 4$~\\Msun\\ per year, and a typical formation redshift of $z \\simeq 19$. The redshift distribution of our sample is not a smoothly declining function towards higher redshift. In particular, our redshift distribution does not extend smoothly beyond $z\\simeq 9$, clustering notwithstanding, Considering the effect of cosmic variance, the number density in our sample is consistent with that derived from other studies, {\\it e.g.}, \\cite{bouwens4}. Given the level of clustering that we see in \\cl, it will be important to average over more HFF, and to perform a luminosity function analysis so that the redshift dependence can be better related to galaxy mass. The work presented in this paper is based on observations made with the NASA/ESA {\\it Hubble Space Telescope}, and has been supported by award AR-13079 from the Space Telescope Science Institute (STScI), which is operated by the Association of Universities for Research in Astronomy, Inc. under NASA contract NAS~5-26555. It is also based on data obtained with the {\\it Spitzer Space Telescope}, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA. This work utilizes gravitational lensing models produced by P.I.s Bradac, Ebeling, Merten \\& Zitrin, Sharon, and Williams, funded as part of the \\HST\\ Frontier Fields program conducted by STScI. These models were calibrated using arcs identified in archival \\HST{} imaging by \\cite{merten}, spectroscopic redshifts of arcs obtained using the VLT/FORS2 spectrograph (J. Richard et~al. 2014, in prep.), and VLT and Subaru/Suprimecam imaging of the Abell~2744 field \\citep{cypriano, okabe08, okabe10, okabe10b}. XS acknowledges support from FP7-SPACE-2012-ASTRODEEP-312725, and NSFC grants 11103017 and 11233002. Support for AZ is provided by NASA through Hubble Fellowship grant HST-HF-51334.01-A awarded by STScI. FEB acknowledges support from Basal-CATA PFB-06/2007, CONICYT-Chile grants (FONDECYT 1141218, ALMA-CONICYT 31100004, Gemini-CONICYT 32120003, Anillo ACT1101), and the Millennium Institute of Astrophysics (Project IC120009). JMD acknowledges support from the Spanish consolider project CAD2010-00064 and AYA2012-39475-C02-01. We thank M. Meneghitti for helpful comments. \\hskip 1.5in" }, "1402/1402.3166_arXiv.txt": { "abstract": "We present optical photometric monitoring of KT Eri (Nova Eridani 2009), a He/N very fast nova which oubursted in November 2009. Our observations include $BVR_cI_c$ brightness estimations as well as monitoring of the rapid brightness variations in V band. The characteristic times of these rapid changes are studied and compared with the observed in other novae. ", "introduction": "KT Eri was discovered by \\citet{2009CBET.2050....1I} in November 2009 at high galactic latitude. Due to this unusual position it is the first classical nova observed in the constellation Eridanus. \\citet{2009ATel.2327....1R} estimated that the maximum brightness $V\\sim5.\\!\\!^\\mathrm{m}6$ occurred on November 15 which shows that the nova is very fast with $t_\\mathrm{2}\\sim8$\\,days and $t_\\mathrm{3}\\sim15$\\,days. \\citet{2009CBET.2053....3M} classified KT Eri as He/N nova. The pre-outburst light curve shows large variations in the optical and an average brightness $\\sim 15^\\mathrm{m}$ \\citep{2009ATel.2331....1D}. Because of the relatively low amplitude ($\\sim 10^\\mathrm{m}$), unusual for very fast classical novae, it was suggested \\citep{2010ApJ...724..480H} that KT Eri is in fact recurrent nova, but no evidence of other outbursts are yet available. Studying the historical light curve of the nova, \\citet{2012A&A...537A..34J} found a possible orbital period of 737 days and suggested that the secondary component could be a RGB star. \\hspace{1em} Quasi-periodic Oscillations (QPOs) are rapid brightness variations often observed in cataclysmic variables. Their typical periods are from 3s to $\\sim$1000s usually unstable and with low amplitudes. QPOs with short periods are more profoundly studied, although all QPOs nature is uncertain (see \\citeauthor{2004PASP..116..115W} \\citeyear{2004PASP..116..115W}, \\citeauthor{2008AIPC.1054..101W} \\citeyear{2008AIPC.1054..101W}). \\hspace{1em} \\citet{2010ATel.2392....1B} were the first to report on possible rapid brightness variations in optical observations of KT Eri obtained from day 17 to day 65 after the maximum. Unfortunately, the frequency of their data sampling was too small for periodogram analysis. Very short $P\\sim35.09$ sec QPO were observed by \\citet{2010ATel.2423....1B} in X-ray between 67 and 79 days after the maximum. Long-term variability of KT~Eri with a period of 56.7 days was reported by \\citet{2011ASPC..451..271H}, interpreted by them as rotation of the hot spot on the accretion disk. ", "conclusions": "The results of our time-resolved photometry show that the KT Eri post-outburst optical QPOs appear and disappear in several day intervals. Taking into account the accuracy of our observations we cannot be sure that the QPOs really are absent. It is possible that we are not able to detect the ones with lower amplitudes. \\hspace{1em} There are several novae which show a variety of QPO periods: IX Vel $\\sim$500 sec, V533 Her $\\sim$1400 sec, BT Mon $\\sim$1800 sec, GK Per $\\sim$5000 sec, V842 Cen $\\sim$750-1300 sec (Warner 2004), V2468 Cyg $\\sim$1260-3000 sec \\citep{2012MmSAI..83..767C}. Most of these periods, excepting GK Per QPOs period are shorter in comparison to ones of KT Eri. \\hspace{1em} It is interesting to note, that about two years after the outburst of the recurrent nova RS Oph in 2006, \\citet{2009ASPC..404...72V} found optical QPOs with a period $\\sim$3000 sec. Very close to the observed in KT Eri on days 133 and 142. Furthermore, \\citet{2011ApJ...727..124O} detected X-ray QPOs with a period $\\sim$35 sec during the supersoft phase of RS Oph, similar to the reported by Beardmoreday et al. (2010) for KT Eri. \\hspace{1em} The resemblance of the KT Eri and RS Oph quasi periodic oscillations may be an additional argument in favor of the suggestion that KT Eri is in fact a recurrent nova. But the large variety of the periods and the low stability of the QPOs observed in novae make this argument insignificant." }, "1402/1402.5095_arXiv.txt": { "abstract": "In this paper we use near-infrared (NIR) spectral observations of Type~Ia supernovae (SNe~Ia) to study the uncertainties inherent to NIR $K$ corrections. To do so, 75 previously published NIR spectra of 33 SNe~Ia are employed to determine $K$-correction uncertainties in the $YJHK_s$ passbands as a function of temporal phase and redshift. The resultant $K$ corrections are then fed into an interpolation algorithm that provides mean $K$~corrections as a function of temporal phase and robust estimates of the associated errors. These uncertainties are both statistical and intrinsic --- i.e., due to the diversity of spectral features from object to object --- and must be included in the overall error budget of cosmological parameters constrained through the use of NIR observations of SNe~Ia. Intrinsic variations are likely the dominant source of error for all four passbands at maximum light. Given the present data, the total $Y$-band $K$-correction uncertainties at maximum are smallest, amounting to $\\pm 0.04$~mag at a redshift of $z = 0.08$. The $J$-band $K$-term errors are also reasonably small ($\\pm 0.06$~mag), but intrinsic variations of spectral features and noise introduced by telluric corrections in the $H$-band currently limit the total $K$-correction errors at maximum to $\\pm 0.10$~mag at $z = 0.08$. Finally, uncertainties in the $K_s$-band $K$~terms at maximum amount to $\\pm 0.07$~mag at this same redshift. These results are largely constrained by the small number of published NIR spectra of SNe~Ia, which do not yet allow spectral templates to be constructed as a function of the light curve decline rate. ", "introduction": "Type Ia supernovae (hereafter SNe~Ia) are standardizable distance indicators at optical wavelengths that provide critical constraints on cosmological parameters \\citep[see][and references therein]{goobar11}. A number of groups have worked diligently to gather optical photometry of homogenous samples of low-, intermediate- and high-$z$ SNe~Ia that, when combined, amount to well over 1000 objects. As the sample size has increased, systematic effects have come to dominate the final uncertainty in the measured value of the equation-of-state parameter of the Universe, $w$ \\citep[e.g., see][]{wood07,astier06,freedman09,kessler09,folatelli10,conley11,suzuki12}. A major systematic that plagues SN~Ia cosmology is our inability to accurately estimate host galaxy dust extinction due to uncertainties in the reddening law and, in particular, variations in the value of the total-to-selective extinction, $R_V$, from object to object \\citep[see][and references therein]{phillips12}. This is further exacerbated by any systematic error in the relative zero points between the nearby and distance SNe~Ia samples, thereby making the latter artificially redder (or bluer) than the former. A sensible way around these problems is to observe in rest-frame near-infrared (NIR) bands instead of rest-frame optical bands, because the effects of dust extinction are minimized and essentially independent of the adopted reddening law \\citep{krisciunas00}. Additional motivation is provided by empirical evidence indicating that the luminosities of SNe~Ia show little or no dependence on decline rate\\footnote[8]{The decline rate of a SN~Ia is traditionally defined as the change in its $B$-band magnitude from the time of maximum brightness to 15 days later, and is denoted as $\\Delta$$m_{15}(B)$. The decline rate is known to correlate with the peak absolute luminosity in such a way that more luminous objects exhibit smaller $\\Delta$$m_{15}(B)$ values \\citep{phillips93}.} in the NIR \\citep{meikle00,kr204,wood08,kriciunas09,mandel09,folatelli10,kattner12}. This translates to a reduced intrinsic dispersion in the NIR Hubble diagram, and at the same time evolutionary effects as a function of redshift on the progenitor populations are potentially minimized. These factors have provided significant impetus for future SN~Ia cosmology studies to construct {\\it homogenous samples} of low- and high-$z$ SN~Ia observed in the rest-frame NIR \\citep[e.g., see][]{green12}. Reaping the benefits afforded by observing SNe~Ia in the NIR requires the development of tools to obtain rest-frame luminosities via $K$~corrections \\citep{oke68}. The $K$~correction is defined as the difference in brightness between an object observed in its rest-frame with a given passband compared to its measured brightness with the same passband when observed at redshift $z$. Specifically, for a given passband, $i$, the $K$~correction is defined as: \\begin{equation} m_{i}=M_{i} + \\mu + K_{i}. \\label{K_corr1} \\end{equation} Here $m_{i}$ is the SN~Ia apparent magnitude observed on Earth, and $M_{i}$ is its absolute magnitude. The magnitude difference encapsulated in the $K$ term is explained by the shifting and stretching of an object's spectral energy distribution (SED), which is inherent to cosmological expansion. The first calculations of SN~Ia $K$~corrections at optical wavelengths were published by \\citet{leibundgut90} and \\citet{ham93}. In the latter paper, optical observations of three nearby SNe~Ia were used to construct a sequence of $B$- and $V$-band $K$~corrections extending to a redshift of $z = 0.5$. The authors found that the temporal variation of the computed values largely mimicked the $(B-V)$ color evolution, implying that the $K$~correction is, to first order, driven by the color of the SN. Shortly thereafter, \\citet{kim96} presented a method to compute cross-band $K$~corrections which is particularly well-suited for SNe~Ia located at $z > 0.2$. \\citet{nugent02} expanded upon these efforts by developing a set of SN~Ia optical spectral templates that enabled $K$~corrections to be computed as a function of temporal phase for any given optical bandpass. Later, \\citet{hsi07} constructed improved optical spectral templates based on a greatly expanded sample of nearby SN~Ia. They found that besides color, spectral diversity also has a significant impact upon the magnitude of the $K$~term. The \\citet{hsi07} template is now routinely used to $K$~correct optical photometry of SNe~Ia, and the statistical uncertainty associated with these corrections is reasonably well understood. In comparison, our knowledge of NIR $K$~corrections is still relatively crude. \\citet{krisciunas04} presented the first calculations of the temporal evolution of the $K$~term in the $JHK$ passbands based on 11 NIR spectra of SN~1999ee \\citep{hamuy02}. Their results indicated that NIR $K$~corrections are non-negligible even at relatively small redshifts. These initial findings emphasized the importance of accurately characterizing NIR $K$~corrections {\\it and} their uncertainties as a function of redshift and temporal phase. Five years later, the publication of 41 spectra by \\citet{marion09} revolutionized the study of the NIR spectral characteristics of SNe~Ia, and \\citet{hsiao09} used these data along with the other published spectra available at that time to extend his spectral template to include NIR wavelengths. In this paper, we refer to this template as the ``Hsiao revised template''. In the present paper, we expand on this work by using the existing library of published NIR SNe~Ia spectra to determine the uncertainties inherent to using the Hsiao revised spectral template to calculate NIR $K$~corrections, particularly those due to intrinsic variations in spectral features. To do this, we first color match\\footnote[9]{Often people tend to use the nomenclature warp or mangle, however, we feel that it is more appropriate and accurate to adopted the term color match.} each observed spectrum to the template, and then calculate $YJHK_s$-band $K$~corrections. An interpolation algorithm based on Gaussian Processes combined with Markov-Chain Monte-Carlo methodology is then used to produce mean $K$~corrections as a function of temporal phase and redshift, along with estimates of both statistical and intrinsic sources of uncertainties. These errors, in turn, will provide important input to future studies utilizing the NIR light curves of SNe~Ia to estimate cosmological parameters. The organization of this article is as follows. In Section~\\ref{kcorrection} the concept of the $K$~correction is briefly reviewed; Section~\\ref{se:data} introduces the data used in our calculations, along with the adopted passbands and methods used to interpolate our computed $K$~corrections; Section~\\ref{se:results} contains the results; and finally, Section~\\ref{conclusions} presents our conclusions. ", "conclusions": "\\label{conclusions} Using a library of publicly available SNe~Ia NIR spectra, we have computed a set of $K$~corrections in the CSP $YJHK_s$ filters at redshifts of $z =0.03$, 0.05, and 0.08. The individual spectra were first color-matched to the Hsiao revised spectral template before calculating the $K$~corrections. A combined Gaussian Process and MCMC method was then employed to derive an empirically-based model of the $K$~terms as a function of temporal phase and redshift. This procedure returns uncertainties in the $K$-correction model due to both statistical noise and intrinsic diversity in the features of the input spectra. \\citet{krisciunas04} published a table of NIR $K$ corrections computed using spectra of SN~1999ee, and a set of filters functions similar to those adopted in this study. A comparison between the results of \\citeauthor{krisciunas04} to those obtained with our expanded set of spectra and more advanced color-matching technique provides good agreement in the $J$ and $H$ bands at $z=0.03$, while for the $K_s$ band we find differences of up to $\\pm$0.08 mag. The discrepancy in the $K_s$ is not surprising owing the the dearth of spectra included in the work of \\citeauthor{krisciunas04}, as well as the difficulties highlighted in this study concerning the difficulties inherent to computing robust $K_s$-band $K$ corrections. Our emphasis in this paper has not been to provide ``lookup\" tables of NIR $K$~corrections, but rather to derive accurate estimates of the uncertainties inherent in the use of spectral templates (like the Hsiao revised template) constructed from the NIR sample of SNe~Ia spectra available today. We find that the $Y$ and $J$ bands currently afford the greatest precision in $K$-correction calculations due to the general weakness of the spectral features and the minimal effect of telluric corrections. The $Y$ band is particularly noteworthy, with the uncertainty due to spectral feature variations increasing from only $\\pm0.02$ to $\\pm$0.04~mag from $z = 0.03$--0.05. The $H$ band is currently more problematic due to noise introduced by telluric corrections and the appearance just after maximum light of strong Fe-peak emission features. Diversity in the strength of this emission most likely explains the strong increase we find in the intrinsic component of the $K$-correction error from $\\pm0.02$ to $\\pm$0.10~mag over the redshift range $z = 0.03$--0.05. This result is a departure from pervious theoretical \\citep{kasen06} and observational \\citep{wood08} findings, which have identified the $H$ band as having potentially the least dispersion among the NIR passbands. Finally, the uncertainties in the $K$~terms for the $K_s$ filter due to intrinsic spectral diversity are similar to those found for the $J$~band, but the currently-available library of spectra covering this wavelength region suffer from both low signal-to-noise and telluric absorption. Further progress in reducing the uncertainties in the $K$~corrections for NIR filter bandpasses demands many more spectral observations over a wider range of redshifts. Figure~\\ref{hist} shows a histogram of the decline rates of the SNe~Ia in our sample with temporal phases within $\\pm 3$~days of \\bmax. For reference, the same figure includes a plot of the decline rates of the spectra employed by \\citet{hsi07} to create their optical spectral template. The discrepancy between these two histograms dramatically illustrates the limitations of the currently-available library of NIR spectra. To address this problem, we have embarked on a four-year, six-months-a-year SN~Ia followup program that builds upon the legacy of the CSP. This project, designated CSP-II, is designed to obtain optical and NIR light-curves of $\\sim$100 SNe~Ia in the smooth Hubble flow ($0.03 < z < 0.08$). Complementary to these observations, frequent NIR spectroscopy is being carried out of nearby SNe~Ia, mainly with the Folded-port InfraRed Echellette (FIRE) spectrograph mounted on the 6.5-m Magellan Baade telescope. Particular emphasis is being placed on obtaining spectral sequences for a sub-sample of the SNe that span a range in decline rate and phase, which are essential to determining correlated errors in the $K$~corrections. We are confident that such observations will not only allow us to more precisely characterize NIR $K$-correction uncertainties, but also to decrease these errors to the low levels now obtained at optical wavelengths. In order to ascertain the contribution of NIR $K$-correction errors to the final error budget of cosmological parameters one must determine whether they are systematic in temporal phase, and therefore propagate to the peak magnitude of a SN~Ia. On the other hand, if the $K$-correction errors are random in phase then the final uncertainty in the peak magnitude will be reduced by template fitting. To determine which of these is the case is, however, beyond the scope of this paper due to the lack of a statistical sample of SNe~Ia with good spectral time series. However, in the near future we will be able to tackle this issue with the high-fidelity spectral series currently being obtained." }, "1402/1402.2483_arXiv.txt": { "abstract": "{ L1642 is one of the two high galactic latitude ($|b| > 30^{\\circ}$) clouds confirmed to have active star formation. } { We examine the properties of this cloud, especially the large-scale structure, dust properties, and compact sources in different stages of star formation. } { We present high-resolution far-infrared and submillimetre observations with the \\emph{Herschel} and AKARI satellites and millimetre observations with the AzTEC/ASTE telescope, which we combined with archive data from near- and mid-infrared (2MASS, WISE) to millimetre wavelength observations (Planck). } { The \\emph{Herschel} observations, combined with other data, show a sequence of objects from a cold clump to young stellar objects at different evolutionary stages. Source B-3 (2MASS J04351455-1414468) appears to be a YSO forming inside the L1642 cloud, instead of a foreground brown dwarf, as previously classified. \\emph{Herschel} data reveal striation in the diffuse dust emission around the cloud L1642. The western region shows striation towards the NE and has a steeper column density gradient on its southern side. The densest central region has a bow-shock like structure showing compression from the west and has a filamentary tail extending towards the east. The differences suggest that these may be spatially distinct structures, aligned only in projection. We derive values of the dust emission cross-section per H nucleon of $\\sigma_e(250 \\mu {\\rm m}) = $ 0.5--1.5 $\\times10^{-25} {\\rm cm}^2/{\\rm H}$ for different regions of the cloud. Modified black-body fits to the spectral energy distribution of \\emph{Herschel} and Planck data give emissivity spectral index $\\beta$ values 1.8--2.0 for the different regions. The compact sources have lower $\\beta$ values and show an anticorrelation between $T$ and $\\beta$. } { Markov chain Monte Carlo calculations demonstrate the strong anticorrelation between $\\beta$ and $T$ errors and the importance of millimetre wavelength Planck data in constraining the estimates. L1642 reveals a more complex structure and sequence of star formation than previously known. } ", "introduction": "\\label{sect:intro} High galactic latitude ($|b| > 30^{\\circ}$) molecular clouds have low background and foreground emission, and are usually quite nearby. They are mostly diffuse or translucent clouds with a low density and typically no signs of star-formation, see~\\citet{McGehee2008} for a review. These features make them ideal for studying the interstellar medium and interstellar radiation field, as well as low-mass objects in selected regions. LDN 1642\\footnote{As a side note, a memory aid for the number: in the year 1642 Galileo Galilei died and Sir Isaac Newton was born.}~\\citep{Lynds1962} (often also called L1642) is one of the two clouds at high galactic latitudes confirmed to have active star-formation, the other one is MBM 12~\\citep{Luhman2001}. Studying this cloud might give constraints of the mechanisms that affect cloud evolution and trigger low-mass star formation. In this article, we use the abbreviation L1642. The cloud is also called MBM 20~\\citep{Magnani1985} and G210.90-36.55~\\citep{Juvela2012}. L1642 is one of the Orion outlying clouds, see~\\citet{Alcala2008} for a review of the area and earlier studies. It forms a blob in the head of a HI cloud with cometary structure, the tail of which extends over 5$^{\\circ}$ in the north-east direction in equatorial coordinates, directly towards the Galactic plane. The cloud is projected on the edge of the Orion-Eridanus Bubble~\\citep{Brown1995}. On the sky, L1642 is $\\sim10^{\\circ}$ from the filamentary reflection nebula IC 2118, or Witch Head nebula, which also contains the molecular clouds MBM 21 and 22~\\citep{Kun2001,Kun2004,Alcala2008}. The different clouds are marked in Fig.~\\ref{fig:Planck_Ak}. The Galactic coordinates of L1642 are $l = 210.9^{\\circ}$ and $b = -36.55^{\\circ}$, and the equatorial coordinates are $\\alpha_{2000} = 4^{\\rm h}35^{\\rm m}$ and $\\delta_{2000} = -14^{\\circ}15'$ ($\\sim 68.75^{\\circ}$ and $-14.25^{\\circ}$, respectively). \\citet{Hearty2000} determined the distance of L1642 to be 112--160 pc. The X-ray observations of~\\citet{Kuntz1997} suggested that the cloud is close to the edge of the Local Bubble, and not necessarily inside the Orion-Eridanus Bubble, indicating a distance of approximately 140 pc according to \\citet{Sfeir1999}. See~\\citet{Welsh2009} for a description of the Local Bubble. We adopted the distance of 140 pc similarly to~\\citet{Russeil2003} and~\\citet{Lehtinen2004}. L1642 has been studied using several molecules, see~\\citet{Liljestrom1991} for a review of the early studies. \\citet{Liljestrom1991} studied L1642 in CO, HCO$^+$ and NH$_3$ and concluded that the cloud is in virial equilibrium and older than $10^6$ years. More recently, \\citet{Russeil2003} studied the morphology and kinematics of L1642 using CO observations obtained with the SEST radio telescope. Their results showed that the cloud consists of a main structure at radial velocity 0.2 km s$^{-1}$ with higher velocity components forming an incomplete ring around it, suggesting an expanding shell. Based on $^{13}$CO data, the peak column density is $N$(H$_2$) $\\sim6\\times10^{21}$ cm$^{-2}$ and the cloud mass $\\sim$59 $M_{\\odot}$. L1642 has also been widely studied using other methods such as dust emission in mid- and far-infrared~\\citep{Laureijs1987,Reach1998,Verter1998,Verter2000,Lehtinen2004,Lehtinen2007} and optical (360--600 nm) surface brightness~\\citep{Laureijs1987,Mattila2012}. \\citet{Mattila2007} studied scattered H{$\\alpha$} light in L1642. The shadowing provided by L1642 has also been used in studies of Galactic diffuse X-ray radiation~\\citep[e.g.,][]{Galeazzi2007,Gupta2009}. The small distance, high galactic latitude with little foreground contamination, and the modest column density make L1642 a good target for studying low-mass star formation and the effect of potential external triggering. \\citet{Sandell1987} identified two IRAS sources, IRAS 04327-1419 = L1642-1 (V* EW Eri, HBC 413) and IRAS 04325-1419 = L1642-2 (HBC 410), found within L1642, to be faint nebulous binary stars with very active secondaries. The primary of L1642-1 was spectroscopically classified to be a K7IV T Tauri star. \\citet{Liljestrom1989} found a weak, bipolar outflow around the binary L1642-2. \\citet{Reipurth1990} discovered a Herbig-Haro object, HH123, also originating from L1642-2. They also concluded that the primary object of L1642-2 is a low-luminosity M0 H$\\alpha$-emission star, and the secondary component is an H$\\alpha$-emission star as well. \\citet{Correia2006} observed L1642-1 in search of high-order multiplicity, but did not find a third component. \\citet{Lehtinen2004} concluded that none of the other four IRAS sources projected within L1642 are likely to be young stellar objects (YSOs) inside the cloud. The 2MASS point source catalogue object 2MASS J04351455-1414468 is situated towards the densest part of L1642 on the sky. \\citet{Cruz2003} classified it as a young ($\\sim$10 Myr) object. They estimated the distance to be probably within 30 pc, meaning that the object would be clearly outside the L1642 star-forming region. Later studies have made the assumption or conclusion that the object is a brown dwarf located in front of the L1642 cloud, at a distance of 14--30 pc~\\citep{Faherty2009,Antonova2013}. In this article, we re-evaluate the classification and distance estimate of this object. The early low-resolution studies treated the L1642 cloud as a single entity. \\citet{Lehtinen2004} presented ISOPHOT far-infrared (FIR) data of L1642 and studied the cloud structure in separate components (named A1, A2, B, and C). Neither of the IRAS point sources nor new YSO candidates were seen in their 200 $\\mu$m maps. \\citet{Lehtinen2007} compared far-IR data with visual extinction and studied the dust grain emissivity properties in L1642. They found that FIR dust emissivity increases by a factor of two between regions with colour temperatures 19 K and 14 K. This might be caused by grain growth in the dense and cold areas. L1642 was recently observed as part of the \\emph{Herschel} open time key programme Galactic Cold Cores~\\citep{Juvela2010} with unprecedented spatial resolution and sensitivity at 100, 160, 250, 350 and 500 $\\mu$m. \\citet{Juvela2012} presented these observations and studied the properties of the cloud on large-scale images, without separating the distinct objects. In Rivera-Ingraham et al. (in prep) we will be carrying out a detailed \\emph{Herschel}-based compact source and environmental study of all high-latitude fields (including L1642) from our programme. Furthermore, we have carried out 1.1 mm dust-continuum observations toward L1642 with the AzTEC/ASTE instrument. In this paper, we use these new data, along with other, already published data, to complete the spectral energy distribution of L1642 in the long-wavelength region and advance with this (i) the study of the large-scale structure of the cloud, (ii) the investigation of general dust properties, and (iii) the characterisation of point and compact sources related to the cloud. The contents of the article are the following: we present the observations and data processing in Sect.~\\ref{sect:observations} and methods in Sect.~\\ref{sect:methods}. The results based on the observations are shown in Sect.~\\ref{sect:results}. We present a radiative transfer model of L1642 in the appendix. We discuss the results in Sect.~\\ref{sect:discussion} and present our conclusions in Sect.~\\ref{sect:conclusions}. ", "conclusions": "\\label{sect:conclusions} We have studied the properties of L1642, one of the two high galactic latitude ($|b| > 30^{\\circ}$) clouds with active star-formation. The high latitude ($-$36.4$^{\\circ}$), indicating no foreground contamination, the small distance ($\\sim$140 pc), the modest column density ($A_V < 12^{\\rm m}$), and the spatially distinct sequence of star-forming cores make L1642 one of the best targets for studying the initial stages of low-mass star formation. We presented in detail high-resolution far-infrared and submillimetre observations with \\emph{Herschel} and AKARI satellites and millimetre observations with AzTEC/ASTE telescope and combined them with archive data from NIR (2MASS) and MIR (WISE) to millimetre (Planck). We studied both larger regions (A1, A2, B, C) and compact sources (both previously known binary T Tauri stars and potential new objects) within the main cloud. Our main conclusions are the following: \\begin{itemize} \\item The high-resolution \\emph{Herschel} observations, combined with other data, show an evolutionary sequence from a cold clump (B-4) to young stellar objects of different spectral classes: B-1 (L1642-1, Class II), B-2 (L1642-2, Flat spectrum), and B-3 (Class III). \\item Based on \\emph{Herschel} FIR--submm data, the point source B-3 (2MASS J04351455-1414468) appears to be a YSO forming inside the L1642 cloud, instead of a foreground brown dwarf, contrary to previous classification. \\item \\emph{Herschel} data reveal striation in the diffuse dust around the cloud L1642, especially in the northern part. Region A shows striation extending towards the NE and has a correspondingly steeper column density gradient on its southern side. Region B has the appearance of being compressed from the west and has a filamentary tail extending eastward. The differences suggest that these may be spatially distinct structures, aligned only in projection. \\item Both \\emph{Herschel} FIR--submm and AzTEC/ASTE mm data show an elongated structure north-east of the binary star B-2 (L1642-2), which was not visible in earlier observations. The structure is aligned with the general striation pattern, but it has a significantly higher column density. \\item We derived values of the dust emission cross-section per H nucleon of $\\sigma_e(250) = $ 0.56, 0.47, 1.45, and 1.04 $\\times10^{-25} {\\rm cm}^2/{\\rm H}$ for the different regions. \\item Modified black-body fits to the SED of \\emph{Herschel} and Planck data give emissivity spectral index $\\beta$ values 1.8--2.0 for the different regions of L1642. \\item Markov chain Monte Carlo fits of the SEDs show a strong anticorrelation between $\\beta$ and $T$ errors and the importance of Planck data in constraining $\\beta$ and $T$ estimates. \\item The compact sources show a $T$-$\\beta$ anticorrelation, but on larger scale the data show no clear correlation or anticorrelation. \\item Radiative transfer modelling suggests that low beta values towards the sources might be caused to a large extent by the LOS temperature variations. On the other hand, the optical depth estimates are expected to be quite accurate (to within $\\sim$ 10 \\%) away from the sources. \\end{itemize}" }, "1402/1402.0509_arXiv.txt": { "abstract": "\\centering Observations of circumbinary planets orbiting very close to the central stars have shown that planet formation may occur in a very hostile environment, where the gravitational pull from the binary should be very strong on the primordial protoplanetary disk. Elevated impact velocities and orbit crossings from eccentricity oscillations are the primary contributors towards high energy, potentially destructive collisions that inhibit the growth of aspiring planets. In this work, we conduct high resolution, inter-particle gravity enabled {\\emph N}-body simulations to investigate the feasibility of planetesimal growth in the Kepler-34 system. We improve upon previous work by including planetesimal disk self-gravity and an extensive collision model to accurately handle inter-planetesimal interactions. We find that super-catastrophic erosion events are the dominant mechanism up to and including the orbital radius of Kepler-34(AB)b, making {\\it in-situ} growth unlikely. It is more plausible that Kepler-34(AB)b migrated from a region beyond 1.5 AU. Based on the conclusions that we have made for Kepler-34 it seems likely that all of the currently known circumbinary planets have also migrated significantly from their formation location with the possible exception of Kepler-47(AB)c. ", "introduction": "\\begin{figure*}[tbph] \\centering \\includegraphics[scale=0.45]{f1.eps} \\caption{Evolution of eccentricity waves at $t_0$ ({\\bf A}), $25\\,P_{AB}$ ({\\bf B}), $100\\,P_{AB}$ ({\\bf C}), $1000\\,P_{AB}$ ({\\bf D}). The instantaneous eccentricity of each planetesimal is shown with a grey value between black ($e=0$) and white ($e=0.02$ {\\bf A}, $e=0.1$ {\\bf B-D}). {\\bf E}-{\\bf H} show eccentricity evolution versus $a$ with $e_{ff}$ \\citep{paardekooper12} shown in grey.} \\label{fig:multi} \\vspace{+10pt} \\end{figure*} Several planets have been discovered by Kepler in `p-type' orbits fully encompassing tight binaries \\citep{doyle11, welsh12, orosz12, schwamb13}. These planets orbit close to their host binaries ($a < 1.1$ AU) and are subject to significant perturbations which cause impact speeds to increase making accretion close to the binary difficult \\citep{marzari00,thebault06,scholl07}. \\cite{paardekooper12} conducted numerical simulations of test particles within circumbinary planetesimal disks including a collision model from \\citet{stewart09} which allowed for accretion and erosion. The collision model provided an alternative mechanism for planetesimal growth: accretion of dust created in catastrophic collisions. However, the authors found that growth at the location of the Kepler-16 and Kepler-34 planets was difficult. Their investigation also revealed the contribution of short-period perturbations on the disk. This fast eccentricity forcing evolves more quickly than the gas damping timescale therefore the disk stays excited. Thus, there remains a degree of uncertainty about whether observed circumbinary planets formed in-situ. In addition, the literature does not agree on the locations that could have supported the planets' growth within a circumbinary disk. For example, \\cite{meschiari12a} identify a narrow range of annuli just outside of 1 AU that could sustain planetesimal growth in Kepler-16 and \\cite{paardekooper12} find the accretion friendly zone is beyond 4 AU. Most previous work omit inter-planetesimal gravity and/or a comprehensive collision model. In this paper, we present high resolution, 3D, inter-particle gravity (IPG) enabled {\\emph{N}}-body simulations of a circumbinary protoplanetary disk in order to address the question of where the circumbinary planets can form. We focus on the orbital dynamics, collisional evolution and physical growth of 100 kilometre-sized planetesimals in the Kepler-34 system. We consider a purely $N$-body case to isolate the effects of inter-planetsimal gravity, leaving the inclusion of a gas disk for future work. To account for oblique, high speed impacts from orbit crossings, we use a state-of-the-art collision model \\citep[][Leinhardt et al.~in prep]{leinhardt12}. This collision model more accurately identifies regions where planetesimal growth can occur than previous work. We use statistical arguments to classify the feasibility of sustained growth events in the disk, addressing the primary question of whether in-situ growth of Kepler-34(AB)b is possible. In section \\ref{sec:theory} we discuss the analytics of circumbinary planetesimals acting under perturbations from the stellar binary. Our numerical method and collision model are outlined in section \\ref{sec:method}. In section \\ref{sec:results} we present our results and section \\ref{sec:dis} discusses the broader implications. Conclusions are drawn in section \\ref{sec:summary}. ", "conclusions": "\\label{sec:dis} Kepler-34(AB)b orbits at 1.09 AU which falls in a regime where only 12$\\%$ of collisions lead to growth (Fig.~\\ref{fig:collspatevo}B). Although the eccentricity forcing falls off with $1/a^2$, $e_{ff}$ only drops to the mean initial eccentricity (e = 0.007) at 1.7 AU. However, it may not be essential to retrieve an identical environment to that of the single star case in order to sustain growth. By 1.5 AU mass changing collisions are split between partial erosion and partial accretion events, therefore, by 1.5 AU prolonged planetesimal growth becomes more sustainable. Thus, a likely scenario is that Kepler-34(AB)b formed beyond 1.5 AU and migrated inwards. It is worth mentioning that we retrieve this highly erosive disk that is unable to support sustained growth, despite our large-sized planetesimals. \\begin{figure}[h!] \\centering \\includegraphics[scale=0.38,angle=0]{f5.eps} \\caption{Eccentricity distribution of planetesimals at 2,000 $P_{AB}$. Green points: $m \\geq 2m_o$, black points: $m<2m_o$. The red line shows contribution from perfect mergers.} \\label{fig:growth} \\vspace{+5pt} \\end{figure} While the visual difference between the IPG and RIPG cases is small, the underlying effects propagate through to the collision outcomes. The lack of gravitational focusing with RIPG accounts for a higher proportion of oblique impacts that generally lead to an increase in hit-and-run events. While the majority of these collision are non-erosive, they do not contribute to the growth of the system. A more subtle point is that the inclusion of inter-planetesimal gravity results in collisions that would not otherwise occur in a scenario without gravitational focusing. This results in a number of low speed and accretion enabling collisions. From these results we find that $e_{ff} > 0.01$ indicates a region of the planetesimal disk that is too perturbed to support planetesimal growth and $e_{ff} < 0.01$ may be calm enough for planetesimal accretion. If we apply this criteria to all known Kepler circumbinary planets, and assuming that this critical value applies to $e_f$ as well, only Kepler 47(AB)c could possibly have formed in-situ and all of the rest: Kepler 34(AB)b, 16(AB)b, 35(AB)b, 38(AB)b, 47(AB)b and 64(AB)b, must have formed further out in the planetesimal disk and migrated inwards to their current location. The effect of a reduced collision rate at large orbital radii on the formation timescale of planets is possibly explained by \\citet{alexander12} who suggests that circumbinary planets have more time to form than planets around single stars due to the longer lifetime of disks around short period binaries. In this paper we presented results from high-resolution N-body simulations investigating the likelihood of in-situ formation of Kepler-34(AB)b. We began with a unperturbed disk using eccentricity and inclination distributions from numerical simulations of planetesimal evolution around single stars and allowed the system to evolve to quasi-steady state while suppressing collisional evolution of the planetesimals. Eccentricity forcing from the binary pumped up the planetesimal eccentricities to 10 times the initial mean value. After 1000 $P_{AB}$ a quasi-steady state was reached where planetesimals with similar orbital radii take on eccentricity extremes, such that both circularised and eccentric bodies can occupy the same space. This is shown by the speckled effect in Figure \\ref{fig:multi}{\\bf D}. This new initial condition was highly perturbed, which meant many orbit crossing events occurred. After the initial condition was reached we allowed the planetesimals to evolve collisionally using our new collision model EDACM based on \\citep{leinhardt12} which allows erosion, accretion, and bouncing events. The high eccentricity observed in the inner circumbinary disk translates into a dominance of super-catastrophic events seen in Figure \\ref{fig:collspatevo}. The impact velocity and parameter decrease with increasing orbital radius; the latter due to a narrower eccentricity dispersion resulting in a reduced number of orbit crossings. The number of high energy, erosive events decrease as a function of increasing semi-major axis. However, planetesimal growth events do not dominate the collision outcomes within the simulation domain ($a<1.5$ AU). We therefore show that the disk is a hostile environment even for our gravitationally strong 120 km planetesimals, suggesting even more difficulties for sustained accretion to occur in simulations that feature much smaller bodies. Using statistical arguments from collisional data, in addition to physical growth rates, we find that Kepler-34(AB)b would struggle to grow in-situ. In addition, we suggest that from all the known Kepler circumbinary planets only Kepler-47(AB)c could have formed in-situ while the rest must have formed at larger $a$ where the protoplanetary disks were less perturbed by the binary stars and migrated inwards to their current location. We also show that inter-planetesimal gravity must be included in planet formation models in order to capture gravitational focusing effects that may be missed otherwise, such as low-velocity, growth-enabling impacts that may influence the outcome of the simulation. Previous work which has attempted to hybridise a protoplanetary disk with a gas counterpart has suggested that the gas disk is similarly perturbed by a stellar binary causing further excitations to the planetesimal eccentricities, which, could further provoke unfavourable impact velocities \\citep{marzari08,marzari13,paardekooper08}. Our simulations, therefore, likely present a best-case scenario for planetesimal accretion." }, "1402/1402.7329_arXiv.txt": { "abstract": "Measurements of the polarized radiation often reveal specific physical properties of emission sources, such as strengths and orientations of magnetic fields offered by synchrotron radiation and Zeeman line emission, and the electron density distribution by free-free emission. Polarization-capable, millimeter/sub-millimeter telescopes are normally equipped with either septum polarizers or ortho-mode transducers (OMT) for the detection of polarized radiation. While the septum polarizer is traditionally conceived to be limited to a significantly narrower bandwidth than the OMT, it does possess advantageous features for astronomical polarization measurements unparalleled by the OMT. Challenging the conventional bandwidth limit, we design an extremely wideband circular waveguide septum polarizer, covering $42\\%$ bandwidth, from 77 GHz to 118 GHz, without any undesired resonance. Stokes parameters constructed from the measured data in between 77 GHz and 115 GHz show that the leakage from $I$ to $Q$ and $U$ is below $\\pm 2\\%$ and the $Q-U$ mutual leakage below $\\pm 1\\%$. Such a performance is comparable to other modern polarizers, but the bandwidth of this polarizer can be at least twice as wide. This extremely wide-band design removes the major weakness of the septum polarizer and opens up a new window for future astronomical polarization measurements. ", "introduction": "Wide-band polarization measurements in millimeter and sub-millimeter Astronomy have always been a challenge. Examples requiring wide-bandwidth polarimetry measurements include the continuum cosmic microwave background (CMB) polarization observation \\citep{kov02, kog03, bar05, mon06, sie07, fri09, chi10, bis11}, the synchrotron polarization observations of compact sources \\citep{dow98, ait00, cul11}, the polarization observation of thermal dust emission \\citep{laz03, bet07, hoa11} and the observation of Zeeman effects via molecular lines, such as CN and SO \\citep{bel89, cru99, shi00}. Except for some particular features of Zeeman effects, most of the above focus on linear polarization measurements. For the continuum observation, it is desirable the available frequency bandwidth to be as wide as possible so that, on one hand, the signal-to-noise ratio can increase, and on the other hand the spectral index may be determined. For the line observation, the emission line can be red-shifted to any unknown frequency when emitted from, or absorbed at, a distant universe, and hence wide frequency coverage has a unique advantages. While the current trend of CMB polarization experiments has shifted to the multi-pixels, silicon wafer-based, incoherent detector approach, such as the microwave kinetic inductance detector (MKID) \\citep{day03, mal10}, the conventional coherent detector approach still has its own merit, for its capability to control, detect and calibrate out the systematics. In this regard, heterodyne polarimeters with wide bandwidths are highly desirable. For synchrotron emitting compact sources, the interferometry array remains to be the means to reach high sensitivity, and it must adopt coherent detectors for polarization measurements. Similar considerations apply to line emission and dust emission from compact molecular cloud cores. All these make the conventional waveguide polarimeter device a time-honored instrument that will always be needed in future forefront millimeter/sub-millimeter telescopes. Polarization measurements require separating the incoming radiation into two orthogonal components for the determination of Stokes parameters. Traditionally two competing devices are available for the separation of polarization, the septum polarizer and the ortho-mode transducer (OMT). An ideal septum polarizer can convert an input linear polarization wave into two circular polarization waves of equal power at two output ports. Interestingly, in a specific arrangement when the electric field of the input wave is perpendicular or parallel to the symmetric axis of the septum polarizer, the two output electric fields will be either in phase or 180 degrees out of phase, with the latter being delayed relative to the former by 90 degrees over a finite frequency interval. It then follows that when the input is a circularly polarized wave, it will exit entirely through only one output port and the other port has a null output. This novel feature of the septum polarizer makes it distinct from the other simpler device, OMT, which separates an input linearly polarized wave into two components of the electric field parallel and perpendicular to the device symmetry axis at two output ports \\citep{wol02,men03}. Traditionally, the septum polarizer has been known to perform well only within a relatively narrow frequency range, limited by the appearance of resonances. On the other hand, the OMT has an advantage of being able to cover a wide bandwidth, and has been installed in modern telescopes, such as Atacama Large Millimeter/sub-millimeter Array. A further comparison for the two kinds of devices reveals that the septum polarizer is good for measurements of Stokes $Q(\\equiv [\\langle E^{x}E^{x*}\\rangle-\\langle E^{y}E^{y*}\\rangle ]/2)$ and Stokes $U(\\equiv [\\langle E^{x}E^{y*} \\rangle+ \\langle E^{y}E^{x*}\\rangle ]/2)$, or the linear polarization, and the OMT good for those of Stokes $U$ and Stokes $V(\\equiv i[\\langle E^{x}E^{y*} \\rangle - \\langle E^{y}E^{x*}\\rangle ]/2)$, or the circular polarization, where $\\langle ...\\rangle$ are the time average. The reasons are as follows. Consider linearly polarized signals and denote the outputs of the septum polarizer to be the right-hand polarization electric field, $E^R\\equiv (E^x+iE^y)/\\sqrt{2}$ and the left-hand polarization electric field $E^L\\equiv(E^x-iE^y)/\\sqrt{2}$. By cross-correlating $E^R$ and $E^L$, we obtain Stokes $Q$ as $\\langle E^R E^{L*}+E^RE^{L*}\\rangle$ and Stokes $U$ as $\\langle E^RE^{L*}-E^RE^{L*}\\rangle$. In practice, one normally needs to amplify the weak incoming signals, with gains $G_R$ and $G_L$, immediately following the polarizer. The constructed Stokes parameters are in effect $\\langle G_R G_L\\rangle Q$ and $\\langle G_R G_L\\rangle U$, assuming the gains are real. (See Section (6) for a discussion of complex gains.) If one has an approximate knowledge of the average gains $G_R$ and $G_L$, the Stokes $Q$ and $U$ can be determined to an acceptable accuracy. This is what can be achieved with a septum polarizer that measures $E^R$ and $E^L$ directly. On the other hand, if one adopts the OMT, the Stokes $U$ becomes $\\langle G_x G_y\\rangle[\\langle E^x E^{y*} \\rangle+\\langle E^y E^{x*} \\rangle]$ and the Stokes $Q$ becomes $\\langle G_x^2\\rangle \\langle|E^x|^2\\rangle-\\langle G_y^2\\rangle \\langle|E^y|^2\\rangle$. While Stokes $U$ can be recovered in a similar manner as that with a septum polarizer, the recovery of Stokes $Q$ has a serious problem. This is because not only the polarized signal is already mixed with the much stronger unpolarized sky, but the amplifiers also introduce substantial unpolarized noise to the signal. Hence $|E^x|^2$ and $|E^y|^2$ contain almost the unpolarized radiation, and the recovery of weak polarized signals is reminiscent of the determination of a very small number by subtracting two big numbers from each other, for which any small error in the two big numbers will render a poorly determined small number. The recovery of Stokes $Q$ is therefore only possible if the amplifier gains $\\langle G_x^2\\rangle$ and $\\langle G_y^2\\rangle$ can be calibrated to a high accuracy. However, due to the presence of gain fluctuations in amplifiers, a telescope equipped with an OMT is often difficult to yield a well-determined Stokes $Q$. On the other hand, when the polarized signal contains only Stokes $U$ and $V$, a similar argument applies, except for replacing $E^x$ by $E^R$, $E^y$ by $E^L$, and the OMT by the septum polarizer. But, there have rarely been pure circular polarization signals in astronomical observations; hence the OMT is normally disfavored for astronomical polarization measurements and used mostly for the measurements of Stokes $I$. (Nevertheless, a sophisticated solution for polarization measurements with the OMT has been proposed \\citep{men03}.) In spite of the great advantage of septum polarizers, they are not widely used in modern telescopes, simply because the polarizer has long been regarded as a narrow-band device. Therefore, It will be a great leap forward in the polarimeter instrumentation if this major weakness of the septum polarizer can be removed. In this paper, we report our work specifically to address this issue. This paper is organized as follows. We introduce prior works on the septum polarizer in Sec.(2) and highlight the novel approach of the present work. Sec. (3) outlines our design principles. In Sec.(4), we report the measurement results of a polarizer fabricated for test. We convert the measurement results to the mutual leakage of Stokes parameters in Sec.(5). The leading-order calibration for reducing the Stokes $I$ leakage to other Stokes parameters is described in Sec.(6). The conclusion is given in Sec.(7). ", "conclusions": "In this paper, we report a novel design of the septum polarizer that has a $42\\%$ well-performed bandwidth, from 77 GHz to 118 GHz, as opposed to the conventional notion of about $20\\%$ maximum bandwidth. The conventionally alleged maximum bandwidth was derived primarily from a consideration of a limit set by the cutoff frequency of the fundamental modes and the excitation frequency of high-order modes. Our polarizer, adopting a circular waveguide and housing a 5-step septum, is able to break the upper bandwidth limit and extends the usable frequency into the range where high-order $TM_{01}$ modes are excited. This is made possible because we have uncovered an under-explored regime in which $TM_{01}$ excitations can be severely suppressed, and $TM_{01}$ resonances be entirely eliminated. Particularly near the $TM_{01}$ excitation frequency, the septum can manage to avoid multiple reflections of the long wavelength modes inside the polarizer. In addition, the mutual leakage among all four Stokes parameters has been measured. It shows that this septum polarizer performs well, having $I$ to $Q$, $U$ leakage less than $2\\%$ and $Q-U$ mutual leakage less than $\\pm 1\\%$ in almost all frequencies. A few dozen of polarizers of our design have been fabricated by conventional precision machining with $\\pm 5$ $\\mu m$ tolerance, and most of them have similar performances as the two modules reported here. Due to the simplicity of the polarizer without any complicated component to assist widening the bandwidth, this polarizer sets a milestone for the instrumentation of polarization measurements. This septum polarizers will be installed in an upgraded National Taiwan University (NTU)-Array prototype, where each receiver is equipped with 19 pixels of coherent detectors. The $80-116$ GHz signals collected by the upgraded NTU-Array are to be processed by digital correlators that are designed to cover simultaneously the 36 GHz bandwidth for all pixels in all receivers with frequency resolution down to 100 kHz using software correlation. In the context of this work, the backend digital processing capability of the telescope can help calibrate these polarizers with fine frequency resolution." }, "1402/1402.4976_arXiv.txt": { "abstract": "In this letter we examine the evolution of the radial metallicity gradient induced by secular processes, in the disk of an $N$-body Milky Way-like galaxy. We assign a [Fe/H] value to each particle of the simulation according to an initial, cosmologically motivated, radial chemical distribution and let the disk dynamically evolve for $\\sim 6$ Gyr. This direct approach allows us to take into account only the effects of dynamical evolution and to gauge how and to what extent they affect the initial chemical conditions. The initial [Fe/H] distribution increases with $R$ in the inner disk up to $R \\approx 10$ kpc and decreases for larger $R$. We find that the initial chemical profile does not undergo major transformations after $\\sim 6$ Gyr of dynamical evolution. The final radial chemical gradients predicted by the model in the solar neighborhood are positive and of the same order of those recently observed in the Milky Way thick disk. We conclude that: \\item(1) the spatial chemical imprint at the time of disk formation is not washed out by secular dynamical processes, and \\item(2) the observed radial gradient may be the dynamical relic of a thick disk originated from a stellar population showing a positive chemical radial gradient in the inner regions. ", "introduction": "Observations suggest that the negative radial metallicity gradient, $d$[Fe/H]$/dR$, measured in the solar neighborhood close to the plane, flattens or even attains positive values if we select stars at higher $|z|$ from the galactic plane or with high [$\\alpha$/Fe] abundances (Allende Prieto et al. 2006; Boeche et al. 2013, Anders et al. 2013; Hayden et al. 2013). This effect can be a signature of a positive radial gradient in the thick disk population, as derived by Marsakov \\& Borkova (2005) and by Carrell et al. (2012), who analyzed kinematically selected thick disk samples. Alternatively, a similar feature can be simply produced by the variation of the fraction of thin disk and thick disk stars due to their different scale-height and scale-length (Anders et al. 2013). This scenario is indeed consistent with the short scale-length, $h_R\\sim 2$ kpc, recently estimated for $\\alpha$-enhanced stars by Bensby et al. (2011) and Cheng et al. (2012). Furthermore, there is observational evidence indicating that the Milky Way (MW) is a barred galaxy. Non-axisymmetric structures in the disk (such as bars and spiral arms) drive galaxy evolution: they redistribute angular momentum among the different galactic components and can enhance radial migration inside the disk. It is still matter of debate the amount of impact that these secular processes can have on the formation of the thick disk and on its chemical and kinematical properties \\citep{roskar2008, loebman2011, brunetti11, minchev12, Kubryk13}. This work is aimed at analyzing whether dynamical effects can be responsible for the non-negative radial metallicity gradients observed today in the MW thick disk, and suggests that such gradient can be regarded as a cosmological feature. We use an $N$-body disk that develops a bar and whose final population displays a radial metallicity gradient; this is obtained from an initial axisymmetric disk with a positive radial metallicity gradient for $R<10$ kpc, which has dynamically evolved for 6.1 Gyr. This secular evolution model was recently proven \\citep{Curir2012} to reproduce the observed velocity-metallicity correlation in the thick disk \\citep{spagna2010, lee2011}. In this letter, we show that the positive slope of the chemical radial gradient in the MW thick disk observed today in the solar neighborhood might be a signature remaining from the primordial population of disk stars. ", "conclusions": "The crucial role of a positive slope $d{\\rm [Fe/H]}/dR$ in the inner disk for the primordial chemical distribution to produce a positive rotation-metallicity correlation in the thick disk \\citep{spagna2010, lee2011} was already shown in \\cite{Curir2012}. In this work we have analyzed how the same cosmologically motivated chemical distribution allows us to constrain the influence of the dynamical evolution on the metallicity gradients observed today in the MW thick disk. We have drawn the radial and vertical chemical distributions from the same simulated MW galaxy used in \\cite{Curir2012}, which includes a {\\it positive} radial chemical gradient in the inner disk. At 6.1 Gyr, within the solar neighborhood ($7.0$ kpc $ < R < 10.5$ kpc), we still obtain a positive slope for the radial chemical distribution in the inner region. This slope decreases in going from $|z| < 0.5$ kpc to the region $2.5$ kpc $< |z| <3$ kpc ($d{\\rm [Fe/H]}/dR = 0.016\\pm0.002$ dex kpc$^{-1}$ and $d{\\rm [Fe/H]}/dR = 0.011\\pm0.007$ dex kpc$^{-1}$, respectively). The radial chemical gradients predicted by the model are at most within a factor of 2 of those observed by \\cite{Carrell12}, with the Photometric Distance Method, and by \\cite{Hayden13}. Our final positive radial gradients are a direct consequence of our initial conditions, that include a positive radial gradient in the inner disk. As shown in Figure \\ref{fig:grad_evol}, the radial gradients in the solar neighborhood are of the same order from 1.8 Gyr to 6.1 Gyr, independently of the layer in $|z|$ (although some redistribution appears to take place). Such a stability seen between $2$ and $6$ Gyr is consistent with the stability of the rotation-metallicity correlation pointed out by \\cite{Curir2012} in the same time range. Furthermore, the vertical gradient in the final configuration of our simulation is very close to zero. This originates from the fact that we do not impose any dependence on $|z|$ in the initial chemical distribution. The lack of a $|z|$-dependence is also responsible from the lack of a trend for the radial gradients to increase with $|z|$. We conclude that the secular disk evolution does {\\it not} appear to be able to modify significantly the disk global chemical profiles, suggesting that the radial and vertical chemical distribution of the thick disk is likely to be a fossil signature of the original distribution. Our results indicate that, if the positive radial metallicity gradient in the solar neighborhood will be confirmed by the observations of the Milky Way thick disk, this is consistent with a thick disk population showing an early positive [Fe/H] radial gradient in the inner disk ($R < 10$ kpc) and negative in the outer disk ($R > 10$ kpc). In the context of the inside-out formation scenarios of the galactic disk \\citep{spit01,Mott13}, such a gradient ``inversion'' derives from the strong infall of primordial gas that can occur at early times in the inner disk. \\begin{figure*} \\includegraphics[angle=0,scale=0.8,trim=0 0 0 0]{fig3.ps} \\caption{Radial chemical distribution of the disk stars in a few kpcs of the MW solar circle, i.e. 7 kpc$ 1.5$ kpc measured by \\cite{Carrell12} through the Isochrone Distance and the Photometric Distance, respectively.} \\label{fig:gradients} \\end{figure*} \\begin{figure*} \\includegraphics[angle=0,scale=0.8,trim=0 0 0 0]{fig5.ps} \\caption{Evolution of the radial gradient in the solar annulus in several layers in $|z|$, namely $0$ kpc $<|z|< 0.5$ kpc (empty circles), $0.5$ kpc $<|z|< 1.0$ kpc (filled circles), $1.0$ kpc $<|z|< 1.5$ kpc (diamonds), $1.5$ kpc $<|z|< 2.0$ kpc (triangles), $2.0$ kpc $<|z|< 2.5$ kpc (squares), $2.5$ kpc $<|z|< 3.0$ kpc (stars).} \\label{fig:grad_evol} \\end{figure*}" }, "1402/1402.3551_arXiv.txt": { "abstract": "{} {We derive the stellar rotation curve of the Galaxy in the range of Galactocentric radii of $R=4-16$ kpc at different vertical heights from the Galactic plane of $z$ between -2 and +2 kpc. With this we reach high Galactocentric distances in which the kinematics is poorly known due mainly to uncertainties in the distances to the sources.} {We used the PPMXL survey, which contains the USNO-B1 proper motions catalog cross--correlated with the astrometry and near-infrared photometry of the 2MASS Point Source Catalog. To improve the accuracy of the proper motions, we calculated the average proper motions of quasars to know their systematic shift from zero in this PPMXL survey, and we applied the corresponding correction to the proper motions of the whole survey, which reduces the systematic error. We selected from the color-magnitude diagram $K$ vs. $(J-K)$ the standard candles corresponding to red clump giants and used the information of their proper motions to build a map of the rotation speed of our Galaxy.} {We obtain an almost flat rotation curve with a slight decrease for higher values of $R$ or $|z|$. The most puzzling result is obtained for the farthest removed and most off-plane regions, that is, at $R\\approx 16$ kpc and $|z|\\approx 2$ kpc, where a significant deviation from a null average proper motion ($\\sim 4$ mas/yr) in the Galactic longitude direction for the anticenter regions can be directly translated into a rotation speed much lower than at the solar Galactocentric radius. In particular, we obtain an average speed of $82\\pm 5$(stat.)$\\pm 58$(syst.) km/s (assuming a solar Galactocentric distance of 8 kpc, and a circular/azimuthal velocity of 250 km/s for the Sun and of 238 km/s for the Local Standard of Rest), where the high systematic error bar is due mainly to the highest possible contamination of non-red clump giants and the proper motion systematic uncertainty.} {A scenario with a rotation speed lower than 150 km/s in these farthest removed and most off-plane regions of our explored zone is intriguing, and invites one to reconsider different possibilities for the dark matter distribution. However, given the high systematic errors, we cannot conclude about this. Hence, more measurements of the proper motions at high $R$ and $|z|$ are necessary to validate the exotic scenario that would arise if this low speed were confirmed.} ", "introduction": "The rotation curve of the Milky Way disk was obtained using many different tracers and within different Galactocentric distance ranges. Dias \\& L\\'epine (2005) determined the rotation speed of the spiral pattern of the Galaxy by direct observation of the birthplaces of open clusters of stars in the Galactic disk as a function of their age, confirming that the spiral arms rotate like a rigid body, as predicted by the classical theory of spiral waves. Bobylev et al. (2008) used space velocities of young open star clusters and the radial velocities of HI clouds and star-forming regions to derive the Galactic rotation curve in the range of $3 27$), iv) visually checked reliable NB$_J$ and $J$ detections and v) $J-K \\leq 0$. We compute photometric redshifts and remove a significant amount of dusty lower redshift line-emitters at $z \\sim 1.4 $ or $2.2$. A total of 13 Ly$\\alpha$ candidates were found, of which two are marked as strong candidates, but the majority have very weak constraints on their SEDs. Using follow-up observations with SINFONI/VLT we are able to exclude the most robust candidates as Ly$\\alpha$ emitters. We put a strong constraint on the Ly$\\alpha$ luminosity function at $z \\sim 9$ and make realistic predictions for ongoing and future surveys. Our results show that surveys for the highest redshift LAEs are susceptible of multiple contaminations and that spectroscopic follow-up is absolutely necessary. ", "introduction": "Finding the first stars and galaxies is one of the most important tasks to test our understanding of galaxy formation in the early Universe. The current theoretical models of when and how these first galaxies were formed can only be tested and improved by reliable detections of galaxies at the highest redshifts. The confirmation of galaxies at a redshift of $z\\sim 9-10$ would also allow the study of the epoch of reionization of the Universe. Measurements of the cosmic microwave background place this epoch at $z \\sim 10.6$ \\citep{Komatsu2011}, while \\cite{Fan2006} located the end of the reionization epoch at a redshift of at least $z\\sim 6$ by studying spectra of quasars at high redshift, where they found a lower limit to the neutral fraction of $\\sim 10^{-3} - 10^{-2}$. A widely used technique to detect very distant galaxies is the Lyman break technique (LBG), pioneered by \\cite{Steidel1996} (see also \\citealt{Guhathakurta1990}), which looks at a distinctive break in the UV spectrum of star-forming galaxies. More generally, one can use deep data in several broadbands to derive a redshift-probability distribution by fitting spectral energy distributions (SED) based on galaxy templates \\citep[e.g.][]{McLure2011}. Using the Lyman Break method, candidate galaxies have been found at very high redshifts \\citep[$z \\sim 7$, e.g.][]{Bouwens2011z7,Finkelstein2012,Oesch2012z8,McLure2012} and even $z\\sim10$ \\citep[]{Ellis2013,Oesch2013,Bouwens2013}, but the great majority of these are too faint to confirm spectroscopically. \\cite{Lehnert2010} claimed the spectroscopic detection of a $z=8.6$ Lyman-$\\alpha$ (Ly$\\alpha$) line of a LBG in the Hubble Ultra Deep Field. However, \\cite{Bunker2013} were unable to reproduce the detection with two independent sets of observations, leading to the suggestion that it could be an artefact. \\cite{Brammer2013} found a tentative emission line that could be Ly$\\alpha$ at $z = 12.12$ using the HST WFC3 grism, but this is only a $<3 \\sigma$ detection and could be a lower redshift interloper. Recently, \\cite{Finkelstein2013} report the detection of a Ly$\\alpha$ emission line in a $z=7.51$ LBG, although the line is very close to a sky-line making identification significantly more difficult. Other attempts have been made, but so far no $z>7.5$ galaxy has been spectroscopically confirmed. There is a spectroscopic redshift determination of a $z=8.2$ Gamma Ray Burst \\citep{Tanvir2009}, but not for its host. Another successful technique to detect very high redshift ($z \\sim 4- 7$) galaxies is the narrow-band technique, which targets Lyman-$\\alpha$ emitters (LAEs; e.g. \\citealt{Pritchet1994,Thompson1994,Thompson1995,Hu1996,Cowie1998,Hu1998,Thommes1998,Rhoads2000,Rhoads2003,Rhoads2004,Fynbo2001,Hu2002,MalhotraRhoads2002,MalhotraRhoads2004,Fynbo2003,Ouchi2003,Hu2004,Taniguchi2005,Iye2006,Kashikawa2006,Shimasaku2006,Ouchi2008,Finkelstein2009,Ota2010,Hibon2011}). Using the narrow-band technique one can search for sources with emission lines at specific redshifts, by looking at the excess the narrow-band has over the broadband. This way sources for which the continuum is too faint to be detected, can still be identified due to the bright emission lines. However, most emission line galaxies detected in narrow-band surveys are lower redshift interlopers such as H$\\alpha$ and [O{\\sc ii}] \\cite[e.g.][]{Sobral2012}, which have to be identified using multiwavelength observations. Because the narrow-band is only sensitive to sources emitting in a small range of wavelengths, they can be used to look at a slice of redshifts and therefore a well-known comoving volume. Moreover, spectroscopic follow-up of high-redshift candidates is a priori easier for candidates detected by the narrow-band technique, as these candidates will have strong emission lines. Currently, the most distant spectroscopically confirmed NB-selected LAE is at a redshift of 6.96 \\citep{Iye2006}, which is detected with narrow-band imaging from the Subaru telescope. \\begin{figure} \\centering \\includegraphics[width=8cm]{./REPORT_FIGURES/transmission.png} \\caption{\\small{The atmospheric transmission in the near-infrared $J$ band, normalised to the maximum transmission for each curve. Atmospheric data is from Mauna Kea, Gemini Observatory \\citep{Lord1992}, the airmass is 1.0 and water vapour column is 1.0 mm. The background (grey) atmospheric (OH) emission lines are also shown. The NB$_J$ filter used in this paper is transparent at wavelengths where there are no strong OH lines and also at wavelengths where the atmosphere is at its maximum transparency, thus allowing us to obtain deep observations in relatively little time.}} \\label{fig:filters} \\end{figure} To observe even higher redshift galaxies, observations in the near-infrared are required. Unfortunately, at these wavelengths there is significant foreground emission due to OH molecules in the Earth's atmosphere. Some OH windows exist at wavelengths where the atmosphere is transparant to radiation. It is possible to observe using narrow-band filters in these windows very effectively and several filters have been developed for this purpose (see Figure $\\ref{fig:filters}$). Recent studies led to the identification of candidate Ly$\\alpha$ emitters at a redshift of $z = 7.7$, but none of these has been spectroscopically confirmed yet \\citep[]{Hibon2010,Tilvi2010,Clement2012,Krug2012,Jiang2013b}. Some attempts at somewhat higher redshifts ($z\\sim9$) were made to detect Ly$\\alpha$ \\citep[]{WillisCourbin2005,Cuby2007,Willis2008,Sobral2009b}. The properties of such galaxies would provide strong tests of current models of galaxy formation and evolution and even the confirmation of just one luminous Ly$\\alpha$ emitter at this redshift will be suitable for the study of these sources way before the next generation of telescopes, such as \\emph{JWST} or the E-ELT. Ly$\\alpha$ radiation is much more attenuated by a neutral intergalactic medium (IGM) than an ionized IGM, so large samples of Ly$\\alpha$ emitters at these redshifts could be used to derive properties of the IGM at these early times. Current simulations \\citep[e.g][]{Iliev2008} suggest that reionization started at the most overdense regions in the Universe, where ionizing sources nurtured expanding shells of ionized gas in the IGM. As Ly$\\alpha$ radiation is easily absorbed by a neutral medium \\citep{MalhotraRhoads2004}, Ly$\\alpha$ emitters can only be observed once the ionized zone around them is large enough for the Ly$\\alpha$ radiation to escape. This is expected to lead to a negative evolution in the Ly$\\alpha$ luminosity function and dropping escape fraction of Ly$\\alpha$ radiation at higher redshifts. Considerable effort has been put in spectroscopically studying the evolution of the Ly$\\alpha$ line in Lyman break galaxies (LBGs) at high redshifts \\citep[e.g.][]{Fontana2010,Pentericci2011,Vanzella2011,Ono2012,Schenker2012,Caruana2013,Finkelstein2013}. Recent non-confirmations and low success-rates at $z>7$ for their spectroscopic confirmation are interpreted as a signature that reionization is not yet completed at these redshifts. \\cite{Treu2013}, for example, find that at $z\\sim8$ Ly$\\alpha$ emission of LBGs is suppressed by at least a factor of three. For LAEs, it is found that up to at least a redshift of $z\\sim6$ the Ly$\\alpha$ luminosity function is remarkably constant \\citep[e.g.][]{Shimasaku2006,Hu2004,Ouchi2008}. This indicates that LAEs are relatively more common and more luminous at earlier epochs, compared to LBGs (as the UV LF drops quickly in this redshift range; \\citealt{Bouwens2007}). At $z\\sim6-8$ there is evidence for evolution of the characteristic luminosity, but these samples, including failed attempts at $z=7.7$, can be significantly affected by cosmic variance, probing $\\lsim 1$ deg$^2$ \\citep[e.g.][]{Ouchi2010,Clement2012}. At the bright end, however, the evolution could plausibly be very different. Luminous sources can ionize their own surroundings to allow Ly$\\alpha$ photons to escape, as they redshift out of restframe-resonance wavelength in about 1 Mpc \\citep[e.g.][]{Barton2004,CenHaiman2000,Curtis-Lake2012}. Furthermore, the observed clustering of Ly$\\alpha$ emitters is expected to increase at higher redshift, as neighbouring sources will have larger overlapping ionized spheres and therefore a higher fraction of escaped Ly$\\alpha$ photons \\citep[e.g.][]{Ouchi2010}. In order to find the most luminous Ly$\\alpha$ emitters in the epoch of reionization which would be suitable for spectroscopic follow-up, we have undertaken the widest area search with a near infrared narrow-band filter to date. This paper is organised in the following way. \\S2 presents the details of the observations, and describes the data reduction, calibrations and source extraction. \\S3 presents the criteria for sources being selected as Ly$\\alpha$ candidates and the results from the narrow-band search. \\S4 presents the spectroscopic follow-up observations and results. \\S5 discusses the results such as constraints on the Ly$\\alpha$ $z=8.8$ luminosity function, and our survey is compared to past and future surveys. Finally, \\S6 outlines the conclusions. A H$_0=70$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_M=0.3$ and $\\Omega_{\\Lambda}=0.7$ cosmology is used and all magnitudes are in the AB system, except if noted otherwise. ", "conclusions": "We have conducted a very wide narrow-band survey over 10 deg$^2$ in the near infrared and identified 6315 line-emitters using a $1.19 \\mu$m narrow-band filter. In this work we identified possible $z = 8.8$ Ly$\\alpha$ candidates in the sample of line-emitters and followed-up the strongest ones spectroscopically. The main conclusions are: \\begin{itemize} \\item A significant fraction ($\\sim 300$) of the line-emitters are consistent with being at high redshift ($z > 3$), of which some might be Ly$\\alpha$ at $z = 8.76$. This narrow-band survey increased the probed volume by half an order of magnitude compared to previous surveys and is thus sensitive to the rarest and most luminous sources. \\item By doing careful visual checks of the robustness of the detections and by excluding line-emitters which are detected in any of the optical bands and which show a red $J-K$ colour, we find 13 possible Ly$\\alpha$ candidates. We order them in different groups based on their broadband photometric constraints. The two most robust candidates have reliable detections in narrow-band, strongest constraints from photometric redshifts, $iz-J$ break and robust $J$ detections. \\item 90\\% of the high redshift candidate line-emitters, selected on having no/very faint flux in the optical, are lower redshift interlopers. By including the $K$ band and computing photometric redshifts we find that approximately 40\\% are H$\\beta$/[O{\\sc iii}] at $z=1.4$, 25\\% [O{\\sc ii}] at $z=2.2$ and 15\\% are candidate AGN emission-lines (e.g. carbon or magnesium) at $z\\sim3-6$, while the remaining are likely very faint lower redshift sources like Pa$\\gamma$ at $z = 0.09$ or He{\\sc ii} at $z = 0.44$. \\item Spectroscopic follow-up of the two most robust Ly$\\alpha$ at $z=8.8$ candidates, two sources with the largest EWs and another with brightest J failed to confirm these sources as line-emitters. This is probably caused by a combination of spurious sources, variability and (although unlikely) solar system objects. This result has very strong implications to current and future candidates for LAEs at $z=7.7$ and $z>8$. \\item After the follow-up, we put the strongest constraints on the bright end of the luminosity function with half an order of magnitude improvement in the probed volume and it could still mean little to no evolution in the luminous end. \\item Using an optimistic upper-limit to the LF and a lower redshift extrapolation, we estimate the number of LAEs that will be detected by the completed VISTA NB118 GTO survey to be between $0.001$ and $1.19$ with the current depth and to be between $0.19$ and $23.35$ for the estimated final depth of the ongoing UltraVista NB118 survey. \\item Because of the lack of a comparably wide surveys, it is difficult to study the evolution of the bright end of the LF and extrapolations from other considerably smaller surveys at lower redshifts are unusable. Although the number density of Ly$\\alpha$ emitters is expected to decline at higher redshifts, this isn't necessary the case for the bright end of the luminosity function, because of the topology of reionization. It is therefore of utmost importance to study the bright end of the Ly$\\alpha$ luminosity function at lower redshifts in order to understand the evolution in the LF completely. \\item As our strongest candidates looked realistic in the images and had realistic physical properties based on the photometry, but still are not confirmed, we highlight the necessity for all other surveys to do this spectroscopic follow-up, especially when candidates are based on just a single-band detection. This has significant consequences for any similar and for deeper surveys, clearly pointing out that despite sources passing all tests, only spectroscopic observations can confirm them. \\end{itemize}" }, "1402/1402.1762_arXiv.txt": { "abstract": "The interest on the solar diameter variations has been primary since the scientific revolution for different reasons: first the elliptical orbits found by Kepler in 1609 was confirmed in the case of the Earth, and after the intrinsic solar variability was inspected to explain the climate changes. The CCD Solar Astrolabe of the Observat\\'orio Nacional in Rio de Janeiro made daily measurements of the solar semi-diameter from 1998 to 2009, covering most of the cycle 23, and they are here presented with the aim to evidence the observed variations. Some instrumental effects parametrizations have been used to eliminate the biases appeared in morning/afternoon data reduction. The coherence of the measurements and the influence of atmospheric effects are presented, to discuss the reliability of the observed variations of the solar diameter. Their amplitude is compatible with other ground-based and satellite data recently published. ", "introduction": "The interest on the solar diameter variations has been prioritary during the last four centuries of astronomy with and without the telescope. During 1600 the verification of the Keplerian hypothesis of elliptical Earth's orbit was possible thanks to measurements of the seasonal variations of the angular diameter of the Sun made by G. D. Cassini and G. B. Riccioli from 1655 to 1666 at the {\\sl Heliometer} of S. Petronio in Bologna. This was a pinhole solar telescope\\footnote{The use of the word {\\sl telescope} for an optics-less instrument like a giant pinhole camera realized into a church provided by a meridian line accurately draft on the ground, may appear awkward. But the Greek word's meaning is related to the capability of seeing at far objects, and the implicit possibility to have magnified images, and this is possible with pinhole instruments.} of the Basilica of St. Petronius in Bologna \\cite{Manfredi}. Later in 1700 the interest of the astronomers was more focused on the Earth's axis orientation, with the variation of the obliquity $\\epsilon$. To measure this parameter it was necessary to know better the position of the solar center, than the diameter. The giant pinhole telescopes \\cite{Heilbron} could operate on a complete yearly range of zenithal distances of $\\sim 47^{o}=2\\times\\epsilon$ without optical distortions, excepted the atmospheric refraction, and for this reason these instruments continued successfully their operations until the end of the century even if their imaging capability was comparatively lower than usual telescopes. The heliometer as we know nowadays was invented in 1743 by S. Savery \\cite{Short1753} and built by J. Dollond ten years later, but the solar diameter has been measured with timing meridian transits at the Greenwich Observatory, starting with N. Maskelyne from 1765 to 1810, and the influence of a variable {\\sl personal equation} with the age, either in reaction time either in contrast sensitivity, has been recognized as the most reliable cause of the recorded variations. The heliometric method, which does not deal with reaction times, but only with the personal contrast sensitivity, was considered more accurate than \\cite{Auwers1890}. The nineteenth century was the golden age of the solar astrometry, and the heliometers built by J. Fraunhofer were so accurate that were used by F. W. Bessel for the measurement of the first stellar parallax \\cite{Bessel1838}. The measurements of the solar diameter variations became a routine with specific instrument designed for, and the observatories involved in the meridian transit measurements were also Neuchatel, Oxford, Washington \\cite{Auwers1890}, and Rome-Campidoglio \\cite{Gething1955}. The current standard value of the solar semi-diameter has been fixed to 959.63 arcsec\\footnote{This angular dimension corresponds to 696000 Km at 1 AU.} by A. Auwers in 1891 \\cite{Auwers1891}. The words of A. Secchi in the book \"Le Soleil\" expressed clearly the status of art at the fall of the century \\cite{Secchi1875}: {\\sl What strikes even more is to see that, despite the variety of methods and instrumental perfection, the measurement of the solar diameter made very little progress.}\\footnote{Ce qui frappe encore davantage, c'est de voir que, malgr\\'e la vari\\'et\\'e des m\\'ethodes et la perfection des instruments, la mesure du diam\\`etre solaire a fait bien peu de progr\\`es (Secchi, 1875 p. 210).} ", "conclusions": "The aforementioned words of A. Secchi in 1875 could be applied also nowadays, even if meanwhile the errorbars of these measurements have been reduced of more than a factor of 10-100. The influence of atmospheric turbulence and the environmental parameters can affect a single measurements with perturbations bigger than the real variations of the solar diameter. The era of satellites opened the way to the milliarcsecond solar astrometry, but for the knowledge of the solar diameter fluctuations on timescales longer than the solar cycle, their lifetime is very short and their operational cost bigger than astrolabes and heliometers. The Solar Heliospheric Observatory SOHO has the longer operating time until now, but its instruments were not designed for astrometry. {\\it Nonetheless its data have been exploited to show that the solar diameter has not changed within a few milliarcsecond per year...} (Delmas et al. (2006) p. 1567 p. 4.) Because of the nature of planetary transit measurements of solar diameter, an excellent angular resolution is achievable by a good timing resolution. Hence the Mercury transits of 2003 and 2006 observed by SOHO provided the occasion to measure the solar diameter based only on timing accuracy independently on angular references \\cite{Emilio12}. Therefore these measurements made with SOHO in 2003 and 2006, 960.08 arcsec, can be fairly compared with the analogous measurements made by PICARD satellite during the transit of Venus of 2012 in the same waveband, 959.86 arcsec, concluding that the variations of the diameter from 2003 to 2012 has been $0.22\\pm0.10$ arcsec. This is in perfect agreement with the variations measured by the Rio Astrolabe during the cycle 23, and this may explain directly the scatter between the yearly averages found in the aforementioned series of Greenwhich, Campidoglio, Rio de Janeiro and Calern. The astrolabe of Rio de Janeiro allowed to monitor the solar during more than a decade without the costs of a space mission and its limited operating lifetime. Future results from Hinode, SDO and PICARD satellite can better elighten this field of research, and instruments like PICARD-Sol and the Reflecting Heliometer of Rio de Janeiro will help to maintain the 0.01 arcsec of accuracy in solar astrometry typical of the satellites to ground-based observations, to prolong the series on the solar diameter for the years to come. The measurements of the solar diameter made with the solar Astrolabes have been debated since many decades. In this paper all original data of the observations made in Rio de Janeiro over the whole cycle 23 have been here presented as appendix from \\cite{Boscardin2011} and discussed. The scope was also to share with the international heliophysics community these data with the problems of their interpretations. Since this kind of problems are common with all ground-based observations \\cite{Gething1955} their final solution will take full advantage also of these data, to verify, for example, the hypotheses on the possibility that the solar activity could influence indirectly the measurements as some scholars claimed \\cite{Rozelot2003}. \\begin{acks} C.S. thanks J.P. Rozelot and W. Dziembowski for fruitful discussions and suggestions. \\end{acks}" }, "1402/1402.2282_arXiv.txt": { "abstract": "{ Comprehensive studies of Wolf-Rayet stars were performed in the past for the Galactic and the LMC population. The results revealed significant differences, but also unexpected similarities between the WR populations of these different galaxies. Analyzing the WR stars in M31 will extend our understanding of these objects in different galactic environments. }{ The present study aims at the late-type WN stars in M31. The stellar and wind parameters will tell about the formation of WR stars in other galaxies with different metallicity and star formation histories. The obtained parameters will provide constraints to the evolution of massive stars in the environment of M31. }{ We used the latest version of the Potsdam Wolf-Rayet model atmosphere code to analyze the stars via fitting optical spectra and photometric data. To account for the relatively low temperatures of the late WN10 and WN11 subtypes, our WN models have been extended into this temperature regime. }{ Stellar and atmospheric parameters are derived for all known late-type WN stars in M31 with available spectra. All of these stars still have hydrogen in their outer envelopes, some of them up to 50\\% by mass. The stars are located on the cool side of the zero age main sequence in the Hertzsprung-Russell diagram, while their luminosities range from $10^{5}$ to $10^{6}\\,L_{\\odot}$. It is remarkable that no star exceeds $10^{6}\\,L_{\\odot}$. }{ If formed via single-star evolution, the late-type WN stars in M31 stem from an initial mass range between $20$ and $60\\,M_{\\odot}$. From the very late-type WN9-11 stars, only one star is located in the S Doradus instability strip. We do not find any late-type WN stars with the high luminosities known in the Milky Way. } ", "introduction": "\\label{sec:intro} M31 or the Andromeda galaxy is the largest member of the Local Group, which contains only two other spiral-type galaxies, M33 and the Milky Way. Because of the very low foreground extinction towards M31 and the known distance, this galaxy is ideal for studying bright, resolvable stellar objects, such as Wolf-Rayet (WR) stars. Analyzing the WR stars in M31 will extend our knowledge about the formation and evolution of massive stars in other galaxies and allow for a comparative analysis of their WR populations. Previous studies chiefly focused on the WR population in the Milky Way and the Magellanic Clouds. One of the reasons for this limitation was the paucity of available spectra of WR stars in other galaxies. The advent of multi-object spectroscopy has greatly facilitated such observations. Good quality spectra of extragalactic WR stars are now becoming available, allowing for qualitative analyses. A large set of WR spectra for M31 have been published by \\citet{NMG2012} and provided the basis for this work. Photometric data of extragalactic WR stars are contained in the Local Group Galaxy Survey (LGGS) by \\citet{Massey+2006}. Spectroscopically, the WR stars are divided into the WN, WC, and WO subclass. The WN stars show prominent nitrogen emission lines in their spectra while the WC and WO stars have prominent carbon and oxygen emission lines. In contrast to the WC stars, the oxygen emission lines are significantly stronger in the WO star spectra, including a prominent \\ion{O}{vi} emission line at $3811$-$34\\,$\\AA. All subclasses are further split into a sequence of subtypes, defined by the equivalent width or peak ratio of certain emission lines \\citep{CHS1995, vdH2001}. For the WN stars, these are the subtypes WN2 to WN11. The WN2 to WN6 subtypes are also called ``early'' types, while the WN8 to WN11 are referred to as ``late'' subtypes. The WN7 subtype is something in between, although it is often included under the late subtypes. The WN9 to WN11 subtypes were introduced by \\citet{CHS1995} as a finer replacement of the former Ofpe/WN9 classification and are sometimes referred to as ``very late'' WN types (WNVL: very late WN). In contrast to the early subtypes, late-type WN stars typically have significant amounts of hydrogen \\citep{HGL2006}. In the past years, hydrogen-rich WN stars, sometimes classified as WNh, have become an interesting topic of research. Studies such as \\citet{Graefener+2011} have pointed out that at least a significant fraction of these stars are not ``classical'' WR stars in the sense that they are stripped cores of evolved massive stars. Instead, WN stars of this kind are most likely core-hydrogen burning, and they form the most luminous group within the WR population. The most luminous WR stars in the Large Magellanic Cloud (LMC) and the Milky Way are in fact such hydrogen-rich WN stars \\citep{BOH2008,Crowther+2010}. In this paper we analyze the late-type WN stars with subtypes ranging from WN7 to WN11. Apart from two stars with insufficient spectra, we cover the whole known sample of late-type WN stars in M31 with available optical spectra. This present paper builds on our recent work that analyzes the WR star populations in the Milky Way \\citep{HGL2006,SHT2012} and the LMC \\citep{Hainich+2014}. Extending the analysis to galaxies beyond the Milky Way may in principle allow study of WR stars at higher metallicities. According to \\citet{ZKH1994}, the metallicity in M31 has a value of $\\log(\\text{O/H}) + 12 = 8.93$, which is a bit higher than the solar value of $8.7$ \\citep{EP1995}. However, as in our Milky Way, the metallicity is not uniform across the whole galaxy but instead increases towards the central bulge. Figure\\,\\ref{fig:wnpos} shows the positions of the analyzed stars in M31. It is significant that none of the known late-type WN stars in M31 are located in the inner part of this galaxy. Instead the stars mostly reside in a zone between 9\\,kpc and 15\\,kpc from the center of M31. It is exactly this region that \\citet{vdB1964} found to be the most active in forming stars. Because of their non-central location it might be that the metallicity of our sample is not significantly higher than the one used in our Galactic WN models \\citep[cf.][]{HG2004}. The distance to M31 is well known \\citep[$d\\,=\\,0.77\\,$Mpc,][]{CNG}. This offers a big advantage for analyzing of stellar populations. Among the Galactic stars, there is only a limited subsample for which distances can be inferred from cluster or association membership. Analyzing the M31 WN stars will therefore allow us to crosscheck our findings from the Galactic WR analyses and compare them with the LMC results where we have a different type of galaxy and lower metallicity. In the next section, we briefly characterize the stellar wind models used in this work. In Sect.\\,\\ref{sec:sample} we report the results and compile the obtained stellar parameters. In Sect.\\,\\ref{sec:evol} we discuss the results and compare them with theoretical stellar evolution tracks. The conclusions are drawn in the Sect.\\,\\ref{sec:conclusions}. In the appendix (Sect.\\,\\ref{appsec:specfits}) we show all spectral fits obtained in this work. \\begin{figure}[ht] \\resizebox{\\hsize}{!}{\\includegraphics[angle=0]{m31wnpos}} \\caption{The positions of the analyzed WN stars in M31. Numbers correspond to the M31WR numbers as in Table\\,\\ref{tab:fitresults}. The background image is taken from Palomar STScI DSS. The size of the image is $\\approx 1.9^\\circ \\times 2^\\circ$. The gray dashed ellipse roughly corresponds to a radius of $9\\,$kpc around the center of M31, corresponding to the inner boundary of the region with the most star formation \\citep{vdB1964}.} \\label{fig:wnpos} \\end{figure} ", "conclusions": " \\begin{itemize} \\item All stars in our sample have luminosities $L$ between $10^5$ and $10^6\\,L_{\\odot}$. Notably, we do not find any star in our sample that exceeds $10^6\\,L_{\\odot}$. \\item The absolute visual magnitude $M_{\\text{V}}$ shows a significant scatter, even within the same WN subtype. This sets limits to the applicability of a subtype magnitude calibration. \\item Since the current catalog of WN stars in M31 does not include the central region of this galaxy, we cannot rule out that this special environment hosts WN stars with luminosities $L > 10^6\\,L_{\\odot}$, especially since we know that such stars exist in the LMC and the Galactic center region. \\item The late-type WN stars in M31 are in good agreement with the latest set of the Geneva tracks with $Z = 0.014$. Only the Geneva tracks with rotation can reproduce the lowest luminosities in our sample. Otherwise, the different sets of tracks \\citep{Georgy+2012,EIT2008} can all explain the observed WNh stars because these stars are not yet in the final WR stages for which the tracks differ more. \\item If formed via single-star evolution, the analyzed late-type WN stars in M31 stem from an initial mass range between $20$ and $60\\,M_{\\odot}$. The lower limit might in fact be higher if the few stars with the lowest luminosities were formed via binary evolution. \\item Our spectral fits reveal that the analyzed late-type WN stars have similar chemical compositions to their Galactic counterparts. This is in line with the fact that the stars are not located in the inner part of M31, where we would expect higher metallicities. The obtained HRD positions of the analyzed stars do not provide additional constraints since they can be reproduced by evolutionary models with both Galactic and higher metallicity. \\item The number of very late WN stars (WN9-11) is low, which could partly be a selection effect of the detection method \\citep{NMG2012}. While a lot of additional WN9 stars have been discovered in the Milky Way in obscured regions with infrared observations, the number of currently known WN10 and WN11 stars in M31 is about the same as in the Milky Way and the LMC. \\item Only one object, the WN10 star \\#148, is located in the S\\,Dor instability strip in the HR-diagram. However, the HRD positions do not rule out that some of the stars of very late-type WN subtypes might be LBVs in the quiescent stage. From the luminosity range, it is hard to tell whether the WN stars have lost parts of their hydrogen via an LBV outbursts or in the red supergiant (RSG) phase. \\end{itemize} In a forthcoming paper we will analyze the early WN subtypes to get a more complete picture of the WN population in M31." }, "1402/1402.5694_arXiv.txt": { "abstract": "We analyzed 3 years of data from the \\kepler\\ space mission to derive rotation periods of main-sequence stars below 6500\\,K. Our automated autocorrelation-based method detected rotation periods between 0.2 and 70 days for 34,030 (25.6\\%) of the 133,030 main-sequence \\kepler\\ targets (excluding known eclipsing binaries and Kepler Objects of Interest), making this the largest sample of stellar rotation periods to date. In this paper we consider the detailed features of the now well-populated period-temperature distribution and demonstrate that the period bimodality, first seen by McQuillan, Aigrain \\& Mazeh (2013) in the M-dwarf sample, persists to higher masses, becoming less visible above 0.6\\,$M_\\odot$. We show that these results are globally consistent with the existing ground-based rotation-period data and find that the upper envelope of the period distribution is broadly consistent with a gyrochronological age of 4.5\\,Gyrs, based on the isochrones of \\citet{bar07}, \\cite{mam+08} and \\cite{mei+09}. We also performed a detailed comparison of our results to those of \\cite{rei+13} and \\cite{nie+13}, who have measured rotation periods of field stars observed by \\kepler. We examined the amplitude of periodic variability for the stars with detected rotation periods, and found a typical range between $\\sim 950$\\,ppm ($5^{\\rm th}$ percentile) and $\\sim 22,700$\\,ppm ($95^{\\rm th}$ percentile), with a median of $\\sim 5,600$\\,ppm. We found typically higher amplitudes for shorter periods and lower effective temperatures, with an excess of low-amplitude stars above $\\sim 5400$\\,K. ", "introduction": "Throughout their main-sequence lifetime, stars lose angular momentum and spin down. For intermediate- and low-mass stars, this angular momentum loss is thought to occur via a magnetized wind that is linked to the stellar outer convection zone \\citep[e.g.,][]{kaw88, bou+97}. Therefore, the present-day rotation period reflects the integrated angular momentum loss history of the star. It has been claimed that for main-sequence stars, the present day rotation period does not depend on the initial conditions, and therefore a tight relationship exists between the period, age and mass, which is known as gyrochronology, for which a range of empirical relations have been derived \\citep[e.g.,][]{kaw89, bar03, bar07, bar10, mam+08}. In recent years there has also been considerable work towards understanding and modeling the underlying physics behind gyrochronology \\cite[][and references therein]{rei+12, mat+12, gal+13}. Therefore, deriving the rotation period for a large number of main-sequence stars has been a long-standing goal in stellar astrophysics, with the potential to shed light on stellar, and even planetary system angular-momentum evolution on one hand, and provide a new method to probe Galactic star formation history and structure on the other hand. In recent years, the study of stellar rotation has seen significant advances thanks to new observational programs and techniques. Previously, measurements of stellar rotation were typically performed using spectroscopy, via the rotational broadening of absorption lines \\cite[e.g.,][]{kal89}. However, this technique yielded only model-dependent constraints on the rotation rate due to the dependence on stellar radius and inclination, and were limited to relatively fast rotators. The era of exoplanet transit surveys provided a new approach to stellar rotation studies, making it possible to measure the rotation period directly for a large number of stars simultaneously. This is done by detecting quasi-periodic brightness variations in the time series photometry, caused by magnetically active regions repeatedly crossing the visible hemisphere as the star rotates \\citep[e.g.,][]{irw+09a}. Open cluster surveys over the past decade have provided thousands of rotation periods for low-mass stars with ages up to $\\sim 650$\\,Myr, an overview of which can be found in \\cite{irw+09} and \\cite{bou+13}. More recent additions to the literature on rotation of young low-mass stars include studies of M35 (310 periods) \\citep{mei+09}, Coma Berenices \\citep{col+09}, M34 (83 periods) \\citep{mei+11}, and h Persei (586) \\cite{mor+13}. Together, these data have provided a relatively complete, but complex picture of pre-main-sequence rotational evolution. Until recently, rotation periods for older stars remained scarce, since their slow rotation and low-amplitude modulation make them difficult to detect with either spectroscopy or ground-based photometry. Notable exceptions include $112$ main sequence F, G and K stars observed as part of the Mount Wilson H-K project \\citep[][and references therein]{bal+96}, $1727$ mid-F to mid-K stars observed by the {\\it CoRoT}\\ satellite \\citep{aff+12}, 41 low-mass ($0.1$--$0.3\\,M_\\odot$) stars from the \\mearth\\ survey \\citep{irw+11}, and $\\sim 2000$ field K and M stars observed by the HATNet survey \\citep{har+11}. The periods of \\cite{aff+12} are based on up to 5 months of CoRoT data and yet include a significant fraction of periods greater than 60 days. Since many of these periods are more than half the length of the dataset, they should be treated with caution and we opted to omit these results from the comparison section of this paper. Asteroseismology studies performed on a small subset of \\kepler\\ and CoRoT targets also provide an insight into stellar rotation, since the oscillation spectrum of a star can reveal information about surface and interior rotation rates \\citep[e.g.,][]{giz02, giz+04}. For a recent review on observational studies of stellar rotation, see \\cite{bou+13}. Data from the Kepler space mission \\citep{bor+10, koc+10} are revolutionizing the study of stellar rotation, providing 4\\,yrs of almost uninterrupted photometry with an unprecedented level of precision and time sampling, for a very large sample of stars. This allows rotation periods to be detected for slowly rotating stars with low-amplitude modulation, for a well-defined large sample. Rotation studies on subsets of the \\kepler\\ data include the measurement of 265 rotation periods for stars with $T_{\\rm eff} \\le 5200\\,$\\,K and $\\log g \\ge 4.0$\\,dex observed by \\kepler\\ for 1--2 quarters through the Cycle 1 Guest Observer program \\citep{har+12}, 1570 M-dwarf rotation periods using quarters 1--4 (Q1--4) \\citep{mcq+13a}, 737 of the Kepler Objects of Interest (KOIs) using Q3--14 data \\citep{mcq+13b}, $\\sim 950$ of the KOIs using Q9 \\citep{wal+13}. Two previous studies focussing on the broader \\kepler\\ sample are those of \\cite{rei+13}, with an emphasis on differential rotation, who derive $\\sim 24,000$ periods using Q3, and \\cite{nie+13}, who measured $\\sim 12,000$ periods from Q2--Q9, and compare to previous spectroscopic studies. These studies, discussed in Section~\\ref{sec:disc}, use Fourier-based detection methods and conservative automatic selection criteria. In this paper we describe the design and application of AutoACF, an automated version of the autocorrelation function (ACF) method introduced by \\cite{mcq+13a} for stellar rotation period measurement. Using AutoACF we derived 34,030 rotation periods from 133,030 main-sequence targets observed by \\kepler. We outline the sample selection in Section~\\ref{sec:data} and the period detection method in Section~\\ref{sec:per_det}. In Section~\\ref{sec:res} we present our results, and examine the mass-period and temperature-period distributions in Section~\\ref{sec:per_mas_teff}, and the amplitude of periodic variability in Section~\\ref{sec:amp_per}. In Section~\\ref{sec:gyro} we compare our results to empirical gyrochronology relations, and in Section~\\ref{sec:comp} we compare our results to those of \\cite{rei+13} and \\cite{nie+13}. We summarize our results in Section~\\ref{sec:disc} and provide full details of the ACF automation method in the Appendix. ", "conclusions": "\\label{sec:disc} The present catalog, of more than 30,000 rotation periods and photometric amplitudes, was derived using a systematic automated search from a well defined sample of more than 133,000 main-sequence stars observed by \\kepler, all of which have stellar temperature estimates. This sample can help us study the mass-age-period relation, and better understand its implication on the amplitude of photometric variability. One clear feature of the sample is the fraction of rotation period detections as a function of temperature, which changes dramatically, from $\\sim 80$\\% for the cool stars to $\\sim 20$\\% for the hot stars of the sample. It seems that the photometric amplitude of the stellar rotation signature for the majority of the hot stars in the sample is smaller than our detection limit. This is supported by Figure~\\ref{fig:teff_amp}, which shows that the detected amplitudes for the hot stars have a range that goes down to $\\sim 300$\\,ppm, below which we do not detect periodicity. For the cool stars, on the other hand, the lower limit of the amplitudes is $\\sim 2000$\\,ppm, well above our detection threshold. We might suggest that the incompleteness of our detections for the cool stars is therefore related to the orientation angle of the stellar axis of rotation relative to our line of sight, which prevents us from detecting the rotation of all cool stars in the sample. Obviously, the photometric amplitude also carries information on the stellar surface inhomogeneity, presumably associated with the stellar magnetic field. Small amplitudes of variability can either result from smaller star spots or from a more even coverage of the stellar surface by spots. Therefore, our sample suggests that the typical inhomogeneity decreases as a function of stellar surface temperature. Further, Figure~\\ref{fig:amp_per} shows that even for a given stellar temperature, the photometric amplitude is a decreasing function of the rotation period. This implies that the stellar inhomogeneity, probably together with the magnetic field, is decreasing as the star is aging, if we adopt the stellar rotation period as an age indicator. The catalog shows that the typical rotation period increases with decreasing stellar mass and temperature. A comparison to empirical gyrochronology models \\citep{bar07, mam+08, mei+09} shows the upper envelope of points is broadly consistent with an age of 4.5\\,Gyrs, but that the models appear to under-predict the stellar ages, with a significant proportion falling below the 1\\,Gyr rotational isochrone. The position of the Sun close to the upper envelope is interesting, since this implies a lack of stars older than the Sun. More slowly rotating stars should still appear on the main sequence and would not have been removed by our exclusion of evolved stars. The `dip' feature in the upper envelope around $\\sim~0.5\\,M_{\\odot}$ ($\\sim 4000$\\,K), mentioned in \\cite{mcq+13a} is still visible in the full sample. This mass corresponds to the transition point in young clusters, where lower mass stars spin up considerably more than stars above this mass \\citep{hen+12}. This feature is not accounted for by any current gyrochronology models. Thorough testing of AutoACF with synthetic LCs containing a range of periods and amplitudes is still required to understand the effect of observational biases on the upper envelope, since low-amplitude, long-period signals are most strongly affected by systematics and quarter joining. A series of period histograms across varying $T_{\\rm eff}$ bins, shown in Figure~\\ref{fig:hists}, displays some of the finer details of the period distribution, including a significant bimodality in the low-mass stars, which becomes less apparent towards higher masses. This bimodality, forming two sequences in the mass- or $T_{\\rm eff}$-period distribution, cannot be explained by any observational or measurement biases, and thus requires a physical explanation. One hypotheses presented in \\cite{mcq+13a} is that this is a possible hint of two epochs of star formation in the solar neighborhood. This effect may become less apparent towards higher masses, since these brighter stars are seen to greater distances, hence probing a larger region of space. An alternative explanation, also presented in \\cite{mcq+13a}, is that the dearth of stars between the two sequences represents a mass-dependent, rapid transition phase in the stellar spin-down law, where few stars are observed. In this scenario the star formation rate is unimodal but the spin-down occurs as two slow stages, with a rapid stage in between, producing a bimodal distribution of periods. Both these hypotheses require further follow-up observations and detailed modeling \\cite[e.g.,][]{gal+13, eps+14}, which will be aided by the larger sample size and wider $T_{\\rm eff}$ coverage of the new data." }, "1402/1402.5970_arXiv.txt": { "abstract": "{Atmospheric dispersion and field differential refraction impose severe constraints on widefield, multiobject spectroscopic (MOS) observations, where the two joint effects cannot be continuously corrected. Flux reduction and spectral distortions must then be minimised by a careful planning of the observations -- which is especially true for instruments that use slits instead of fibres. This is the case of VIMOS at the VLT, where MOS observations have been restricted, since the start of operations, to a narrow two-hour range from the meridian to minimise slit losses -- the so-called two-hour angle rule. } {We revisit in detail the impact of atmospheric effects on the quality of VIMOS-MOS spectra with the aim of enhancing the instrument's overall efficiency, and improving the scheduling of observations.} {We model slit losses across the entire VIMOS field-of-view as a function of target declination. We explore two different slit orientations at the meridian: along the parallactic angle (North-South), and perpendicular to it (East-West).} {We show that, for fields culminating at zenith distances larger than 20 deg, slit losses are minimised with slits oriented along the parallactic angle at the meridian. The two-hour angle rule holds for these observations using North-South orientations. Conversely, for fields with zenith angles smaller than 20 deg at culmination, losses are minimised with slits oriented perpendicular to the parallactic angle at the meridian. MOS observations can be effectively extended to plus/minus three hours from the meridian in these cases. In general, night-long observations of a single field will benefit from using the East-West orientation. All-sky or service mode observations, however, require a more elaborate planning that depends on the target declination, and the hour angle of the observations. } {We establish general rules for the alignment of slits in MOS observations that will increase target observability, enhance the efficiency of operations, and speed up the completion of programmes -- a particularly relevant aspect for the forthcoming spectroscopic public surveys with VIMOS. Additionally, we briefly address the (non-negligible) impact of field differential refraction on future widefield MOS surveys.} ", "introduction": "\\begin{table}[!h] \\caption{Characteristics of operational and planned/proposed multiobject optical spectrographs with FOVs larger than 100 arcmin$^2$.} \\begin{tabular}{llccccc} \\hline\\hline Telescope & Instrument & o/p\\,\\tablefootmark{a} & Aperture & s/f\\,\\tablefootmark{b} & FOV & N$_{max}$\\\\ & & & [m] & & [deg$^2$] & \\\\ \\hline SDSS\t& \tSDSS \t\t& o\t &\t2.5\t & f \t&\t7.07\t\t&\t1000\\\\ WIYN\t& \tHydraW\t\t& o &\t3.5\t & f\t&\t0.78\t\t&\t100\\\\ Blanco\t& \tHydraB\t\t& o &\t3.9\t & f\t&\t0.35\t\t&\t138\\\\ AAT\t \t&\t2dF\t\t \t& o &\t3.9\t & f\t&\t3.14\t\t&\t392\\\\ LAMOST\t& \tLAMOST\t\t& o &\t4.0\t & f\t&\t19.6\t\t&\t4000\\\\ WHT\t \t&\tWYFFOS\t\t& o &\t4.2\t & f\t&\t0.35\t\t&\t150\\\\ Magellan &\tIMACS\t\t& o &\t6.5\t & s\t&\t0.2\t\t&\t600\\\\ MMT\t \t&\tHectospec\t\t& o &\t6.5\t & f\t&\t0.78\t\t&\t300\\\\ VLT\t \t&\tVIMOS\t\t& o &\t8.2\t & s\t&\t0.08\t\t&\t600\\\\ VLT\t \t&\tFLAMES\t\t& o &\t8.2\t & f\t&\t0.14\t\t&\t132\\\\ Mayall\t& \tDESI\t\t\t& p &\t3.9\t & f\t&\t7.07\t\t&\t5000\\\\ VISTA\t& \t4MOST\t\t& p &\t4.1\t & f\t&\t7.07\t\t&\t3000\\\\ WHT\t \t&\tWEAVE\t\t& p &\t4.2\t & f\t&\t3.14\t\t&\t1000\\\\ VLT\t \t&\tMOONS\t\t& p &\t8.2\t & f\t&\t0.14\t\t&\t500\\\\ Subaru\t& \tPFS\t\t\t& p &\t8.3\t & f\t&\t1.33\t\t&\t2400\\\\ ngCFHT\t& \tngCFHT\t\t& p &\t10.0\t & f\t&\t1.5\t\t&\t4000\\\\ GMT\t \t&\tGMACS\t\t& p &\t21.9\t & s\t&\t0.05\t&\t400\\\\ \\hline \\end{tabular} \\label{table:facilities} \\tablefootmark{a}{(o)perational/(p)lanned/(p)roposed}\\\\ \\tablefootmark{b}{(s)lits/(f)ibres}\\\\ \\end{table} The effects of atmospheric dispersion on spectrophotometric observations were first tackled in a seminal paper by \\citet{Filippenko1982}, and have since been addressed by many other authors \\citep{Cohen1988,Donnelly1989,Cuby1994,Szokoly2005}. Two different components contribute to slit losses: a chromatic dispersion caused by the wavelength variation of the index of refraction of air; and an achromatic differential refraction due to airmass variations across the field-of-view (FOV). As pointed out by \\citet{Cuby1998}, the chromatic effect is almost constant for a given field, and can thus be counterbalanced with an atmospheric dispersion compensator (ADC). On the other hand, field differential refraction cannot be continuously corrected, and this is especially problematic for optical instruments with large FOVs. In the case of multiobject spectroscopic (MOS) observations the two joint effects cannot be compensated, so that aperture losses must be minimised by a careful planning of the observations \\citep[e.g.,][]{Cuby1998,Szokoly2005}. This may range from frequent reconfiguration of the fibres, to imposing limited observability windows. The latter is actually the only alternative for instruments that use slits instead of fibres, even more so because field rotation prevents the alignment of all slits along the parallactic angle. VIMOS\\,\\footnote{http://www.eso.org/sci/facilities/paranal/instruments/vimos.html} \\citep{LeFevre2003} is a widefield (4x7x8 arcmin$^{2}$) instrument with imaging, integral field, and MOS capabilities mounted at the Nasmyth B focus of VLT UT3. The instrument operates in the optical wavelength range (360-1000 nm), and is equipped with six sets of grisms, six sets of broad-band filters, plus three additional filter sets specifically designed to be used in combination with the grisms to block the second order spectra. The unique combination of instrument FOV, large collecting power, and very high multiplexing (up to $\\sim$\\,600 targets in the low resolution modes; see Table\\,\\ref{table:facilities} and Fig.\\,\\ref{fig:vimos}) has made VIMOS a particularly efficient instrument for large spectroscopic surveys of cosmological fields \\citep[e.g.,][]{LeFevre2005,Lilly2007,Popesso2009,Guzzo2013}. VIMOS-MOS observations are carried out using multislit masks, which provide very accurate sky subtraction and high instrumental throughput. However, the lack of ADCs makes atmospheric dispersion and field differential refraction factors that need to be taken into account. The atmospheric effects in instruments like VIMOS were studied by \\citet{Cuby1998}. They show that i) atmospheric dispersion dominates at shorter wavelengths, while differential refraction is not negligible at the red end of the visible spectrum; ii) in the former case, image drifts occur along the meridian; iii) differential refraction can be almost neglected for zenith distances smaller than 25 deg for exposure times up to two hours from the meridian (HA = 0\\,h). In view of these results they recommended that, in order to minimise slit losses, the slits be positioned along the dispersion direction at mid-exposure, and observations be limited to a narrow two-hour range from meridian crossing -- the so-called two-hour angle rule. This guarantees that losses remain below 20 per cent for zenith angles $<50$ deg at culmination. These rather limiting guidelines had always been in place for all MOS observations since the start of operations in 2003, and in practice translate into mandatory airmass constraint limits for the VIMOS-MOS observing blocks (OBs). During the last few years the instrument performance has been significantly enhanced (see \\citealt{Hammersley2010,Hammersley2013}) by changing the detectors to red-sensitive, low-fringing CCDs; replacing the HR-blue grism set with higher throughput VPH grisms; introducing an active flexure compensation system; redesigning the focusing mechanism and mask cabinet; and introducing a new pre-image-less MOS mode \\citep{Bristow2012}. All these improvements have made VIMOS a much more stable instrument, and have extended its lifetime in order to prepare it for the start of the spectroscopic public surveys for which ESO has recently issued a call. Further work to improve the operational efficiency of the instrument includes the present study, which has as main goal to revisit the need for restricted observability of targets only within plus/minus two hours from the meridian in the MOS mode. Increasing the observability of targets in the MOS mode provides more flexibility to operations, because the number of masks that can be loaded in the instrument before the beginning of each night is limited. As we previously noted, VIMOS is very often used for deep observations of cosmological fields, where very long integrations are taken for the same field. By increasing the target visibility (relaxing the two-hour angle rule), the programmes can be completed faster. \\begin{figure} \\centering \\includegraphics[width=.5\\textwidth]{survey_spectrographs_eff_vimos} \\caption{Survey efficiency (telescope area $\\times$ maximum number of allocated targets) and sky mapping efficiency (telescope area $\\times$ instrument field-of-view, or \\'etendue) of all operational (black), and planned/proposed (grey) multiobject optical spectrographs with FOVs larger than 100 arcmin$^2$. All figures are normalised to those of the SDSS. Note the extraordinary survey efficiency of VIMOS, only comparable to that of next generation instruments.} \\label{fig:vimos} \\end{figure} In this paper we revisit in detail the impact of atmospheric dispersion on the quality -- in terms of slit losses and spectrophotometric distortions -- of VIMOS observations. % Our aim is to establish general rules for the alignment of slits in MOS observations that enhance the efficiency of operations, not only for VIMOS, but for other operational and future multislit spectrographs as well. We note that the parameter space of this problem is large. Irrespective of image quality and atmospheric conditions, slit losses depend on slit size, orientation, and position within the FOV, observed wavelength range, target declination, total exposure time, and hour angle (HA) of the observations. We build upon the previous work by \\citet{Cuby1998}, but include specific calculations for all the grisms currently available in VIMOS, and present a more detailed analysis of the optimal slit orientation as a function of target declination. This paper is organised as follows. In Section\\,\\ref{sect:model} we describe the model with which we compute slit losses in VIMOS. Section\\,\\ref{sect:slitlosses} presents the main results of our analysis, while in Section\\,\\ref{sect:fdr} we briefly address the specific contribution of field differential refraction, and its importance for widefield MOS observations. Finally, in Section\\,\\ref{sect:conclusions} we discuss our main results, and summarise our findings. ", "conclusions": "\\label{sect:conclusions} We have shown that the optimal slit alignment in widefield MOS observations is a complex function of target declination and hour angle (Fig.\\,\\ref{fig:nsew}). The traditional orientation of slits along the parallactic angle at the meridian provides very stable results for observations within two hours from the meridian. A setup where slits are oriented \\emph{perpendicular} to the parallactic angle at the meridian is however preferred in the case of fields culminating at low zenith distances, or generally when observations extend to $|$HA$| >$ 2\\,h. This result can readily be understood from the following qualitative line of arguments. As discussed by \\citet{Szokoly2005}, slit losses depend on the interplay between two effects. The first one is independent of the slit orientation, as both atmospheric dispersion and differential refraction are larger at increasing airmasses. On the other hand, the angle between the slit and the parallactic angle is constantly changing. When slits are oriented along the parallactic angle at the meridian (N-S), the angle between the slits and the parallactic angle tends to increase, and the two effects amplify each other. Conversely, when slits are oriented perpendicular to the parallactic angle at the meridian (E-W) --and dispersion is always minimal--, the slits get systematically closer to the parallactic angle at higher airmasses, so that a large fraction of the dispersed light falls back into the slits. This is what creates the opposing trends with declination and hour angle depicted in Figs.\\,\\ref{fig:hrblue}-\\ref{fig:lrred}: the N-S orientation depends strongly on HA, but not so much on $\\delta$, while the strong dependence of the E-W orientation on $\\delta$ is almost invariant to changes on HA. \\begin{figure}% \\includegraphics[width=.5\\textwidth]{vimos_airmass} \\caption{VIMOS airmass constraints for MOS observing blocks. The two different shaded areas correspond to the limits for fields that can be observed with slits having N-S orientations at meridian (grey), or E-W orientations (purple). The lower curve in each case corresponds to the physical limit for targets of that declination.} \\label{fig:airmass} \\end{figure} This has important implications for the scheduling of observations. Night-long observations of a single field (visitor mode) will generally benefit from using the E-W orientation for slits at the meridian. All-sky or service mode observations, however, require a more elaborate planning that depends on the target and the hour angle of the observations. This is a particularly relevant aspect for the forthcoming spectroscopic public surveys with VIMOS, where observations will be carried out in visitor mode. We conclude that the two-hour angle rule, together with the default N-S slit orientation, provide the most stable results for service observations with VIMOS, with slit losses and spectral distortions below 20 per cent -- and almost independent of target declination. This should always be the preferred option for users having targets at $\\delta \\gtrsim -5$ deg or $\\delta \\lesssim -45$ deg. However, for targets within the $-45 \\lesssim \\delta \\lesssim -5$ deg range, the E-W orientation results in comparable or even reduced slit losses. This slit orientation allows for observations to go past the two-hour angle rule, and be effectively extended up to $|$HA$| = 3$ hours, hence making it a preferred option. This holds for all grisms currently offered in VIMOS, provided the acquisition is done with a filter that closely matches the grism wavelength range. Figure\\,\\ref{fig:airmass} shows the new airmass constraint limits for MOS OBs. They have been significantly relaxed for fields culminating at small zenith distances, thus increasing target observability. This shall enhance the efficiency of operations, and speed up the completion of programmes. Even though the results presented here are specifically tailored to VIMOS, these general recommendations apply to all current and future widefield multislit spectrographs (e.g., Magellan/IMACS, \\citealt{dressler2011}; GMT/GMACS, \\citealt{depoy2012}). They are already in place since September 2013 for VIMOS-MOS observations \\citep{rsj2013}." }, "1402/1402.0849_arXiv.txt": { "abstract": "We report on the discovery of SN~2014J in the nearby galaxy M~82. Given its proximity, it offers the best opportunity to date to study a thermonuclear supernova over a wide range of the electromagnetic spectrum. Optical, near-IR and mid-IR observations on the rising lightcurve, orchestrated by the intermediate Palomar Transient Factory (iPTF), show that SN~2014J is a spectroscopically normal Type Ia supernova, albeit exhibiting high-velocity features in its spectrum and heavily reddened by dust in the host galaxy. Our earliest detections start just hours after the fitted time of explosion. We use high-resolution optical spectroscopy to analyze the dense intervening material and do not detect any evolution in the resolved absorption features during the lightcurve rise. Similarly to other highly reddened Type Ia supernovae, a low value of total-to-selective extinction, $R_V \\lsim 2$, provides the best match to our observations. We also study pre-explosion optical and near-IR images from HST with special emphasis on the sources nearest to the SN location. ", "introduction": "Type Ia supernovae (SNe~Ia) are among the most luminous transient events at optical wavelengths and extremely valuable tools to measure cosmological distances, see \\citet{2011ARNPS..61..251G} for a recent review. Yet, SNe~Ia close enough to allow for detailed scrutiny of their physical properties are very rare, especially in a galaxy like M~82, the host of several recent core-collapse supernovae \\citep{2013MNRAS.431.2050M,2013MNRAS.431.1107G}. At an estimated distance to M~82 of 3.5 Mpc \\citep{2009ApJS..183...67D}, SN~2014J is the closest identified Type Ia SN in several decades, possibly rivaled by SN~1972E in NGC~5253 \\citep{1973A&A....28..295A} and SN~1986G in NGC~5128 \\citep{1987PASP...99..592P}. Thus, SN~2014J is exceptionally well-suited for follow-up observations in a wide range of wavelengths, from radio to gamma-rays. These have the potential to yield transformational new clues into the progenitor systems of SNe~Ia, as well as the detailed properties of dust along the line of sight, key astrophysical unknowns for the study of the accelerated expansion of the universe. There is strong evidence that SNe~Ia arise from thermonuclear explosions of carbon-oxygen white dwarfs (WD) in binary systems \\citep{2011Natur.480..344N,2012ApJ...744L..17B}. However, the nature of the second star remains unclear. For a long time, the preferred scenario was the single degenerate (SD) model \\citep{1973ApJ...186.1007W}, where a WD accrets mass from a hydrogen or helium rich donor star, thus becoming unstable while approaching the Chandrasekhar mass. The double-degenerate (DD) model involving the merger of two WDs \\citep{1981NInfo..49....3T,1984ApJS...54..335I,1984ApJ...277..355W} has gained considerable observational support in recent years, see e.g., \\citet{2012NewAR..56..122W}. In this work, we search for potential signatures of a SD progenitor system, such as variable Na~D lines, precursor nova eruptions, features in the early lightcurve, radio emission, or a coincident source in pre-explosion in HST images. ", "conclusions": "The discovery of SN~2014J presents us with a unique opportunity to explore the physics of Type Ia SNe and the line-of-sight effects due to intervening matter. Further understanding in these areas is of utmost importance for the use of SNe~Ia in cosmology. The early data from P48, starting as early as only hours from the explosion, and the multi-wavelength follow-up by the iPTF team covers an important range of the available windows in the electromagnetic spectrum. Just as the lightcurve reaches its maximum, we have learned that the SN has suffered non-standard extinction. We have searched for, but not detected, any time variation in our high-resolution spectra of the Na~I~D doublet. Similarly, we do not detect any pre-explosion activity in the $\\sim$1500 days of P48 monitoring. In a study of pre-explosion HST images in the near-IR, the nearest resolved source is found $0.2{\\arcsec}$ away from the SN location. The source brightness and offset from the SN makes it unlikely as a donor star in a single-degenerate scenario. Further, we make a first study of the spectral features of SN~2014J and find that it exhibits high-velocity features from intermediate mass material but lacks C and O often seen in very early spectra. Otherwise, it is a very similar to several well-studied normal SNe Ia." }, "1402/1402.0880_arXiv.txt": { "abstract": "Atmospheric neutrinos are produced in interactions of cosmic rays with Earth's atmosphere. At very high energy, the contribution from semi-leptonic decays of charmed hadrons, known as the prompt neutrino flux, dominates over the conventional flux from pion and kaon decays. This is due to the very short lifetime of the charmed hadrons, which therefore do not lose energy before they decay. The calculation of this process is difficult because the Bjorken-$x$ at which the parton distribution functions are evaluated is very small. This is a region where QCD is not well understood, and large logarithms must be resummed. Available parton distribution functions are not known at such small $x$ and extrapolations must be made. Theoretically, the fast rise of the structure functions for small $x$ ultimately leads to parton saturation. This contribution describes the ``ERS'' \\cite{ers} calculation of the prompt neutrino flux, which includes parton saturation effects in the QCD production cross section of charm quarks. The ERS flux calculation is used by e.g.\\ the IceCube collaboration as a standard benchmark background. We are now updating this calculation to take into account the recent LHC data on the charm cross section, as well as recent theoretical developments in QCD. Some of the issues involved in this calculation are described. ", "introduction": "When cosmic rays collide with nuclei in the air in Earth's atmosphere, hadrons are produced in inelastic collisions. These hadrons interact, lose energy, and finally decay, leading to a cascade of particles in the atmosphere that will finally hit the ground. Semileptonic decays of hadrons in the atmosphere generate a flux of neutrinos known as the atmospheric neutrino flux. This is an irreducible background for neutrino observatories such as Super-Kamiokande, IceCube, Antares or the planned KM3NeT. The fluxes at lower energies are pretty well understood, and come mainly from decays of long-lived charged pions and kaons, which are produced in essentially every inelastic collision. This ``conventional'' component of the flux falls steeply with increasing energy, both due to the spectral shape of the incoming cosmic ray flux, and due to the energy loss experienced by the mesons before they decay. Charged pions, for example, have a proper decay length of 8 meters, and with time dilation, their interaction lengths are much smaller than their decay lengths so that they have plenty of time to lose energy. The resulting neutrino energies are therefore downgraded compared to the incoming cosmic ray flux. Theoretical predictions by Honda and collaborators~\\cite{honda,gaisserhonda} of the conventional flux agree very well with measurements up to energies of roughly $10^{5}$~GeV~\\cite{Schukraft:2013ya}. At even higher energies, production of hadrons containing charm and bottom quarks leads to another component of the flux known as the ``prompt flux.'' These hadrons are produced much more rarely, but since they decay promptly without losing much energy, the resulting flux is harder and has more or less the same energy dependence as the incoming cosmic ray flux. At energies $\\sim 10^{5}-10^{6}$~GeV, this flux is believed to start to dominate the conventional flux, as shown in Fig.~\\ref{fig:fluxes}. \\footnote{The importance of charm production in atmospheric cascades has been known a long time. We have also recently included charm production in the calculation of neutrino fluxes from astrophysical sources \\protect\\cite{Enberg:2008jm}.} Coincidentally, this is roughly the same energy that is starting to be probed by the IceCube experiment. The prompt flux is much more poorly known than the conventional flux, due to the theoretical difficulties involved in calculating charm quark production at very high energies. The calculation is sensitive to very small Bjorken-$x$ and very forward rapidities, and this kinematic region is not probed by present collider or fixed target data. One of the most recent predictions of the prompt flux is the ``ERS'' flux of Ref.~\\cite{ers}, which was computed before the start of the LHC. This prediction is used by the IceCube experiment as a benchmark for the prompt flux. The IceCube experiment has recently observed two very high energy neutrino events at about 1~PeV~\\cite{Aartsen:2013bka}, and later 26 more events at slightly lower energies~\\cite{Aartsen:2013jdh}, all at and above the energies where the prompt flux becomes important. If these events are to be interpreted as (partly) due to a flux of astrophysical neutrinos, it must be clear that they can not be explained by the background coming from atmospheric neutrinos alone. The theoretical uncertainty in the prompt flux would probably not allow assigning the IceCube observations to atmospheric neutrinos---one would need a prompt flux larger than the ERS flux by a factor 15 in order to get a 10\\% probability to observe two atmospheric events at 1~PeV~\\cite{Aartsen:2013bka}. However, the significance of their observation depends on the normalization of the atmospheric neutrino flux; the significance of the first two events was $2.8\\sigma$, but with an increase in the prompt flux by a factor 3.8, the significance is reduced to $2.3\\sigma$. Similarly, the significance of the 28 events to not be completely of atmospheric origin is $4.1\\sigma$, which is reduced to $3.6\\sigma$ if the prompt flux background is increased by a factor 3.8. (The factor 3.8 is the level of the measured upper limit on the prompt flux.) Such an increase in the prediction if the various theoretical uncertainties would be better understood is not completely unlikely. On the other hand, if the prediction would become smaller, the significance would increase. In any case, to understand these events it is necessary to understand the background. The LHC experiments have now measured the charm production cross section at several energies~\\cite{LHCcharm}, providing some constraints on the input to the flux calculation. There have also been theoretical developments in QCD after Ref.~\\cite{ers}, that can be used to constrain the calculation~\\cite{NLL,Albacete:2010sy,Kutak:2012qk}. It is therefore time to improve the calculation to get a better handle on the prediction. In this contribution I will discuss some of the ingredients in the calculation, and in particular I will discuss why there are large theoretical uncertainties involved. I will necessarily have to skip a lot of details of the calculation, but everything that is not discussed here can be found in our paper \\cite{ers} or in the earlier calculations of the prompt flux \\cite{Lipari:1993hd,Gondolo:1995fq,Bugaev:1998bi,Pasquali:1998ji,Martin:2003us}. See also the book by Gaisser~\\cite{GaisserBook} for details on the description of the cascade in the atmosphere. ", "conclusions": "" }, "1402/1402.6696.txt": { "abstract": "We study a class of simplified dark matter models in which dark matter couples directly with a mediator and a charged lepton. This class of Lepton Portal dark matter models has very rich phenomenology: it has loop generated dark matter electromagnetic moments that generate a direct detection signal; it contributes to indirect detection in the cosmic positron flux via dark matter annihilation; it provides a signature of the same-flavor, opposite-sign dilepton plus missing transverse energy at colliders. We determine the current experimental constraints on the model parameter space for Dirac fermion, Majorana fermion and complex scalar dark matter cases of the Lepton Portal framework. We also perform a collider study for the 14 TeV LHC reach with 100 inverse femtobarns for dark matter parameter space. For the complex scalar dark matter case, the LHC provides a very stringent constraint and its reach can be interpreted as corresponding to a limit as strong as two tenths of a zeptobarn on the dark matter-nucleon scattering cross section for dark matter masses up to 500 GeV. We also demonstrate that one can improve the current collider searches by using a Breit-Wigner like formula to fit the dilepton MT2 tail of the dominant diboson background. ", "introduction": "\\label{sec:intro} %__________________________________________________% The search for thermal relic Weakly Interacting Massive Particle (WIMP) dark matter has a long history, particularly within models of weak-scale supersymmetry (SUSY)~\\cite{Haber:1984rc,Jungman:1995df}. Such models can furnish both signatures of new physics at the TeV scale and a viable candidate for dark matter. Collider, direct detection and indirect detection searches for Minimal Supersymmetric Standard Model (MSSM) dark matter particles have limited the vanilla parameter space~\\cite{CahillRowley:2012kx,Cahill-Rowley:2013dpa} and weaken the strong tie between WIMP dark matter and the SUSY framework. Outside of weak-scale SUSY models, there is no specific reason for dark matter to have mass near 100 GeV. Fortunately, the ``WIMP miracle'' provides guidance for the plausible region of dark matter mass and interaction strength~\\cite{Kolb:1990vq}. Since even the discovery of dark matter from multiple experimental probes is unlikely to immediately tell us the underlying framework, in this paper we concentrate on a class of simplified dark matter models, which serves as a phenomenological bridge between experiments and a deep underlying theory. There have been a number of recent studies of simplified dark matter models with the emphasis on the complimentarity from different experimental searches~\\cite{Chang:2013oia,An:2013xka,Bai:2013iqa,DiFranzo:2013vra,Buchmueller:2013dya,Cheung:2013dua,Papucci:2014iwa,Simone:2014}. Most of those studies have concentrated on dark matter interactions with the quarks of the Standard Model (SM), which leads to a new framework for interpretation of LHC and direct detection searches in terms of dark matter properties. For instance, in Ref.~\\cite{Bai:2013iqa} the signature of two jets plus missing transverse energy has been studied within the context of Quark Portal dark matter models, which is a class of simplified models in which dark matter particles and mediators interact with a single quark. In this paper, following our previous study in Ref.~\\cite{Bai:2013iqa}, we concentrate on the lepton sector and study a class of Lepton Portal dark matter models. In these models, there are two new particles in the dark matter sector with the lightest one being the dark matter candidate, which must be a singlet under electromagnetism and color. The other particle plays the role of mediator and connects the dark matter particle to the leptons. Obviously, to conserve SM gauge symmetry, the mediator particle should be charged under the electroweak symmetry. For the dark matter interactions to be renormalizable, the mediator must have the same quantum numbers as the left-handed lepton weak doublet or the right-handed charged leptons. In our study, we consider only the latter case for simplicity. Compared to Quark Portal dark matter models, Lepton Portal dark matter models have totally different phenomenology at the three frontiers of the search for WIMP dark matter. For direct detection, unlike the Quark Portal case, dark matter particles do not directly couple to target nuclei at tree level. At one loop, the dark matter can couple to the photon through various electromagnetic moments, which generates the dominant interaction with the target nucleus. The latest LUX results from Ref.~\\cite{Akerib:2013tjd} can constrain a large portion of parameter space for Dirac fermion or complex scalar dark matter. For indirect detection, dark matter particle annihilation can generate electrons or positrons with a harder spectrum than the Quark Portal case. Hence, the electron and positron flux measurement from AMS-02 in Refs.~\\cite{Aguilar:2013qda,AMS-electron-positron} becomes relevant for the Lepton Portal models. At colliders, the Quark Portal models have a larger signal production but also a larger QCD background. In the Lepton Portal models, the dark matter mediator particles can be pair produced via off-shell photons or $Z$ bosons. The corresponding collider signature is two same-flavor charged leptons plus missing transverse energy. Because both ATLAS and CMS collaborations at the LHC can make very good measurements of charged lepton momenta, the signature of dilepton plus missing transverse energy could serve as the \\emph{discovery channel} for dark matter particles. Therefore, we pay more attention to understanding and optimizing the key kinematic variables and work out the sensitivity at the 14 TeV LHC. Colliders can cover the light dark matter mass region beyond the direct and indirect detection sensitivity. This is simply due to different kinematics for different probes. For the three categories of dark matter particles: Majorana fermion, Dirac fermion and complex scalar, we have found that the 14 TeV LHC has a much better reach than the direct detection experiments for the Majorana fermion and complex scalar cases. For the Majorana case, the dark matter scattering cross section is suppressed by the dark matter velocity and predicts a very small rate for direct detection experiments. For the complex scalar case, the dark matter fermion partner has a large production cross section at the LHC and a high discovery probability at the LHC. Our paper is organized as follows. In Section \\ref{sec:lepton}, we introduce the Lepton Portal class of simplified models. We determine the allowed parameter space for dark matter to be a thermal relic in Section~\\ref{sec:relic-abundance}. The direct detection will be covered in Section~\\ref{sec:direct-detection}, where we perform loop-level calculations to determine the dark matter elastic scattering cross section. In Section~\\ref{sec:indirect-detection}, we work out constraints on model parameter space from the AMS-02 positron and electron flux measurement. We then perform a collider study for the sensitivity at the 14 TeV LHC with 100 fb$^{-1}$ and present summary plots in Section~\\ref{sec:collider}. We conclude in Section~\\ref{sec:conclusion}. %__________________________________________________% ", "conclusions": "\\label{sec:conclusion} %__________________________________________________% We want to first emphasize the importance of colliders for discovering or excluding the Lepton Portal dark matter. The signature with the same-flavor and opposite-sign dilepton plus missing energy is a pretty clean one. The $M_{T2}$ cut can be imposed to make almost background free. As a result, the discovery reach is purely determined by the signal production cross section times the acceptance. For a large mass splitting between dark matter and its partner, the signal acceptance is large, so the discovery reach is limited by the signal cross section. From Fig.~\\ref{fig:feyn-diagram}, one can see a large increase of the mediator production cross sections from 8 TeV to 14 TeV and a discovery of dark matter signals at the LHC may happen in the near future. In our analysis, we have considered both the electron and muon cases and neglected the tau lepton case. We anticipate a slightly weaker limit from the LHC because of the tau-tagging and mis-tagging efficiencies. Another parameter region that we have ignored is the co-annihilation region. The collider searches become less sensitive because the leptons from the mediator decays are either too soft to pass the basic cuts or generate insufficient $M_{T2}$ and would be buried in the SM backgrounds. In the extremely degenerate region, one could include an additional jet, photon, $W$ and $Z$ gauge bosons from initial state radiation to gain sensitivity. In summary, we have studied Lepton Portal dark matter for three cases: Majorana fermion, Dirac fermion and complex scalar dark matter. For direct detection, the majorana fermion case has a very small predicted event rate because of the leading operator of the dark matter coupling to photon has an additional velocity suppression. On the other hand, the direct detection signals for the Dirac fermion and complex scalar cases are not suppressed. In terms of indirect detection, since only the Dirac fermion case has non-zero $s$-wave annihilation, AMS-02 has the best coverage for its model parameter space. At colliders, the LHC has better reaches for the light dark matter mass region than the direct detection experiments. For the complex scalar case, the 14 TeV LHC with 100 fb$^{-1}$ can cover mediator masses up to 800 GeV and provides a constraint on spin-independent dark matter-nucleon scattering cross section as low as $2\\times10^{-46}$~cm$^2$ for dark matter masses up to 500 GeV and a unit coupling. \\subsection*" }, "1402/1402.2631.txt": { "abstract": "Massive stars can alter physical conditions and properties of their ambient interstellar dust grains via radiative heating and shocks. The \\ion{H}{2} regions in the Large Magellanic Cloud (LMC) offer ideal sites to study the stellar energy feedback effects on dust because stars can be resolved, and the galaxy's nearly face-on orientation allows us to unambiguously associate \\ion{H}{2} regions with their ionizing massive stars. The \\emph{Spitzer Space Telescope} survey of the LMC provides multi-wavelength (3.6--160 $\\mu$m) photometric data of all \\ion{H}{2} regions. To investigate the evolution of dust properties around massive stars, we have analyzed spatially-resolved IR dust emission from two classical \\ion{H}{2} regions (N63 and N180) and two simple superbubbles (N70 and N144) in the LMC. We produce photometric spectral energy distributions (SEDs) of numerous small subregions for each region based on its stellar distributions and nebular morphologies. We use \\texttt{DustEM} dust emission model fits to characterize the dust properties. color--color diagrams and model fits are compared with the radiation field (estimated from photometric and spectroscopic surveys). Strong radial variations of SEDs can be seen throughout the regions, reflecting the available radiative heating. Emission from very small grains drastically increases at locations where the radiation field is the highest, while polycyclic aromatic hydrocarbons (PAHs) appear to be destroyed. PAH emission is the strongest in the presence of molecular clouds, provided that the radiation field is low. ", "introduction": "\u00ca\\label{sec:intro} %================================================================= Far infrared (FIR) emission of interstellar dust has been commonly used to gauge star formation rates \\citep[e.g.,][]{Cetal10}. \u00caIn a star-forming region, interstellar dust can be heated, sputtered, and shattered by energy feedback from massive stars via ultraviolet (UV) radiation, fast stellar winds, and supernova remnant (SNR) shocks. \u00caIt would thus be prudent to investigate how dust properties are modified in the harsh environment around massive stars. \u00ca The Large Magellanic Cloud (LMC) provides a nearby, ideal laboratory to study the influence of massive stars on dust properties because of the following advantages over our Galaxy: (1) the LMC has a nearly face-on orientation, mitigating the confusion and extinction along the Galactic plane; (2) the LMC is at a known distance, $\\sim$50 kpc \\citep{fea99}, so stars can be resolved and studied in conjunction with the interstellar gas and dust; and (3) there exist many complementary multi-wavelength observations of stars, gas, and dust in the LMC. For these reasons, the LMC has been targeted by every space-based IR observatory to study dust properties and calibrate dust emission as star formation indicators. The first report of \\emph{Spitzer Space Telescope} \u00caobservations of an LMC \\ion{H}{2} complex was made by \\citet{Getal04} for LHA\\,120-N206 \\citep[N206 for short; designation from][]{hen56}; its dust emission was qualitatively compared with that of the Orion Nebula. \u00caSubsequently, the entire LMC was surveyed by \\emph{Spitzer} in the Legacy program Surveying the Agents of a Galaxy's Evolution \\citep[SAGE;][]{Meixner06}. \u00ca\u00caUsing the SAGE \u00caand additional data of 32 \\ion{H}{2} complexes in the LMC and Small Magellanic Cloud (SMC), \\citet{Letal10} found that the 70 $\\mu$m fluxes appear to be the best star formation indicator, although the correlation between the 70 $\\mu$m flux and the bolometric IR flux in the 30 Dor giant \\ion{H}{2} region shows a large degree of scatter. \u00caThe anomaly in 30 Dor could be attributed to the intense energy feedback from the large number of massive stars encompassed by this region. The SAGE data have also been used to analyze the global FIR emission in the LMC and spectral energy distributions (SEDs) of three large regions selected to represent different levels of star-formation activity \\citep{Betal08}. A 70 $\\mu$m excess in the SEDs is found, and this excess increases along the sequence ``Milky Way--LMC--SMC'', suggesting an influence of decreasing metallicity. \u00ca\\citet{Betal08} fit the SEDs with an improved version of the dust emission model by \\citet{Detal90} and suggested that the 70~$\\mu$m excess can be explained by modifying the size distribution of very small grains (VSGs) to have an increased abundance of larger VSGs. Since large VSGs can be produced through erosion of large grains in the diffuse medium, the presence of VSGs suggests effects of energy feedback. To investigate the effects of massive stars on dust emission, \\citet{Slater11} studied the dust properties of 4 classical and 12 superbubble \\ion{H}{2} regions in the LMC and found little correlation between dust emission properties and the spectral type of the hottest star, the bolometric luminosity of the underlying massive stars, or the evolutionary status of the \\ion{H}{2} regions. \u00caThis result is surprising and is at odds with the anomalous properties of 30 Dor and the abundant presence of large VSGs. We have examined the \\ion{H}{2} region sample of \\citet{Slater11} and find that their division between classical and superbubble cases may not be consistent with observations. \u00caFor example, DEM\\,L34 \\citep[= N11; DEM designation from][]{DEM76} and DEM\\,L152 (= N44) both contain multiple OB associations with a central superbubble surrounded by compact \\ion{H}{2} regions, but are classified as ``classical'' and ``superbubble'' \\ion{H}{2} regions, respectively. \u00caFurthermore, they have used the integrated dust emission from entire \\ion{H}{2} regions, while it is expected that the influence of stellar energy feedback ought to depend on radial distances from the stars. To investigate the influence of massive stars on interstellar dust grains with the least ambiguity, we have chosen to study two clearly defined classical \\ion{H}{2} regions, N63 (= DEM\\,L243) and N180 (= DEM\\,L323), and two relatively simple superbubbles, N70 (= DEM\\,L301) and N144 (= DEM\\,L199), whose stellar content is well studied. We have analyzed \\emph{Spitzer} observations of these four regions, measuring flux densities in radially distributed subregions, constructing SEDs, and using dust emission model fits to determine the dust properties. \u00caIn this paper, we describe our methodology, target selection, and data sets used in Section \\ref{sec:hii}, discuss the stellar content and assess the stellar energy feedback in Section \\ref{sec:SCRF}, describe the creation and modeling of SEDs in Section \\ref{sec:SEDs}, examine the results of each \\ion{H}{2} region in Section \\ref{sec:dust_analysis}, discuss the relationship between stellar energy feedback and dust properties in Section \\ref{sec:disc}, and summarize the findings in Section \\ref{sec:sum}. % \u00caGalliano et al. 2011, astro-ph \u00ca[Herschel data] \u00ca\\\\ % \u00caI find Galliano et al paper not very useful for our intro section. %================================================================ ", "conclusions": "\\label{sec:disc} %=========================================================== We have used model fits to spatially resolved SEDs of two classical and two superbubble \\ion{H}{2} regions to determine the dust composition and properties with respect to the radiation field. In this section we discuss the individual dust components and compare the dust properties in differing \\ion{H}{2} regions and SNRs. \\subsection{Dust} \\subsubsection{PAHs} Radiation is needed to excite PAHs; however, PAHs also can be destroyed in the presence of a strong radiation field \\citep[e.g.,][]{Madden06,Lebout07}, primarily due to high energy photons \\citep[e.g.,][]{Aitken85,Voit92}. Moreover, the location of PAH and H$_2$ emission are strongly correlated in cool-PDRs, such as the filaments in the Horsehead Nebula \\citep{Hab05,com07} and $\\rho$ Ophiuchus \\citep{Hab03}. However, the locations become anti-correlated with higher radiation field, as seen in the Orion bar \\citep{Tie93} and in Monoceros R2 \\citep{Ber09}. The results in this paper are consistent with these studies. We find that PAH emission is enhanced toward locations of molecular clouds except where the radiation is high. We also find that PAH emission is diminished at locations that are optically thin to UV radiation. Since our regions are relatively young (no known SNRs in N144 and N180, one known in N63, and a history of one or multiple in N70), the radiation field is dominated by massive stars that produce harsh photons. Although high energy photons are the main mechanism for destroying PAHs rather than the radiation field, we will assume that the stellar content of the regions follow the same initial mass function and have similar spectral shape in their radiation field and attempt to provide an estimate of the typical $U$ required for significant PAH destruction. In particular, we focus on N144 and N180 since N63 lacks large-scale CO observations and the dust emission from N70 is thought to come from the surrounding shell. For these two regions, Figure \\ref{fig:Upah} both $Y_{\\rm{PAH}}/Y_{\\rm{DUST}}$ and the color [8.0]--[24.0] versus the radiation field as ascertained by massive stars ($U$). We only show subregions when at least half the region contains a 2$\\sigma$ Magellanic Mopra Assessment \\citep[MAGMA;][]{Wong11} CO detection. Note that higher values of the color [8.0]--[24.0] indicates less PAH emission. For both panels, subregions for N180 indicate that PAHs are less coincident with CO molecular gas for high $U$, particularly above $U \\approx 30$. N144 subregions have points in the same general vicinity as N180, but an obvious correlation between PAHs and radiation field is not apparent. This could be due to the lack of data (particularly at low values of $U$) or physical differences between N144 and N180. Our calculations of the radiation field may be highly uncertain since they can be quite dissimilar (sometimes by about an order of magnitude) to the fitted results. However, the fitted results of $U$ are typically below those of the massive stars, so $U \\approx 30$ is a fair estimate of an upper limit of the typical radiation field needed to destroy PAHs in N180. Though the $U$ limit of N180 may not be representative of other molecular clouds, we note that maps in \\citet{Galametz13} also show a reduction at PAHs with similar modeled $U$ values in the LMC classical \\ion{H}{2} region N159, with a drastic depletion in the fraction of PAHs for $U \\gtrsim$ 100. %Plots in \\citet{Galametz13} indicated a value consistent with ours; the modeled radiation field in the classical \\ion{H}{2} region N159 in the LMC has a drastic depletion in the fraction of PAHs for $U \\gtrsim$ 100. %This value is roughly consistent with classical \\ion{H}{2} LMC region N159 where the modeled fraction of PAHs is drastically reduced for $U \\gtrsim$ 100 \\citep{Galametz13}. %In \\citet{Galametz13}, it is seen that in some regions of the LMC, the modeled fraction of PAHs is drastically reduced for $U \\gtrsim$ 100. Our results push this limit a bit lower, though this limit is certainly more dependent on the quantity of high energy photons rather than the value of $U$. %Dust modeling in \\citet{draine2007} showed that with a fix fraction of dust that are PAHs, for $U \\lesssim 10^3$ and for $\\lambda<20 \\mu$m, the \\emph{normalized} emission for dust grains is independent of $U$. Thus, given a constant PAH mass fraction throughout our subregions, we would expect an increase in PAH emission where the radiation is stronger. However, we certainly see a decrease for increasing $U$, which may suggest that the PAH fraction is not constant across our subregions due to their destruction via radiation. %As highlighted in \\citet{Ber09}, several studies have analyzed the spatial distribution of PAH emission and rotational H$_2$ lines in PDRs. The location of PAH and H$_2$ are strongly correlated in cool-PDRs, such as the filaments in the Horsehead Nebula \\citep{Hab05,com07} and $\\rho$ Ophiuchus \\citep{Hab03}. However, the locations becomes anti-correlated with higher radiation field, as seen in the Orion bar \\citep{Tie93} and in Monocerous R2 \\citep{Ber09}. Indeed, observations have shown that these PAHs are destroyed due to radiation \\citep[e.g.,][]{Madden06,Lebout07}, particularly by high energy photons \\citep[e.g.,][]{Aitken85,Voit92}. %\\cite{Ber09} suggested that for a dense PDRs with a high UV field, the rotational H$_2$ lines is caused by collisional excitation and thus highly dependent on gas temperature, causing the locations of PAHs and warm H$_2$ lines to be different. %\\cite{Ber09} suggested that this spatial discrepancy may be due to a selection bias; for a dense PDRs with a high UV field, the rotational H$_2$ lines is caused by collisional excitation and thus highly dependent on gas temperature, causing the locations of PAHs and warm H$_2$ lines to be different. %\\citet{Ber09} attributed this fact that in dense PDRs, high UV causes the rotational H$_2$ lines to be caused by collisional excitation \\subsubsection{VSGs and BGs} \\label{sec:VSGBG} In all of our \\ion{H}{2} regions except N70 (whose dust is uniformly distributed about its shell), $Y_{\\rm VSG}$/$Y_{\\rm BG}$ is highest at locations where the radiation field is the strongest, as evident in Figure \\ref{fig:U_vs_vsg_bg_ratio-eps-converted-to.pdf}. We stress that the VSG population is highly dependent on each SED's 24~$\\mu$m excess and is typically anti-correlated with the BG population. Grain populations typically change due to both destruction (i.e., dust changes back to gas) or disruption (large grains break into smaller grains). \\citet{Jones96} showed that fast shocks ($>$150 km~s$^{-1}$) likely destroy grains via sputtering, while the velocity requirement for shattering BGs with VSGs depends on the dust species, with velocities ranging from $\\sim$1--3 km~s$^{-1}$. In our regions, dust-dust collisions are probably uncommon due to low densities ($n_{\\rm{H}} \\sim 100$~cm$^{-3}$). However, O-stars are known to have wind-blown bubbles expanding at velocities of 10--15 km~s$^{-1}$ \\citep[with observations of such bubbles in N180,][]{Naze01}, which could lead to grain shattering. Other possibilities of an increased $Y_{\\rm VSG}$ include dust coagulation and dust erosion. Dust coagulation of the smallest grains into slightly larger VSGs could increase the $24~\\mu$m emission, but the dust in our regions are not in environments conducive to coagulation \\citep[see, e.g.,][]{koh12}. \\citet{Betal08} suggested erosion of BGs into VSGs likely explains the LMC's 70~$\\mu$m excess in neutral/ionized regions. Moreover, dust erosion is thought to be increased during star formation due to increased dust processing \\citep{par11}. The central OB associations of our \\ion{H}{2} regions have a rich history of previous star formation, which could certainly lead to more dust processing/erosion than the areas analyzed at the macro-scale (several 100~pc; \\citealt{Betal08}), leading to the increased 24~$\\mu$m emission. Thus, at scales of $\\sim$10~pc, BGs near OB associations may have gone through substantial erosion that has caused this increase of VSGs. %Radiation pressure on a VSG depends on the dust absorption efficiency. For grains smaller than the photon wavelengths, this efficiency is very low, making the radiation pressure negligible. However, for the massive stars in our \\ion{H}{2} regions, the dominate radiation wavelengths becomes very similar to the sizes of the VSGs, allowing for the larger VSGs to have dust absorption efficiency near unity \\citep[e.g.,][]{com2011}. A calculation shows that during a VSG's mean free path between collisions, radiation pressure can accelerate VSGs to shattering velocities (a few km/s) without reaching catastrophic speeds \\citep[$\\gtrsim 100 km~s^{-1}$,][]{Jones96}. Moreover, O-stars are shown to have wind-blown bubbles expanding at velocities of 10-15 km~s$^{-1}$ \\citep[including N180,][]{Naze01}. Therefore, at locations where there is a large radiation field in these \\ion{H}{2} regions, it is sensible to see this substantial increase in $Y_{\\rm VSG}$/$Y_{\\rm BG}$. On the other hand, \\citet{Pilleri2012} saw a definite decrease of ``evaporating\" VSGs with increasing radiation field. The sources they analyzed typically had a higher radiation field ($U \\gtrsim 100$) than our regions, though we still see an increase in VSGs from $U=100$ to $U=1000$. However, their dust model of VSGs follows more closely to the smallest grains of \\citet{Li01} while ours follows \\citet{Detal90}. Since \\citet{Pilleri2012} modeled VSGs that are more similar to PAHs and thus have an emission spectrum at much shorter wavelengths, these VSGs are not comparable to the VSGs modeled in this paper. %Although radiation can cause the primordial VSGs to be preferentially driven away over BGs in less than 1 Myr, we observe more VSG emission with increasing ISRF. VSGs are potentially replenished by BGs, whose shattering threshold may be achieved via stellar winds from nearby massive young stars. % What makes the ``shocks'' so important? % High enough dust velocities that can cause shattering? % Can dust be considered an ideal gas? % Shocks in the past? % % (1) radiation pressure drives smaller grains out faster than big grains, % \u00ca\u00ca\u00ca\u00ca making differential speed % (2) fast stellar wind can vaporize the dust grains, if hit % \u00ca\u00ca\u00ca\u00ca I made some crude calculation, and it seems plausible % (3) ionization front drives shocks into the ambient medium % (1) doesn't make sense to me since more radiation we see more smaller grains % (2) ditto % Meeting notes from Dec 12, 2012 -- back of the envelope calculation says a 100 angstrom % grain w/ density 2.2 g / cm^3 can be pushed as far as 38 pc (sans collisions) in 1 Myr from a % 10^38 erg / s luminosity. \u00caPossibly useful sentence: Although primordial VSGs can be % radiatively driven from HII regions by massive ionizing stars on the order of 1 Myr, we % observe more VSG emission with increasing ISRF. \u00caVSGs are potentially \u00ca % replenished by BGs, whose shattering threshold may be achieved via stellar winds % from nearby massive young stars. %\\subsection{Dust properties and massive stars} %\\subsection{Radiation Fields of Regions Analyzed} %\\subsubsection{Spectroscopically Classified Massive Stars and MCPS stars} %(1) spectroscopically classified massive stars \\\\ %(2) blue stars found from MCPS via CMDs (this description needs to happen in sec 4)\\\\ %Plot these stars and mark them in IR/Ha images. %Spectral typing can further quantify the stellar radiation field; \\citet{Garmany94} and \\citet{Oey96} provide spectral typing for the most massive stars in these regions. \u00ca\\citet{Oey96} reports 130 typed objects within the LMC, the locations of which (will be plotted soon) are plotted in Figs ... %\\subsection{SEDs and Stellar Feedback} %To quantify the effects of stellar feedback on the ISM, we pair the overall SED trend and location of each SED region to nearby stellar sources. \\subsection{Dust Properties of Evolved \\ion{H}{2}~Regions} We find that N70, the most evolved \\ion{H}{2} region, likely has its dust mass located in the shell. However, major differences in dust properties between classical \\ion{H}{2} regions and superbubbles are not obvious with this small sample size. \\citet{Relano13} analyzes the SEDs of 119 \\ion{H}{2} regions in M33, categorizing them in a likely evolutionary sequence (filled, mixed, shell, and clear shell morphologies). They found that each category had similar SED features, though differed slightly between each classification. Specifically, they found that the FIR peak is located at longer wavelengths for older regions (shells and clear shells), indicating that colder dust is likely present. Uniform cold dust temperatures are expected since gas has been pushed far away from the central heating source. This result is consistent with our analysis of N70, which shows a uniform, low fraction of $Y_{\\rm{VSG}}/Y_{\\rm{BG}}$ compared to our less evolved \\ion{H}{2} regions that do not have the same shell geometry. \\subsection{Comparison to SNR Studies} Changes in dust populations are commonly studied in SNRs due to their dynamic properties. Fast shocks ($\\gtrsim150$ km~s$^{-1}$) allow dust to reach high velocities that may cause their destruction via catastrophic shattering or sputtering \\citep[e.g.,][]{Jones96,Micelotta10}, though VSGs and PAHs can also be created through low-velocity shattering \\citep{Jones96}. While PAHs can survive or even be produced in slow shocks, fast shocks are generally considered to destroy all PAHs in a region except in circumstances where they can be recreated or protected by their environment \\citep[e.g.,][]{Micelotta10,Seok12}. Additionally, harsh photons provide another mechanism of destroying the smallest grains; the radiation field in SNRs can vary drastically, though it is typically about 10-100 times the local ISRF \\citep[e.g.,][]{Andersen11}. With these different mechanisms for creation and destruction of dust, it is important to observe and model a wide range of SNRs with different shock velocities, densities, and radiation fields to test theoretical models. Moreover, dust properties of SNRs can be compared to our \\ion{H}{2} regions to provide a diagnostic of how dust changes in varying radiation fields in the absence of large shock velocities. %Even at the youngest ages ($<300$ yr), dust is known to be produced in the ejecta of these SNRs \\citep{Rho09}. However, since PAHs were only recently discovered with $Spitzer$ \\citep{Tappe06}, studies of the dust properties have been limited. Since then, \\citet{Andersen11} used \\texttt{DustEM} (also using the \\citealt{Detal90} dust models) to fit BGs, VSGs, and PAHs for 14 Galactic SNR, representing the largest survey to model these properties in SNRs. The \\emph{Spitzer} IR Spectrograph (IRS) has been used to study a large number of SNRs. Dust produced in supernova ejecta has been detected in young ($<$300 yr) SNRs \\citep{Rho09}, and PAHs are detected in SNRs \\citep[e.g.,][]{Tappe06}. Using a large spectral survey of 14 Galactic SNRs, \\citet{Andersen11} used \\texttt{DustEM} (also using the \\citealt{Detal90} dust models) to model the contributions of emission from BGs, VSGs, and PAHs. They found that the ratio of VSGs to BGs is 2--3 times larger in SNRs than the plane of the Galaxy, and they attribute dust shattering as the mechanism for creating the surplus of VSGs. Two of the 14 SNRs had relatively high radiation fields; fits of these regions had relatively low PAHs and virtually no VSGs. \\citet{Andersen11} suggested that these regions have high shock velocities and low densities, which can lead to destruction of VSGs and PAHs through sputtering. With high radiation, we find that the VSG population can increase and PAHs are generally destroyed; however, fast shocks can destroy both grains. Unlike SNR regions, we find that VSG population increases where there is higher $U$, and there are no known fast shocks in our regions. Indeed, while observations from \\emph{XMM Newton} show X-ray emission from N70, there is no X-ray emission toward N180 and N144, and for N63, X-ray emission is only associated with the known SNR N63A (S. L. Snowden et al. in preparation). It has been suggested that N70 has recently been heated by an interior SNR \\citep{CM90,rod11,zha14}. If shocks are the main mechanism that are heating the grains in N70 rather than the ISRF, our \\texttt{DustEM} modeling would be invalid (since $U$ is our free parameter rather than temperature). Regardless, the shape of the SEDs of each subregion is the same with a dip at 24~$\\mu$m, which agree with our \\texttt{DustEM} fit of a uniform, low abundance of VSG. Therefore, it is most likely that a past history of fast shocks has destroyed the population of VSGs, and no mechanism has been available for creation of new VSGs (due to, e.g., the lack of centrally located gas since it has been blown away). %and that knowledge of shock velocities is crucial when analyzing the dust populations in a region. %=====================================================================" }, "1402/1402.6554_arXiv.txt": { "abstract": "{} {In previous studies a very hot plasma component has been diagnosed in solar active regions through the images in three different narrow-band channels of SDO/AIA. This diagnostic from EUV imaging data has also been supported by the matching morphology of the emission in the hot Ca XVII line, as observed with Hinode/EIS. This evidence is debated because of unknown distribution of the emission measure along the line of sight. Here we investigate in detail the thermal distribution of one of such regions using EUV spectroscopic data.} {In an active region observed with SDO/AIA, Hinode/EIS and XRT, we select a subregion with a very hot plasma component and another cooler one for comparison. The average spectrum is extracted for both, and 14 intense lines are selected for analysis, that probe the $5.5 < \\log T < 7$ temperature range uniformly. From these lines the emission measure distributions are reconstructed with the MCMC method. Results are cross-checked with comparison of the two subregions, with a different inversion method, with the morphology of the images, and with the addition of fluxes measured with from narrow and broad-band imagers.} {We find that, whereas the cool region has a flat and featureless distribution that drops at temperature $\\log T \\geq 6.3$, the distribution of the hot region shows a well-defined peak at $\\log T = 6.6$ and gradually decreasing trends on both sides, thus supporting the very hot nature of the hot component diagnosed with imagers. The other cross-checks are consistent with this result.} {This study provides a completion of the analysis of active region components, and the resulting scenario supports the presence of a minor very hot plasma component in the core, with temperatures $\\log T > 6.6$. } ", "introduction": "It is accepted that the energy source that sustains the high temperature of solar corona is in the magnetic field \\citep{2006SoPh..234...41K}. Regarding the way in which this energy is converted to thermal energy, we can broadly distinguish between two scenarios that depend on the timescale of the energy release: one is continuous heating, the other is in the form of discrete, rapid pulses. The latter is consistent with the nanoflare model proposed by \\cite{1988ApJ...330..474P}. According to this model, the magnetic field tubes are displaced by the photospheric motions, and can approach and interact. When two flux tubes are almost in contact they will form a current sheet, where the field lines can reconnect. The reconnection can release a large quantity of energy in impulsive events called nanoflares. Signatures of these events are difficult to observe for several reasons, one of which is the fine structuring of the magnetic loops, that is hardly resolved with present-day instruments (e.g., \\citealt{2013ApJ...770L...1T}). One feature to discriminate the heating release is the presence or absence of very hot plasma. In active regions, the mean coronal temperature is typically 2-3 MK. If heat pulses occur, we expect that a small amount of plasma hotter than the average (6-10 MK) will be ever-present. Recently, a number of studies, mostly based on data from the Hinode and the Solar Dynamics Observatory (SDO) missions, have shown increasing evidence for such small very hot components in active regions (\\citealt{2009ApJ...704L..58R}, \\citealt{2009ApJ...698..756R}, \\citealt{2009ApJ...697...94M}, \\citealt{2009ApJ...693L.131S}, \\citealt{2009ApJ...697.1956K}, \\citealt{2009ApJ...696..760P}, \\citealt{2010AstL...36...44S}, \\citealt{2010A&A...514A..82S}), but the issue is still debated (\\citealt{2012ApJ...754L..40T}, \\citealt{2011ApJ...734...90W}). In SDO images taken with a channel sensitive also to emission of plasma at 6 MK (94 \\AA), cores of active regions contain bright strands \\citep{2011ApJ...736L..16R}, as predicted by models of nanoflaring loops \\citep{2010ApJ...719..576G}. It remains to be proven whether this plasma is really at such high temperatures or not. Spectroscopic observations should help greatly. In this work, we analyse an active region that shows evidence for this very hot component in SDO data, but for which spectroscopic data are also available from the EUV spectrometer EIS on-board the Hinode mission. In a previous work (\\citealt{2012ApJ...750L..10T}, hereafter Paper I), emission in the CaXVII line, which forms around temperature of 6-8MK, was detected in the hot structures identified with SDO/AIA data. In that work they built a 3 color image, to highlight the presence of a very hot component of emitting plasma inside the active region. The AIA 94~\\AA\\ band is known to be multi-thermal. It is sensitive to hot plasma, due to the presence of an Fe XVIII line, formed around 6~MK, but is also sensitive to plasma at 1 MK, because of the presence of a Fe X line and cooler Fe IX and Fe VIII (see, e.g., \\citealt{2011A&A...535A..46D}, \\citealt{2012ApJ...745..111T}, \\citealt{2011ApJ...743...23M}, \\citealt{2011ApJ...740L..52F}, \\citealt{2012A&A...537A..22O}, \\citealt{2012A&A...546A..97D}). Recently, also a Fe XIV line was identified in \\cite{2012A&A...546A..97D}. This line is normally stronger than the other cool components in active region cores, as shown in \\cite{2013A&A...558A..73D}. It is therefore not simple to assess if the hot emission seen in the AIA 94~\\AA\\ band is really due to Fe XVIII. To clarify this point, \\cite{2012ApJ...750L..10T} compared the AIA 94~\\AA\\ image with the Ca XVII image obtained from the EIS spectrometer. They showed a strong correlation between the hot CaXVII and the emission in the 94\\AA\\ AIA band, so concluded that the hot emission seen in the AIA 94~\\AA\\ band is effectively due to very hot plasma (6-8MK). More direct evidence has been found from observations of another Fe XVIII line by the SUMER spectrometer on board the SOHO mission \\citep{2012ApJ...754L..40T}. However, even if the indication is rather strong, it is still not enough to establish that the plasma is actually so hot, since in theory it is possible that plasma at a lower temperature, but with very high emission measure, can give the same line intensity, as suggested by \\cite{2012ApJ...754L..40T}. Indeed \\cite{2013A&A...558A..73D} used simultaneous EIS and AIA observations of active regions cores to show that a significant fraction of the Fe XVIII 94~\\AA\\ intensity can be due to plasma at 3~MK and not 6~MK. In fact, \\cite{2013A&A...558A..73D} showed that often Fe XVIII 94~\\AA\\ emission is present in the cores of active regions, but Ca XVII (which has a narrower formation temperature, hence is sensitive to hotter plasma) is not. The only way to disentangle the various contributions to the AIA 94~\\AA\\ band is therefore to perform an emission measure modelling. To this purpose, here we use the same observations from Hinode and SDO as in Paper I, that include both high-resolution spectroscopic data, over a wide spectral window, and images with high spatial resolution. All this information provides simultaneous constraints on the plasma thermal structure along the line of sight. To further support the analysis we replicate the same analysis on the same data set but taking a region outside of the core that shows no evidence of these very hot components.We also compare two different inversion methods. In Section~\\ref{sec:obs} we describe the observation and the data analysis, and in Section~\\ref{sec:discuss} the results are discussed. ", "conclusions": "\\label{sec:discuss} We analysed the thermal distribution of the plasma along the line of sight in two different regions where the narrow-band imagers diagnosed very different dominant temperatures, a hotter and a cooler one. The analysis was mostly based on the spectral data from Hinode/EIS. Our major attention was to the hottest components that may be signature of impulsive heating at work, and here our aim was to reconstruct the whole EM distribution and check the coherence of the overall scenario, including the hot component. At variance from many other similar analyses, our approach was to consider a limited number of spectral lines ($\\sim 1$ per temperature bin), and we picked the most intense lines uniformly covering the temperature range. In this way we simplified our control and interpretation of the reconstruction results, and all measured fluxes had the same weight in determining the global emission measure distribution. We also had a particular approach regarding the assessment of the uncertainties. The typical choice is to associate the same percentage error to all measured fluxes; 20\\% is the most commonly assumed value (e.g. \\citealt{2011ApJ...734...90W}). This uncertainty safely includes possible systematic unknown effects due to the instrument, atomic data and chemical abundances, but weights the same all measured values, independently of whether a line is strong or weak. Our choice is different. The measured fluxes were assigned exclusively the poissonian error, dictated only by the photon statistics. This allowed us to weight more the intense lines and to minimise the uncertainties of the solutions of the DEM reconstructions. As a consequence, the DEM reconstruction with the MCMC technique leads to solution clouds with a narrow distribution around the best solution for many temperature bins (factor 2-3), significantly narrower than in previous works. Only a minority of temperature bins show a spread solution cloud. The constraint on the hot components is tighter. The error estimate is important in this analysis, and we are aware that the real uncertainties are surely larger than those that we assume. The uncertainties can influence considerably the error on the DEM solution, and sometimes even the solution itself, as thoroughly discussed in \\cite{2012ApJ...745..111T}. We should also consider that the cloud spread may underestimate the real error on the DEM solution. The small uncertainties typically lead to a larger number of iterations with the MCMC method for better convergence \\citep{2012ApJ...745..111T}, but our 400 iterations is certainly an appropriate number for our cases. Tests show that DEM structures are reliably recovered on scales larger than $\\Delta \\log T = 0.2$, so we do not discuss narrower features. The good agreement between predicted and observed line intensities confirms this, although we note that some significant discrepancies are present, in particular with the Ca XV line, as also found previously \\citep{2013A&A...558A..73D}. The good agreement between the two inversion methods, already found in \\cite{2011A&A...535A..46D}, is confirmed. This suggests that the main source of uncertainty resides in the choice of parameters, atomic data and instrument calibration. Although our assumptions probably underestimate the errors, the coherent support from other instruments (AIA, XRT), the agreement with the results from another reconstruction method and the coherence with the morphology seen in the images spanning the different temperature regimes makes us confident that other errors should not affect our results considerably. Overall, the analysis has confirmed the two general characteristics anticipated by the imagers. The comparison of the DEM distributions of the cold and hot regions has, on the one hand, revealed substantial thermal components for $logT < 6.3$ in the cold region, without showing prominent features. The reconstruction of the cold region also shows minor components for $logT > 6.3$. Since the images show no bright features in the hot channels and lines, we may use the values of these emission measure components as sensitivity limits of our analysis, i.e. we may not trust emission measure components below $10^{26}$ cm$^{-5}$. The hot region shows a much more peaked thermal structure, with the positive gradient typically found in previous studies in coronal loops of active region cores \\citep{2011ApJ...734...90W}. The peak is at a rather high temperature ($\\log T = 6.6$) and beyond that the emission measure declines, but not very steeply and still showing a significant fraction of emission measure at temperature $\\log T \\leq 6.8$. Some components might be present at even higher temperature, although with higher uncertainties, up to the limit of the thermal sensitivity of our analysis ($\\log T \\sim 7$). We believe that the joint use of hot spectral lines, AIA 94 \\AA\\ channel and XRT filterbands, helps to partially remove the so-called \"blind spot\" for $\\log T > 6.8$ \\citep{2012ApJ...746L..17W}. Although the presence of the very hot component looks confirmed, still it is based on a very limited amount of information, and therefore some care should still be used. Some further feedback is provided by the comparison with the images, and between the images. The morphology of the region in the Ca XVII line only partially overlaps the morphology in immediately cooler lines (Fe XVI), while it matches well the X-ray and AIA 94 \\AA\\ morphology. This confirms that in the hot region selected in the present analysis there is indeed significant emission at 6~MK, resulting in strong Ca XVII, so that the Fe XVIII emission in the AIA 94 \\AA\\ band is also due to this component, unlike other cases in the cores of active regions where Ca XVII is not observed, and the Fe XVIII emission is due to a large emission measure at 3~MK \\citep{2013A&A...558A..73D}. Our analysis has remarked how spectral data are by far more constraining than the data from imagers, because the spectral lines are much more sensitive to temperature variations that the broader bands of the imagers (in agreement with the findings of \\cite{2012ApJ...758...54T}. Still it stresses that a quantum leap in the diagnostics of the hottest DEM components needs constraints from more lines sensitive to emission from high temperature plasma. These might be, for instance, easily accessible to broad-band X-ray spectrometers, to which we look forward in future space missions." }, "1402/1402.0691_arXiv.txt": { "abstract": "Ultraviolet observations are the best tool to study hot stellar populations which emit most of their light at short wavelengths. As part of a large project devoted to the characterization of the UV properties of Galactic globular clusters (GGCs), we collected mid/far UV and optical images with the WFPC2@\\hst\\ for 31 GGCs. These clusters cover a wide range in metallicity and structural parameters, thus representing an ideal sample for comparison with theoretical models. Here we present the first results from an ongoing analysis aimed at deriving the temperature distribution of Horizontal Branch stars in GGCs. ", "introduction": "\\label{sec:intro} In old stellar populations like Galactic globular clusters (GGCs), the hottest stars are Blue Stragglers (BSS), Horizontal Branch (HB), Post-Early Asymptotic Giant Branch (PEAGB) and AGB-manqu\\'e whose spectral emission peaks at effective temperatures higher than $\\sim 6,000\\,K$ \\citep{Fer12,Moe07,Fer97,Swe87}. Therefore these stars mostly emit in the ultraviolet (UV) part of the electromagnetic spectrum \\citep{Moe95,deB85}. Since the Earth's atmosphere is almost completely opaque to this wavelength regime, any campaign for the observation of hot stars in UV must be performed using space facilities like, for instance, the Hubble Space Telescope (\\hst) or the Galaxy Evolution Explorer (GALEX). Unfortunately, the theoretical models describing the evolution of stars of the earliest spectral types still suffer from many uncertainties which are also due to the absence of a complete and homogeneous sample of UV emitters. Therefore, the UV observation of hot stars in GGCs is extremely important in order to compare data and evolutionary models in observational planes not hampered by critical color - effective temperature (\\teff) transformations \\citep{Dal13,Sch12,Dal11,Fer06-2,Cas04,Fer03,Fer98,Fer97-2}. One of the most interesting stellar evolutionary sequences that still wait to be fully understood is the HB. It is commonly accepted that metallicity is the first parameter driving the Horizontal Branch morphology. Indeed, metal-rich GGCs typically have red HBs while metal-poor ones have more extended and blue HBs. However, there are several clusters with the same metallicity showing remarkable differences in Horizontal Branch morphology: nice examples are the couples \\ngc6388 - \\ngc5927, M\\,3 - M\\,13 and M\\,15 - M\\,92. Therefore, metallicity alone is not able to explain the observation of the complex HB zoology in GGCs \\citep{Fre81}. This issue, known as the `2nd parameter' problem, has focused the attention of several authors in the last years \\citep{Gra10,Dot10,Cat09,Lee94,Fus93}, but the solution is not obvious because many mechanisms play a role in shaping the color distribution of HB stars, such as mass loss, age and He abundance \\citep{Roo73}. Nonetheless, now there is a general consensus about the fact that age is the main global 2nd parameter \\citep{Gra10,Dot10}. \\par The HB temperature extension can be an efficient way to parametrize the HB morphology and therefore to study the `2nd parameter' problem. This approach has been already attempted by \\citet{Rec06} and, later on, by \\cite{Bab09}, who used optical CMDs for their analysis. However, temperature derived from optical filters only, can be underestimated by more than $10,000\\,K$ in the case of very extended HBs \\citep[see the case of \\ngc2808, ][]{Dal11}. This demonstrates the difficulties of deriving properties of hot HB populations from pure optical data. Indeed in the optical plane, stars move down to the blue HB tail either because the bolometric correction increases or because stellar luminosity decreases. For very blue HB stars (like Blue Hook stars) both the above quantities are changing and optical colors cannot provide a reliable estimate of \\teff. \\par As a part of a large project aimed at characterizing the UV properties of GGCs \\citep{San14,Dal13-2,Dal13,Con12,Dal12,San12,Sch12,Dal11,Dal09} we present here some preliminary results coming from the photometric analysis of UV observations of 31 GGCs obtained with the Wide Field and Planetary Camera 2 (WFPC2) and finalized to accurately derive the HB temperature distribution and extension. ", "conclusions": "We have analyzed far/mid UV-optical images for ten out of 31 GGCs collected within the WFPC2 UV survey. The main task of this analysis is to derive the extension and the temperature distribution of the HB of the targeted clusters taking advantage of the UV data. The use of UV bands is the best tool to accomplish our task since, for stars with high \\teff, the bolometric correction decreases and the stellar luminosity increases with respect to the optical bands. It has been shown, indeed, that the temperatures of extended HBs derived from a combination of pure optical filters can be underestimated by $10,000 - 15,000\\,K$ \\citep{Dal11}. We used \\ngc6293 as a starting point in our analysis, determining the temperature distribution of its HB stars. We are now applying this analysis to other GGCs in our sample with extended HBs, like $\\omega\\,Cen$ and M\\,15." }, "1402/1402.0002_arXiv.txt": { "abstract": "Disc-winds originating from the inner parts of accretion discs are considered as the basic component of magnetically collimated outflows. The only available analytical MHD solutions to describe disc-driven jets are those characterized by the symmetry of radial self-similarity. However, radially self-similar MHD jet models, in general, have three geometrical shortcomings, (i) a singularity at the jet axis, (ii) the necessary assumption of axisymmetry, and (iii) the non-existence of an intrinsic radial scale, i.e. the jets formally extend to radial infinity. Hence, numerical simulations are necessary to extend the analytical solutions towards the axis, by solving the full three-dimensional equations of MHD and impose a termination radius at finite radial distance. We focus here on studying the effects of relaxing the (ii) assumption of axisymmetry, i.e. of performing full 3D numerical simulations of a disc-wind crossing all magnetohydrodynamic critical surfaces. We compare the results of these runs with previous axisymmetric 2.5D simulations. The structure of the flow in all simulations shows strong similarities. The 3D runs reach a steady state and stay close to axisymmetry for most of the physical quantities, except for the poloidal magnetic field and the toroidal velocity which slightly deviate from axisymmetry. The latter quantities show signs of instabilities, which, however, are confined to the region inside the fast magnetosonic separatrix surface. The forces present in the flow, both of collimating and accelerating nature, are in good agreement in both the 2.5D and the 3D runs. We conclude that the analytical solution behaves well also after relaxing the basic assumption of axisymmetry. ", "introduction": "Astrophysical jets are observed in association with a wide range of objects, from Brown Dwarfs and young stellar objects to supermassive Black Holes in active galactic nuclei; however, there are still open several questions concerning the launching and acceleration mechanisms of jets. In all cases, jets and discs are inter-related, while magnetic fields play a key role in accelerating the outflows. \\citet{BlP82} studied the magneto-centrifugal acceleration along magnetic field lines threading an accretion disc. They showed the braking of matter in the azimuthal direction inside the disc and the outflow acceleration above the disc surface guided by the poloidal magnetic field components. Toroidal components of the magnetic field then collimate the outflow. Numerous semi-analytic models extended the work of Blandford \\& Payne along the guidelines of radially self-similar solutions of the full magnetohydrodynamic (MHD) equations \\cite[e.g.][]{CoL94, Li95, Li96, Fer97, VlT98}. Several numerical studies exist, which have focused on the magnetic launching of disc-winds. In most models a polytropic equilibrium accretion disc has been used as a boundary condition \\citep[e.g.][]{UKR95, KLB99, KLB03, OCP03, NaM04, ALK05, ALK06, PRO06}. The magnetic feedback on the disc structure was therefore not calculated self-consistently. Only in recent years have been carried out the first simulations which include self-consistently the accretion disc in the calculations of jet launching \\citep[e.g.][]{CaK02, CaK04, KMS04, ZFR07, TFM09, MFZ10, SFP12, TFM13, FeS13}. Often analytical solutions and their integrals of motion are used for connecting observed quantities far away from the jet source with properties of the jet-driving accretion disc. The behaviour of integrals of motions has been also tested in numerical simulations. Several 2.5D simulations of disc winds from the disc as a boundary condition \\citep[e.g.][]{RUK97, OuP97, KLB99, UKR99} found super-fast magnetosonic flows with properties which are expected from self-similar theory, although they have not used self-consistent analytical solutions for describing the boundary conditions. Recently, \\citet{SNO10} performed 3D simulations extending the calculations of \\citet{OuP97} and found good agreement between both. In addition, \\citet{SNO10} also simulated configurations closer to self-similar solutions (their run BP) and compared synthetic emission maps derived from them with HST observations. A similar comparison is presented in \\citet[][hereafter paper~II]{SGT10} using simulations from \\citet[][hereafter paper~I]{STV08}. Most self-similar models have three serious limitations, (1) the flow often does not cross all critical points (especially not the fast-magnetosonic limiting characteristic), with the result that the terminal wind solution is not causally disconnected from the disc, (2) singularities exist at the jet axis in radial self-similar models, and (3) for deriving self-similar models the assumption of axisymmetry is necessary. \\citet[][V00 hereafter]{VTS00} showed that a terminal wind solution can be constructed that is causally disconnected from the disc and hence any perturbation downstream of the superfast transition cannot affect the upstream structure of the steady outflow. The other two limitations can only be solved using numerical simulations extending the analytical solution of e.g. V00 as done by \\citet[]{GVT06} using the MHD code NIRVANA \\citep{Zie98}, \\citet[]{MTV08} using the MHD code PLUTO\\footnote{publicly available at http://plutocode.ph.astro.it/} \\citep{MBM07} and again with PLUTO in paper I for comparison with models with finite jet-emitting disc radii. \\citet{CGV08} extended the solution by adding the effects of resistivity. However, all of those models still assumed axisymmetry. The aim of this paper is to investigate numerically how relaxing the assumption of axisymmetry, i.e. performing full 3D numerical simulations of a disc-wind crossing all magnetohydrodynamic critical surfaces, affects the topology, structure and stability of this particular radial self-similar analytical solution. This paper is organized as follows: the numerical setup is briefly described in Sec.~\\ref{sec_num_models}. In Sec.~\\ref{sec_results} we describe the results of our simulation. We close with a summary and discussion of the results in the last section. ", "conclusions": "We have shown the results of the first 3D simulation of a disc-wind crossing all magnetohydrodynamic critical surfaces. This result is important in order to assure that the main flow is causally disconnected from its source, which is a prerequisite for a jet with the observed long lifetime. We have compared these results with previous axisymmetric 2.5D simulations. The structure of the flow in all simulations exhibits strong similarities. In the outer part of the flow, its structure is almost identical in all cases. Near and at the position of the fast magnetosonic separatrix surface (FMSS) which shows as a shock, some minor deviations of the 3D run from the 2.5D runs are present. The 3D runs reach a steady state and stay close to axisymmetry for most of the variables, except for the poloidal magnetic field and the toroidal velocity which deviate considerably from axisymmetry, but are not dynamically important. The latter quantities show signs of instabilities, which, however, are confined to the region inside the FMSS. It is important to emphasize that the crossing of the FMSS, which represents the ``event horizon'' for the propagation of MHD waves, does not allow any disturbance at large distances to reach the base of the flow, contributing to its stability. The forces present in the flow, both of collimating and accelerating nature, are in good agreement in both the 2.5D and the 3D runs. The main goal of the present paper is to check whether the axisymmetric, radially self-similar MHD solution for a polytropic disc-wind which crosses all appropriate critical surfaces to satisfy causality \\citep{VTS00}, behaves ``well'' also after (i) removing the axial singularity and (ii) relaxing the assumption of axisymmetry. We have found that this solution is structurally stable to non-axisymmetic perturbations. The next step will be to further configure and improve this solution such as it may describe realistic astrophysical disc-winds. As in paper~I, this further extension will involve the truncation of the solution at some arbitrary radii and use of this truncated solution to study the temporal evolution of jets within an MHD model and further comparison with observations. \\citet{SNO10} argue that models exhibiting a slowly varying poloidal field component in the accretion disc (their model OP) match observations better than those resembling self-similar MHD solutions (as their model BP). Note, however, that in paper~II we showed that radially truncated self-similar solutions may very well describe real HST observations." }, "1402/1402.0833_arXiv.txt": { "abstract": "In the theory of resonant mode coupling, the parent and child modes are directly related in frequency and phase. The oscillations present in the fast rotating $\\delta$\\,Sct star KIC\\,8054146 allow us to test the most general and generic aspects of such a theory. The only direct way to separate the parent and coupled (child) modes is to examine the correlations in amplitude variability between the different frequencies. For the dominant family of related frequencies, only a single mode and a triplet are the origins of nine dominant frequency peaks ranging from 2.93 to 66.30 cycles day$^{-1}$ (as well as dozens of small-amplitude combination modes and a predicted and detected third high-frequency triplet). The mode-coupling model correctly predicts the large amplitude variations of the coupled modes as a product of the amplitudes of the parent modes, while the phase changes are also correctly modeled. This differs from the behavior of ``normal'' combination frequencies in that the amplitudes are \\emph{three orders of magnitude} larger and may exceed even the amplitudes of the parent modes. We show that two dominant low frequencies at 5.86 and 2.93 cycles day$^{-1}$ in the gravity-mode region are not harmonics of each other, and their properties follow those of the almost equidistant high-frequency triplet. We note that the previously puzzling situation of finding two strong peaks in the low-frequency region related by nearly a factor of two in frequency has been seen in other $\\delta$ Sct stars as well. ", "introduction": "\\begin{figure*}[!t] \\centering \\includegraphics[bb=40 70 750 550, width=\\textwidth, clip]{Fig1.eps} \\caption{The dominant 9 frequencies of the T family of frequencies. The frequency numbers, e.g. f26, denote their approximate values in cycles day$^{-1}$ and this notation will be used in the rest of the paper. Note that the frequencies are both in the low-frequency g-mode and high-frequency p-mode regions. The bottom panel illustrates their relation to all the detected frequencies.} \\end{figure*} The recent spaced-based studies of $\\delta$\\,Sct and $\\gamma$\\,Dor stars have revealed that most of these stars pulsate with a multitude of pulsation modes in both the low and high-frequency regions (Uytterhoeven et al. 2011), where the gravity (g) and pressure (p) modes are found. The numerical relationships between many low and high frequencies indicate that it is not possible to treat these regions independently of each other. For example, the low-frequency region contains g-mode pulsations as well as rotation peaks (Balona 2011, Breger et al. 2011), while the high-frequency region contains p-modes as well as high-degree g-modes (Monnier et al. 2010, Breger et al. 2013). Regrettably, pulsation mode identifications are generally not available for these fainter stars studied from space; consequently, frequency patterns have been the most valuable tool. In the present investigation, we add amplitude and phase variations as an additional method to determine the modal origin. The rapidly rotating ($V \\sin i = 300\\,$km/s) hybrid $\\delta$\\,Sct/$\\gamma$\\,Dor pulsator KIC\\,8054146 has been measured extensively by the {\\it Kepler} spacecraft. Altogether, 349 statistically significant frequencies were determined from three years of short-cadence {\\it Kepler} data (Breger et al. 2012, hereafter Paper 1). The amplitudes ranged from about 2 ppm (parts-per-million) to 200 ppm. The excellent frequency resolution of three years of data for KIC\\,8054146 also revealed that many frequencies are related over and beyond the expected simple harmonics and combinations. In fact, three separate families of frequencies spanning a 200 cycles day$^{-1}$ frequency range have been discovered so far. Within each family, the amplitude variations of the low-frequency members correlate with those of the high-frequency members. The unprecedented accuracy of the {\\it Kepler} allows us examine the physical origin of these families in detail. One of these families could already be interpreted as high-degree prograde Kelvin modes (Breger et al. 2013). \\begin{table} \\caption{Dominant frequencies of the ``T'' family} \\begin{tabular}{llrl} \\hline \\noalign{\\smallskip} Name & \\multicolumn{2}{c}{Frequency} & Comment\\\\ & cycles day$^{-1}$ & $\\mu$Hz \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} f3 & 2.9308 & 33.92 & Low frequency\\\\ f3b & 2.9334 & 33.95 & Low frequency\\\\ f6 & 5.8642 & 67.87 & =f3+f3b\\\\ f26 & 25.9509 & 300.36 \\\\ f34 & 34.4836 & 399.12 & Triplet 2\\\\ f37 & 37.4170 & 433.07 & Triplet 2\\\\ f40 & 40.3479 & 466.99 & Triplet 2\\\\ f60 & 60.4346 & 699.47 & Triplet 1\\\\ f63 & 63.3680 & 733.43 & Triplet 1\\\\ f66 & 66.2988 & 767.35 & Triplet 1\\\\ \\hline \\end{tabular} \\end{table} This paper investigates the dominant set of modes excited in KIC\\,8054146, which we have called the T family. The tools for this investigation are not only the frequency values, but also their large and steady amplitude and phase changes over the three years. The bottom panel of Fig. 1 shows the amplitude spectrum of KIC 8054146 in the 0 to 100 cycles day$^{-1}$ range. For more observational information we refer to Paper I, which was extended to include the {\\it Kepler} short-cadence data from quarters 11 to 14 (Q11 to Q14). Since the frequency resolution of the combined Q5-Q14 data exceeds the time scales of the (small) frequency variations of the star, the diagram actually shows the sum of four subsets. The top panel of Fig. 1 presents the T family of dominant modes with additional properties of these modes listed in Table 1. While equidistant frequency spacings and a numerical relationship between the low and high frequencies are also seen in other $\\delta$ Sct stars studied by $\\it Kepler$, the number and sizes of such occurrences in KIC 8054146 is unusual. In addition, the continuous short-cadence coverage over three years makes a study of the correlations of amplitude variability of the different modes possible: this allows us to uniquely separate the parent and child (coupled) modes. ", "conclusions": "For KIC 8054146 we identify coupled modes and can uniquely distinguish which are the parent and child modes. This is made possible by the fact that for the simplest, lowest-order mode coupling the amplitudes of the coupled modes are given by the product of the amplitudes of the parents. We speculate that the simple mode-coupling model is applicable due to the relatively small amplitudes and therefore lowest-order nonlinear interaction of the modes involved. Consequently, the present study is only made possible by the excellent time coverage and precision of the data from the {\\it Kepler} spacecraft. We note that the generally unknown mode-coupling constants do not present a problem for KIC 8054146 due to the large amount of amplitude variability over the three years of data; a single short stretch of data allows us to derive the mode-coupling coefficients, which subsequent data groups show to be constant. The state of nonlinear or combination frequencies detected in KIC\\,8054146 is the following: \\begin{enumerate} \\item We detect ``normal'' nonlinearities in this star; these are combination frequencies for which the amplitudes are related by $A_3\\sim \\mu * A_1 * A_2$, where $\\mu$ is $\\sim 4$. These small nonlinearities are commonly seen in most if not all Delta Scuti stars. \\item KIC 8054146 also possesses larger-amplitude nonlinear signals, and we are able to use the measured amplitude and phase variations of these signals to determine which modes are the ``parent'' modes and which are the ``child'' modes. The derived value of $\\mu$ relating the amplitudes of child and parent modes is $\\mu \\sim1000$ to 10,000 or more. \\end{enumerate} A generic nonlinear model (such as the nonlinear relation between temperature and flux at the surface) is not able to simultaneously produce both these ``small'' and ``large'' nonlinearities. On the other hand, if a subset of the combination frequencies lie close to eigenfrequencies in the star, then these modes and only these modes can be resonantly driven to larger amplitudes. While this is far from the last word on the subject, the data are fully consistent with this ``resonant mode coupling'' scenario. In particular, the present investigation showed that in KIC 8054146, two dominant low frequencies are actually coupled modes with amplitudes larger or equal to those of the dominant parents in the higher-frequency domain. The low-frequency region of these so-called hybrid pulsators contains a large number of linearly driven gravity modes, as was confirmed by Chapellier et al. (2012) from searching for equidistant period spacings in ID 105733033. Furthermore, in a number of hybrid pulsators, some frequency values in the gravity-mode and pressure-mode domains are numerically not independent of each other. Given the existence of these gravity modes, we speculate that these could be the ``child'' modes that are resonantly excited to larger amplitudes. A confirmation of this resonance explanation would be provided if additional data could reveal that these particular low frequencies are part of a sequence of low-frequency modes equidistantly spaced in period." }, "1402/1402.3479_arXiv.txt": { "abstract": "We present a study of age-related spectral signatures observed in 25 young low-mass objects that we have previously determined as possible kinematic members of five young moving groups: the Local Association (Pleiades moving group, age=20 - 150 Myr), the Ursa Major group (Sirius supercluster, age=300 Myr), the Hyades supercluster (age=600 Myr), IC 2391 supercluster (age=35--55 Myr) and the Castor moving group (age=200 Myr). In this paper we characterize the spectral properties of observed high or low resolution spectra of our kinematic members by fitting theoretical spectral distributions. We study signatures of youth, such as lithium~{\\sc i} 6708 \\AA, H$\\alpha$ emission and other age-sensitive spectroscopic signatures in order to confirm the kinematic memberships through age constraints. We find that 21 (84\\%) targets show spectroscopic signatures of youth in agreement with the age ranges of the moving group to which membership is implied. For two further objects, age-related constraints remain difficult to determine from our analysis. In addition, we confirm two moving group kinematic candidates as brown dwarfs. ", "introduction": "Identifying members of known moving groups (MG) or open clusters can provide an important constraint on their age and composition. Low-mass stars and brown dwarfs (BDs) members of known MGs or open clusters can provide important feedback to atmospheric and evolutionary models of such objects at young ages, which currently await good calibration from constraints on age and composition. Therefore, there exists an imperative to compile a sample of MG members in this mass range that can be used as anchor points or test-beds for these models. In this paper we present the last in a series of results of a major survey of Ultra-Cool Dwarfs (UCDs; objects with a spectral classification of M7 or later corresponding to Teff$<$2500 K), in young moving groups which began with the results presented in Clarke et al. (2010), hereafter Paper I, and G\\'alvez-Ortiz et al. (2010), hereafter Paper II. We focus here on the study of age-sensitive spectroscopic signatures such as gravity sensitive features (e.g. Gorlova et al. 2003; McGovern et al. 2004), H$\\alpha$ emission (e.g. West et al. 2004, 2008, 2011) and the presence of Li~{\\sc i} 6708 \\AA \\ in a sample of objects previously classified as members of a MG based on kinematic or astrometric criteria. Despite the recent disagreement about the origin of moving groups (Famaey et al. 2007, 2008; Antoja et al. 2008; Klement et al. 2008; Francis \\& Anderson 2009; Zhao et al. 2009; L\\'opez-Santiago et al. 2009; Bovy \\& Hogg 2010; Murgas et al. 2013), and as we discussed in Paper II, we are assuming the classical concept: a moving group is a young stellar population that shares a common space motion (e.g. Pinfield et al. 2006) whose members have a common origin, and therefore, age and composition. We focused our studies on well-documented groups: the Hyades supercluster (HY; 600 Myr), the Ursa Major group (Sirius supercluster) (Si; 300 Myr), IC 2391 supercluster (IC; 35-55 Myr), the Castor Moving Group (CA; 200 Myr), and the Local Association (20-150 Myr) or Pleiades (PL) moving group (See Paper I and II for details of MG properties and references thereof). Many efforts have been made to obtain accurate models to understand the cool and complex atmospheres of low mass stellar and substellar objects (e.g. Kirkpatrick et al. 1993; Rajpurohit et al. 2011; Reyle et al. 2011; Rajpurohit et al. 2012). But the models still show discrepancies, for example in the strength of some absorption bands: discrepancies likely due to inaccurate atomic parameters and/or missing molecular opacities. Also, dust formation and its behaviour with temperature changes spectral characteristics in many ways. Models that describe Very Low-Mass stars (VLMS) and BD atmospheres cannot hope to be in agreement with observations without the inclusion of dust. The onset of iron and silicates dust grain formation suspended in the photospheres of late-type M dwarfs through the L dwarf spectral sub-types is accompanied by spectral reddening, especially in near-IR (based upon the size, quantity and distribution of this dust). The T dwarf spectral sub-type sequence demonstrates a reversal to bluer colours due to the dust settling below the photosphere, thus reducing the reddening. Consequently, atmospheric models that describe dwarf objects with T$_{\\rm eff}$ $\\leqslant$ 2800 K should include treatment for the effects of atmospheric dust evolution. Up to date, there are several families of stellar model atmospheres of very low mass stars. These models are still incomplete or approximate in some physical properties (opacities, oscillator strengths for some lines and molecular bands, etc), showing few or considerable discrepancies depending on the atmospheric regions, temperatures, etc, and there are also non covered areas or \"gaps\". A comprehensive study of models tested by observations is needed to fully understand this cool atmospheres. Since BDs are objects occupying an intermediate position between stars and giant planets, studying their atmospheres can help us to better understand the processes in giant exo-planet atmospheres. The determination of the physical properties of VLMS and BDs is also important for understanding a broad range of topics including stellar and planetary formation, circumstellar disks, dust formation in cool atmospheres, and the initial mass function. By comparing high and low resolution optical spectra of our 25 low-mass stars and BD kinematic MG candidates with a grid of atmospheric models, we can obtain effective temperature and surface gravity estimates. In Section~2 we describe the sample characteristics and selection criteria, while in Section~3 we present details of our observations and data reduction techniques. In Section~4 we explain the model atmosphere fitting process, and in Section~5 we describe the age indicators and the analysis used to assess the sample memberships. Finally, in Section~6 we present a brief discussion and summary of our results. ", "conclusions": "We have presented a study of the spectral signatures of 25 low-mass objects that were candidate members of five young moving groups. We studied different typical age-constraining spectroscopic criteria utilising high and low resolution spectra, and combined the results to extract a final membership assessment for each target. We took into account spectral classification and thus approximated atmospheric temperature to apply the most appropriate model or models in the search of the best physical parameter constraints. When we used the synthetic stellar atmospheric models, the good agreement between observed and theoretical energy distributions suggests that semi-empirical models describe well the impact of dust in the atmospheres of M dwarfs. $S$-function analysis shows that the inclusion of dust effects makes it possible to achieve better fits for objects with $T_{\\rm eff}\\leqslant2800$~K. In both samples, although youth can generally be established, $log~g$ and $EW$(Na~{\\sc i}) were not useful for discriminating between different MGs. Some targets of sample A that were candidates to the HY moving group having lower $log~g$ values than candidates to younger MGs, and similarly in $EW$(Na~{\\sc i}) values. In sample B the values are more consistent but most of the targets have several candidatures, making the situation more complicated. However, looking at the final membership results, only one target of sample A is classified as a member of a MG with age $<$200 Myr, which makes this criterion non applicable for most of them. For sample B, in spite of the multiple membership candidatures in the final results, we can discriminate between the lower values of $log~g$ (or lower values of $EW$s(Na~{\\sc i})) for targets that are members of younger MGs, medium values for targets that are members of intermediate age MGs and the higher values for targets that are members of oldest MGs, which finally is consistent with the $log~g$-age relation. The comparison of $EW$(Na~{\\sc i})s is probably useful to discriminate very young targets (less than 100 Myr) in a homogeneous sample (as suggested by Schlieder et al. 2012) while $log~g$ values are probably useful to discriminate very young (less than 100 Myr) and very old (more that 600 Myr) objects, serving as complementary study to other criteria. The activity-age relation is a reasonably useful age criterion up to $\\approx$M7 spectral type but activity variability, binarity, and the comparison of different instruments/resolutions can introduce considerable uncertainties in age estimation. The presence of Lithium discriminates between young and old low mass stars but an ambiguity between young low mass stars and BDs limits its use in this cool temperature region. Gravity sensitive features and rotational velocity can be useful youth indicators in support of activity and lithium diagnostics, but clearly have large dispersion and they need to be applied in conjunction with similar data. For a reliable constraint of cool object ages, a combination of different criteria are needed. We find 10 of the 13 objects from sample A to be probable members of one of the MGs, and also that 2MASS0020-2346 is a young disk member. SIPS0007-2458, a candidate member of the IC2391 MG, shows positive candidature under most criteria but does not show the expected lithium (associated with the age of the IC2391 MG). Due to the age constraint from $H\\alpha$ emission and surface gravity, we are inclined to think that SIPS0007-2458 is probably a young object from the YD class, but is not in the IC2391 MG, although it could also be a contaminant field object. 2MASS0334-2130 is a similar case, with an M4.5-6 spectral type, it should contain lithium if it were an IC2391 MG member. Except for the $H\\alpha$ criterion, that could be reflecting a maximum in activity level or the influence of an unseen companion, everything indicates that 2MASS0334-2130 is older than the IC2391 MG. Thus we also conclude that 2MASS0334-2130 is not an IC2391 member, although it could be a young object from the YD, or a contaminant target from the field. In sample B, 10 objects could be MG members, although more information is needed to discriminate between the possible MG membership of objects with several candidatures. We could not obtain any results for 2MASS1909-1937, so this remains an astrometric candidate to the Castor MG. 2MASS1734-1151 gave contradictory results so it also remains an astrometric Pleiades MG candidate until further analysis can be performed. We also confirm two moving group candidates, SIPS2045-6332 and SIPS2039-1126, as BDs. With all the acquired information, we find that 85\\%, 83\\% and 84\\% of samples A, B, and both combined, show spectroscopic signatures of youth in agreement with the age of the moving group to which they present kinematic membership. This result suggests that there is a fairly low rate of contamination in such kinematic candidate samples. Also, there are candidates that cannot be confirmed or dismissed with the information obtained, which might (in the future) increase the final confirmation rate. We note that additional diagnostics to assess membership may be measured in the future, such as chemical tagging. Recently, in a kinematical-chemical investigation of the AB Dor moving group \u201cStream\u201d, Barenfeld et al. (2013) shows that kinematics, color-magnitude positions, and stellar youth indicators alone can be insufficient for testing whether a kinematic group of stars actually shares a common origin. Future chemical study of our MG candidates will complement this study. In addition, this study has also helped to test and improve the atmosphere models for cool dwarfs (see Kuznetsov et al. 2012 and Kuznetsov et al. 2013). Youth can be an advantage for various reasons. Because young objects are brighter they can be detected and studied out to greater distance than older, fainter objects. Also, the proximity of the MGs can allows the exploration of the faint circumstellar environment at relatively small distances from the star. Indeed, we will target our strongest MG candidates to search for lower mass companions using high resolution imaging techniques (e.g. adaptive optics). The discovery of planetary systems around young and low-mass objects provides crucial information for the understanding of planetary and stellar formation. From our complete study of the 80 objects considered (within papers I, II and this paper), we have found a total of 45 new possible MG members (with 8 of them showing more than one MG candidature) and around 21 possible young disk objects with no clear membership of the 5 MGs considered. Of these we find 26, 13, 6, 4 and 5 possible members of the Hyades, Castor, Ursa Major, Pleiades and IC2391 MGs respectively. Tables~\\ref{tab:comptod} and \\ref{tab:compothers} of the appendix compile all the M-L dwarfs investigated in our study. \\begin{figure*} \\centering \\includegraphics[width=8.5cm,angle=0,clip]{Figura1_1.ps} \\includegraphics[width=8.5cm,angle=0,clip]{Figura1_2.ps} \\caption{Left: Observed SEDs of objects from sample A, the 13 targets plus reference objects in the 7600 to 8400 \\AA \\ range. Right: Observed SEDs of objects from sample B, the 12 targets plus one reference object in the 6500 to 9000 \\AA \\ range. They are ordered by spectral type derived from spectroscopy.} \\label{fig:todashl} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=6.0cm,angle=270,clip]{Figura2_1.eps} \\includegraphics[width=6.0cm,angle=270,clip]{Figura2_2.eps} \\caption{Sample A: random example of fits of theoretical spectra to the observed SEDs.} \\label{fig:fight} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=6.0cm,angle=270,clip]{Figura3_1.eps} \\includegraphics[width=6.0cm,angle=270,clip]{Figura3_2.eps} \\caption{Sample B: random example of the best fits to the observed SEDs. } \\label{fig:figlt} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=6.0cm,angle=270,clip]{Figura4_1.eps} \\includegraphics[width=6.0cm,angle=270,clip]{Figura4_2.eps} \\caption{Sample A: random example of the best fits to the spectral regions across K~{\\sc i} lines. } \\label{fig:fighk} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=6.0cm,angle=270,clip]{Figura5_1.eps} \\includegraphics[width=6.0cm,angle=270,clip]{Figura5_2.eps} \\caption{Sample A: random example of the best fits to the spectral regions across Na~{\\sc i} lines. } \\label{fig:fighna} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=7.0cm,angle=270,clip]{Figura7.eps} \\caption{ This figure (updated Figure~9 from Paper I), illustrates how an object would be selected for the various methods of age constraining. Candidates that appear to the right of the lithium edge (blue continuous line) can be followed up with a lithium test programme. Objects that appear younger than 200 Myr (horizontal red dashed line), are eligible for follow-up using spectroscopic gravity sensitive features. Vertical and horizontal orange dot-dashed line represents the Schlieder et al. (2012) limits for the Na~{\\sc i} diagnostic applicability. We here took into account the 200 Myr limit. Candidates that fall to the left left of the spectral type=M7 limit (vertical black dotted line), would thus be suitable for age/activity relation follow-up, although candidates with a spectral type close to M7 may be subject to large uncertainties on their age. Some candidate cannot be tested by any of these methods but will be eligible for age testing using $v\\sin{i}$. Sample A targets are plot as blue circles and sample B as red circles. We plotted up to two candidatures for targets with multiple MG.} \\label{fig:modelo} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=8.0cm,clip]{Figura8n.ps} \\includegraphics[width=8.0cm,clip]{Figura8nb.ps} \\caption{ The right and left panels show two versions of the same information with different legends in order to highligh different information. Left: From Mart\\'in et al. (2010), 65 high-gravity field objects are plotted as six pointed asterisk symbol, 6 low-gravity objects as open triangles, 12 reference field stars as open circles and 7 Upper Sco candidates as solid hexagons. We overplot sample A as blue crosses, and sample B as red filled circles. The red star marks LP944-20. Errors are approximately the size of plot symbols. Right: Here we overplot to Mart\\'in et al. (2010) data, sample A (crosses and filled squares), and sample B (filled circles and triangles), plotted in different colours depending on their final MG membership candidature (Table~\\ref{tab:resultfinal}). When objects present more than one candidature, we plotted the membership to the youngest MG. To better discriminate young targets, we use triangles and squares for candidates finally classified as possible member of MGs with ages $\\geq$200 Myr while crosses and circles are candidates finally classified as possible member of MGs with ages $<$200 Myr (see Sect. 6). } \\label{fig:ewnacomp} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=8.5cm,clip]{Figura9_1.ps} \\includegraphics[width=8.5cm,clip]{Figura9_2.ps} \\caption{ $EW$(H$\\alpha$) versus spectral type. In the same way as in Fig.~7, both panels show two versions of the same information with different legends in order to highligh different information. Left: Sample A is plotted as blue crosses, Sample B as red circles and literature known-age objects in different symbols according to age: red asterisk for objects with age between 1-12 Myr, blue diamonds for 40-50 Myr, violet asterisk for 90-100 Myr, blue asterisk for 120 Myr, red diamonds for 150 Myr, black asterisk for 300 Myr, black diamonds for 400 Myr and green asterisk for objects with ages over 1200 Myr. Literature data has been obtained from Terndrup et al. (2000), Mohanty \\& Basri (2003) and Shkolnik et al. (2009) where ages were calculated by known-age MG or Cluster membership or by other age constraining methods. Right: Similar to left panel where to favor the age discrimination we plot only three age intervals. Sample A targets are plotted as filled squares and sample B as filled triangles, where targets classified with ages $\\geq$300 Myr are plotted in blue and targets with ages $<$300 Myr are plotted in red. } \\label{fig:ewhacomp} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=6.0cm,angle=270,clip]{Figura10_1.eps} \\includegraphics[width=6.0cm,angle=270,clip]{Figura10_2.eps} \\caption{Fit to the observed spectra across Li~{\\sc i} resonance doublet for SIPS2045-6332 (left) and SIPS2039-1126 (right). Best synthetic fits are overplot with different Li abundances. } \\label{fig:lithium} \\end{figure*} \\begin{table*} \\caption[]{A: Properties for sample A \\label{tab:samplea}} \\begin{flushleft} \\begin{center} \\small \\begin{tabular}{lllccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} Name & $\\alpha$ (2000) & $\\delta$ (2000) & J-mag & Moving group$^{1}$ & $V\\sin{i}$$^{2}$ \\\\ & (h m s) & ($^{\\rm o}$ ' '') & & candidature & (km s$^{-1}$) \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} SIPS0007-2458 & 0 7 7.800 & -24 58 3.80 & 13.11$\\pm$0.02 & IC & 18 \\\\ 2MASS0020-2346 & 0 20 23.155 & \u221223 46 5.38 & 12.35$\\pm$0.02 & YD & 12 \\\\ DENIS0021-4244 & 0 21 5.896 & -42 44 43.33 & 13.52$\\pm$0.03 & IC, CA & 11 \\\\ SIPS0027-5401 & 0 27 23.240 & \u221254 1 46.20 & 12.36$\\pm$0.02 & HY & 27 \\\\ SIPS0153-5122 & 1 53 11.430 & \u221251 22 24.99 & 13.45$\\pm$0.03 & IC, HY & 14 \\\\ SIPS0214-3237 & 2 14 45.440 & -32 37 58.20 & 14.01$\\pm$0.02 & HY & 18 \\\\ SIPS0235-0711 & 2 35 49.470 & \u22127 11 21.90 & 12.45$\\pm$0.03 & HY & 22 \\\\ 2MASS0334-2130 & 3 34 10.657 & \u221221 30 34.35 & 11.91$\\pm$0.02 & IC & $<$10$^{a}$ \\\\ SIPS2039-1126 & 20 39 13.081 & -11 26 52.30 & 13.79$\\pm$0.03 & PL & $>$15 \\\\ SIPS2045-6332 & 20 45 2.278 & -63 32 5.30 & 12.62$\\pm$0.03 & PL, CA & $>$15 \\\\ LEHPM4908 & 22 36 42.656 & -69 34 59.30 & 12.68$\\pm$0.02 & HY & - \\\\ 2MASS2254-3228 & 22 54 58.110 & \u221232 28 52.20 & 13.58$\\pm$0.03 & PL, CA & 19 \\\\ LEHPM6542 & 23 57 54.822 & \u221219 55 1.89 & 13.31$\\pm$0.02 & HY & 16 \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} $^{1}$ MG to which targets are kinematic candidate (Paper II). HY= Hyades MG; SI= Ursa Major group; CA= Castor MG; PL= Pleiades; IC = IC 2391 MG; YD= Other young disk object.\\\\ $^{2}$ From paper II.\\\\ $^{a}$ Measused with low S/N.\\\\ \\end{flushleft} \\end{table*} \\begin{table*} \\caption[]{B: Properties for sample B \\label{tab:sampleb}} \\begin{flushleft} \\begin{center} \\small \\begin{tabular}{lllcccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} Name & $\\alpha$ (2000) & $\\delta$ (2000) & J-mag & Moving group \\\\ & (h m s) & ($^{\\rm o}$ ' '') & & candidature \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} 2MASS0814-4020 & 08 14 35.46 & -40 20 49.26 & 14.356$\\pm$0.023 & HY \\\\ 2MASS1146-4754 & 11 46 51.04 & -47 54 38.17 & 14.897$\\pm$0.042 & SI \\\\ 2MASS1236-6536 & 12 36 32.38 & -65 36 35.6 & 15.277$\\pm$0.057 & PL, CA, IC, SI \\\\ 2MASS1326-5022 & 13 26 53.48 & -50 22 27.04 & 14.715$\\pm$0.037 & IC, CA \\\\ 2MASS1433-5148 & 14 33 41.95 & -51 48 03.70 & 14.206$\\pm$0.034 & CA, PL, SI, IC \\\\ 2MASS1557-4350 & 15 57 27.39 & -43 50 21.47 & 14.224$\\pm$0.028 & PL, CA, IC \\\\ 2MASS1618-3214 & 16 18 08.92 & -32 14 36.17 & 14.920$\\pm$0.040 & IC, CA, SI \\\\ 2MASS1734-1151 & 17 34 30.53 & -11 51 38.83 & 13.110$\\pm$0.028 & PL \\\\ 2MASS1736-0407 & 17 36 56.09 & - 4 07 25.84 & 15.516$\\pm$0.070 & SI, CA \\\\ 2MASS1745-1640 & 17 45 34.66 & -16 40 53.81 & 13.646$\\pm$0.026 & SI, CA, HY \\\\ 2MASS1756-4518 & 17 56 29.63 & -45 18 22.47 & 12.386$\\pm$0.019 & CA \\\\ 2MASS1909-1937 & 19 09 08.21 & -19 37 47.96 & 14.520$\\pm$0.026 & CA \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} \\end{flushleft} \\end{table*} \\begin{table*} \\caption[]{Details of observing runs \\label{tab:obs}} \\begin{flushleft} \\begin{center} \\small \\begin{tabular}{ccccccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} Number & Date & Telescope & Instrument & Spect. range & Dispersion \\\\ & & & & (\\AA) & (\\AA) \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} 1 & 28/03-21/06/08 & ESO-VLT-U2 & UVES$^{1}$ & 6700-10425$^{2}$ & 0.027-0.041 \\\\ 2 & 01-09-2009-09/01/2010 & ESO-VLT-U2 & UVES$^{1}$ & 5700-7530 \\& 7650-9470 & 0.027-0.041\\\\ 3 & 03-05/05/2010 & 6.5~m Baade-Maguellan & IMACS Short-Camera$^{3}$ & 6550-10000 & 1.98 \\\\ 4 & 15-16/02/2011 & 6.5~m Baade-Maguellan & IMACS Short-Camera$^{3}$ & 4300-10800 & 1.97 \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} $^{1}$ UVES: Ultraviolet and Visual Echelle Spectrograph.\\\\ $^{2}$ Effective range.\\\\ $^{3}$ IMACS: The Inamori Magellan Areal Camera and Spectrograph.\\\\ \\end{flushleft} \\end{table*} \\begin{table*} \\caption[]{Sample A: Model fit results for M3-M7.5 objects. Spectroscopic spectral types are from Paper II. The equivalent effective temperatures for the spectral types were taken from Reyl\\'e et al. (2011). $\\Delta$T$_{\\rm eff}$= 100 K, $\\Delta$$log~g$= 0.5 cm s$^{-2}$. \\label{tab:resulta}} \\begin{flushleft} \\begin{center} \\small \\begin{tabular}{lcccccccccccccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} & \\multicolumn{3}{c}{$SpT$} & \\multicolumn{5}{c}{NextGen} & \\multicolumn{4}{c}{Semi-empirical model} & \\\\ \\noalign{\\smallskip} Object & $SpT_{Spec}$ & $T_{\\rm eff}SpT$ & & $T_{\\rm eff}$ & $log~g$ & [M/H] & $S_{min}$ & & $T_{\\rm eff}$ & $log~g$ & [M/H] & $S_{min}$ & best fit$^{1}$ \\\\ & & (K) & & (K) & (cm s$^{-2}$) & dex & & & (K) & cm s$^{-2}$ & dex & & \\\\ \\hline \\noalign{\\smallskip} GL876 & M4.0 & 3100 & & 3400 & 5.5 & 0.0 & 2.9 & & - & - & - & - & NextGen \\\\ SIPS0007-2458 & M7.5 & 2550 & & 2700 & 4.5 & -1.0 & 7.0 & & 2700 & 4.5 & 0.0 & 6.2 & s.-e. \\\\ 2MASS0020-2346 & M6.0 & 2800 & & 2800 & 5.5 & -0.5 & 4.8 & & 2800 & 5.0 & 0.0 & 4.56 & s.-e. \\\\ SIPS0027-5401 & M6.5 & 2650 & & 2700 & 4.5 & -1.0 & 3.35 & & 2700 & 4.5 & 0.0 & 2.70 & s.-e. \\\\ SIPS0153-5122 & M6.0 & 2800 & & 2800 & 5.5 & -0.5 & 2.42 & & 2800 & 5.0 & 0.0 & 2.31 & s.-e. \\\\ SIPS0214-3237 & M6.5 & 2650 & & 2800 & 5.0 & -1.0 & 1.44 & & 2800 & 5.0 & 0.0 & 1.2 & s.-e. \\\\ SIPS0235-0711 & M6.0 & 2800 & & 2700 & 5.0 & -1.0 & 0.97 & & 2700 & 4.5 & 0.0 & 0.91 & s.-e. \\\\ 2MAS0334-2130 & M4.5-M6.0 & 3000-2800 & & 2800 & 5.5 & -0.5 & 4.65 & & 2800 & 5.0 & 0.0 & 5.65 & $^{a}$ \\\\ LEHPM4908 & M6.0 & 2800 & & 2800 & 4.5 & -1.0 & 7.02 & & 2900 & 5.0 & 0.0 & 5.70 & s.-e. \\\\ 2MASS2254-3228 & M5.5 & 2850 & & 2700 & 5.0 & -1.0 & 1.37 & & 2700 & 4.5 & 0.0 & 1.25 & s.-e. \\\\ LEHPM6542 & M6.0 & 2800 & & 2800 & 5.0 & -0.5 & 13.5 & & 2700 & 4.5 & 0.0 & 4.03 & s.-e. \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} $^{1}$ NextGen or semi-empirical (s.-e.) models.\\\\ $^{a}$ both models are equally likely. \\\\ \\end{flushleft} \\end{table*} \\begin{table*} \\caption[]{Sample B: Model fit results for M3-M7.5 objects. The equivalent effective temperature for the spectral types were taken from Reyl\\'e et al. (2011). $\\Delta$T$_{\\rm eff}$= 100 K, $\\Delta$$log~g$= 0.5 cm s$^{-2}$. \\label{tab:resultb}} \\begin{flushleft} \\begin{center} \\scriptsize \\begin{tabular}{lccccccccccccccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} & \\multicolumn{4}{c}{$SpT$} & \\multicolumn{5}{c}{NextGen} & \\multicolumn{4}{c}{Semi-empirical model} & \\\\ \\noalign{\\smallskip} Object & $SpT_{Phot}$ & $SpT_{Spec}$ & $T_{\\rm eff}Sp$ & & $T_{\\rm eff}$ & $log~g$ & [M/H] & $S_{min}$ & & $T_{\\rm eff}$ & $log~g$ & [M/H] & $S_{min}$ & best fit \\\\ & & (K) & (K) & & (K) & (cm s$^{-2}$) & dex & & & (K) & cm s$^{-2}$ & dex & & \\\\ \\hline \\noalign{\\smallskip} 2MASS1146-4754 & M8.5 & M7-8 & $\\sim$2500 & & 2800 & 5.5 &-0.5 & 0.39 & & 2700 & 4.5 & 0.0 & 0.18 & s.-e. \\\\ 2MASS1236-6536 & M7.5 & M4.0 & 3100 & & 3200 & 5.5 & 0.0 & 0.64 & & 2900 & 4.0 & 0.0 & 1.14 & NextGen \\\\ 2MASS1326-5022 & M9.0 & M7.5 & 2550 & & 2800 & 5.0 & 0.0 & 0.57 & & 2700 & 4.0 & 0.0 & 0.22 & s.-e. \\\\ 2MASS1433-5148 & M9.0 & M6-7 & $\\sim$2650 & & 3000 & 5.5 & 0.0 & 0.58 & & 2900 & 5.0 & 0.0 & 0.27 & s.-e. \\\\ 2MASS1618-3214 & M7.0 & M6.5 & 2650 & & 2800 & 5.5 &-0.5 & 0.69 & & 2700 & 4.5 & 0.0 & 0.41 & s.-e. \\\\ 2MASS1736-0407 & M8.5 & M4-5 & $\\sim$3000 & & 3200 & 5.0 & 0.0 & 0.49 & & 2900 & 4.0 & 0.0 & 0.55 & NextGen \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} \\end{flushleft} \\end{table*} \\begin{table*} \\caption[]{Sample A: Model fit results for $\\sim$M8 object. First line refers to fit results when use NextGen, DUSTY and COND models while second line refers to fit results when use the semi-empirical models based on NextGen, DUSTY and COND models respectively. Spectroscopic spectral types are from Paper II. $\\Delta$T$_{\\rm eff}$= 100 K, $\\Delta$$log~g$= 0.5 cm s$^{-2}$. \\label{tab:resultab}} \\begin{flushleft} \\begin{center} \\scriptsize \\begin{tabular}{lccccccccccccccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} & \\multicolumn{2}{c}{$SpT$} & \\multicolumn{4}{c}{NextGen} & \\multicolumn{4}{c}{DUSTY} & \\multicolumn{4}{c}{COND} & \\\\ Object & $SpT_{Spec}$ & $T_{\\rm eff}SpT$ & & $T_{\\rm eff}$ & $log~g$ & $S_{min}$ & & $T_{\\rm eff}$ & $log~g$ & $S_{min}$ & & $T_{\\rm eff}$ & $log~g$ & $S_{min}$ & best fit \\\\ & & (K) & & (K) & (cm s$^{-2}$) & & & (K) & cm s$^{-2}$ & & & (K) & cm s$^{-2}$ & & \\\\ \\hline \\noalign{\\smallskip} SIPS2039-1126 & M8.0 & ~2500 & & 2800 & 5.5 & 22.45 & & 2600 & 5.0 & 39.96 & & 2600 & 4.5 & 40.12 & \\\\ Semi-empirical & & & & 2700 & 4.5 & 22.90 & & 2600 & 4.5 & 20.00 & & 2600 & 4.5 & 20.01 & s.-e. DUSTY \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} \\end{flushleft} \\end{table*} \\begin{table*} \\caption[]{Sample B: Model fit results for $\\sim$M8 objects. First line refers to fit results when use NextGen, DUSTY and COND models while second line refers to fit results when use the semi-empirical models based on NextGen, DUSTY and COND models respectively. $\\Delta$T$_{\\rm eff}$= 100 K, $\\Delta$$log~g$= 0.5 cm s$^{-2}$. \\label{tab:resultbb}} \\begin{flushleft} \\begin{center} \\scriptsize \\begin{tabular}{lccccccccccccccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} & \\multicolumn{2}{c}{$SpT$} & \\multicolumn{4}{c}{NextGen} & \\multicolumn{4}{c}{DUSTY} & \\multicolumn{4}{c}{COND} & \\\\ Object & $SpT_{Phot}$ & $SpT_{Spec}$ & & $T_{\\rm eff}$ & $log~g$ & $S_{min}$ & & $T_{\\rm eff}$ & $log~g$ & $S_{min}$ & & $T_{\\rm eff}$ & $log~g$ & $S_{min}$ & best fit \\\\ & & & & (K) & (cm s$^{-2}$) & & & (K) & cm s$^{-2}$ & & & (K) & cm s$^{-2}$ & & \\\\ \\hline \\noalign{\\smallskip} 2MASS0814-4020 & L4.0 & M7-8 & & 2700 & 5.0 & 0.74 & & 2600 & 5.5 & 1.15 & & 2600 & 5.0 & 1.22 & \\\\ Semi-empirical & & & & 2700 & 4.5 & 0.35 & & 2600 & 4.5 & 0.28 & & 2600 & 4.5 & 0.28 & s.-e. DUSTY $^{1,2}$\\\\ 2MASS1557-4350 & L0.0 & M7.5-8.5 & & 2800 & 4.5 & 0.74 & & 2600 & 5.5 & 0.93 & & 2400 & 5.5 & 1.11 \\\\ Semi-empirical & & & & 2800 & 4.0 & 0.31 & & 2500 & 4.0 & 0.29 & & 2600 & 4.0 & 0.29 & s.-e. DUSTY $^{1}$\\\\ 2MASS1756-4518 & M7.5 & M8-9 & & 2600 & 5.5 & 0.21 & & 2600 & 5.5 & 0.18 & & 2600 & 5.0 & 0.20 & \\\\ Semi-empirical & & & & 2600 & 4.0 & 0.09 & & 2600 & 5.5 & 0.08 & & 2600 & 5.0 & 0.10 & s.-e. DUSTY \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} $^{1}$ DUSTY and COND have the same $S_{min}$, but DUSTY model is more appropriate in the temperature range.\\\\ $^{2}$ We could appreciate parts of the spectrum were NextGen was the best fit.\\\\ \\end{flushleft} \\end{table*} \\begin{table*} \\caption[]{Sample A: Model fit results for M8.5-M9.5 objects. First line refers to fit results when use DUSTY and COND models while second line refers to fit results when use the semi-empirical models based on DUSTY and COND models respectively. Spectroscopic spectral types are from Paper II. The equivalent effective temperature for the spectral types were taken from Dahn et al. (2002). $\\Delta$T$_{\\rm eff}$= 100 K, $\\Delta$$log~g$= 0.5 cm s$^{-2}$. \\label{tab:resultac}} \\begin{flushleft} \\begin{center} \\small \\begin{tabular}{lcccccccccccccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} & \\multicolumn{3}{c}{$SpT$} & \\multicolumn{5}{c}{DUSTY} & \\multicolumn{4}{c}{COND} & \\\\ \\noalign{\\smallskip} Object & $SpT_{Spec}$ & $T_{\\rm eff}SpT$ & & $T_{\\rm eff}$ & $log~g$ & [M/H] & $S_{min}$ & & $T_{\\rm eff}$ & $log~g$ & [M/H] & $S_{min}$ & best fit \\\\ & & (K) & & (K) & (cm s$^{-2}$) & dex & & & (K) & cm s$^{-2}$ & dex & & \\\\ \\hline \\noalign{\\smallskip} DENIS0021-4244 & M9.5 & 2300-2400 & & 2000 & 5.5 & 0.0 & 7.85 & & 2400 & 5.5 & 0.0 & 13.40 & \\\\ Semi-empirical & & & & 2400 & 4.0 & 0.0 & 5.89 & & 2500 & 4.0 & 0.0 & 6.62 & s.-e. DUSTY \\\\ SIPS2045-6332 & M8.5 & 2300-2400 & & 2000 & 5.5 & 0.0 & 11.39 & & 2400 & 4.0 & 0.0 & 17.19 & \\\\ Semi-empirical & & & & 2400 & 4.0 & 0.0 & 8.14 & & 2500 & 4.0 & 0.0 & 10.30 & s.-e. DUSTY \\\\ LP944-20 & M9.0 & 2300-2400 & & 2000 & 5.5 & 0.0 & 0.42 & & 2400 & 5.5 & 0.0 & 0.57 & \\\\ Semi-empirical & & & & 2400 & 4.0 & 0.0 & 0.29 & & 2400 & 4.0 & 0.0 & 0.35 & s.-e. DUSTY \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} \\end{flushleft} \\end{table*} \\begin{table*} \\caption[]{Sample B: Model fit results for $>$M8.5 objects. First line refers to fit results when use DUSTY and COND models while second line refers to fit results when use the semi-empirical models based on DUSTY and COND models respectively. The equivalent effective temperature for the spectral types were taken from Reyl\\'e et al. (2011). $\\Delta$T$_{\\rm eff}$= 100 K, $\\Delta$$log~g$= 0.5 cm s$^{-2}$. \\label{tab:resultbc}} \\begin{flushleft} \\begin{center} \\tiny \\begin{tabular}{lccccccccccccccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} & \\multicolumn{3}{c}{$SpT$} & \\multicolumn{5}{c}{DUSTY} & \\multicolumn{4}{c}{COND} & \\\\ Object & $SpT_{Phot}$ & $SpT_{Spec}$ & $T_{\\rm eff}SpT$ & & $T_{\\rm eff}$ & $log~g$ & [M/H] & $S_{min}$ & & $T_{\\rm eff}$ & $log~g$ & [M/H] & $S_{min}$ & best fit \\\\ & & & (K) & & (K) & (cm s$^{-2}$) & dex & & & (K) & cm s$^{-2}$ & dex & & \\\\ \\hline \\noalign{\\smallskip} 2MASS1734-1151 & L0.0 & M9.0 & 2300-2500 & & 2600 & 5.5 & 0.0 & 1.34 & & 2600 & 5.0 & 0.0 & 1.43 & \\\\ Semi-empirical & & & & & 2400 & 4.0 & 0.0 & 0.79 & & 2400 & 4.0 & 0.0 & 0.86 & s.-e. DUSTY \\\\ 2MASS1745-1640 & L8.0 & M9-L2 & $\\sim$2300-2000 & & 2000 & 5.5 & 0.0 & 0.06 & & 2200 & 5.5 & 0.0 & 0.07 & \\\\ Semi-empirical & & & & & 2000 & 5.5 & 0.0 & 0.05 & & 2200 & 5.5 & 0.0 & 0.05 & s.-e. DUSTY \\\\ 2MASS1909-1937 & M9.0 & L0.0 & $\\sim$2200 & & 2100 & 3.5 & 0.0 & 0.55 & & - & - & - & - & \\\\ Semi-empirical & & & & & 2000/2100 & 6.0/3.5 & 0.0 & 0.45 & & - & - & - & - & s.-e. DUSTY ? \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} \\end{flushleft} \\end{table*} \\begin{table*} \\centering \\caption[]{MG membership parameters. For each target, we give the result obtained for each criterion used. N/A= non applicable criterion in this case; Y= the age parameter agrees with the MG membership; N= the age parameter does not agree with the MG membership; when more than one membership was possible in column 5, the MG name that the criteria determined as possible is given. Where HY is Hyades, SI is Ursa Major, IC is IC 2391, CA is Castor and PL is Pleiades MG and YD is other young disk object. When interrogation appears after MG name we indicate that the membership probability is fewer than for other MGs or in the case of 2MASS1734-1151 and 2MASS1909-1937, that the criteria are not conclusive.\\\\ \\label{tab:resultfinal}} \\begin{flushleft} \\begin{center} \\scriptsize \\begin{tabular}{lcccccccccc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} Object & $EW$(Na~{\\sc i})$^{1}$ & $EW$(H$\\alpha$)$^{2}$ & age$^{2}$ & MG memb. & M. from & M. from & M. from & M. from & M. from & Final MG \\\\ & (\\AA) & (\\AA) & (Myr) & (Kinem/Astrom)$^{3}$ & $EW$(Na~{\\sc i})$^{1}$ & $EW$(H$\\alpha$)$^{2}$ & $log~g$$^{4}$ & Li~{\\sc i}$^{5}$ & $v\\sin{i}$$^{6}$ & \\\\ \\hline \\noalign{\\smallskip} SIPS0007-2458 & 8.2 & 11.0 & 1-300 & IC & N & Y & Y & N & Y & {\\it YD} \\\\ 2MASS0020-2346 & 8.1 & 6.1 & $\\sim$300 & YD & Y & Y & Y & ? & Y & {\\it YD} \\\\ DENIS0021-4244 & 8.4 & $^{a}$ & N/A & IC, CA & CA & N/A & Y & - & N & {\\it CA} \\\\ SIPS0027-5401 & 7.7 & 8.1 & $\\sim$300 & HY & N/A & Y & Y & Y & Y & {\\it HY} \\\\ SIPS0153-5122 & 8.0 & 7.5 & $\\sim$300 & IC, HY & HY & HY & HY & HY & Y & {\\it HY} \\\\ SIPS0214-3237 & 7.9 & 8.4 & $\\sim$300 & HY & N/A & Y & Y & Y & Y & {\\it HY}\\\\ SIPS0235-0711 & 7.9 & 7.7 & $\\geq$300 & HY & N/A & Y & Y & Y & Y & {\\it HY} \\\\ 2MASS0334-2130 & 7.3 & 11.5 & 1-300 & IC & N & Y & N & N & ? & {\\it YD }\\\\ SIPS2045-6332 & 6.3 & 2.1 & N/A & PL, CA & CA & N/A & Y & Y & ? & {\\it CA }\\\\ SIPS2039-1126 & 7.3 & - & - & PL & N & - & Y & Y & Y & {\\it PL}\\\\ LEHPM4908 & 7.3 & 6.5 & $\\sim$300 & HY & N/A & Y & Y & Y & - & {\\it HY }\\\\ 2MASS2254-3228 & 8.5 & 5.8 & $\\sim$300 & PL, CA & CA & CA & Y & - & Y & {\\it CA }\\\\ LEHPM6542 & 7.6 & 4.9 & $\\geq$300 & HY & N/A & Y & Y & Y & Y & {\\it HY }\\\\ & & & & & & & & & \\\\ 2MASS0814-4020 & 9.0 & 7.2 & $\\geq$300 & HY & N/A & Y & Y & N/A & N/A & {\\it HY}\\\\ 2MASS1146-4754 & 7.9 & 5.2 & $\\geq$300 & SI & N/A & Y & Y & N/A & N/A & {\\it SI}\\\\ 2MASS1236-6536 & 5.8 & 1.0 & 300-1200 & PL, IC, CA, SI & CA, SI & CA, SI & N & N/A & N/A & {\\it CA, SI} \\\\ 2MASS1326-5022 & 4.2 & 23.6 & 1-100 & IC, CA & IC & IC & Y & N/A & N/A & {\\it IC}\\\\ 2MASS1433-5148 & 7.9 & 19.8 & 1-100 & PL, IC, CA, SI & CA, SI & PL, IC & SI & N/A & N/A & {\\it CA?, SI} \\\\ 2MASS1557-4350 & 4.0 & 16.4 & 1-100 & PL, IC, CA & PL, IC & N/A & Y & N/A & N/A & {\\it PL, IC} \\\\ 2MASS1618-3214 & 8.0 & 7.8 & 1-300 & IC, CA, SI & CA, SI & Y & Y & N/A & N/A & {\\it IC?, CA, SI}\\\\ 2MASS1734-1151 & 7.7 & 5.5 & N/A & PL & N & N/A & Y & N/A & N/A & {\\it PL?}\\\\ 2MASS1736-0407 & 6.1$^{c}$ & 4.2 & 50-300 & CA, SI & N/A & Y & Y & N/A & N/A & {\\it CA, SI}\\\\ 2MASS1745-1640 & 5.6 & 1.4 & $\\geq$300 & CA, SI, HY & Y & N/A & HY & N/A & N/A & {\\it CA?, SI?, HY} \\\\ 2MASS1756-4518 & 7.5 & 3.2 & N/A & CA & N/A & N/A & Y & N/A & N/A & {\\it CA}\\\\ 2MASS1909-1937 & 6.3 & $^{a}$ & N/A & CA & N/A & N/A & - & N/A & N/A & {\\it CA?}\\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} \\end{tabular} \\end{center} $^{1}$ Equivalenth width of the Na~{\\sc i} doublet as expained in Section 5.1.\\\\ $^{2}$ See Section 5.2.\\\\ $^{3}$ MG membership from kinematic (sample A) or from photometric and astrometric criteria (sample B). See Section 2.\\\\ $^{4}$ From value obtained in the synthetic fits, given in columns 5 and 9 of Tables~\\ref{tab:resulta}, ~\\ref{tab:resultb} and ~\\ref{tab:resultab}.\\\\ $^{5}$ Only applicable for high resolution data. See Section 5.3.\\\\ $^{6}$ From Paper II.\\\\ $^{a}$ Absorption line filled in with emission.\\\\ $^{b}$ Absorption line.\\\\ $^{c}$ Measured with a cosmic ray in the line.\\\\ \\end{flushleft} \\end{table*}" }, "1402/1402.1967_arXiv.txt": { "abstract": "We use a dynamical systems approach based on the method of orthonormal frames to study the dynamics of a two-fluid, non-tilted Bianchi Type I cosmological model. Such a universe is anisotropic, spatially homogeneous, and spatially flat. In our model, one of the fluids is a fluid with bulk viscosity, while the other fluid assumes the role of a cosmological constant and represents nonnegative vacuum energy. We begin by completing a detailed fixed-point analysis of the system which gives information about the local sinks, sources and saddles. We then proceed to analyze the global features of the dynamical system by using topological methods such as finding Lyapunov and Chetaev functions, and finding the $\\alpha$- and $\\omega$-limit sets using the LaSalle invariance principle. The fixed points found were a flat Friedmann-LeMa\\^{\\i}tre-Robertson-Walker (FLRW) universe with no vacuum energy, a de Sitter universe, a flat FLRW universe with both vacuum and non-vacuum energy, and a Kasner quarter-circle universe. We also show in this paper that the vacuum energy we observe in our present-day universe could actually be a result of the bulk viscosity of the ordinary matter in the universe, and proceed to calculate feasible values of the bulk viscous coefficient based on observations reported in the Planck data. We conclude the paper with some numerical experiments that shed further light on the global dynamics of the system. ", "introduction": "In this paper, we use a dynamical systems approach to investigate in detail the dynamics of a Bianchi Type I universe with a bulk viscous fluid and cosmological constant. Such a universe is spatially flat, spatially homogeneous, and anisotropic. Such a model may have considerable importance in present studies of cosmology given the recent results of the Planck measurements \\cite{planckdata}, which suggest that the curvature of the spatial sections of the present-day universe is in agreement with spatial flatness. Moreover, Bianchi models are more general than the Friedmann-Lema\\^{\\i}tre-Robertson-Walker (FLRW) models and therefore can provide better descriptions of the early universe where viscous effects may have been dominant \\cite{hervik}. One can then study the effects of viscosity on the dynamical evolution of the universe which we observe to be of FLRW-type today. As discussed by Coley and Wainwright \\cite{coleywainwright}, cosmological models with single fluids are necessarily a simplification in the sense that they can only describe one epoch during the evolution of the universe. More general models can be constructed using two fluids with barotropic equations of state that are also comoving. One can then use these models to describe the transitions between different epochs in the universe's evolution, such as going from a radiation-dominated phase $(w = -1/3)$ to a matter-dominated phase $(w = 0)$. As discussed by Gr{\\o}n and Hervik (Chapter 13, \\cite{hervik}), viscous models have become of general interest in early-universe cosmologies largely in two contexts. Firstly, in models where bulk viscous terms dominate over shear terms, the universe expands to a de Sitter-like state, which is a spatially flat universe neglecting ordinary matter, and including only a cosmological constant. Such models isotropize indirectly through the massive expansion. Secondly, in the absence of any significant heat flux, shear viscosity is found to play an important role in models of the universe at its early stages. In particular, neutrino viscosity is considered to be one of the most important factors in the isotropization of our universe. By also including a nonnegative cosmological constant in our model, and interpreting it to represent vacuum energy, we are also able to give a detailed description of the roles played by both viscosity and vacuum energy in the isotropization of our universe. Bianchi cosmological models which contain a viscous fluid matter source in addition to a cosmological constant have been studied in detail several times. Lorenz-Petzold \\cite{lorenzp} examined Bianchi Type I and V models in the the presence of perfect fluid matter with bulk viscosity and a nonzero cosmological constant. Pradhan and Pandey \\cite{pradhanpandey} studied Bianchi Type I magnetized cosmological models in the presence of a bulk viscous fluid in addition to a monotonically decreasing cosmological constant. Saha \\cite{sahaI} studied the evolution of a Bianchi Type I universe with a viscous fluid and a cosmological constant. Pradhan, Srivastav, and Yadav \\cite{pradhan2} studied Bianchi Type IX viscous models with a time-dependent positive cosmological constant. Belinch\\'on \\cite{belinchon} investigated the dynamics of a locally rotationally symmetric (LRS) Bianchi Type I universe with a bulk viscous fluid and a time-dependent cosmological constant. Pradhan, Jotania, and Rai \\cite{pradhan3} studied Bianchi Type V cosmological models with bulk viscous fluid and a time-dependent cosmological constant. They also discussed some physical and geometrical aspects of such models. Pradhan and Pandey \\cite{pradhan4} studied Bianchi Type I cosmological models with both shear and bulk viscosity and a monotonically decreasing cosmological constant. The authors considered the special case in which the expansion tensor only had two components. Saha and Rikhvitsky \\cite{saha} analyzed a Bianchi Type I universe with a cosmological constant and dissipative processes due to viscosity. They showed that a positive cosmological constant leads to an ever-expanding universe. Singh and Kale \\cite{singhkale2} studied Bianchi Type I, Kantowski-Sachs, and Bianchi Type III cosmological models containing as matter sources a bulk viscous fluid and non-constant gravitational and cosmological constants. Pradhan and Kumhar \\cite{pradhankumhar} studied LRS Bianchi Type II models with bulk viscous fluid and a decaying cosmological constant. Mostafapoor and Gr\\o n \\cite{mostafgron} studied a Bianchi Type I universe with a cosmological constant and nonlinear viscous fluid. Sadeghi, Amani, and Tahmasbi \\cite{sadeghi} investigated a Bianchi Type-VI cosmological model with a cosmological constant and viscous fluid. Barrow \\cite{barrownuc2} showed that models of an inflationary universe driven by Witten strings in the very early universe are equivalent to the addition of bulk viscosity to perfect fluid cosmological models with zero curvature. In this work, Barrow considered the case where the bulk viscosity has a power-law dependence upon the matter density. It was shown that if the exponent is greater than $1/2$, there exist deflationary solutions which begin in a de Sitter state and evolve away from it asymptotically in the future. On the other hand, if this exponent is less than $1/2$, then solutions expand from an initial singularity towards a de Sitter state. Barrow \\cite{barrownuc3} also estimated the entropy production associated with anisotropy damping in the early universe by considering a Bianchi type I metric with an equilibrium radiation gas and anisotropic stresses produced by shear viscosity. It was shown that the shear viscosity based on kinetic theory has the general form of being proportional to the matter density and that the entropy production due to collisional transport is negligible in such a model. All of the aforementioned papers use the metric approach (Page 39, \\cite{ellis}) to obtain the dynamical evolution of the Bianchi model under consideration. The alternative approach which is based on the method of orthonormal frames pioneered by Ellis and MacCallum \\cite{ellismac} in conjunction with dynamical systems theory is the path we take in this paper. Belinkskii and Khalatnikov \\cite{belkhat} used phase-plane techniques to study a Bianchi Type I model under the influence of both shear and bulk viscosity. Goliath and Ellis \\cite{goliathellis} used dynamical systems methods to study FLRW, Bianchi Type I, Bianchi Type II, and Kantoswki-Sachs models with a positive cosmological constant. Coley and van den Hoogen \\cite{coleyvanV} analyzed in detail a Bianchi Type V model with viscosity, heat conduction, and a cosmological constant. They showed that all models that satisfy the weak energy condition isotropize. Coley, van den Hoogen, and Maartens \\cite{coley} examined the full Israel-Stewart theory of bulk viscosity applied to dissipative FLRW models. Coley and Dunn \\cite{coleydunn} used dynamical systems methods to study the evolution of a Bianchi Type V model with both shear and bulk viscosity. Burd and Coley \\cite{burdcoley} examined using dynamical systems methods the effects of both bulk and shear viscosities upon the FLRW, Bianchi Type I, Bianchi Type V, and Kantowski-Sachs models. They found that these models were structurally stable under the introduction of bulk viscosity. Kohli and Haslam \\cite{isk1} used dynamical systems methods to study the future asymptotic behavior of a Bianchi Type IV model containing both bulk and shear viscosity. Kohli and Haslam \\cite{isk2} used dynamical systems methods to study a Bianchi Type I model containing bulk and shear viscosity in addition to a homogeneous magnetic field. With respect to dynamical systems methods in multi-fluid models, Stabell and Refsdal \\cite{stabell} considered the dynamics of a two-fluid FLRW system consisting of dust and a cosmological constant. Phase plane methods were used by Madsen, Mimoso, Butcher, and Ellis \\cite{madsenmimoso} to study the evolution of FLRW models in the presence of an arbitrary mixture of perfect fluids. Coley and Wainwright \\cite{coleywainwright} examined orthogonal Bianchi and FLRW models in the presence of a two-fluid system. Ehlers and Rindler \\cite{ehlersrind} studied in great detail three-fluid models, containing radiation, dust, and a cosmological constant. Recently, Barrow and Yamamoto\\cite{barrowyama} considered a two-fluid system with one of the fluids representing a cosmological constant in their study of the instabilities of Bianchi Type IX Einstein static universes. For more details on the history of multi-fluid models, the interested reader should see Pages 53-55, 60-62, 171-172 and references therein of \\cite{ellis}. Despite all of the important aforementioned contributions, we feel it will be of considerable value to consider the dynamics of a Bianchi Type I universe with a viscous fluid and cosmological constant with respect to dynamical systems theory following the methods outlined in \\cite{ellis} and \\cite{barrowyama}. To the best of the authors' knowledge at the time of writing this paper, such an investigation has not been carried out in the literature. Throughout this paper, we assume a metric signature of $(-,+,+,+)$ and use geometrized units, where $8\\pi G = c = 1$. ", "conclusions": "We have presented in this paper a comprehensive analysis of the dynamical behavior of a Bianchi Type I two-fluid model with bulk viscosity and a cosmological constant. We began by completing a detailed fixed-point analysis of the system which gave information about the local sinks, sources and saddles. We then proceeded to analyze the global features of the dynamical system by using topological methods such as finding Lyapunov and Chetaev functions, and finding the $\\alpha$- and $\\omega$-limit sets using the LaSalle invariance principle. The fixed points found were a flat FLRW universe with no vacuum energy and only energy due to ordinary matter, a de Sitter universe, a mixed FLRW universe with both vacuum and non-vacuum energy, and a Kasner quarter-circle universe. We found conditions for which the former three were local sinks of the system, that is, future asymptotic states, and where the latter was a source of the system, that is, a past asymptotic state. The flat FLRW universe solution we found with both vacuum and non-vacuum energy is clearly of primary importance with respect to modelling the present-day universe, especially in light of the recently-released Planck data. In fact, using this Planck data we gave possible conditions for which a non-zero bulk viscosity in the early universe could have lead to some of the conditions described in the Planck data in the present epoch. In particular, since we found that this equilibrium point is a local sink of the dynamical system, all orbits approach this equilibrium point in the future. Therefore, there exists a time period for which our cosmological model will isotropize and be compatible with present-day observations of a high degree of isotropy of the cosmic microwave background in addition to the existence of both vacuum and non-vacuum energy." }, "1402/1402.4129_arXiv.txt": { "abstract": "We present the largest search to date for Y-band dropout galaxies ($z\\sim8$ Lyman break galaxies, LBGs) based on 350 arcmin$^2$ of HST observations in the V-, Y-, J- and H-bands from the Brightest of Reionizing Galaxies (BoRG) survey. In addition to previously published data, the BoRG13 dataset presented here includes approximately 50 arcmin$^2$ of new data and deeper observations of two previous BoRG pointings, from which we present 9 new $z\\sim8$ LBG candidates, bringing the total number of BoRG Y-band dropouts to 38 with $25.5\\leqslant m_{J} \\leqslant 27.6$ (AB system). We introduce a new Bayesian formalism for estimating the galaxy luminosity function, which does not require binning (and thus smearing) of the data and includes a likelihood based on the formally correct binomial distribution as opposed to the often used approximate Poisson distribution. We demonstrate the utility of the new method on a sample of $97$ Y-band dropouts that combines the bright BoRG galaxies with the fainter sources published in \\cite{Bouwens:2011p8082} from the Hubble Ultradeep Field (HUDF) and Early Release Science (ERS) programs. We show that the $z\\sim8$ luminosity function is well described by a Schechter function over its full dynamic range with a characteristic magnitude $M^\\star = -20.15^{+0.29}_{-0.38}$, a faint-end slope of $\\alpha = -1.87^{+0.26}_{-0.26}$, and a number density of $\\log_{10} \\phi^\\star [\\textrm{Mpc}^{-3}] = -3.24^{+0.25}_{-0.24}$. Integrated down to $M=-17.7$ this luminosity function yields a luminosity density, $\\log_{10} \\epsilon [\\textrm{erg}/\\textrm{s/Hz/Mpc}^{3}] = 25.52^{+0.05}_{-0.05}$. Our luminosity function analysis is consistent with previously published determinations within 1$\\sigma$. The error analysis suggests that uncertainties on the faint-end slope are still too large to draw firm conclusion about its evolution with redshift. We use our statistical framework to discuss the implication of our study for the physics of reionization. By assuming theoretically motivated priors on the clumping factor and the photon escape fraction we show that the UV luminosity density from galaxy samples down to $M=-17.7$ can ionize only 10-50\\% of the neutral hydrogen at $z\\sim8$. Full reionization would require extending the luminosity function down to $M=-15$. The data are consistent with a substantial fraction of neutral hydrogen at $z>7$, in agreement with recent suggestions based on deep spectroscopy of redshift 8 LBGs. ", "introduction": "\\label{sec:intro} Characterizing the epoch of reionization, i.e., the epoch at which the first stars and galaxies in the Universe reionized the vast majority of the neutral hydrogen, is a key outstanding issue in the continued effort to map the formation and early evolution of galaxies. For almost four years, since the installment of the Wide Field Camera 3 (WFC3) onboard the Hubble Space Telescope (HST), the frontier of this knowledge has been pushed further and further back towards the dawn of cosmic reionization. At present, several hundred high redshift galaxy candidates have been found at redshift $z\\sim6$ \\citep[e.g.,][]{Stark:2010p27668,Bouwens:2007p29848,Bouwens:2012p31891,Ouchi:2010p27936,Bradley:2013p32053}, and with the improved near-IR efficiency of WFC3 the search for candidates has been pushed to $z\\gtrsim 8$ using the Lyman Break technique. In particular the Hubble Ultra Deep field efforts in 2009 \\citep{Oesch:2010p30140,Oesch:2010p30134,Lorenzoni:2011p12992,Bouwens:2010p30142,Bouwens:2011p8082,McLure:2010p31124} and 2012 \\citep{Koekemoer:2012p26719,Ellis:2012p26700,Dunlop:2012p26717,Schenker:2013p26914,McLure:2013p27183,Ono:2012p26883,Oesch:2013p27877,Illingworth:2013p31165} have revealed the faintest samples of galaxy candidates at $z\\gtrsim8$. Simultaneously larger area observations are targeting brighter and rarer candidates, either in legacy fields, such as GOODS/CANDELS \\citep{Grogin:2011p9457,Koekemoer:2011p9456}, or in pure-parallel random pointings, like those of our \\emph{Brightest of Reionizing Galaxies Survey}\\footnote{\\url{https://wolf359.colorado.edu}} \\citep[hereafter BoRG,][]{Trenti:2011p12656,Trenti:2012p13020,Bradley:2012p23263}. Our ongoing BoRG survey has two key goals. The first goal is to provide bright targets that can potentially yield spectroscopic confirmation of $z\\sim8$ galaxies by follow-up observations \\citep{Treu:2012p12658,Treu:2013p32132}. In fact, while $z\\sim6$ dropout samples have extensive spectroscopic redshifts, \\citep[e.g.,][]{Vanzella:2009p29479,Stark:2010p27668,Stark:2013p28962}, only a handful of $z\\gtrsim 7$ galaxies \\citep[e.g.,][]{Ono:2012p27651} have currently confirmed redshifts with the highest being at $z=7.5$ \\citep{Finkelstein:2013p32467}. So far no Y-band dropout at $z\\sim8$ has been spectroscopically confirmed. Only upper limits on Ly$\\alpha$ flux have been provided to date \\citep{Caruana:2012p27502,Caruana:2013p32713,Capak:2013p31627,Treu:2013p32132,Faisst:2014p34184} and those leave open the interpretation of whether the photometric selection technique breaks down at $z\\gtrsim 7$ (which would be surprising given the small change in magnitudes and filters) versus the more interesting physical explanation of an increase in the intergalactic medium (IGM) optical depth to Ly$\\alpha$ arising from a higher neutral hydrogen fraction at $z\\sim8$ \\citep{Treu:2013p32132} with respect to redshift 7 \\citep{Fontana:2010p29506} and 6 \\citep{Stark:2010p27668,Stark:2013p28962}. The second goal of the BoRG survey is to improve the determination of the $z\\sim8$ luminosity function, by identifying rare and bright dropouts to extend the dynamic range of observations in smaller area deep fields, which are dominated by fainter sources. An accurate measure of the luminosity function is necessary not only to study how galaxies evolve across time, but also to quantify the photon budget available for hydrogen ionization \\citep{Trenti:2010p29335,Zaroubi:2013p24088,Dunlop:2013p23759}. At lower redshift, it is well established \\citep{Bouwens:2007p29848} that the luminosity function is accurately described by a Schechter function \\citep{Schechter:1976p29330} so it is natural to expect a similar form at higher redshift. However, data covering a wide dynamic range are needed to establish that this is indeed the case, and to resolve the degeneracy between the Schechter function parameters in the luminosity function fit \\citep{Bradley:2012p23263,Oesch:2012p30149}. In this paper we have two goals. The first is to present the complete sample of Y-band dropouts from the BoRG cycle-19 data, and to use these in combination with the literature to determine the galaxy luminosity function at $z\\sim8$. The second goal is to study the consequences for cosmic reionization the inferred luminosity function has. To determine the luminosity function we develop and implement a rigorous statistical Bayesian method to infer the posterior distribution function of the parameters of the luminosity function from the data. The method supersedes those commonly adopted in this field \\cite[e.g][]{Bradley:2012p23263,McLure:2013p27183,Schenker:2013p26914,Oesch:2012p30149} in several ways: the data are not binned, thus avoiding smearing the luminosity function \\citep{Trenti:2008p32309}; the flux uncertainties are correctly taken into account; the counts are modeled using the formally correct binomial distribution \\citep{Kelly:2008p29070}, instead of the Poisson approximation; the full posterior probability distribution function is computed using Markov Chain Monte Carlo methods instead of relying on maximum likelihood estimators based on the asymptotic covariance matrix from the observed Fisher information for uncertainties. By applying this framework to a large sample of $z\\sim 8$ galaxies, consisting of $N=97$ objects both bright (from BoRG) and faint (from the Hubble UDF/ERS fields), we show that the credible intervals include previous best-fit estimates By treating the problem in a self-consistent statistical manner we carry out an inference about reionization by combining the inferred observational uncertainties with various theoretical priors. The paper is organized as follows. We start by briefly describing the BoRG survey in Section~\\ref{sec:borg}. The current sample of Y-band dropouts containing 9 new and 2 improved $z\\sim8$ galaxy candidates (BoRG13) with respect to those previously published by our team \\citep[BoRG09; BoRG12][]{Trenti:2011p12656,Bradley:2012p23263} is described in Section~\\ref{sec:ydrop}. Two Appendixes (\\ref{app:borg58} and \\ref{app:noise}) take advantage of the follow-up observations of one field and of the large number of BoRG pointings to characterize and discuss the statistics of detections and contaminants in dropout searches. In Section~\\ref{sec:LF} we apply our inference of the Schechter luminosity function parameters using our rigorous Bayesian framework, discussed in detail in Appendix~\\ref{sec:BF}. The results are presented and discussed in the context of cosmic reionization in Section~\\ref{sec:results}. A brief summary is given in Section~\\ref{sec:conc}. All magnitudes are AB magnitudes and a standard concordance cosmology with $\\Omega_m=0.3$, $\\Omega_\\Lambda=0.7$, and $h=0.7$ is assumed. ", "conclusions": "\\label{sec:conc} The BoRG survey has carried out the largest-area search to date for Y-band dropouts (HST F098M-dropouts). We present new observations from 12 parallel fields not included in our previous studies and additional deeper datasets for two fields. We combine our BoRG sample with a sample of fainter dropouts taken from the literature and use the combined sample of $97$ dropouts to carry out a rigorous study of the luminosity function at $z\\sim8$ and its implications for reionization. Our main results can be summarized as follows: \\begin{enumerate} \\item We present 9 new Y-band dropouts from the $\\sim$50 arcmin$^2$ of new data. Furthermore, we re-confirm two dropouts previously presented by \\citet{Trenti:2011p12656} and \\citet{Bradley:2012p23263}. Combining these 11 dropouts with out previously published $z\\sim8$ LBGs from BoRG gives a sample of 38 bright ($25.5\\leqslant m_{J} \\leqslant 27.6$) redshift 8 LBGs from the BoRG survey. \\item We develop and implement an improved method for estimating the luminosity function parameters for a sample of $n$ galaxies. Using a Bayesian framework the posterior distribution of the population based on a binomial distribution is given. Combining the BoRG Y-band dropouts with the faint $z\\sim8$ LBGs from HUDF/ERS presented by \\cite{Bouwens:2011p8082} and sampling over this posterior distribution enables a robust recovery of the faint-end slope, $\\alpha$, the characteristic magnitude of the Schechter function, $M^\\star$, the normalizing number density of $z\\sim8$ LBGs, $\\phi^\\star$, and the luminosity density, $\\epsilon$, of the redshift 8 luminosity function. \\item The inferred luminosity function at $z\\sim8$ is described by the parameters: \\begin{itemize} \\item[] $M^\\star= -20.15^{+0.29}_{-0.38}$, \\item[] $\\alpha= -1.87^{+0.26}_{-0.26}$, \\item[] $\\log_{10} \\phi^\\star [\\textrm{Mpc}^{-3}]= -3.24^{+0.25}_{-0.34}$, \\item[] $\\log_{10} \\epsilon \\left[\\frac{\\textrm{erg}}{\\textrm{s Hz Mpc}^{-3}}\\right]= 25.52^{+0.05}_{-0.05}$ \\end{itemize} Here $\\epsilon$ is the inferred UV luminosity density integrated down to the HUDF limit M$_{\\rm lim}=-17.7$. Our inferred credible intervals include recent estimates of the same parameters. \\item We show that for the BoRG13 $z\\sim8$ luminosity function the average fraction of ionized hydrogen $Q$ is only of the order 10-50\\% for samples down to the HUDF limit of $M=-17.7$ assuming standard values of the clumping factor and the photon escape fraction. To sustain a fully ionized Universe at redshift 8 with the presented luminosity function it is necessary to account for the radiation of objects as faint as $M=-15$. \\item The inferred ionization fractions suggest a relatively late reionization scenario where a considerable fraction of neutral hydrogen is still present at $z\\sim8$ in good agreement with the results of our recent spectroscopic MOSFIRE campaign (and others from the literature) where we followed up a subsample of the BoRG redshift 8 LBGs presented here. \\end{enumerate} The inference on the implications of the BoRG13 luminosity function for reionization presented here are still limited by the sizable error bars at redshift 8. To reduce the uncertainties on the inferred quantities bright and faint galaxy samples from e.g., the Frontier Fields and future parallel HST campaigns are crucial." }, "1402/1402.4403_arXiv.txt": { "abstract": "While Population III stars are typically thought to be massive, pathways towards lower-mass Pop III stars may exist when the cooling of the gas is particularly enhanced. A possible route is enhanced HD cooling during the merging of dark-matter halos. The mergers can lead to a high ionization degree catalysing the formation of HD molecules and may cool the gas down to the cosmic microwave background (CMB) temperature. In this paper, we investigate the merging of mini-halos with masses of a few 10$^5$~M$_\\odot$ and explore the feasibility of this scenario. We have performed three-dimensional cosmological hydrodynamics calculations with the \\verb|ENZO| code, solving the thermal and chemical evolution of the gas by employing the astrochemistry package \\verb|KROME|. Our results show that the HD abundance is increased by two orders of magnitude compared to the no-merging case and the halo cools down to $\\sim$60 K triggering fragmentation. Based on Jeans estimates the expected stellar masses are about 10 M$_\\odot$. Our findings show that the merging scenario is a potential pathway for the formation of low-mass stars. ", "introduction": "The formation modes of the first generations of stars have been a challenging subject for many years and today still represent a central topic of research in modern cosmology. The current understanding is that these stars were formed in mini-halos of about 10$^6$ M$_\\odot$ around $z$~=~20-30, which is strongly supported by many numerical simulations performed over the years \\citep{Abel2002,Bromm2002,Ciardi05,Yoshida2006,Oshea2007,Greif2008,Greif2012}. Since \\citet{Saslaw1967}, it is widely accepted that the collapse process leading to the formation of these early objects is mainly driven by H$_2$ cooling \\citep{Peebles1968,Palla1983}. Due to the lack of a permanent dipole moment the hydrogen molecule becomes an ineffective coolant below $\\sim$200 K when it reaches the local thermodynamic equilibrium (LTE). Considering these thermal conditions, it has been postulated that the final mass of the first generation of stars should be of the order of $\\sim$100~M$_\\odot$ or more leading to a resultant short life \\citep{Abel2002,Bromm2002,BrommRev2013}. However, recent numerical simulations \\citep{ClarkGlover2011,Greif2012,Stacy2012} suggest that the characteristic mass scale of the first stars should be lowered to several ten solar masses, if fragmentation occurs. The large parameter study by \\citet{Hirano2013} indicates a significant spread around this mass scale, and extreme values up to 1000 M$_\\odot$. Comprehending the conditions under which the first stars were formed is very important to study the early history of the Universe as they initiated cosmic reionization and the enrichment of the intergalactic medium \\citep[e.g.][]{Barkana2001,Schneider06,Schleicher08}, and may also have seeded the formation of supermassive black holes \\citep{Li2007,Volontieri2012,LatifBH,Hosokawa2013,LatifBH2}. Many authors \\citep{Bromm2002,Johnson2006MNRAS.366..247J,Yoshida2006,Ripamonti2007,Greif2008,McGreer2008} also discussed the possible role of HD as an additional coolant which may have led to the formation of lower-mass stars. They found that in the presence of a higher electron abundance, HD cooling can become efficient allowing the gas to reach the temperature of the CMB. The need of a high ionization fraction suggested the introduction of a second mode to form primordial stars named Pop~III.2 to be distinguished from Pop~III.1 in which H$_2$ cooling dominates throughout \\citep[see][]{Tan2008,BrommRev2013}. The Pop III.2 stars are expected to be assembled from a gas which is still metal-free but embedded in ionized environments. The latter may be triggered by the explosions of the first supernovae or HII regions or strong shocks \\citep[e.g.][]{Mackey2003,Yoshida2007,Greif2008}. Lower-mass stars may also be formed within the Pop III.1 paradigm as shown by \\citet{Hirano2013}. They suggested that in a slowly collapsing cloud the compressional heating is reduced and may create the conditions for the HD cooling to become effective. In some of the above mentioned papers \\citep{Bromm2002,Ripamonti2007,McGreer2008} the authors claimed that in mini-halos of a few $\\sim$10$^5$~M$_\\odot$ HD can efficiently be formed and overcome the H$_2$ cooling rate if the temperature drops below $\\sim$200~K at densities of at least 10$^4$ cm$^{-3}$. \\citet{Stacy2013} performed cosmological simulations of the formation and growth of Pop III stellar systems and argued that a mini-halo formed at redshift 15 and supported by high baryonic angular momentum can lead to the formation of 1-5 M$_\\odot$ stars even in the absence of HD cooling. Lower-mass stars are thus also included in the expected scatter of the Pop III.1 distribution. \\citet{McGreer2008} concluded that in halos with masses $>$10$^6$ M$_\\odot$ the cooling is dominated by H$_2$ and the final stars could have a mass of $\\sim$100~M$_\\odot$, while for lower mass halos HD cooling becomes important and leads to stellar masses between 10-40~M$_\\odot$. However, the occurrence of such low-mass systems was found to be less frequent. These conclusions are in overall agreement with the 1D Lagrangian simulation reported by \\citet{Ripamonti2007} and with the robust statistical study presented by \\citet{Hirano2013}. The main idea behind the possible role of HD in primordial star formation comes from the series of chemical reactions which leads to the formation of HD. This is strongly related to the ionization fraction because the main formation path for HD is \\begin{equation} \\mathrm{H_2 + D^+ \\rightarrow HD + H^+}\\\\, \\end{equation} as largely discussed in earlier papers \\citep{Stancil1998,Galli2002,Stancil2011}. The only way to trigger the formation of HD is then to boost the formation of H$_2$ pumping the main reactive channel which is directly linked to H$^-$ via the following reactions: \\begin{eqnarray} &\\mathrm{H} + \\mathrm{e^-} \\rightarrow \\mathrm{H^-} + \\gamma\\\\ &\\mathrm{H^-} + \\mathrm{H} \\rightarrow \\mathrm{H_2} + \\mathrm{e^-}. \\end{eqnarray} In addition the formation and destruction of D$^+$ should be considered \\begin{equation} \\mathrm{D^+} + \\mathrm{H} \\rightleftharpoons \\mathrm{H^+} + \\mathrm{D}. \\end{equation} This leads us to the conclusion that a high ionization fraction boosts the formation of HD via a chain of chemical events that includes the key species in the following way $\\mathrm{e^-\\rightarrow H^-\\rightarrow H_2 \\rightarrow HD}$. In addition, we note that also an increase of the gas temperature will boost the formation of H$^{-}$, and thus H$_2$ and HD. Another possible route for primordial star formation (Pop III.2) which also involves the HD cooling has been proposed by \\citet{Merging2006}, and \\citet{Prieto2012,Prieto2013}. The idea suggested by \\citet{Merging2006} is based on the merging of dark-matter halos within the context of the hierarchical scenario of structure formation \\citep{Barkana2001,Ciardi05}. Merging induces compression and the consequent formation of shock-waves which may enrich the baryonic component of the gas by free electrons catalyzing the HD formation. The authors consider a simplified one-dimensional model where two identical halos collide and follow the non-equilibrium evolution of the HD behind the shock-waves which formed during the merging. In these halos, they further assume a uniform density equal to the virial density. They found that to have a higher ionization degree and to boost the formation of HD the merger masses should be $M>8\\times 10^6[(1+z)/20]^{-2}$~M$_\\odot$. In a real merger, we however expect a more complex density structure, and the strength of the shocks is likely enhanced in the central regions. Their mass scale can therefore just serve as a rough estimate. In a recent study, \\citet{Prieto2012,Prieto2013} performed cosmological N-body simulations and computed the statistic of the halos fulfilling the \\citet{Merging2006} criterion. Their findings suggest that the fraction of halos going through this phase is significant. In this paper we investigate the role of HD as a coolant in post-shocked environments generated by dark-matter halo merging. In particular we consider a merger of more than two halos with masses well below the threshold set by \\citet{Merging2006} and run self-consistent cosmological simulations, following both the merging and the collapsing process with enough accuracy to capture important chemo-dynamical features. The paper is organized as follow: we first introduce the methodology and the chemical model involved in the cosmological simulations, then discuss the main results, and finally give our conclusions. ", "conclusions": "\\label{sect:conclusions} In this paper we follow the evolution of a mini-halo of $\\sim$7$\\times$10$^5$ M$_\\odot$ formed from the merging of a number of halos with masses of a few 10$^5$ M$_\\odot$. It was in fact suggested \\citep{Merging2006} that if two massive enough halos collide and merge the gas can go through a post-shocked ionized stage which boosts the formation of the HD molecule then allowing the gas to cool below 200 K. As the final mass of the first luminous objects is given by the Jeans mass \\begin{equation} M_J = 500\\,\\mathrm{M_\\odot}\\left(\\frac{T}{200 \\mathrm{K}}\\right)^{3/2}\\left(\\frac{10^4 \\mathrm{cm^{-3}}}{n}\\right)^{1/2} \\end{equation} it is clear that in an environment dominated by H$_2$ cooling the expected mass is of about 500 M$_\\odot$ (assuming $T=200$ K and $n = 10^4$ cm$^{-3}$), while for the HD cooling case the mass is lowered to $\\sim$~10 M$_\\odot$ (e.g. assuming a temperature of 60~K at $n\\sim10^6$ cm$^{-3}$). A first attempt to investigate the merger-induced HD cooling was made by \\citet{Prieto2012} who performed hydro-cosmological simulations of 3$\\times$10$^7$ M$_\\odot$ halo resulting from the merging of two minihalos of a few 10$^6$ M$_\\odot$. A non-equilibrium treatment of the chemistry was included as well as cooling from H$_2$ and HD. They followed the evolution of the halo until densities of 10$^3$ cm$^{-3}$. The main results from their work was mainly focused on development of turbulence during the merging process. However, they also noticed that few cells of the entire simulations were able to reach temperatures of 100 K with HD fraction around 10$^{-6}$. The densities reached in their work are nevertheless too low to catch possible features due to HD overcooling. \\citet{Prieto2013} continued their investigation studying the statistics of the halo going through a merging process within a cosmological framework but performing DM-only simulations. They suggested that $\\sim$30\\% of the halos which generates from mergers are able to overcool the gas. Here we go beyond these studies by coupling the solution of dark matter and hydrodynamics (solved with \\verb|ENZO|) with an accurate treatment of the non-equilibrium chemistry (solved with \\verb|KROME|) following the collapse until a gas density of 10$^8$ cm$^{-3}$, thus extensively covering the density regime where HD cooling is known to usually dominate \\citep{Lipovka2005,GloverAbel08}. We performed high-resolution cosmological simulations for a box of 1 comoving Mpc and follow the evolution of the halo starting from $z=99$, going through the end of the merging process around $z=12$, and following the collapse of our final halo of $\\sim$7$\\times$10$^5$ M$_\\odot$ until $z=11.42$. Our results clearly show that the shock-waves generated by the merging process enhance H$^-$ and D$^+$ fractions which catalyze the formation of HD. We found a core HD abundance two orders of magnitudes larger than the one obtained with an isolated halo setup. As a consequence, the temperature in the core dropped down to 60-70 K, much lower compared to the isolated halo run where the gas is hotter (around $\\sim$500 K). Such over-cooling in the halo has induced fragmentation and multiple clumps are observed, out of which two are rotationally supported and one is gravitationally bound. \\begin{figure*} \\begin{center} \\includegraphics[width=.8\\textwidth]{fig4OK} \\caption{Density, temperature, and HD mass fraction projections, for the run where merging occurs at a scale of 10 pc for two different redshifts, $z=11.44$ (upper panel), and $z=11.43$ (lower panel).}\\label{fig:figure4} \\end{center} \\end{figure*} It is quite interesting to note that a merging process involving several halos with masses of few 10$^5$~M$_\\odot$ can similarly catalyze the required HD fraction via the efficient production of H$^-$. The differences in the masses of the mergers rise from the fact that we performed realistic three-dimensional calculations compared to the simplified one-dimensional estimates by \\citet{Merging2006} that provide only a qualitative picture of the process. In particular simplified one-dimensional models underestimate the strength of the merger shocks and do not take into account the collapse dynamics and the hydrodynamical effects (e.g. compressional heating). Another strong simplification is the assumption of a virial density while realistic halo are partially collapsed, enhancing the formation of molecules. Based on the halo finder results from the simulations we found a relative velocity of $\\sim$7.0 km s$^{-1}$ between the dark matter halos which is about a factor of 1.5 larger than the estimate by \\citet{Merging2006} from their Eq. 2. In a previous study, \\cite{McGreer2008} found that the fraction of HD-cooled halos is negligible in primordial \\emph{unperturbed} environments for halos of masses $\\geq$10$^6$~M$_\\odot$, but can become relevant for masses $<$10$^6$~M$_\\odot$ or in ionized media. The authors explored a series of physical conditions that should favour the formation of HD, i.e. gas efficiently cooled by H$_2$ to a critical temperature which should hold for time long enough to allow the build up of HD molecules. These conditions are more suitable in lower mass halos which collapse more uniformly. On the other hand, \\citet{Ripamonti2007} and \\citet{Bromm2002} have shown that the critical masses should be of a few 10$^5$~M$_\\odot$ and attribute the increase in HD fraction to a longer collapsing time. This was recently confirmed by the comprehensive study by \\citet{Hirano2013}. However, in the study by \\cite{McGreer2008}, it is not clear if mergers have occurred during the halo formation then a direct comparison with our results is difficult. Here, we simulate a situation where the halo formed via the occurance of major mergers and we find that HD abundance is enhanced up to 10$^{-5}$ in the aftermath of the merging process. Our findings show that the merger-driven shocks enhance the temperature on large scales, thus favouring the formation of H$^-$ and increasing the H$_2$ fraction. Also the HD abundance is slightly enhanced at that stage, even though it is unable to produce a runaway effect at those densities. When the collapse occurs, the enhanced H$_2$ and HD abundance however allows to trigger a runaway effect as also described by \\citet{McGreer2008}. Our findings suggest that the fraction of these halos might be larger than previously estimated as merger events are much more common in the early Universe ($\\sim$30\\% as reported in \\citet{Prieto2012}). The present work indicates a potential formation route for low-mass primordial stars due to merger-enhanced HD cooling. This pathway should be explored further in future studies, including a larger range of initial conditions. Finally, the presence of high turbulent energy produced in the aftermath of merging event may also contribute to the amplification of magnetic fields through the so-called small scale dynamo process \\citep{Schober2012ApJ,Bovino2013NJP,Schleicher2013AN,LatifBH3,Schleicher2013NJPh,Latif2013MNRAS}. The presence of such fields may further influence fragmentation and disk formation at higher densities \\citep[e.g.][]{Machida2013,Latif2013}." }, "1402/1402.3824_arXiv.txt": { "abstract": "{This article is the second in a two part series introducing r-Java 2.0, a nucleosynthesis code for open use that performs r-process calculations and provides a suite of other analysis tools.} {The first paper discussed the nuclear physics inherent to r-Java 2.0 and in this article the astrophysics incorporated into the software will be detailed. } {R-Java 2.0 allows the user to specify the density and temperature evolution for an r-process simulation. Defining how the physical parameters (temperature and density) evolve can effectively simulate the astrophysical conditions for the r-process. Within r-Java 2.0 the user has the option to select astrophysical environments which have unique sets of input parameters available for the user to adjust. In this work we study three proposed r-process sites; neutrino-driven winds around a proto-neutron star, ejecta from a neutron star merger and ejecta from a quark nova. The underlying physics that define the temperature and density evolution for each site is described in this work.} {In this paper a survey of the available parameters for each astrophysical site is undertaken and the effect on final r-process abundance is compared. The resulting abundances for each site are also compared to solar observations both independently and in concert. R-Java 2.0 is available for download from the website of the Quark-Nova Project: \\url{quarknova.ucalgary.ca} } {} ", "introduction": "\\label{intro} The rapid neutron capture process (r-process) is believed to be the mechanism for the nucleosynthesis of about half of the stable nuclei heavier than iron \\citep{Burbidge,Cameron}. Explosive and neutron-rich astrophysical environments present ideal conditions for the r-process to take place. Possible candidate sites discussed in the literature include the neutrino-driven neutron-rich wind from proto-neutron stars \\citep{qian96,Qian03b}, prompt explosions of collapsed stellar cores, \\citep{Sumiyoshi,Wanajo,sar12}, neutron star decompression \\citep{meyer97,Goriely}, tidal disruption in binary merger events \\citep{FRT}, outflows in gamma-ray bursts \\citep{Surman}, the LEPP process in low metalicity stars \\citep{trav04}, supernova fallback \\citep{fryer06}, etc. Most importantly, abundance data on r-process elements in metal-poor stars \\citep{Sneden03,Truran} and certain radio-nuclides in meteorites \\citep{QW08} point toward the distinct possibility of multiple r-process sites. An important paper by \\cite{beth85} on supernova neutrinos brought the high-entropy environment in neutrino-driven winds from Type II Supernovae (SNe) to the forefront of the discussion about astrophysical sites for the r-process. Since then, much progress has been made in the modelling of type II SNe and neutrino winds of nascent neutron stars \\citep{Woosley,Takahashi,qian96,Cardall,Otsuki,Wanajo01,TBM}. The most natural explanation in the neutrino-driven wind scenario is that the observed r-process pattern follows from a superposition of neutron capture events with differing neutron-to-seed ratios and exposure time-scales. A particular challenge for high-entropy winds as an r-process site is that producing the third peak requires extreme values of entropy and dynamic time-scale that are not supported by current hydrodynamic models of Type II SNe explosions \\citep[e.g.][]{arc07,QW08,fis10, arc11}. Neutron star mergers can provide a much larger neutron-to-seed ratio than type II SNe, which makes for an appealing r-process site. Recent relativistic hydrodynamic simulations of neutron star mergers have shown that a significant amount of r-process enriched matter can be ejected \\citep{janka99,ross04,oec07,gori11}, implying that the solar abundance of r-process nuclei may be influenced by neutron star mergers. However the neutron star (NS) coalescences time scale which likely ranges from 1 - 1000 Myr$^{-1}$ per Milky Way Equivalent Galaxy \\citep{kalo,kaloErr} puts the neutron star merger explanation at odds with enrichment of r-process elements relative to iron observed in metal-poor stars \\citep{qian00, arg04}. A possible new site for the r-process that is conjectured to be effective in producing the heavy elements beyond the second peak is the quark nova \\citep{ODD}. If this scenario occurs in nature, it presents a new possibility for explaining the origin of the heavy elements. \\cite{meyer89} and \\cite{Goriely} have studied the r-process in decompressing {\\it cold} neutron star matter \\citep{latt77}, although no specific mechanism for decompression was proposed. In previous work \\citep{jaik07,nieb10}, it was suggested that the dynamics of a quark-hadron phase transition inside a neutron star could be sufficiently strong to power the decompression and subsequent ejection of the neutron-rich crust, this is essentially the quark nova scenario. Clearly there is further study needed to better understand r-process in the Universe and to this end we present r-Java 2.0 an r-process code that is transparent and freely available\\footnote{Software and user manual can be downloaded at quarknova.ucalgary.ca} to the nuclear physics and astrophysics community. This article is meant to display the astrophysics incorporated in r-Java 2.0 and should be considered in conjunction with \\cite{kostka} which covered the nuclear physics inherent to the software. This paper is organized as follows; an overview of r-Java 2.0 is presented in section \\ref{rJavaNuc}, the high-entropy wind (HEW) module is described in section \\ref{hewSect}, the neutron star merger (NSM) module is laid out in section \\ref{nsmSect} and the quark nova (QN) module is discussed in section \\ref{qnSect}. Within each astrophysical module section simulation results are presented. In section \\ref{discuss} the results from each astrophysical scenario is compared to the solar observations. Finally a summary is presented in section \\ref{conc}. ", "conclusions": "\\label{conc} For the HEW simulations done in this work when the initial nuclei abundances were calculated under the assumption of NSE the r-process was not capable of producing nuclei significantly heavier that A = 80. When an $\\alpha$-rich freeze-out was assumed to compute the initial abundances a strong A = 130 peak was synthesized for all parameters surveyed. Starting from an $\\alpha$-rich freeze-out only the largest entropy (S = 270 and thus neutron-to-seed ratio = 100) studied was capable of producing transuranic elements. The HEW scenario reproduced the europium rare earth peak observed in solar r-process abundances. Fission recycling had the most impact on final r-process abundances in the NSM of all the scenarios studied. For all parameters surveyed (with the exception of the initial abundance of only neutrons, protons and alpha particles) r-process in the context NSM was capable of synthesizing the A = 195 peak and in most cases transuranic elements as well. When compared to the solar abundances NSM robustly reproduced the A = 130 and A = 195 peaks, however the europium rare earth peak was not formed. Recent work by \\cite{gori13} studying fission fragmentation of neutron-rich nuclei has addressed the europium rare earth peak in the context of NSM. Simulations done in this work showed that an ejecta mass of $10^{-4} M_{\\odot}$ produced the strongest A =195 of all QN scenarios studied. The lighter ejecta mass ($10^{-5} M_{\\odot}$) studied cooled too rapidly for the r-process to be able to produce transuranic elements. The final r-process abundances in heavier ($10^{-3} M_{\\odot}$) ejecta case produced a relatively flat distribution of nuclei. When compared to solar abundances, the QN scenario under-produced the breadth of the A = 130 peak but was capable of synthesizing the rare earth peak and also the A = 195 peak. For the simulations performed in this work no single parameter set for a chosen astrophysical site could reproduce the solar r-process abundance distribution. Either a combination of different astrophysical sites or different parameter sets within an astrophysical site was needed in order to re-produce solar abundances. Within the parameters surveyed in this work each scenario is capable of producing both a A = 130 and A = 195 peak comparable to those observed in the solar abundances. For the parameters studied in this work, the resultant r-process abundances for the NSM and QN were quite similar. This similarity is due to the high initial neutron-to-seed ratio for both scenarios which causes fission recycling to become a driving factor that shapes the final nuclei distribution. The similarity in r-process yields from these two scenarios could be an indication as to a cause for the apparent universality of the r-process nuclei distribution \\citep{qianUni}. The study of r-process nucleosynthesis remains a challenging topic which requires a deeper understanding of both the underlying nuclear physics as well as the astrophysics that shapes the r-process environment. With r-Java 2.0 we have provided a robust and easy-to-use platform that is capable of tackling both sides of the r-process problem. In \\cite{kostka} we discussed the cutting-edge nuclear physics incorporated in r-Java 2.0 which includes; temperature-dependant neutron capture cross-sections and photo-dissociation rates, realistic fission recycling and beta-delayed neutron emission. In this work we have shown the capabilities of r-Java 2.0 to model the evolution of density and temperature of three different proposed astrophysical r-process site; HEWs around a proto-neturon star, ejecta from NSMs, and ejecta from QNe. A fundamental tenet followed during the development of r-Java 2.0 was maximizing flexibility. For this reason we ensured that the software is cross-platform compatible and gave the user the ability to change any nuclear property both quickly and easily. In this work we have used the three built-in astrophysical modules in order to highlight the fact that with r-Java 2.0 for the first time anyone can run a comparison of proposed astrophysical r-process sites with their dynamical evolution and ejecta conditions. R-Java 2.0 goes beyond existing codes by including the option to define a custom density evolution profile in order to study other possible r-process scenarios. As many of the nuclear inputs used in r-Java 2.0 are only theoretically known, the results from rare-isotope beam facilities and upcoming sensitivity studies will provide much needed experimental insight into r-process nuclei \\citep{dillman,hosmer,vanscelt}. The flexibility of r-Java 2.0 allows for researchers to easily include experimental results in order to test their impact on the r-process. The areas of development that we are undertaking for the next release of r-Java are; the inclusion of neutrino-induced and spallation reactions, the development of a charged-particle reaction network module and adding the ability to study nuclear isomers. \\onecolumn \\clearpage \\begin{figure} \\includegraphics[scale = 0.75]{HEW.pdf} \\caption{\\textbf{High-entropy wind.} See section \\ref{hewRes} for details. \\textit{Top:} The black solid line denotes an entropy (S) of 140 and expansion velocity ($v_{\\rm exp}$) of 3$\\times10^4$ km s$^{-1}$. The red dashed line denotes S = 175 and $v_{\\rm exp}$ = 1.5$\\times10^4$ km s$^{-1}$. The green dotted line denotes S = 270 and $v_{\\rm exp}$ = 3.8$\\times10^3$ km s$^{-1}$. NSE determined initial abundances. \\textit{Middle:} Same as top panel with $\\alpha$-rich freeze-out initial abundances. \\textit{Bottom:} The black solid line denotes a neutron-to-seed ratio of 25, the red dashed line a neutron-to-seed ratio of 75 and the green dotted line a neutron-to-seed ratio of 100. Each simulation started from $\\alpha$-rich freeze-out abundances.} \\label{HEWcomp} \\end{figure} \\clearpage \\begin{figure} \\includegraphics{NSM_rhos.pdf} \\caption{Using the neutron star merger module the density evolution of ejecta with two different polytropic indices (n) are compared. The black solid line denotes relativistically degenerate matter (n = 1) and the red dashed line represents non-relativistically degenerate matter (n = 3). } \\label{NSMrho} \\end{figure} \\clearpage \\begin{figure} \\includegraphics[scale = 0.75]{NSM_compPhys.pdf} \\caption{\\textbf{Neutron star merger.} Final r-process abundances are compared for varying physical parameters. See section \\ref{NSMphysRes} for details of simulations. \\textit{Top:} Varying polytropic indexes (n); n = 1 (black solid line), n = 2 (red dashed line) and n = 3 (green dotted line). \\textit{Middle:} Varying initial densities ($\\rho_0$); $\\rho_0 = 2\\times10^{11}$g cm$^{-3}$ denoted by the black solid line, $\\rho_0 = 3\\times10^{11}$g cm$^{-3}$ by the red dashed line and $\\rho_0 = 4\\times10^{11}$g cm$^{-3}$ by the green dotted line. \\textit{Bottom:} Varying expansion velocity ($v_{\\rm exp}$); $v_{\\rm exp} = 10^2$km s$^{-1}$ denoted by the black solid line, $v_{\\rm exp} = 10^3$km s$^{-1}$ by the red dashed line and $v_{\\rm exp} = 10^4$km s$^{-1}$ by the green dotted line } \\label{NSMcompPhys} \\end{figure} \\clearpage \\begin{figure} \\includegraphics[scale = 0.75]{NSM_CompInitial.pdf} \\caption{\\textbf{Neutron star merger.} The final nuclei abundances are plotted for different initial electron fractions ($Y_{\\rm e,0} = 0.1$ denoted by the black solid line, $Y_{\\rm e,0} = 0.15$ by the red dashed line and $Y_{\\rm e,0} = 0.2$ by the green dotted line. See Sect. \\ref{NSMiniRes} for details on initial physical parameters. \\textit{Top:} Initially only neutron, protons and $\\alpha$-particles present. \\textit{Middle:} An initial seed of iron group isotopes. \\textit{Bottom:} The initial abundances determined using NSE. } \\label{NSMcompInit} \\end{figure} \\clearpage \\begin{figure} \\includegraphics[scale = 0.75]{initialNSE_compMejZeta.pdf} \\caption{\\textbf{Quark nova.} Comparison of the effect on r-process yield of varying the percentage of quark nova energy transformed into kinetic energy of the ejecta. In each panel the black solid line denotes 1$\\%$, the red dashed line denotes 2$\\%$ and the green dotted line 5$\\%$. For each panel the initial nuclei abundance distribution is determined by NSE. \\textit{Top:} Quark nova ejecta of $10^{-3}$M$_{\\odot}$. \\textit{Middle:} Quark nova ejecta of $10^{-4}$M$_{\\odot}$. \\textit{Bottom:} Quark nova ejecta of $10^{-5}$M$_{\\odot}$. } \\label{QNcompMejZeta} \\end{figure} \\clearpage \\begin{figure} \\includegraphics[scale = 0.75]{QN_compInitial.pdf} \\caption{\\textbf{Quark nova.} Comparison of the final r-process abundance yield of different masses of quark nova ejecta ($10^{-3}$M$_{\\odot}$ denoted by the black solid line, $10^{-4}$M$_{\\odot}$ by the red dashed line and $10^{-5}$M$_{\\odot}$ by the green dotted line). See section \\ref{QNiniRes} for details of simulations. \\textit{Top:} Initially only neutrons, protons and $\\alpha$-particles. \\textit{Middle:} Initially a seed of iron group isotopes. \\textit{Bottom:} Initial abundance determined by NSE.} \\label{QNcompInitial} \\end{figure} \\clearpage \\begin{figure} \\includegraphics[scale = 0.85]{CompSolar.pdf} \\caption{In each panel the superposition of the r-process yields (red solid line) from each astrophysical site is compared to the observed solar r-process abundances (black crosses). The constituent components are as well plotted for each site. \\textit{Top:} The r-process abundances from high-entropy wind events; $Y_{\\rm n} / Y_{\\rm seed} = 25$ (green dash-dot line), $Y_{\\rm n} / Y_{\\rm seed} = 50$ (blue dashed line) and $Y_{\\rm n} / Y_{\\rm seed} = 100$ (black dotted line). \\textit{Middle:} The r-process abundances from neutron star merger events; $Y_{\\rm e} = 0.3$ and $\\rho = = 3 \\times 10 ^{11}$g cm $^{-3}$ (green dash-dot line), $Y_{\\rm e} = 0.1$ and $\\rho = 8 \\times 10 ^{11}$g cm $^{-3}$ (blue dashed line). \\textit{Bottom:} The r-process abundances from quark novae; $m_{\\rm ej} = 10^{-4} M_{\\odot}$ and $\\zeta = 1\\%$ (green dash-dot line), $m_{\\rm ej} = 10^{-4} M_{\\odot}$ and $\\zeta = 5\\%$ (black dotted line) and $m_{\\rm ej} = 10^{-5} M_{\\odot}$ and $\\zeta = 1\\%$ (blue dashed line). } \\label{sol} \\end{figure} \\clearpage \\begin{figure} \\includegraphics{allSites.pdf} \\caption{All sites scaled by predicted rates compared to solar abundances. } \\label{allSites} \\end{figure} \\clearpage \\begin{table} \\caption{Quark Nova Ejecta Parameters} % \\label{qnTable} % \\centering % \\begin{tabular}{c | c c c c} % \\hline\\hline % Mass & T$_{0}$ & $\\rho_0$ & Y$_{\\rm e,0}$ & $\\tau$ \\\\ % [M$_{\\odot}$]& [10$^9$K] & [10$^{11}$g cm$^{-3}$] & &[$\\mu$s] \\\\ \\hline % 10$^{-5}$ & 2.4 & 1.0 & 0.18 & 0.1 \\\\ % 10$^{-4}$ & 3.8 & 4.0 & 0.13 & 0.5 \\\\ 10$^{-3}$ & 7.5 & 30 & 0.07 & 6.0 \\\\ \\hline % \\end{tabular} \\end{table} \\clearpage \\appendix" }, "1402/1402.7016_arXiv.txt": { "abstract": "White-light observations by the Solar Dynamics Observatory's Helioseismic and Magnetic Imager of a loop-prominence system occurring in the aftermath of an X-class flare on 2013 May 13 near the eastern solar limb show a linearly polarized component, reaching up to $\\sim$20\\% at an altitude of $\\sim$33 Mm, about the maximal amount expected if the emission were due solely to Thomson scattering of photospheric light by the coronal material. The mass associated with the polarized component was 8.2$\\times$10$^{14}$ g. At 15 Mm altitude, the brightest part of the loop was 3($\\pm$0.5)\\% linearly polarized, only about 20\\% of that expected from pure Thomson scattering, indicating the presence of an additional unpolarized component at wavelengths near Fe I (617.33 nm), probably thermal emission. We estimated the free electron density of the white-light loop system to possibly be as high as 1.8$\\times$10$^{12}$ cm$^{-3}$. ", "introduction": "\\citet{Martinez2014} (thereafter Paper I) have already reported on the observations by the Helioseismic and Magnetic Imager \\citep[HMI,][]{Schou2012b,Scherrer2012} of the Solar Dynamics Observatory \\citep[SDO,][]{Pesnell2012} of coronal emission from two flares occurring on 2013 May 13. Both of these also showed white-light (WL) footpoint sources at the level of the photosphere. The gradual coronal emissions can be identified as visual counterparts of the classical loop-prominence system, but were brighter than expected and possibly seen in the continuum rather than line emission, as inferred from high-resolution HMI spectra. In this interpretation, the coronal sources detected by HMI in these flares represent flare loops, initially heated to X-ray temperatures, and detected in the process of cooling. The authors found the HMI flux to exceed the radio/X-ray interpolation of the bremsstrahlung produced in the flare soft X-ray sources by at least one order of magnitude, implying the participation of cooler sources that could produce free-bound continua and possibly line emission detectable by HMI. \\begin{figure*}[ht!] \\centering \\includegraphics[width=16.6cm]{Fig1.eps} \\caption{The gradual-phase sources, with image times: 16:25:22.7 (WL), 16:25:28.1 (160 nm), and 16:25:21.5 (19.3 nm). The AIA 19.3 nm image shows loop-shaped absorption features, corresponding loosely to the WL and UV loops. The purple contours are from RHESSI observations of thermal X-rays, and the red line is the 20\\% contour level of the HMI image. Reproduction of Fig. 5 from Paper I, with the addition of a green box used to accumulate the time profiles displayed in Figure~\\ref{fig:HMIlc}.} \\label{fig:FigX} \\end{figure*} Historically, the loop-prominence phenomenology played a major role in establishing the standard model of large-scale magnetic reconnection as a mechanism for the formation of flare loops projected against the corona, initially through observations in chromospheric lines \\citep[e.g.,][]{Svestka1972}. Observations of polarization in broad-band continuum by coronagraphs typically do not extend low enough to study the arcade development; for example, the Mk4 K-coronameter at Mauna Loa only observes above about 1.12 R$_\\odot$, some 80 Mm above the photosphere. Nevertheless some direct broad-band intensity observations in the lower corona have been reported (Hiei et al., 1992; Leibacher et al., 2004). These did not include the polarization signatures that HMI provides. In this paper, we will concentrate on one of the two events discussed in Paper I, SOL-2013-05-13T16:01 (X2.8), and we present evidence of a Thomson-scattered component in the HMI emission, which allows us to discriminate between emission mechanisms contributing to the observed HMI emission. We describe the interpretation of the polarization signatures in the Appendix. \\begin{figure}[ht!] \\centering \\includegraphics[width=7.5cm]{Fig2.eps} \\caption{ 2013 May 13 GOES X-ray (black: 1--8 \\AA\\ ; orange: 0.5--4 \\AA\\ ) and HMI Stokes I, Q, U, V time profiles. The HMI lightcurves are generated by averaging over all six filters the pixels in the green box of Figure~\\ref{fig:FigX}. Each pixel within the box was temporally median-filtered (with window size 3). No other data alterations were performed. } \\label{fig:HMIlc} \\end{figure} ", "conclusions": "\\label{sect:ccl} We have reported what we believe to be the first detection of linearly polarized scattered white-light of an evolving flare loop system in the vicinity of the Fe I line, and probably the first unambiguous mass estimate of such a system, owing to the linear dependence on density and temperature-independence of Thomson scattering. It is only due to SDO/HMI's remarkable dynamic range and polarimetric capabilities that such a faint signature of Thomson scattering could be identified. This has enabled us to dynamically 1) identify the fraction of WL emission in a flare loop at the limb that is due to Thomson scattering, and 2) estimate the free electron content and the mass of the scattering sources, as well as likely limits to their free electron densities. Such an approach brings a powerful new diagnostic tool to the study of limb flares, and possibly the flare/CME connection as well." }, "1402/1402.2858_arXiv.txt": { "abstract": "{We present deep \\ion{H}{i} observations of three compact high-velocity clouds (CHVCs).} {The main goal is to study their diffuse warm gas and compact cold cores. We use both low- and high-resolution data obtained with the 100 m Effelsberg telescope and the Westerbork Synthesis Radio Telescope (WSRT). The combination is essential in order to study the morphological properties of the clouds since the single-dish telescope lacks a sufficient angular resolution while the interferometer misses a large portion of the diffuse gas.} {Here single-dish and interferometer data are combined in the image domain with a new combination pipeline. The combination makes it possible to examine interactions between the clouds and their surrounding environment in great detail.} {The apparent difference between single-dish and radio interferometer total flux densities shows that the CHVCs contain a considerable amount of diffuse gas with low brightness temperatures. A Gaussian decomposition indicates that the clouds consist predominantly of warm gas.} {} ", "introduction": "\\label{sec:introduction} High-velocity clouds (HVCs) are neutral atomic hydrogen (\\ion{H}{i}) clouds with radial velocities incompatible with simple Galactic rotation models \\citep{1997ARA&A..35..217W}. After their initial discovery by \\citet{1963CRAS..257.1661M}, they have been observed all around the Milky Way. \\citet{1999ApJ...514..818B} suggested that HVCs are distributed throughout the Local Group, but this theory has been disproven several times by searching for HVCs around other galaxies: all these observations have either failed to find any HVCs (\\citet{2004ApJ...610L..17P, 2007ApJ...662..959P}) or located them very close to their host galaxies \\citep{2004ApJ...601L..39T}. High-velocity clouds are found either as compact and isolated objects or as part of large complexes \\citep{1997ARA&A..35..217W}. Recent studies propose three main hypotheses for the origin of HVCs \\citep{2004ASSL..312..341B}. First, HVCs consist of primordial gas that is accreted onto galaxies. It can be primordial gas flows from the filaments or have its origin in gaseous dark matter haloes. Second, HVCs originate from the tidal and ram-pressure interaction of the Milky Way Galaxy with dwarf galaxies. Third, HVCs have been formed as the result of a galactic fountain, i.e., by gas flows driven by supernovae. The most serious problem is the limited information on the HVC distances \\citep{2000ASPC..218..407V, 2004ASSL..312..195V, 2005A&A...440..775K, 2008ASPC..393..179B}. Investigations of the physical conditions of HVCs indicate that some show signs of interaction with the Galactic halo gas \\citep{2000A&A...357..120B, 2001A&A...370L..26B, 2005A&A...432..937W, 2011MNRAS.418.1575P, 2012A&A...547A..12V}. The aim of our study is to investigate the morphological properties of compact high-velocity clouds (CHVCs) as well as their radial velocity and line width. In the past, the existence of two different gradients either in density \\citep{2011MNRAS.418.1575P} or velocity \\citep{2000A&A...357..120B} were required within the cloud for the designation head-tail (HT). The objects studied in the present paper appear to conform to both these criteria. The appearance of the HT structure is interpreted as follows. As the cloud moves through the hot and ionized halo, part of the cloud is compressed. This higher density area constitutes the head of the cloud. Part of the gas is stripped off the cloud to form a less dense and thin tail structure that follows the head with velocities lower than those of the cloud bulk motion \\citep{2002A&A...391..713K, 2009ApJ...698.1485H}. The cloud HVC\\,125+41--207 is prototypical for interacting HVCs denoted accordingly as head-tail (HT) HVCs. \\citet{2000A&A...354..853B} used the Westerbork Synthesis Radio Telescope (WSRT) to study the small--scale structure of compact HVCs (CHVCs). Towards HVC\\,125+41--207, they discovered an extremely narrow \\hi line unresolved by their spectral resolution of $2.47\\,{\\rm km\\,s^{-1}}$. The peak brightness temperature of this line is exceptionally high with $T_{\\rm B} = 75$\\,K indicating the existence of a cold and dense cold neutral medium (CNM) core. Their maps reveal a highly structured \\hi CNM distribution. \\citet{2000A&A...357..120B} used Effelsberg observations of HVC\\,125+41--207 to deduce that this dense core is embedded within a warm neutral medium (WNM) envelope. Moreover, they found evidence that the WNM shows signs of deceleration not only towards the cloud's tail but also along the rims. Proportional to this deceleration of the radial velocity component, the WNM gets warmer (see their Fig.\\,2). Only the combined analysis of the WSRT and Effelsberg data allowed the deduction of such a homogenous view of HVC\\,125+41--207 as an interacting CHVC. To investigate the gaseous structures of HVCs and their dynamics accurately, the combination of radio interferometric and single-dish data is essential. Radio interferometers are insensitive to structures on the scale of tens of arc minutes and beyond, the WNM. Single-dish observations are unable to resolve the small-scale structure of the CNM. The combination of \\ion{H}{i} single-dish and interferometric observations provides the possibility of studying HVCs in the necessary detail. In this paper, we present high-resolution \\ion{H}{i} observations of three CHVCs, using the 100 m Effelsberg telescope and the WSRT. Section\\,\\ref{sec:observations} presents the details of the data, and the \\ion{H}{i} observations are introduced. Section\\,\\ref{subsec:method} deals with the method that has been used to combine these two data sets. Section\\,\\ref{sec:CHVCs} discusses the physical and morphological properties of the clouds and the results of the Gaussian decomposition of the integrated spectral lines. Section\\,\\ref{sec:conclusions} summarizes our findings and gives an outlook on future work. ", "conclusions": "\\label{sec:conclusions} We present deep integrated observations of the warm neutral medium for three CHVCs using the 100 m Effelsberg telescope and high-resolution observations of the more compact regions from the WSRT, as well as the results of their combination. Our approach shows that the combination in the image domain meets the expectations. The critical step in the algorithm is the regridding, where a flux inconsistency can occur as a result of interpolation inaccuracies. The combination results demonstrate the importance of the zero-spacing correction regarding determination of the physical and morphological properties of the objects. Here in particular the warm neutral medium is of major interest. High-velocity clouds in higher ambient pressure environments might show complex spatial and spectral structures. The results of the Gaussian decomposition suggest strongly that the three clouds mostly consist of WNM. The analysis of \\citet{2011A&A...533A.105W} also confirms the lack of cold components in HVCs in the complex known as Galactic centre negative (GCN). However, high Doppler temperatures above $10^4$ K are an indication of the existence of turbulence within the clouds. The results demonstrate that the gas gets warmer or more turbulent in the tail region and at the edges of the clouds. This also suggests that the warm gas is floating away to the direction of the tail, which is an indication of existing ram pressure. Another remarkable aspect is the decreasing velocity gradient at the regions with lower column densities. The relatively slow movement of the clouds also reveals the lack of cold cores in the clouds, and the measured high Doppler temperatures demonstrate that the clouds are not in equilibrium. The linear approach used in the image domain is easy to implement and is not very CPU-intensive. It could be very helpful for computation of huge amounts of data coming from telescopes like ASKAP and WSRT/APERTIF." }, "1402/1402.5363_arXiv.txt": { "abstract": "We have conducted a survey of 17 wide ($>$ 100 AU) young binary systems in Taurus with the Atacama Large Millimeter Array (ALMA) at two wavelengths. The observations were designed to measure the masses of circumstellar disks in these systems as an aid to understanding the role of multiplicity in star and planet formation. The ALMA observations had sufficient resolution to localize emission within the binary system. Disk emission was detected around all primaries and ten secondaries, with disk masses as low as 10$^{-4}$ M$_{\\odot}$. We compare the properties of our sample to the population of known disks in Taurus and find that the disks from this binary sample match the scaling between stellar mass and millimeter flux of F$_{mm} \\propto$ M$_{\\ast}^{1.5-2.0}$ to within the scatter found in previous studies. We also compare the properties of the primaries to those of the secondaries and find that the secondary/primary stellar and disk mass \\emph{ratios} are not correlated; in three systems, the circumsecondary disk is more massive than the circumprimary disk, counter to some theoretical predictions. ", "introduction": "\\label{intro} Most stars are formed in binary or multiple systems and remain in such systems for their main sequence lifetimes \\citep [e.g.][]{mon07}. Therefore, understanding the causes and effects of multiplicity is an essential ingredient of complete models of both star and planet formation. Circumstellar disks play a crucial role in both processes, tracing effects of different binary formation mechanisms, providing conduits for material to accrete onto the stars, and serving as the reservoir of raw material for planet formation. At a given point in time, the distribution of observed disk masses is a function of the initial disk masses and disk evolution. For multiple systems, dynamical interactions between the stars, the circumstellar disks, and any circumbinary material will also impact both the disk formation and evolution. Models of binary star formation by \\citet{bat00} predict that the circumprimary disk, i.e. the disk around the more massive star, will have more mass than the circumsecondary disk; however, these models do not follow the viscous evolution of the disk after the formation stage. Observations to date largely support the prediction of a more massive circumprimary disk, although the sample of systems observed is relatively small and generally comprise only the brightest sources. \\citet{jen03} found that the primary star had the most massive disk in all four young binaries they observed; indeed in only one system was the secondary's disk detected at all, despite most of the secondaries showing signs of accretion. More recent work by \\citet{har12} has expanded the number of observed binary systems and also found that when both components were detected, the primary had higher flux, but with sensitivity levels of a few mJy, many secondaries remained undetected. Planet formation in these systems may also be impacted as models of the interactions of binary stars with their associated disks predict that the disks will be truncated somewhere between 0.2 and 0.5 times the binary separation, depending on the eccentricity of the system \\citep{art96}. However, these models do not address the surface density and evolution of the remaining disk material. If the secondary disks retain roughly the same surface density as the inner parts of disks around single stars, then they may still retain enough mass to form planets. Previous observations have not had the sensitivity to distinguish between disks that are simply truncated, and those that have been significantly depleted by further accretion. Disk models show that truncation effects can affect the observed flux for separations up to a few hundred AU \\citep{jen96}. The essential question for planet formation, then, is whether or not the disks around individual components of close binary stars are similar to the inner regions of disks around single stars. Early observations demonstrated that the unresolved millimeter emission, which traces the dust in the outer regions of the disk, is indeed reduced, consistent with truncation \\citep{bec90,ost95,jen96}. But most observations of binaries with separations in the ranges of 50-100 AU have yielded upper limits rather than detections, and indeed only about half of all young binaries in Taurus have been detected at all at millimeter wavelengths, despite the fact that many more than half of them were detected by IRAS at 60 $\\mu$m. With the advent of ALMA observations, which provide a substantial increase in sensitivity at the required resolution, it is now possible to reach much lower disk surface densities, and possibly to detect very low mass protoplanetary disks. To address these issues, we have obtained ALMA Cycle 0 observations of 17 young binary systems in Taurus for which the components can be resolved. In \\S \\ref{sample} we describe the sample selection and properties, in \\S \\ref{obs} we describe the ALMA observations and data reduction, in \\S \\ref{results} we present the results, and we give our conclusions in \\S \\ref{conclusions}. \\begin{deluxetable*}{llllll}[b!] \\tabletypesize{\\footnotesize} \\tablewidth{0pt} \\tablecaption{Taurus Binary Sample} \\tablehead{ \\colhead{Source Name} & \\colhead{Additional names} & \\colhead{Separation} & \\colhead{mm flux\\tablenotemark{a}} & \\colhead{Spectral types\\tablenotemark{b}} & \\colhead{Previous mm} \\\\ & & \\colhead{(AU)} & \\colhead{(mJy)} & & \\colhead{detection\\tablenotemark{c}} } \\startdata FV Tau & & 101 & 48$\\pm$5 & K5/K6 & 1,2\\\\ HBC 387 & FV Tau/c & 104 & $<$25 & M2.5/M3.5 & --\\\\ FQ Tau & & 110 & 28$\\pm$7 & M3/M3.5 & 1\\\\ UY Aur & & 120 & 102$\\pm$6 & M0/M2.5 & 1,2\\\\ FX Tau & & 130 & 17$\\pm$3 & M1/M4 & 1\\\\ HBC 411 & CoKu Tau/3 & 290 & $<$8 & M1/M4.5 & --\\\\ IRAS 05022+2527 & CIDA 9 & 320 & 71$\\pm$7 & K8/M1.5 & 1\\\\ HK Tau & & 340 & 130$\\pm$2 & M0.5/M2 & 1,2\\\\ IT Tau & & 340 & 22$\\pm$3 & K3/M4 & 1,2\\\\ DK Tau & & 350 & 80$\\pm$10 & K8/M1 & 1\\\\ GK Tau & & 340 & 33$\\pm$7 & K7 & 1\\\\ HN Tau & & 430 & 29$\\pm$3 & K5/M4 & 1 \\\\ V710 Tau & & 450 & 152$\\pm$6 & M0.5 & 1\\\\ IRAS 04113+2758 & MHO 1/2& 550 & 380$\\pm$3 & M2.5/M2.5 & 1,2\\\\ IRAS 04298+2246 & JH 112 & 920 & 30$\\pm$10 & K6/M8.5 & 1\\\\ HO Tau & & 970 & 44$\\pm$6 & M0.5 & 1 \\\\ DS Tau & & 990 & 39$\\pm$4 & K5 & 1\\\\ \\enddata \\label{tab:sample} \\tablenotetext{a}{Millimeter fluxes are single-dish fluxes at 850 $\\mu$m taken from \\citet{and05} except for HK Tau \\citep[850 $\\mu$m, interferometry]{har12} and IRAS 04113+2758 \\citep[1.3 mm, interferometry]{har12} } \\tablenotetext{b}{Spectral types from \\citet{and13}} \\tablenotetext{c}{1 = primary, 2 = secondary} \\end{deluxetable*} \\section {Sample} \\label{sample} We selected targets from a single star formation region, Taurus (distance $\\sim$ 140 pc), so that effects such as age and cluster environment are kept constant as much as possible. Taurus is ideal in having a significant population of YSOs that have evolved into the disk-only state (with no remaining envelope) and is very well studied, containing both a well known set of single stars with disks for comparison and a significant population of binaries where both stellar components have been characterized in the optical or near-infrared. We started with the list of known Taurus binaries \\citep{and05,kra09} and selected those with separations in the range of 0\\farcs7 (100 AU) to 10\\arcsec\\ (1400 AU). The inner cutoff was selected such that the two components could be clearly resolved with the resolution offered in Cycle 0, while the outer cutoff was chosen to ensure that the systems are likely to be physically associated. We eliminated systems classified as Class I from their spectral energy distributions, as these systems often contain substantial envelope emission that must be disentangled from the disk emission, and we eliminated systems with no active accretion signatures that had not been previously detected at millimeter wavelengths. The resulting sample contains 17 systems (Table \\ref{tab:sample}) and includes all Class II Taurus binaries with separations of 100 to 1400 AU from \\citet{and05} and \\citet{kra09}. Higher-order multiple systems were excluded where known at the time of our sample selection, although two were observed (see notes in \\S \\ref{individ}). ", "conclusions": " \\begin{itemize} \\item The majority of our new detections were for secondary sources and for these wide binaries ($> 100$ AU), the secondary disk fraction is somewhat higher than shown in previous studies. We found 11 of the 15 bona fide young stellar binaries have disks with masses $\\geq 10^{-4}$M$_{\\odot}$ around both stars, while \\citet{jen03} detected a circumsecondary disk in 1 of 4 systems and \\citet{har12} detected a circumsecondary disk in 6 of 12 systems where the components were resolved. There is significant overlap in the samples between these studies and the new detections are primarily due to the higher sensitivity of the ALMA observations. The newly detected primary disk masses and most of the secondary disk masses are considerably smaller than the minimum mass solar nebula, but this is not surprising given that the host stars are generally less than 1 M$_{\\odot}$. While it may be difficult for massive planets to form in these less massive disks, models of core accretion around lower-mass stars show that they may be able to form cores for lower-mass planets \\citep[e.g.][]{lau04} and if the disk mass has evolved, larger planets may have formed earlier when the disk was more massive. \\item In two binary systems the secondary disk has a higher mm flux than the primary disk. This has not been seen in previous, smaller surveys and is counter to predictions of formation models where the infalling material is directly accreted onto the primary or secondary disk as opposed to accreting onto a circumbinary structure. This result could be explained by faster dissipation of the primary disk, which has been shown to be a function of stellar mass. \\item For this sample of wide binaries, the secondary/primary disk mass ratio is not correlated with the secondary/primary stellar mass ratio. This suggests that for these binary systems, any environmental factors shared between the two components that could affect the initial disk mass and disk evolution are not the dominant factor in determining the range of disk masses for a given stellar mass. \\end{itemize} From these conclusions, it is clear that binaries do not follow a simple pattern of primary/secondary disk mass distribution; therefore, care should be taken when assigning flux to components in unresolved systems." }, "1402/1402.4479_arXiv.txt": { "abstract": "{} {We investigate whether the cylindrical (galactocentric) radial velocity gradient of $\\sim-3\\kmseckpc$, directed radially from the Galactic center and recently observed in the stars of the solar neighborhood with the \\emph{RAVE} survey, can be explained by the resonant effects of the bar near the solar neighborhood.} {We compared the results of test particle simulations of the Milky Way with a potential that includes a rotating bar with observations from the \\emph{RAVE} survey. To this end we applied the \\emph{RAVE} selection function to the simulations and convolved these with the characteristic \\emph{RAVE} errors. We explored different ``solar neighborhoods\" in the simulations, as well as different bar models.} {We find that the bar induces a negative radial velocity gradient at every height from the Galactic plane, outside the outer Lindblad resonance and for angles from the long axis of the bar compatible with the current estimates. The selection function and errors do not wash away the gradient, but often make it steeper, especially near the Galactic plane, because this is where the \\emph{RAVE} survey is less radially extended. No gradient in the vertical velocity is present in our simulations, from which we may conclude that this cannot be induced by the bar.} {} ", "introduction": "Many of the past efforts in modeling the mass distribution of the Milky Way have assumed that the Galaxy is axisymmetric and in a steady state. However, there is a wealth of evidence that these assumptions are not really valid. The two most important deviations from axisymmetry are the spiral arms and the bar. These features are not only apparent as non-axisymmetric density enhancements, but they also have long-range gravitational effects. In particular, the bar modifies the kinematics of the outer parts of the Galactic disks, far beyond its extension, through resonant interactions. That the velocity distribution of stars very near to the Sun is not smooth (as one would expect in a steady state axisymmetric system), but instead rich in substructures, has been established observationally thanks to data from the \\emph{Hipparcos} satellite and other surveys (\\citealt{Dehnen1998,Famaey2005,Antoja2008}). Several authors have explained these substructures as being due to orbital resonant effects of the bar (\\citealt{Dehnen2000, Fux2001}), of the spiral arms (\\citealt{Mayor1970,DeSimone2004}; \\citealt{Antoja2011}), or both (\\citealt{Antoja2009,Quillen2011}). Using data from the \\RAVE survey (\\citealt{Steinmetz2006}), \\cite{Antoja2012} discovered that some of the kinematic substructures detected in the vicinity of the Sun can be traced further, both on and above the plane of the Galaxy, up to $\\sim 0.7\\Kpc$. But \\emph{RAVE} also made it possible to discover large scale streaming motions. \\cite{Siebert2011grad} (in the rest of the paper S11) used a sample of $213,713$ stars to discover a gradient in the mean galactocentric radial velocity that decreases outward with Galactic radius. S11 show that this gradient was also present when using only the $29,623$ red clump stars in their sample, whose distances are more accurate. \\cite{Siebert2012} modeled the gradient as caused by a long-lived spiral pattern. \\cite{Williams2013} (hereafter W13) studied the 3D velocity distribution of red clump stars in \\emph{RAVE} in detail, confirmed the existence of the radial velocity gradient and also discovered a more complicated vertical velocity distribution than expected, attributing it to secular phenomena in the Galaxy. \\cite{Faure2014} generalized to 3D the model for the spiral arms presented in \\cite{Siebert2012}, which now also depends on the distance from the Galactic plane. This model nicely predicts a behavior for the mean vertical velocity that is similar to what is observed in W13 (i.e., resembling ``rarefaction-compression'' waves), together with the radial velocity gradient. On the other hand, \\cite{Monari2013} (hereafter M13), used 3D test particle simulations to show that the gravitational effects of the bar can significantly affect the kinematics of stars near the Sun, even at distances from the Galactic plane up to at least $z\\sim1\\Kpc$ for the thin disk and $z\\sim2\\Kpc$ for the thick disk. These results imply that some of the substructures found in \\cite{Antoja2012} could also be caused by the bar. In this paper we investigate an alternative explanation for the observed radial velocity gradient, beyond that caused by the spiral arms, by suggesting that it can be created by the bar. To do so, we compare the results of the test particle simulations in M13 with the \\emph{RAVE} data. The paper is organized as follows. In Sect.~\\ref{sect:sim} we summarize the salient characteristics of the simulations from M13, and in Sect.~\\ref{sect:sf} we describe how we apply the \\RAVE selection function and error convolution to them, to mimic a \\RAVE catalog. In Sect.~\\ref{sect:res} we present the results. In Sect.~\\ref{sect:expl} we explain how the bar can create a radial velocity gradient as observed. In Sect.~\\ref{sect:disc} we discuss the similarities between our results and the ones in W13 and in Sect.~\\ref{sect:concl} we conclude. ", "conclusions": "\\label{sect:concl} In this work we have proposed a new explanation for the recent discovery (S11, W13) of a negative $R$ gradient of the (galactocentric) radial velocity. We found that the bar can create a negative gradient if the Sun is placed just outside the outer Lindblad resonance and at angles from the long axis of the bar similar to the current estimates from the literature. The velocity gradients become steeper when increasing the angle from the bar and also for the Long Bar model. On the other hand, in the less massive bar case they become shallower. Moreover, such gradients do not depend strongly on the height from the Galactic plane. This happens because the bar affects the kinematics of the Galaxy almost in the same way from $z=0$ to $z\\sim2\\Kpc$, as explained in M13. Because of this, the bar provides a natural mechanism for the observed gradients at different heights. We compared the 3D test particle simulations presented in M13 with the findings of \\emph{RAVE}, after applying the \\emph{RAVE} selection function and proper error convolution. The gradients exist in our simulations for all bar parameters and positions of the Sun explored (all outside the outer Lindblad resonance). These gradients are never completely washed out by the selection function and the errors, but rather they are enhanced in some cases. In fact, the gradients in the solar neighborhood spheres considered are in general shallower than those observed in the Milky Way, but the selection function can enhance them to the level of $\\sim3-4\\kmseckpc$ (as e.g., happens for all the studied simulation slices with $|z|<0.5\\Kpc$ and the Sun centered at $R=8\\Kpc$). However, none of the models that we explored in this work accurately describes the behavior in \\RAVE of $\\avvR$ at every $z$: the gradients are too shallow for $z<0$. Some models resemble \\RAVE for $z>0$, especially our default bar case at $R=8\\Kpc$ and $\\phi\\leq-20\\degr$. We conclude from this that the bar should at least contribute to the negative gradient observed, for position angles with respect to the bar $\\phi < 0$ and for locations of the Sun near but outside the outer Lindblad resonance ($R > \\ROLR$). Furthermore, our simulations do not show any kind of vertical velocity gradient as seen in the data for \\RAVE by W13. This result is consistent with the distribution function of the simulated disks being an even function of $v_z$. On the other hand, the recent paper by \\cite{Faure2014} shows that a 3D model for spiral arms is successful in reproducing radial and vertical velocity gradients similar to those observed in W13. In reality both effects of bar and spiral arms probably coexist and shape the velocity distribution of the solar neighborhood. However, while in the case of the bar the slope of the radial velocity gradient depends significantly on the angular location of the observer in the Galaxy, in the case of tightly wound spirals the angle is much less important (Fig.~6 and 7 in \\citealt{Faure2014}). Future observations of the Galactic disk (e.g., obtained with the \\emph{Gaia} satellite) are expected to be sufficiently extended to distinguish whether the main cause of the radial velocity gradient is the bar or the spiral arms. A natural future development of this work is to fit the kinematics of the extended solar neighborhood with the analytic predictions from the bar perturbation theory, in the same fashion as in \\cite{Siebert2012} for the spiral arms, in order to retrieve the best fit values for the bar pattern speed, bar angle, and bar strength." }, "1402/1402.6336_arXiv.txt": { "abstract": "Gamma Ray Bursts (GRBs) are characterized by ultra-relativistic outflows, while supernovae are generally characterized by non-relativistic ejecta. GRB afterglows decelerate rapidly usually within days, because their low-mass ejecta rapidly sweep up a comparatively larger mass of circumstellar material. However supernovae, with heavy ejecta, can be in nearly free expansion for centuries. Supernovae were thought to have non-relativistic outflows except for few relativistic ones accompanied by GRBs. This clear division was blurred by SN 2009bb, the first supernova with a relativistic outflow without an observed GRB. Yet the ejecta from SN 2009bb was baryon loaded, and in nearly-free expansion for a year, unlike GRBs. We report the first supernova discovered without a GRB, but with rapidly decelerating mildly relativistic ejecta, SN 2012ap. We discovered a bright and rapidly evolving radio counterpart driven by the circumstellar interaction of the relativistic ejecta. However, we did not find any coincident GRB with an isotropic fluence of more than a sixth of the fluence from GRB 980425. This shows for the first time that central engines in type Ic supernovae, even without an observed GRB, can produce both relativistic and rapidly decelerating outflows like GRBs. ", "introduction": "\\subsection{Ordinary Supernovae} The optical lightcurves of supernovae have been well described by \\citet{1982ApJ...253..785A} as a nearly black body photosphere which recedes into the ejecta, heated by $\\gamma$-rays from nuclear decay and cooled by rapid expansion with a characteristic velocity of $v\\sim 10^4$ km s$^{-1}$. The interaction of the ejecta with the circumstellar medium set up by the stellar wind of the progenitor star has been described by \\citet{1982ApJ...258..790C} using self similar solutions. Such solutions have also successfully described the radio emission from type Ic supernovae \\citep{1998ApJ...499..810C}. The combination of nearly free expansion leading to decreasing optical thickness from free-free or synchrotron self absorption (SSA) and decreasing magnetic fields produces a decreasing peak frequency (the frequency at which the peak in the radio spectrum occurs), but a nearly constant flux density at the peak frequency \\citep{1998ApJ...499..810C}. In the self-similar solution, this interaction produces a shockfront which expands in a powerlaw fashion, with $R \\propto t^m$, where $R$ is the radius of the shockfront, $t$ is the time since shock breakout and $m$ is known as the expansion parameter (also sometimes called the deceleration parameter). In the case of Type Ib/c supernovae, which generally have a relatively tenuous circumstellar medium, this interaction produces little deceleration, and $m$ is close to 1. The ejecta would slow down significantly only after encountering a mass of external medium comparable to the $\\sim1$ M$_\\odot$ of ejecta expected for Type I b/c supernovae. This Sedov time \\citep{1950RSPSA.201..159T} is expected to be $\\gtrsim 10^2$ years for supernovae. Therefore young supernovae are usually found in a Newtonian phase of nearly free expansion. \\subsection{Gamma Ray Bursts} GRBs were discovered by \\cite{1973ApJ...182L..85K} using the Vela satellites, designed for the detection of nuclear tests in space. Ultra-relativistic blast waves had already been described by fluid dynamical \\citet{1976PhFl...19.1130B} solutions before they were implicated in the production of GRBs \\citep{1986ApJ...308L..47G,1986ApJ...308L..43P}. The production of GRBs require the ejecta to have initial bulk Lorentz factors of $\\Gamma \\gtrsim 10^2$ \\citep{1999PhR...314..575P} in order to overcome the pair production opacity \\citep{1975NYASA.262..164R,1978Natur.271..525S}. Furthermore, even a small number of baryons in the initial ejecta can soak up most of the explosion energy available for $\\gamma$-rays \\citep{1990ApJ...365L..55S}. This is called baryon poisoning. So GRB jets must have $\\lesssim 10^{-6}$ M$_\\odot$ of relativistic ejecta \\citep{1999PhR...314..575P} after breakout from the stellar progenitor surface, so as not to be baryon poisoned. In a short time, this relatively small mass of ejecta encounters a larger mass of external matter. Therefore GRB afterglows are found in a relativistic but rapidly decelerating phase and have $\\Gamma \\lesssim 20$. The broad-lined (high velocity) type Ic SN 1998bw was discovered in the direction of GRB980425 \\citep{1998Natur.395..670G}, and was characterized by the bright radio emission from its relativistic ejecta which had $\\Gamma \\sim 2-3$ \\citep{1998Natur.395..663K}. This association between a Type Ic supernova and a long GRB was followed by other supernovae like that of GRB030329 with SN 2003dh \\citep{2003Natur.423..847H} and XRF060218 with SN 2006aj \\citep{2006Natur.442.1011P,2006Natur.442.1014S} also associated with GRBs. Some of these supernovae were marked by broad lines or by asphericity. Yet the definitive property, of this small subset of Ic supernovae, that let them produce relativistic outflows remains elusive. \\subsection{Search for relativistic ejecta} This evidence, for a supernova-GRB connection, inspired a systematic search for relativistic ejecta from nearby type Ic supernovae using radio observations, leading to the discovery of SN 2009bb by \\citet{2010Natur.463..513S} with $\\gtrsim10^{49}$ erg of energy in radio emitting relativistic ejecta. Like GRBs, SN 2009bb had relativistic ejecta \\citep{2010Natur.463..513S} but this ejecta continued to be in nearly free expansion for $\\sim 1$ year \\citep{2011NatCo...2E.175C,2010ApJ...725....4B}, leading to the suggestion, that unlike GRBs, it is baryon loaded \\citep{2011ApJ...729...57C}. It has been suggested by \\citet{2011NatCo...2E.175C} that such engine-driven relativistic supernovae can even accelerate ultra-high-energy cosmic rays. Such a baryon loaded fireball has also been implied in PTF11agg \\citep{2013ApJ...769..130C}. A rapid decline in flux density, faster than $\\propto t^{-1}$ at a given radio frequency, or a decreasing peak flux density may signal deceleration of the ejecta. Recently the rapidly declining PTF12gzk was observed with $\\lesssim10^{46}$ erg of energy in fast ejecta \\citep{2013ApJ...778...63H}. So far, however, highly energetic ($\\gtrsim10^{49}$ erg) relativistic ejecta combined with rapid deceleration has never been observed until now in a supernova unassociated with a GRB. ", "conclusions": "Even though no GRB counterpart was found for SN 2012ap, its radio afterglow shares remarkable characteristics with those of GRB associated supernovae; it has a relativistic ejecta component with relatively few baryons. While SN 2009bb had a relativistic outflow, it was clearly baryon loaded which is not the case here. A rapid radio decline has been observed in PTF12gzk which may also have had fast ejecta. But SN 2012ap had orders of magnitude more energy in relativistic ejecta, putting its energetics firmly in the class of GRB-associated supernovae. SN 2012ap pushes the boundaries of explosion parameters observed from stripped core progenitors with central engines (see Figures \\ref{vm_schem} and \\ref{momemass}). The parameters that separate the outflows, from GRB associated supernovae and ordinary Type Ib/c supernovae, according to \\citet{2011ApJ...729...57C} are the velocity and baryon loading of their fastest ejecta. According to \\citet{2014arXiv1402.6344M} the duration of central engine activity drives the diversity of explosion outcomes. By bridging the gap between ordinary supernovae and GRB associated supernovae in terms of its high velocity and low ejecta mass, SN 2012ap demonstrates the role of CEDEXs in understanding the supernova-GRB connection." }, "1402/1402.1499_arXiv.txt": { "abstract": "We study the relationship between the stability level of late-type galaxy disks and their star-formation activity using integral-field gaseous and stellar kinematic data. Specifically, we compare the two-component (gas$+$stars) stability parameter from \\citeauthor{2011MNRAS.416.1191R} ($\\qrw$), incorporating stellar kinematic data for the first time, and the star-formation rate estimated from 21cm continuum emission. We determine the stability level of each disk probabilistically using a Bayesian analysis of our data and a simple dynamical model. Our method incorporates the shape of the stellar velocity ellipsoid (SVE) and yields robust SVE measurements for over 90\\% of our sample. Averaging over this subsample, we find a meridional shape of $\\sigma_z/\\sigma_R = 0.51^{+0.36}_{-0.25}$ for the SVE and, at 1.5 disk scale lengths, a stability parameter of $\\qrw = 2.0\\pm 0.9$. We also find that the disk-averaged star-formation-rate surface density ($\\Ssfre$) is correlated with the disk-averaged gas and stellar mass surface densities ($\\Sge$ and $\\Sse$) and anti-correlated with $\\qrw$. We show that an anti-correlation between $\\Ssfre$ and $\\qrw$ can be predicted using empirical scaling relations, such that this outcome is consistent with well-established statistical properties of star-forming galaxies. Interestingly, $\\Ssfre$ is not correlated with the gas-only or star-only \\citeauthor{1964ApJ...139.1217T} parameters, demonstrating the merit of calculating a multi-component stability parameter when comparing to star-formation activity. Finally, our results are consistent with the \\citeauthor{2010ApJ...721..975O} model of self-regulated star-formation, which predicts $\\Ssfre/\\Sge\\propto\\Sse^{1/2}$. Based on this and other theoretical expectations, we discuss the possibility of a physical link between disk stability level and star-formation rate in light of our empirical results. ", "introduction": "\\label{sec:intro} Stars are formed by the collapse of gas. In galaxy disks, a gas cloud should be gravitationally unstable if (1) it cannot adjust its internal pressure to balance the local gravitational pressure on timescales shorter than a free-fall time and (2) it occupies an area smaller than the scale on which differential rotation will shear it apart. \\citet{1964ApJ...139.1217T} codified these concepts into a criterion for the stability of an infinitely thin, rotating, self-gravitating, fluid disk: \\begin{equation} Q = \\frac{\\kappa \\sigma}{\\pi G \\Sigma} > 1, \\label{eq:q} \\end{equation} where \\begin{equation} \\kappa^2 = 2\\frac{v_c}{R}\\left( \\frac{v_c}{R} + \\frac{\\partial v_c}{\\partial R}\\right) \\label{eq:efreq} \\end{equation} is the epicyclic frequency, $v_c$ is the circular speed of the potential, $\\sigma$ is the radial velocity dispersion, $\\Sigma$ is the mass surface density, and $G$ is the gravitational constant. However, self-gravity is only one major player in the star-formation process, with chemodynamical processes, turbulence, and magnetic fields also significantly affecting the dynamics \\citep[see][and references therein]{2007ARA&A..45..565M}. The relative importance of these physical properties to the star-formation law is a matter of ongoing debate. Empirical studies of the star-formation law in disk galaxies have predominantly focused on observations of their gaseous components. In a seminal article, \\citet[][hereafter \\citetalias{1998ApJ...498..541K}]{1998ApJ...498..541K} demonstrated that the star-formation rate per unit area ($\\Sigma_{\\rm SFR}$), or equivalently the time derivative of the stellar mass surface density ($\\Ssfr \\equiv \\Sigma_{\\rm SFR}$), is well correlated with the surface density of the hydrogen gas, $\\Sigma_{\\rm H} = \\sdhi+\\sdmh$, following the star-formation law suggested by \\citet{1959ApJ...129..243S}. Considering the proportionality from equation \\ref{eq:q}, one might expect such a relation if star-formation is driven by self-gravity. Owing much to the flood of relevant data, the quantitative details of the star-formation law in galaxy disks and its relation to the Kennicutt-Schmidt (KS) law \\citepalias[$\\Ssfr \\propto \\Sigma_{\\rm H}^{1.4}$;][]{1998ApJ...498..541K} have been greatly scrutinized. This scrutiny has lead to a number of alternatives to this paradigm; see compilations by, e.g., \\citet{2008AJ....136.2782L} and \\citet{2013MNRAS.434.3389Z}. Many of these alternatives involve consideration of gas properties, such as accounting for the dust-to-gas ratio \\citep{2013AJ....146...19L}, or use of specific gas tracers. \\citet{2002ApJ...569..157W} and \\citet{2008AJ....136.2846B} have shown that $\\Ssfr$ is more tightly correlated with $\\sdmh$ than with $\\Sigma_{\\rm H}$. Therefore, the star-formation law will also be affected by the ability of the interstellar medium to convert \\hone\\ to H$_2$ \\citep{2004ApJ...612L..29B, 2006ApJ...650..933B}, which can be related to the hydrostatic pressure in the disk plane \\citep{1989ApJ...338..178E, 1993ApJ...411..170E}. As an inherently dynamical process, the star-formation law should also be influenced by relevant dynamical timescales, such as the local free-fall time \\citep{2012ApJ...745...69K}. \\citetalias{1998ApJ...498..541K} considered a star-formation law that incorporated the orbital timescale \\citep{1997ApJ...481..703S, 1997RMxAC...6..165E}, which is relevant in a scenario where dynamical processes in the disk (such as bars and spiral arms) are a primary driver of star formation. Here, we explore the role of the disk stability level in the star-formation process. Although the star-formation law may be written to show an explicit dependence on the {\\it gas} stability parameter \\citep[$Q_g$; see, e.g.,][]{2005ApJ...630..250K}, the most relevant assessment of the stability level includes both the gas and the stars \\citep{2005ApJ...626..823L}. Indeed, \\citet{2003MNRAS.346.1215B} have shown that the two-component (gas$+$stars) stability parameter of 16 galaxies is a better estimator of the star-formation {\\em threshold} than one incorporating the gas alone. In general, however, the effect of the stellar component on the star-formation {\\em law} is not well-understood: \\citet{2003MNRAS.346.1215B} have also shown that a star-formation law that considers only the gas component (the KS law) is statistically indistinguishable from one proposed by \\citet{1994ApJ...430..163D} that incorporates the {\\em total} disk mass surface density. More recently, \\citet{2011ApJ...733...87S} have shown an explicit dependence of $\\Ssfr$ on $\\Sigma_\\ast$. Their ``extended Schmidt law'' --- one that incorporates the dependence on $\\Sigma_\\ast$ --- is consistent with the self-regulation model proposed by \\citet[][hereafter \\citetalias{2010ApJ...721..975O}]{2010ApJ...721..975O}, who find $\\Ssfr \\propto \\Sigma_g \\Sigma_\\ast^{1/2}$ for galaxies with a constant-scale-height stellar disk\\footnote{ In detail, the star-formation law from \\citetalias{2010ApJ...721..975O} allows for star formation in starless systems, which would be prohibited by a law with an explicit dependence on the stellar mass of a galaxy. } \\citep[see also][]{2011ApJ...743...25K, 2013ApJ...776....1K}. In their model, the explicit dependence of the star-formation law on $\\Sigma_\\ast$ is via its contribution to the vertical gravitational field of the disk. Therefore, we also consider the correlation between the star-formation activity of a disk and its stellar mass surface density. Previous studies considering the relation of the two-component stability level of disks and/or stellar mass surface density to the star-formation law have lacked the kinematic data necessary to measure either of these quantities dynamically. Instead, they have used stellar mass estimates from stellar-population-synthesis modeling, which have not been directly calibrated by dynamical mass measurements in external disk galaxies \\citep[see discussion in][hereafter \\citetalias{2010ApJ...716..198B}]{2010ApJ...716..198B}. However, with its unparalleled stellar kinematic data in the dynamically cold regime of galaxy disks and its ancillary gas data, the DiskMass Survey \\citepalias{2010ApJ...716..198B} is well suited to studying the effect of the stellar component (via its mass surface density and stability level) on star formation in galaxy disks. Our paper is organized as follows: We briefly discuss the relevant observational data in Section \\ref{sec:data}. We describe our dynamical modeling in Section \\ref{sec:genmodel}; however, a more detailed discussion of this modeling approach will be presented in a forthcoming paper. For now, we provide a brief summary of the equations used in the dynamical model in Appendix \\ref{app:dynmodel}, and we discuss the details of our sampling of the posterior probability of the model in Appendix \\ref{app:converge}. Our probabilistic modeling is the basis for our calculations of the disk stability parameter and stellar mass surface density. These calculations, the stability results, and a comparison of the star-formation properties of our galaxy sample with the ``Normal Spirals'' from \\citetalias{1998ApJ...498..541K} are discussed in Section \\ref{sec:stability}. We explore any correlations among disk stability level, stellar mass surface density, and star-formation rate in Section \\ref{sec:sfrcorr}. Among other findings, we show that the two-component disk stability parameter is anti-correlated with the star-formation activity of the disk. In Section \\ref{sec:scaling}, we show that this anti-correlation can be predicted by considering a closed system of empirical scaling relations. Finally, we summarize and briefly discuss our results in Section \\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} \\subsection{ Summary of Analyses and Empirical Findings } In this paper, we have used data from the DiskMass Survey, briefly described in Section \\ref{sec:data}, to study the relationship between dynamical properties of galaxy disks and their star-formation activity. Unlike other local-universe surveys, the DiskMass Survey has directly measured the stellar kinematics in a sample of galaxy disks, which are critical to dynamical calculations of both stellar surface mass density and disk stability level. Our calculations are based on a generative --- fully probabilistic --- model of the relevant gaseous and stellar observations, adopting a simple analytical model for the disk dynamics. We introduce our probabilistic modeling approach in Section \\ref{sec:genmodel}, we briefly outline the dynamical model in Appendix \\ref{app:dynmodel}, and we discuss our usage of a sophisticated MCMC algorithm to sample the probabilistic model in Appendix \\ref{app:converge}. A more complete discussion of our modeling approach will be the subject of a forthcoming paper. We calculate the star-formation rate, $\\sfr$, for each of our galaxies using measurements of the 21cm radio continuum \\citep[NVSS;][]{TPKMPhD} and the calibrations from \\citet{2001ApJ...554..803Y} (see Section \\ref{sec:qeqns}). These calculations are performed by including appropriate probability distributions for the relevant quantities in our probabilistic model (equation \\ref{eq:pfull}). We calculate an effective star-formation-rate surface density, $\\Sigma_{\\rm SFR}\\equiv\\Ssfre$, where the surface area of the disk is determined by $R_{25}$ measurements from NED (Table \\ref{tab:data}). Combined with effective gas mass surface densities from our probabilistic model (equation \\ref{eq:avsig}), we compare our galaxy sample to the ``Normal Spirals'' from \\citetalias{1998ApJ...498..541K} in Section \\ref{sec:KS} (see Figure \\ref{fig:ks}). We find that the galaxies in our sample have low star-formation rates relative to the KS law, but are fully consistent with the low star-formation end of the data presented by \\citet{2008AJ....136.2846B}. We calculate the disk stability parameter derived by \\citet{2011MNRAS.416.1191R}, $\\qrw$, which incorporates both the gaseous and stellar components (see Section \\ref{sec:qeqns}) and includes corrections for the disk thickness. In Section \\ref{sec:result}, we find a stability parameter of $\\qrw=2.0\\pm0.9$ at 1.5 $h_R$ marginalized over all galaxies in our sample. This result is comparable to other empirical assessments from the literature. In particular \\citet{2013MNRAS.433.1389R} similarly find $\\qrw\\sim2$ for their sample; however, they also find that the stellar component most often dominates the disk stability level. This is contrary to our results, which show that the gas-only stability parameter is lower than the star-only stability parameter ($Q_g < \\mathcal{Q}_\\ast$) for 65\\% of our sample. These different findings are more likely because of differences in the data used and the detailed analysis methods, as described in Section \\ref{sec:result}, not because of an intrinsic difference in our galaxy samples. A stability parameter of $\\qrw\\sim2$ is also comparable to expectations from N-body simulations; however, it should be noted that these theoretical studies typically calculate the nominal (infinitely thin, single-component, collisionless) \\citet{1964ApJ...139.1217T} criterion ($\\mathcal{Q}_\\ast$), as opposed to our calculations for a multi-component, non-zero thickness disk. An important avenue for numerical simulations of galaxy disks is to study the stability levels of disks that include realistic gas components. The primary goal of this paper has been to explore any dependence of the star-formation rate on two dynamical properties of our sample, $\\Sse$ and $\\qrw$ (see Section \\ref{sec:sfrcorr}). Our primary findings are: \\begin{itemize} \\item There is a clear correlation between $\\Sse$ and $\\Ssfre$ and a clear anti-correlation of $\\qrwmin$ with $\\Ssfre$ (Figure \\ref{fig:sigsfr}). \\item The anti-correlation between $\\qrwmin$ and $\\Ssfre$ is expected given a theoretical study by \\citet{2006ApJ...639..879L}. However, our galaxies exhibit significantly higher stability levels than in their simulations and our data show a steeper power-law slope in the relation. \\item We find that the star-formation efficiency (SFE; $\\Ssfre/\\Sge$) is correlated with $\\Sse$ (Figure \\ref{fig:sigsfg}), which is expected both observationally \\citep{2011ApJ...733...87S} and theoretically \\citepalias{2010ApJ...721..975O}. In detail, our data are consistent with the proportionality $\\Ssfre/\\Sge \\propto \\Sse^{1/2}$, which is a limiting behavior of the theory derived by \\citetalias{2010ApJ...721..975O}; see also \\citet{2011ApJ...743...25K, 2013ApJ...776....1K}. \\item If star formation in our galaxy sample is not strongly affected by other physical properties, the quantity $\\Ssfre \\Sge^{-1} \\Sse^{-1/2}$ should be roughly constant. Indeed, we find that this quantity is effectively uncorrelated with the large number of physical quantities we have calculated (listed in Section \\ref{sec:sfrcorr}; see the examples of $\\Sse$ and $\\qrwmin$ in Figure \\ref{fig:sigsfgs}). However, the scatter in the data is large. We find an error-weighted geometric mean of $\\langle\\log(\\Ssfre \\Sge^{-1}\\Sse^{-1/2})\\rangle = -3.25\\pm0.27$ in units of $(G/{\\rm pc})^{1/2}$ and a range of $-3.7\\leq\\log(\\Ssfre \\Sge^{-1}\\Sse^{-1/2})\\leq-2.7$. \\item Although the correlation is weak, we also find an indication that the SFE is anti-correlated with $\\qrwmin$ (Figure \\ref{fig:sigsfg}). This contradicts previous observational studies \\citep{2008AJ....136.2782L}, but is roughly consistent with the expectation provided by \\citet{2006ApJ...639..879L}. \\end{itemize} The anti-correlation between star-formation rate and disk stability parameter from Figure \\ref{fig:sigsfr} is {\\em not} seen if we instead consider the gas-only ($Q_g$) or star-only ($\\mathcal{Q}_\\ast$) stability parameters. In terms of the effect on star formation, this result is reasonable in that neither $Q_g$ nor $\\mathcal{Q}_\\ast$ includes the gravitational effects of the other component. Thus, our results show the importance of considering {\\em both} components in assessing the effect of the disk stability level on star formation. In Section \\ref{sec:scaling}, we quantitatively predict $\\qrw$ at 1.5$h_R$ for the galaxies in our sample based on a closed system of scaling relations and the following global quantities for our galaxies: $\\sfr$, $\\Upsilon_K$, $\\mu_{0,K}$, $h_R$, $R_{25}$, $\\sigma_{cg}$, and $\\alpha=\\sigma_z/\\sigma_R$. The accuracy of the prediction depends on the details of the assumed mass distributions; however, all four approaches we discuss exhibit systematic errors of roughly 35\\% (68\\% confidence interval). Assuming our galaxy sample is representative of the overall population \\citepalias{2010ApJ...716..198B}, our calculations demonstrate that one should {\\em expect} an anti-correlation between $\\Ssfre$ and $\\qrw$, particularly for $\\qrw^{1.5h_R} < 2$. \\subsection{ A Physical Link Between Disk Stability Level and Star Formation? } Our use of empirical scaling relations to predict the anti-correlation between $\\Ssfre$ and $\\qrw^{1.5h_R}$ suggests that this outcome is consistent with the physical drivers of other morphological and dynamical outcomes of late-type-galaxy evolution. However, does the anti-correlation imply a physical link between disk stability level and star formation? The studies of \\citet{2005ApJ...626..823L, 2006ApJ...639..879L} are particularly relevant because the star-formation in their simulations is, in fact, driven by gravitational instabilities, and our data are roughly consistent with their predictions (Figures \\ref{fig:sigsfr} and \\ref{fig:sigsfg}). However, most of our galaxies are very stable, in the regime in which \\citet{2005ApJ...626..823L} find that it is difficult to form stars ($\\qrwmin\\gtrsim1.6$). A comparison of Figure 10 from \\citet{2006ApJ...639..879L} with our Figure \\ref{fig:sigsfr} shows that our galaxies exhibit higher stability levels at the low star-forming end. Part of this discrepancy is due to systematic differences in our calculation of the stability parameter (e.g., our applied corrections for disk thickness); however, our data should yield relatively large stability values even after removing these systematic differences. Thus, just as seen via ultraviolet radiation in the very extended parts of disks \\citep{2007ApJS..173..538T}, our measurements suggest that star-formation occurs in locales with high stability levels. One theory that addresses this phenomenon is provided by \\citetalias{2010ApJ...721..975O}. \\citetalias{2010ApJ...721..975O} \\citep[see also][]{2011ApJ...743...25K} present a model for star formation where the interstellar medium (ISM) is divided into self-gravitating and diffuse components. The pressure in the diffuse gas --- regulated by heating, cooling, and supernova-driven turbulence --- is assumed to be balanced by the vertical gravitational forces of the disk. A fundamental assumption is that the surface density of gravitationally bound clouds is converted to stars over a star-formation timescale; however, this timescale ($\\sim$2 Gyr) is much longer than the free-fall time. The proportionality $\\Ssfr/\\Sigma_g \\propto \\Sigma_\\ast^{1/2}$ is a limiting behavior of their model, assuming the star-formation rate in all galaxies is similarly affected by chemical composition, turbulence, and magnetic fields. We have shown in Figure \\ref{fig:sigsfg} that this proportionality is consistent with our data, albeit with significant scatter. We have not assessed the variations in metallicity, turbulent pressure, or magnetic field strength in our sample, such that this may explain some of the scatter seen in our data. The correspondence of our data with the limiting behavior of the \\citetalias{2010ApJ...721..975O} prediction suggests that our galaxies form stars at a rate that maintains the pressure balance in the diffuse ISM. Such star formation does not require large-scale gravitational instabilities, but it does require the vertical gravitational field, largely effected by the stellar component in most cases, to maintain the total pressure in the disk midplane. Thus, star-formation can be active in the disks of our galaxies, despite our rather large measurements of $\\qrw$. This may argue against a physical link between $\\qrw$ and $\\Ssfre$ at large $\\qrw$. However, the consistency of the trend seen in Figure \\ref{fig:sigsfr} with the prediction of \\citet{2006ApJ...639..879L} at low $\\qrw$ is suggestive of the role gravitational instabilities play in this small subset ($\\sim$15\\%) of our sample. Therefore, a direct physical link between the stability level of galaxy disks and their star-formation activity may only be relevant to a small subset of the galaxies in the local universe; however, the story is most likely very different at earlier cosmic epochs \\citep[see, e.g.,][]{2009MNRAS.397L..64A}. \\subsection{ Spiral Structure Effects } A detailed analysis of the spiral-arm strength in our galaxy sample is beyond the scope of this paper; however, Figure 9 from \\citetalias{2010ApJ...716..198B} shows that spiral arms are easily discernible in all of our sample galaxies. From N-body simulations, we expect spiral structure to be more easily generated in disks with lower $\\mathcal{Q}_\\ast$. If we assume that the same is true for $\\qrwmin$, we might expect effects related to spiral arms to be more evident in galaxies with lower $\\qrwmin$. In their appendix, \\citetalias{2010ApJ...721..975O} discuss the effects of spiral structure on their equilibrium model. They find that the azimuthally averaged (or disk-averaged) star-formation-rate surface density should be significantly lower than the true value if the contrast between the arm and inter-arm gas mass surface density is sufficiently high. If this effect were evident in our data, we may expect the residuals of our galaxy sample about $\\Ssfre \\Sge^{-1}\\Sse^{-1/2} = 10^{-3.25}$ $(G/{\\rm pc})^{1/2}$ to be related to the strength of the spiral structure and therefore to $\\qrwmin$. However, Figure \\ref{fig:sigsfgs} shows no such relation. This may be because (1) $\\qrwmin$ is not a good proxy for the mass-loading of spiral-arm density waves, (2) our galaxies have all been similarly affected by spiral structure such that these properties have only changed the mean value of $\\Ssfre\\Sge^{-1}\\Sse^{-1/2}$, (3) the timescales involved in the passage or lifetime of spiral arms is such that the equilibrium is never fully realized or substantially altered within the spiral-arm regions, or (4) the scatter in $\\Ssfre \\Sge^{-1}\\Sse^{-1/2}$ caused by spiral arms or other physical processes has obscured the relation. Some of the scatter in our measurements of $\\Ssfre\\Sge^{-1}\\Sse^{-1/2}$ can be attributed to systematic error; however, there is room for intrinsic scatter as well. It is of great interest to understand the scatter in $\\Ssfre\\Sge^{-1}\\Sse^{-1/2}$, as it relates to spiral structure and/or other physical properties that can affect how stars form." }, "1402/1402.0383_arXiv.txt": { "abstract": "{{\\em Fermi}/LAT observations of star-forming galaxies in the $\\sim$0.1-100\\gev\\ range have made possible a first population study. Evidence was found for a correlation between \\gray\\ luminosity and tracers of the star formation activity. Studying galactic cosmic rays (CRs) in various global conditions can yield information about their origin and transport in the interstellar medium (ISM).}{This work addresses the question of the scaling laws that can be expected for the interstellar \\gray\\ emission as a function of global galactic properties, with the goal of establishing whether the current experimental data in the GeV range can be constraining.}{I developed a 2D model for the non-thermal emissions from steady-state CR populations interacting with the ISM in star-forming galaxies. Most CR-related parameters were taken from Milky Way studies, and a large number of galaxies were then simulated with sizes from 4 to 40\\,kpc, several gas distributions, and star formation rates (SFR) covering six orders of magnitude.}{The evolution of the \\gray\\ luminosity over the 100\\kev-100\\tev\\ range is presented, with emphasis on the contribution of the different emission processes and particle populations, and on the transition between transport regimes. The model can reproduce the normalisation and trend inferred from the \\textit{Fermi}/LAT population study over most of the SFR range. This is obtained with a plain diffusion scheme, a single diffusion coefficient, and the assumption that CRs experience large-scale volume-averaged interstellar conditions. There is, however, no universal relation between high-energy \\gray\\ luminosity and star formation activity, as illustrated by the scatter introduced by different galactic global properties and the downturn in \\gray\\ emission at the low end.}{The current \\textit{Fermi}/LAT population study does not call for major modifications of the transport scheme for CRs in the Milky Way when extrapolated to other systems, probably because the uncertainties are still too large. Additional constraints may be expected from doubling the {\\em Fermi}/LAT exposure time and later from observing at TeV energies with the \\textit{Cherenkov Telescope Array}.} ", "introduction": "\\label{intro} \\indent Star-forming galaxies harbour large-scale populations of energetic particles accelerated in violent phenomena associated with stellar evolution, such as supernova explosions, colliding-wind binaries, or pulsar wind nebulae. These galactic cosmic rays (hereafter CRs) interact with the various components of the interstellar medium (ISM): gas, turbulence, radiation, and magnetic fields. These interactions determine the transport of CRs from sources to intergalactic medium, and are responsible for their confinement and accumulation in the galactic volume over a large number of acceleration episodes (of the order of $\\sim10^5$ for the Milky Way). The existence of such large-scale populations is attested by various extended emissions, notably in radio through synchrotron emission, but also in \\grays\\ through inverse-Compton scattering, Bremsstrahlung, and inelastic nuclear collisions. These emissions are signatures of the combined processes of CR acceleration and transport, and their interpretation in the frame of other astronomical data can provide us with a better picture of the life cycle of CRs \\citep[see][for a review]{Strong:2007}. \\indent Understanding CRs is relevant to many fields beyond high-energy astrophysics. This is mostly because they are more than a simple side effect of the end stages of stellar evolution. Galactic cosmic rays interact with the ISM and can significantly modify the environment in which they are flowing. They can heat and ionise the gas, especially at the centre of dense molecular clouds, thereby impacting star formation conditions \\citep{Papadopoulos:2010}; through their diffusion, CRs can alter the turbulence of the interstellar medium, with consequences on their own transport \\citep{Ptuskin:2006}; they can contribute to large-scale outflows, an important feedback effect on star formation over cosmic times \\citep{Hanasz:2013}; CRs may play a role in shaping large-scale galactic magnetic fields through a dynamo process \\citep{Hanasz:2004}; and last, CRs from classical astrophysical objects can hide the signatures of more exotic components of the Universe such as dark matter \\citep{Delahaye:2010}. For all these reasons, understanding the non-thermal content and emission of star-forming galaxies is crucial. \\indent The current generation of high-energy (HE) and very-high-energy (VHE) \\gray\\ telescopes has enabled us to detect several external galaxies that shine at photon energies $>100$\\mev\\ as a result of CRs interacting with their ISM (and not because of a central black hole activity, as is the case for the vast majority of detected \\gray-emitting external galaxies). The {\\em Fermi}/LAT space-borne pair-creation telescope has permitted the detection of five external galaxies at GeV energies (SMC, LMC, M31, M82, and NGC253), with two other systems being possible candidates \\citep[NGC4945 and NGC1068, whose emission may be contaminated by central black hole activity; see][]{Abdo:2009l,Abdo:2010d,Abdo:2010e,Abdo:2010f,Ackermann:2012}. In the same time, the HESS and VERITAS ground-based Cherenkov telescopes have permitted the detection at TeV energies of NGC253 and M82, respectively \\citep{Acero:2009,Abramowski:2012,Acciari:2009}. Star-forming galaxies can now be considered as an emerging class of \\gray\\ sources, with a notable increase of the number of detected objects from the previous generation of instruments \\citep[only the LMC was detected by CGRO/EGRET; see][]{Sreekumar:1992}. \\indent Even with such a limited sample, it is tempting to perform a population study to find clues about the main parameters regulating CR populations. From the {\\em Fermi}/LAT sample of five objects plus an estimate for the global emission of the Milky Way and an upper limit for M33, the work of \\citet{Abdo:2010f} showed first hints for a slightly non-linear correlation between \\gray\\ luminosity and star formation rate (SFR). A more systematic study by \\citet{Ackermann:2012}, including upper limits for about sixty luminous and ultra-luminous infrared galaxies, found further evidence for a quasi-linear correlation between \\gray\\ luminosity and tracers of the star formation activity. Such a relationship can be expected to first order, because the production of CRs is expected to scale with the rate of massive star formation; but it is surprising that the correlation is so close to linear, over such a wide range of galactic properties, from dwarfs like the SMC to interacting systems like Arp 220. \\indent The interpretation of this correlation therefore raised the question of the scaling laws that can be expected for the diffuse \\gray\\ emission as a function of global galactic properties. Clarifying this at the theoretical level is required to establish whether the current experimental data can be constraining in any way. This is all the more needed because the sample of detected objects is small, the coverage over the GeV-TeV range is uneven, and the spectral characterization is still rather limited. In addition, the determination of global properties of external systems, such as total mass or star formation rate, is in itself a challenge, and there is most likely some intrinsic scatter in the actual physical correlations. \\indent That star-forming galaxies might become a class of \\gray\\ sources was anticipated early, well before the launch of the \\textit{Fermi} satellite \\citep[see for instance][]{Volk:1996, Pavlidou:2001,Torres:2004}. The interest in the subject extended beyond the study of nearby objects, because this emission cumulated over cosmic time might account for a significant fraction of the isotropic diffuse \\gray\\ background \\citep{Pavlidou:2002,Thompson:2007, Ackermann:2012}. Before and after the launch of the \\textit{Fermi} satellite, the \\gray\\ emission from individual starbursts was studied by several authors, with the focus being mainly on M82 and NGC253 \\citep{Persic:2008,deCeadelPozo:2009,Lacki:2011,Paglione:2012}. \\indent The purpose of this paper is to provide a more global picture of the evolution of the \\gray\\ interstellar emission from star-forming galaxies, in the context of the recent population study conducted on the basis of {\\em Fermi}/LAT observations. Instead of trying to fit a model to a particular object, I defined a global framework by fixing as many cosmic-ray-related parameters as possible from Milky Way studies, leaving only a few global quantities as independent variables. Here, I present a generic model for the transport of CRs and the associated non-thermal interstellar emissions in star-forming galaxies. Simulating systems with various sizes and gas distributions thus allows investigating the impact of global galactic parameters on the \\gray\\ output. In a first part, I introduce the different components and related assumptions of the model, together with the set of synthetic galaxies that was considered. Then, the \\gray\\ emission over 100\\kev-100\\tev\\ for this series of synthetic galaxies is presented, and its evolution with global properties is analysed. In a subsequent part, the scaling of \\gray\\ luminosity with global galactic properties is derived for soft, high-energy, and very-high-energy \\gray\\ ranges, and is compared with observations in the case of the high-energy band. Last, the impact of the model parameters is discussed, and the connection to the far-infrared - radio correlation is addressed. ", "conclusions": "\\label{conclu} \\indent The interaction of galactic cosmic rays (CRs) with the interstellar medium (ISM) produces \\grays\\ through hadronic interactions, inverse-Compton scattering, and Bremsstrahlung. This interstellar emission has been observed in the Milky Way for decades and provided a way to indirectly probe CRs outside the solar system, especially the dominant nuclei component. The current generation of high-energy (HE) and very-high-energy (VHE) \\gray\\ instruments has enabled us to extend this kind of study to several other star-forming galaxies. Five external systems, possibly seven, were recently detected at GeV energies with the \\textit{Fermi}/LAT. Combined with upper limits on about sixty other star-forming galaxies, evidence was found for a quasi-linear correlation between \\gray\\ luminosity and tracers of the star formation activity. This raised the question of the scaling laws that can be expected for the interstellar \\gray\\ emission as a function of global galactic properties, with the aim of knowing whether this early population study can constrain the origin and transport of CRs. \\indent In the present work, I introduced a model for the non-thermal emissions from steady-state CR populations interacting with the ISM in star-forming disk galaxies. The model is two-dimensional and includes realistic CR source distributions as well as a complete treatment of their transport in the galactic disk and halo (in the diffusion approximation). This contrasts with most previous works, which relied on a one-zone approach and typical time-scale estimates. Prescriptions were used for the gas distribution, the large-scale magnetic field, and the interstellar radiation field. Cosmic-ray-related parameters such as injection rates and spectra or diffusion properties were taken from Milky Way studies. The model is set to reproduce the average interstellar conditions of the Galaxy, its global non-thermal emission, and the far-infrared - radio correlation. A series of star-forming galaxies with different sizes, masses, and gas distributions was simulated to assess the impact of global galactic parameters on the \\gray\\ output. The link between gas content and star formation was made from the Schmidt-Kennicutt relation. The full set covered almost six orders of magnitude in star formation rate (SFR). The emission was examined over 100\\kev-100\\tev, with dedicated discussions for instrumental bands: 100\\kev-10\\mev\\ (band I, soft \\grays), 100\\mev-100\\gev\\ (band II, HE \\grays), and 100\\gev-100\\tev\\ (band III, VHE \\grays). \\indent In band I, the emission is dominated by inverse-Compton emission, with Bremsstrahlung contributing at the $\\sim20$\\% level. With increasing average gas surface density, the luminosity rises more than linearly, with a power-law index on SFR $\\sim$1.2-1.3, depending on the gas distribution. This is the result of the growing contribution of secondary leptons. The latter eventually accounts for 70-80\\% of the emission. \\indent In band II, the emission is dominated by pion-decay emission, with Bremsstrahlung contributing at the $\\sim20$\\% level, and inverse-Compton being equally important for small galaxies with low average gas surface density. With increasing average gas surface density, the luminosity rises more than linearly, with a power-law index on SFR initially in the range $\\sim$1.4-1.7 and then decreasing to $\\sim$1.1-1.2. This behaviour is the result of the progressive shift of the CR nuclei population from a diffusion-dominated regime towards a loss-dominated regime. \\indent In band III, the emission is dominated by inverse-Compton emission for small galaxies with low average gas surface density. For larger and denser systems, pion decay is the main emission mechanism, accounting for more than 95\\% of the luminosity for the densest ones. With increasing average gas surface density, the luminosity rises more than linearly, with a power-law index on SFR in the range $\\sim$1.4-1.5. The behaviour is the same as in band II, with emitting particles being closer to diffusion-dominated, and with inverse-Compton becoming increasingly important at the low end. \\indent Without any specific tuning, the model was able to account for the normalisation and trend inferred from the {\\em Fermi}/LAT population study over almost the entire range of SFRs tested here. This suggests that the \\textit{Fermi}/LAT population study does not call for major modifications of the scheme inferred for CRs in the Milky Way when extrapolated to other systems, probably because uncertainties are still too large. A good match can be obtained with a plain diffusion scheme, a single diffusion coefficient, and the assumption that CRs experience large-scale volume-averaged interstellar conditions. There is no need for a strong dependence of the diffusion properties on interstellar conditions, and no impact of strong galactic winds in the most active star-forming systems. The only requirement in the framework used here is that small galaxies with a disk radius of a few kpc have halos of at least 2\\,kpc in half height. There is, however, no universal relation between high-energy \\gray\\ luminosity and star formation activity, as illustrated by the downturn in \\gray\\ emission at low SFR values and the scatter introduced by different galactic global properties. Varying the galaxy size, gas distribution, and gas density introduces a variation in the 100\\mev-100\\gev\\ luminosity at a given SFR by up to a factor of 3, while varying the SFR for a given gas layout was found to add another 50\\% of scatter. Interestingly, the same model accounted pretty well for the far-infrared - radio correlation over most of the range of galactic properties tested here, with a remarkably small scatter of $\\leq 30$\\% at any infrared luminosity and a decrease in synchrotron emission at the low end. \\indent Progress on this kind of population study is expected from the doubling of the {\\em Fermi}/LAT exposure, especially at low SFRs with the detection of M33 and the resolving of the Magellanic Clouds. An additional test of our understanding of interstellar \\gray\\ emission from star-forming galaxies may come from the 100\\gev-100\\tev\\ range. In this band, the scaling of luminosity with SFR has a higher non-linearity favourable to distant but more active objects, and the next-generation {\\em Cherenkov Telescope Array} with its ten-fold improvement in sensitivity is expected to perform constraining observations. \\begin{acknowledgement} I acknowledge support from the European Community via contract ERC-StG-200911 for part of this work. I wish to thank Olivier Bern\\'e and D\\'eborah Paradis for their assistance on interstellar radiation field models, and Andy Strong for his help with the GALPROP package. I also thank Keith Bechtol and Guillaume Dubus for reading and commenting on a draft version of the manuscript. \\end{acknowledgement}" }, "1402/1402.2386_arXiv.txt": { "abstract": "For the first time we reconstruct the magnetic helicity density of global axisymmetric field of the Sun { using} method proposed by \\citet{2003AdSpR..32.1835B} and \\citet{pip13M}. To determine the components of the vector potential, we apply the gauge which is typically employed in mean-field dynamo models. This allows for a direct comparison { of reconstructed helicity} with the predictions from the mean-field dynamo models. We apply the method to two different data sets: the synoptic maps of line-of-sight (LOS) magnetic field from the Michelson Doppler Imager (MDI) on board of Solar and Heliospheric Observatory (SOHO) and vector magnetic field measurements from Vector Spectromagnetograph (VSM) on Synoptic Optical Long-term Investigations of the Sun (SOLIS) system. Based on the analysis of MDI/SOHO data, we find that in solar cycle 23 the global magnetic field had positive (negative) magnetic helicity in the northern (southern) hemisphere. This hemispheric sign asymmetry is opposite to helicity of solar active regions, but it is in agreement with the predictions of mean-field dynamo models. { The data also suggest that the hemispheric helicity rule may have reversed its sign in early and late phases of cycle 23.} Furthermore, the data indicate an imbalance in magnetic helicity between the northern and southern hemispheres. This imbalance seem to correlate with the total level of activity in each hemisphere in cycle 23. Magnetic helicity for rising phase of cycle 24 is derived from SOLIS/VSM data, and qualitatively, its latitudinal pattern is similar to the pattern derived from SOHO/MDI data for cycle 23. ", "introduction": "The generation of the magnetic field in the Sun is tightly related with the convective helical motions. In the framework of axisymmetric dynamos, the magnetic field is typically decomposed into toroidal and poloidal components. \\citet{P55} suggested that solar dynamo can be represented as a periodic transformation of poloidal magnetic field, $\\bar{\\mathbf{B}}^{(p)}={\\bar{B}}_{r}\\mathbf{e}_{r}+{\\bar{B}}_{\\theta}\\mathbf{e}_{\\theta}$, into toroidal field $\\bar{\\mathbf{B}}^{(t)}={\\bar{B}}_{\\phi}\\mathbf{e}_{\\phi}$ (via the differential rotation) and the reverse transformation of $\\bar{\\mathbf{B}}^{(t)}$ to $\\bar{\\mathbf{B}}^{(p)}$ by helical convective motions. Further development of dynamo theory showed that the two processes produce helical magnetic fields on both small and large spatial scales \\citep{pouquet-al:1975a,pouquet-al:1975b}, and that the conservation of magnetic helicity is an important factor for the dynamical quenching of large-scale magnetic field generation \\citep{kleruz82,1991ApJ...376L..21C,1992ApJ...393..165V,kle-rog99,2000A&A...361L...5K,2005PhR...417....1B}. Early observations of various proxies of magnetic/current helicity established what is now known as the hemispheric helicity rule: magnetic fields of active regions exhibit preferentially negative (positive) helicity in the northern (southern) hemisphere \\citep[][and references therein]{see1990SoPh,pev95,zetal10}. On the other hand, some researchers \\citep[e.g.,][]{2003AdSpR..32.1835B,bl-br2003,warn2011,pip2013ApJ,pip13M} argued that magnetic helicity of large-scale (global) axisymmetric field should be positive/negative in the northern/southern hemisphere. Furthermore, the mean-field dynamo models predict reversals of the sign of helicity in association with the propagation of dynamo wave inside the convection zone \\citep[e.g.,][]{warn2011,pip13M}. Reversals of the hemispheric helicity rule have been reported in observations, but the results seems inconclusive, with some researchers reporting the presence of such reversals in early/late phases of solar cycle \\citep{bao2000,hag05}, while others are questioning their existence \\citep{pev2001,pev2008,2013ApJ...772...52G}. {Such apparent controversy may be resolved via direct comparison of observations with the model predictions. Since dynamo models can provide a detailed information about the distribution of magnetic helicity of global axisymmetric field in the convection zone and near the photosphere, it is highly desirable to directly compare these model estimates with the observations.} Theory suggests that magnetic helicity on small and large scales should have opposite sign \\citep[e.g.,][]{pouquet76,seehafer96,bl-br2003,2005PhR...417....1B,pip2013ApJ,pip13M}. Then, the small-scale helicity may dissipate on small spatial scales subject to the Ohmic dissipation \\citep[e.g.,][]{pouquet76} or the helicity of both signs could emerge through the solar photosphere. Early measurements of vertical component of large-scale current helicity density, \\citep[e.g.,][]{2000ApJ...528..999P,2010ApJ...720..632W} found that in its sign, the large-scale magnetic fields follow the same hemispheric helicity rule as the active regions. These early studies concentrated on spatial scales larger than the active regions but smaller then the solar hemisphere. One should note that in the framework of mean-field dynamo, the magnetic fields of active regions represent the ``small-scale'' fields, while the large-scale fields refer to spatial scales comparable with the size of solar hemisphere. In this article, we address the helicity determination for large-scale magnetic fields as defined by mean-field dynamo theory. To avoid confusion with previous studies, we use the terms ``global'' and ``large-scale'' to refer to magnetic fields on spatial scales much larger then active regions. We reconstruct the magnetic helicity of global axisymmetric field using the approach suggested by \\citet{2003AdSpR..32.1835B} and \\citet[][see Section 2 below]{pip13M}. Section 3 describes the data sets and the reduction procedure. Section 4 presents our main results, and in Section 5 we discuss our findings. ", "conclusions": "Using synoptic charts from SOHO/MDI, we, for the first time, reconstruct the magnetic helicity density of global axisymmetric field of the Sun. In solar cycle 23 the global axisymmetric magnetic field exhibits positive magnetic helicity in the northern hemisphere, and negative one in the southern. In general, such reconstructions require a knowledge of the $\\bar{B}_{r}$ and $\\bar{B}_{\\phi}$ components of the axisymmetric magnetic field. In the past, vector global magnetic field components were reconstructed via various approaches \\citep{2000ApJ...528..999P,2005ApJ...620L.123U,hoeks10,2012SoPh..280..379M}. Here we used synoptic charts of LOS magnetic field corresponding to different longitudinal offsets relative to central meridian to compute $\\bar{B}_{r}$ and $\\bar{B}_\\phi$, see equation (\\ref{eq:torf}). Derived $\\bar{B}_{r}$ agrees well with the $\\bar{B}_{r}$ provided by the MDI team, which we see as indirect validation of our method. Based on the analysis of dynamo equations, \\citet{2003AdSpR..32.1835B} suggested that the magnetic helicity of global magnetic field in solar cycle 23 should be positive/negative in the northern/southern hemisphere. \\citet{pip13M} analyzed the distributions of magnetic helicity for large- and small-scale magnetic fields in the axisymmetric mean-field dynamo taking into account the conservation of the total magnetic helicity in the dynamo processes. They concluded (see their Figures 2e,d and 5) that magnetic helicity density of large-scale field should have positive sign in the northern hemisphere and negative sign in the southern hemisphere during the most part of the magnetic cycle. Our present results provide observational support to these early theoretical predictions. We find that during most of cycle 23, global magnetic fields exhibited a persistent pattern of positive/negative helicity in the northern/southern hemispheres. In respect to helicity of active region magnetic fields (small-scale in the framework of this discussion), the hemispheric helicity rule is negative/positive in the northern/southern hemispheres \\citep[][and references therein]{see1990SoPh,pev95,zetal10}. Taken together, these two results support the notion that the solar dynamo creates helicity of two opposite signs as was suggested in early papers. However, helicity of both signs seem to cross the solar photosphere. We further found that the hemispheric helicity rule for global magnetic fields exhibits sign-reversals in early and late phases of cycle 23. If the helicities of small- and large-scale fields are tied together, this should imply a need for similar reversals in the hemispheric helicity rule for active regions. Alas, while some researchers claimed observing reversals in the hemispheric helicity rule near the minimum of solar cycle 22 and 23 \\citep{bao2000,hag05}, others were not able to find them \\citep{pev2001,pev2008,2013ApJ...772...52G}. Clearly, this question about possible reversals of the hemispheric helicity rule for active region magnetic fields needs to be re-examined. Although, the predictions of the model by \\cite{pip13M} agree qualitatively with the results reported in our paper, there are some differences related to the shape of the helicity density patterns. For example, our present results suggest that magnetic helicity density pattern of the same sign can extend from the equator to the poles which is not seen in the model. Similarly, the pattern of $\\bar{\\mathbf{A}}\\cdot\\bar{\\mathbf{B}}$ of reverse sign penetrates to equatorial regions during the minima of cycle around year 1997 and 2009. If the reversals of the hemispheric helicity rule are real, this will pose a challenge for some proposed mechanisms of helicity generation \\citep[e.g., helicity generation by the differential rotation,][]{2000JGR...10510481B}. { As an alternative explanation, our results could be interpreted in the framework of helicity of axisymmetric and non-axisymmetric parts of global magnetic fields. In that model, the axisymmetric component of helicity (derived in this article) follows the hemispheric helicity rule of opposite in sign to the non-axisymmetric component (associated with active regions). Such a possibility was raised by Zhang (2006), who used Berger \\& Ruzmaikin (2000) data to show that helicity flux of non-axisymmetric modes was opposite in sign to helicity of axisymmetric (m=0) mode.} { One may also question the importance of vector magnetic field measurements for studies of global helicity, when the line-of-sight data seem to provide reasonable results. Here we presented the first ever derivations of magnetic helicity density of global field based on vector synoptic maps. While we see similarities in distribution of global helicity derived from LOS and vector data, there are also some differences. For example, synoptic maps of toroidal field derived from LOS data show more or less uniform distribution of the magnetic field polarity of one sign suggested by the the Hale polarity law. In addition to that pattern, vector field maps show mix of two different polarities: one is more concentrated and other is somewhat diffused (Figure 3b). The diffuse component (of toroidal field) seems to correspond to a trailing polarity field. Such component of large-scale field is not present in the maps of toroidal flux derived from the LOS magnetic fields. The vector field data are limited to cycle 24, and thus, are insufficient to conclude if there are any changes in this diffuse component of toroidal field with solar cycle. This and other differences between derivations based on LOS or vector field data require further investigation.} Our findings indicate that helicity of the large-scale magnetic fields is imbalanced between the northern and the southern hemispheres in different phases of solar cycle. However, when taken over the entire cycle, the positive and negative helicity of large-scale magnetic field is well-balanced. Indirectly, this is in agreement with \\citet{geor2009}, who found that helicity injection through the solar photosphere associated with active region magnetic fields is well-balanced over the solar cycle 23. On the other hand, \\citet{2012ApJ...758...61Y} reported significant imbalance between helicity fluxes of northern and southern hemispheres. Our findings (Figure 8c) allow to reconcile \\citet{geor2009} and \\citet{2012ApJ...758...61Y} conclusions. {Due to limitations of the existing datasets, the observational studies of helicity often refer to proxies of current helicity density. It is usually assumed that these proxies represent magnetic helicity sufficiently well. Contrary to that, we find that while the general tendencies are similar in magnetic and current helicity densities, there are differences, for example, in small-scale patterns, which may be present in one helicity proxy but are absent in the other. For example,} proxy of current helicity, $\\bar{B}_{r}\\left(\\nabla\\times\\bar{\\mathbf{B}}\\right)_{r}$, exhibits a distinct ``zebra'' pattern, but no such pattern is present in the distribution of magnetic helicity. Early, \\citet{2000ApJ...528..999P} and \\citet{2013ApJ...772...52G} reported similar pattern in current helicity density of large-scale magnetic fields. The pattern could also be expected from spatial structure of the dynamo wave of the large-scale magnetic field components $\\bar{B}_{r}$ and $\\bar{B}_{\\phi}$, which are illustrated in Figure \\ref{fig:Components-of-the}a. We note that in the equatorial regions the inequality $\\bar{B}_{r}\\bar{B}_{\\phi}<0$ holds for the most part of the sunspot cycle (see, Figure \\ref{fig:Components-of-the}a). We also found that modes $b_{r}^{(3)}$ and $b_{\\phi}^{(2)}$ dominate, which means that $\\bar{B_{r}}\\left(\\boldsymbol{\\nabla}\\times\\mathbf{\\bar{B}}\\right)_{r}\\sim b_{\\phi}^{(2)}b_{r}^{(3)}P_{2}P_{3}$, here the sign of the $b_{\\phi}^{(2)}b_{r}^{(3)}$ defines the hemispheric sign rule and the product $P_{2}P_{3}$ defines that zebra pattern as illustrated in Figure \\ref{fig:tot}c and Figure \\ref{fig:tot-1}. Thus, the results shown in Figure \\ref{fig:tot}c are expected for any dynamo model that qualitatively reproduces Figure \\ref{fig:Components-of-the}a. Finally, keeping in mind the approximations which were used in the reconstruction of components of the global magnetic field of the Sun, our results should be considered as preliminary. Further development in this direction is likely to shed more light on the role of magnetic helicity in global solar and astrophysical dynamos." }, "1402/1402.5429_arXiv.txt": { "abstract": "\\noindent An intriguing discrepancy emerging in the concordance model of cosmology is the tension between the locally measured value of the Hubble rate, and the `global' value inferred from the cosmic microwave background (CMB). This could be due to systematic uncertainties when measuring $H_0$ locally, or it could be that we live in a highly unlikely Hubble bubble, or other exotic scenarios. We point out that the global $H_0$ can be found by extrapolating $H(z)$ data points at high-$z$ down to $z=0$. By doing this in a Bayesian non-parametric way we can find a model-independent value for $H_0$. We apply this to 19 measurements based on differential age of passively evolving galaxies as cosmic chronometers. Using Gaussian processes, we find $H_0=64.9 \\pm 4.2$ km s$^{-1}$ Mpc$^{-1}$ $(1\\sigma)$, in agreement with the CMB value, but reinforcing the tension with the local value. An analysis of possible sources of systematic errors shows that the stellar population synthesis model adopted may change the results significantly, being the main concern for subsequent studies. Forecasts for future data show that distant $H(z)$ measurements can be a robust method to determine $H_0$, where a focus in precision and a careful assessment of systematic errors are required. ", "introduction": "There is a strong tension, recently quantified by \\cite{verde2013}, between the value of the Hubble constant $H_0$ derived by {\\it Planck} \\citep{planck} from anisotropies in the cosmic microwave background (CMB): $67.3 \\pm 1.2$ km s$^{-1}$ Mpc$^{-1}$, and the value from local measurements: $73.8 \\pm 2.4$ km s$^{-1}$ Mpc$^{-1}$ \\citep{H0riess}. While the latter measurement is based on local measurements, the former infers a global value for the Hubble constant within a cosmological model. There remains disagreement about the local value of $H_0$ depending on the distance indicator used to measure it, which hints the discrepancy with {\\it Planck} could be the result of systematic errors. \\cite{H0riess} calibrated the SNe Ia distances with three indicators: distance to NGC 4258 based on a megamaser measurement, parallax measurements to Milk Way cepheids (MWC) and cepheids observations and a revised distance to the Large Magellanic Cloud (LMC). Contrarily, calibrating the SNe Ia with the tip of red-giant branch, \\cite{tammann} provides $H_0$ $=$ 63.7 $\\pm$ 2.3 km s$^{-1}$ Mpc$^{-1}$. This shows how crucial is the first-step calibration in the distance ladder to measure $H_0$. However, there are several local $H_0$ measurements with higher values. \\cite{riess2012} found $H_0$ $=$ 75.4 $\\pm$ 2.9 km s$^{-1}$ Mpc$^{-1}$ by using cepheids in M31. With a mid-infrared calibration for the cepheids, \\cite{freedman2012} derived $H_0$ $=$ 74.3 $\\pm$ 2.1 km s$^{-1}$ Mpc$^{-1}$, and with 8 new classical cepheids observed in galaxies hosting SNe Ia \\cite{fiorentino2013} got $H_0$ $=$ 76.0 $\\pm$ 1.9 km s$^{-1}$ Mpc$^{-1}$. By using HII regions and HII galaxies as distance indicators, \\cite{chavez2012} obtained $H_0$ $=$ 74.3 $\\pm$ 3.1(random) $\\pm$ 2.9 (syst.) km s$^{-1}$ Mpc$^{-1}$. Some of these are over 4$\\sigma$ away from the CMB-derived value. See Fig. \\ref{figH0} for a plot of different measurements of $H_0$. \\begin{figure} \\centerline{\\epsfig{figure=graf_values_H0.eps,width=0.5\\textwidth} \\hskip 0.1in} \\caption{{\\it Different measurements of $H_0$.} The figure shows how the result obtained in this work (GaPP) is compared to other determinations of $H_0$. The points refer to the following references: Planck \\citep{planck}, TR \\citep{tammann}, Efs1 \\citep{efstathiou} with one anchor, Efs3 with three anchors, R11 \\citep{H0riess}, R12 \\citep{riess2012}, Freed \\citep{freedman2012}, Fior \\citep{fiorentino2013} and Chavez \\citep{chavez2012}.} \\label{figH0} \\end{figure} A variety of different physical effects could explain such a discrepancy. It could just be cosmic variance: as we can observe the Universe from only one position, we are not able to realize the global parameters from the local parameters, as in the local expansion rate for instance. If we live in an locally underdense region, a ``Hubble bubble'', a higher value for $H_0$ is obtained compared to the global value. This effect was carefully addressed by \\cite{cosmic_variance} through a modelling of the statistics of matter distribution which provides the distribution of the gravitational potential at the observer. The outcome is that cosmic variance can alleviate the tension, but a complete elimination requires a very rare fluctuation \\citep{cosmic_variance,cosmic_variance2}. Another way to look at the problem is to consider that the discrepancy may indicate new physics, such as massive neutrinos \\citep{hu}, or alternative dark energy models \\citep{cde,xcdm}. Recently, some analyses were performed trying to identify sources of systematic errors in order to remove or alleviate the tension. For example, by using only the geometric maser distance to NGC 4258 of \\cite{humphreys} as an anchor, \\cite{efstathiou} revisited \\cite{H0riess} analysis and derived $H_0=70.6 \\pm 3.3$ km s$^{-1}$ Mpc$^{-1}$, while combining with LMC and MWC anchors the value is $72.5 \\pm 2.5$ km s$^{-1}$ Mpc$^{-1}$, alleviating the tension. The {\\it Planck} data were also reanalysed by \\cite{spergel}, where it was claimed that the 217 GHz $\\times$ 217 GHz detector is responsible for some part of the tension. Their new Hubble constant without the 217 GHz $\\times$ 217 GHz detector is slightly higher: $H_0 = 68.0 \\pm 1.1$ km s$^{-1}$ Mpc$^{-1}$. With so many alternatives, progress can be achieved by developing new ways to address the issue. We point out here that $H(z)$ data which are not calibrated on a $H_0$ estimate can be extrapolated to z=0 to provide an independent measurement of the global $H_0$. Here, the Hubble function is reconstructed in order to derive $H_0$ from 19 $H(z)$ measurements of passively evolving galaxies as cosmic chronometers \\citep{jimloeb}. Many of these are at relatively moderate and high redshifts so intrinsically probe the global value for $H_0$ rather than the local one. We use Gaussian Processes (GP), which is a non-parametric method, to obtain the value of the Hubble constant in a completely cosmological model-independent way, which is in principle not affected by the local systematics. We show the value of the Hubble constant derived in this way is lower than the standard local measurements. We obtain $H_0$ $=$ $64.9 \\pm 4.2$ km s$^{-1}$ Mpc$^{-1}$ $(1\\sigma)$, in agreement with the CMB-inferred value. A better understanding of systematic errors, especially the adopted stellar population synthesis model, is required: we show that to improve this result a big effort is necessary to decrease the errors substantially in future, and a focus on precision is worthier than the number of data. The paper is organized as follows: in Sec. \\ref{meth} we describe GP as well as standard parametric methods adopted to constrain $H_0$. In Sec. \\ref{const} the bounds derived for the Hubble constant are displayed, followed by forecasts of constraints in Sec. \\ref{fore}. We finish the paper in Sec. \\ref{conc} with the conclusions. ", "conclusions": "\\label{conc} We have applied GaPP, a non-parametric smoothing method based on Gaussian Processes, to 19 $H(z)$ measurements in order to constrain the Hubble constant $H_0$. This method does not rely on a cosmological model, so its results can be used to infer the impact of systematic errors as well as the underlying cosmological framework. We have obtained $H_0$ to be $64.9 \\pm 4.2$ km s$^{-1}$ Mpc$^{-1}$ $(1\\sigma)$, a value which is in agreement with {\\it Planck}, but in disagreement with local measurements. This supports the notion that either there are unidentified systematic errors in the local $H_0$ data, or the local value is indeed different from the global value. A better comprehension of systematic errors, especially a thorough analysis of the impact of SPS models, can improve the robustness of our results. Simulations have shown that improvements in distant $H(z)$ measurements can help pin down the global value of $H_0$." }, "1402/1402.0510_arXiv.txt": { "abstract": "We present results from near-infrared spectroscopy of 26 emission-line galaxies at $z\\sim2.2$ and $z\\sim1.5$ obtained with the Folded-port InfraRed Echellette (FIRE) spectrometer on the 6.5-meter Magellan Baade telescope. The sample was selected from the WFC3 Infrared Spectroscopic Parallels (WISP) survey, which uses the near-infrared grism of the Hubble Space Telescope Wide Field Camera 3 to detect emission-line galaxies over $0.3\\lesssim z \\lesssim2.3$. Our FIRE follow-up spectroscopy (R$\\sim$5000) over 1.0--2.5~$\\mu$m permits detailed measurements of physical properties of the $z\\sim2$ emission-line galaxies. Dust-corrected star formation rates for the sample range from $\\sim$5--100~$\\mathrm{M}_{\\odot}$~yr$^{-1}$ with a mean of 29~$\\mathrm{M}_{\\odot}$~yr$^{-1}$. We derive a median metallicity for the sample of 12~+~log(O/H)~=~8.34 or $\\sim$0.45~Z$_{\\odot}$. The estimated stellar masses range from $\\sim$$10^{8.5}-10^{9.5}$~M$_{\\odot}$, and a clear positive correlation between metallicity and stellar mass is observed. The average ionization parameter measured for the sample, log~$U\\approx-2.5$, is significantly higher than what is found for most star-forming galaxies in the local universe, but similar to the values found for other star-forming galaxies at high redshift. We derive composite spectra from the FIRE sample, from which we infer typical nebular electron densities of $\\sim$100-400~cm$^{-3}$. Based on the location of the galaxies and composite spectra on BPT diagrams, we do not find evidence for significant AGN activity in the sample. Most of the galaxies as well as the composites are offset in the BPT diagram toward higher [\\oiii]/\\hb\\ at a given [\\nii]/\\ha, in agreement with other observations of $z\\gtrsim1$ star-forming galaxies, but composite spectra derived from the sample do not show an appreciable offset from the local star-forming sequence on the [\\oiii]/\\hb\\ versus [\\sii]/\\ha\\ diagram. We infer a high nitrogen-to-oxygen abundance ratio from the composite spectrum, which may contribute to the offset of the high-redshift galaxies from the local star-forming sequence in the [\\oiii]/\\hb\\ versus [\\nii]/\\ha\\ diagram. We speculate that the elevated nitrogen abundance could result from substantial numbers of Wolf-Rayet stars in starbursting galaxies at $z\\sim2$. ", "introduction": "The rest-frame optical spectra of star-forming galaxies at all redshifts exhibit emission lines from which detailed physical properties can be inferred. For galaxies at the peak of cosmic star formation at $z\\sim2$, these emission lines are shifted into the near-infrared, which, combined with the intrinsic faintness of the sources, makes them difficult to observe from the ground. For this reason, relatively few near-infrared spectra of galaxies at $z\\sim2$ that cover all of the important rest-frame optical emission lines have been published to date (e.g., \\citealp{Erb06, Hainline09, Erb10, Rigby11, Belli13}). The available near-infrared spectra of star-forming galaxies at $z\\sim2$ have revealed differences in comparison with counterparts in the local universe \\citep{Liu08, Newman13}. For example, star-forming galaxies at $z\\sim2$ tend to have higher [\\oiii]/\\hb\\ ratios at a given [\\nii]/\\ha\\ ratio than local star-forming galaxies. This observation has been attributed to more extreme interstellar medium (ISM) conditions, on average, in galaxies at high redshift, possibly as a result of higher nebular electron densities, harder ionizing radiation fields, different gas volume filling factors, or some combination of these \\citep{Shapley05, Brinchmann08, Shirazi13, Kewley13a, Kewley13}. The clumpy morphology and relatively high velocity dispersions observed in many of these sources \\citep{Pettini01, Forster06, Genzel08, Law09} may support the conjecture that star-formation in the early universe generally occurs in denser and higher pressure environments than those found in local star-forming galaxies. Significant contribution of active galactic nuclei (AGN) to emission line fluxes for $z\\sim2$ galaxies has also been suggested as a source of the elevated line ratios \\citep{Trump11,Trump13}. The slitless grism spectroscopy provided by the Wide Field Camera 3 (WFC3) on the Hubble Space Telescope (HST) has enabled the discovery of large numbers of star-forming galaxies near the peak of cosmic star formation \\citep{Atek10, Atek11, Straughn11, Vanderwel11, Trump11, Brammer12}. Grism surveys such as the WFC3 Infrared Spectroscopic Parallels (WISP) survey \\citep{Atek10} are well-suited to finding low-mass star-forming galaxies at intermediate redshifts through their optical emission lines. While WFC3 grism spectroscopy detects large numbers of emission-line galaxies, it is not ideal for extracting the physical information encoded in their optical spectra. The spectral resolution ($R\\sim130$ in G141 and $R\\sim210$ in G102) is insufficient to resolve \\ha\\ from [\\nii]$\\lambda$6548, 6583, or detect line broadening due to AGN activity. Moreover, despite the broad wavelength coverage ($\\sim$0.8--1.7~$\\mu$m) of the grism, \\ha\\ is not detected for galaxies at $z\\gtrsim1.6$ and important metallicity diagnostics such as $R_{23}$~$\\equiv$~([\\oiii]$\\lambda4959,5007$+[\\oii]$\\lambda$3727)/H$\\beta$ are often inaccessible. For these reasons, ground-based spectroscopy with the new generation of near-infrared spectrometers is required to constrain the physical properties of these galaxies. Here we present rest-frame optical spectroscopy of 26 emission-line galaxies from WISP obtained with the Folded-port Infrared Echellette (FIRE, \\citealp{Simcoe08, Simcoe10}) on the Magellan Baade 6.5 meter telescope. The sample consists of 13 sources at $z\\sim2.2$ and 13 at $z\\sim1.5$. Galaxies in the sample were selected from the WISP survey based on the detection of strong [\\oiii]$\\lambda5007$ emission (at $z\\sim2.2$) or \\ha\\ (at $z\\sim1.5$) in the G141 grism data. Follow-up near-IR spectroscopy with FIRE enables us to: (1) detect \\ha\\ for sources at $z>1.6$ and split \\ha\\ and [\\nii] in order to determine star formation rates (SFRs), (2) get accurate dust reddening estimates using the Balmer decrement, (3) infer metallicities and ionization parameters from strong lines, (4) measure diagnostic line ratios to test for nuclear activity, and (5) resolve emission line velocity dispersions. We also construct composite spectra from our sample, which allow us to investigate the statistical properties of strongly star-forming galaxies at $z\\sim2.2$ and $z\\sim1.5$ in greater detail. This paper is structured as follows. In \\S2 we provide an overview of the WISP survey and the emission-line galaxies selected for follow-up spectroscopy. In \\S3 we discuss the Magellan FIRE spectroscopy and data reduction. In \\S4 we present the physical properties measured from the FIRE spectra, including dust obscuration, metallicity, ionization parameter, star formation rate and kinematics. In \\S5 we discuss the composite spectrum derived from the sample. In \\S6 we analyze the results, exploring the implications of our findings for the nature of starbursting galaxies at $z\\sim2$. In \\S7 we conclude with a discussion. We adopt a cosmology with $\\Omega_{M}=0.3$, $\\Omega_{\\Lambda}=0.7$ and $H_{0}=70~\\mathrm{km~s}^{-1}~\\mathrm{Mpc}^{-1}$. ", "conclusions": "We have presented rest-frame optical spectra taken with Magellan FIRE of a sample of 26 emission-line galaxies at $z\\sim1.5$ and $z\\sim2.2$ selected from the HST-WFC3 grism spectroscopy of the WISP survey. The FIRE spectra provide significant additional information about these galaxies, allowing us to measure metal abundances, dust reddening, kinematics, star formation rates and important diagnostic line ratios. We also created composite spectra that reveal further details about the average properties of the sample. Direct WFC3 imaging shows the sources to be compact, with many ($\\gtrsim$40\\%) exhibiting clumpiness or asymmetry. The galaxies are low mass ($\\sim$10$^{8.5}$--10$^{9.5}$~M$_{\\odot}$), with a median metallicity of 0.45~Z$_{\\odot}$. A clear mass-metallicity relation is found for the sample. As seen in other high-redshift samples, the ionization parameters of the galaxies in our sample are significantly higher, on average, than those found for typical star-forming galaxies in the local universe. The velocity dispersions of the emission lines for the sample range from $\\sim$50--200~km~s$^{-1}$ but usually are around 75~km~s$^{-1}$. Assuming that the dispersions are almost entirely due to gravitational motions (which is probably not the case), this implies that stars make up roughly 30\\% of the dynamical mass of the galaxies. The composite spectra we generated from the FIRE sample reveal more detail about the average properties of emission-line galaxies at $z\\sim2.2$ and $z\\sim1.5$. From the overall composite we infer an average electron density for the star-forming regions of 100-400~cm$^{-3}$ and put a limit on the average temperature in these regions of $T_{e}<16,800$~K from the non-detection of [\\oiii]$\\lambda4363$. The locations of galaxies and composites on the BPT diagnostic diagrams favor a starburst classification, and we find little evidence for substantial AGN contribution. Our sample includes galaxies with unusually high [\\oiii]/\\ha\\ and [\\oiii]/\\hb\\ line ratios that are not reproduced with existing photoionization models. Composite spectra sorted on the ratio [\\oiii]/\\ha\\ indicate that those sources with high [\\oiii]/\\ha\\ ratios show stronger [\\oiii]$\\lambda$5007 and [\\neiii]$\\lambda$3869 emission relative [\\oii]$\\lambda$3727, indicating higher ionization parameters. The high [\\oiii]/(\\ha,\\hb) sources tend to be lower metallicity, which seems to be the main factor driving their higher ionization parameters. Reproducing the line ratios observed in such objects may require modifications to photoionization models to better account for the star-forming conditions in high redshift galaxies. The well-known offset of high redshift star-forming galaxies from the local star-forming sequence in the [\\oiii]/\\hb\\ versus [\\nii]/\\ha\\ diagnostic diagram is observed in our sample, but we do not find a similar offset in the [\\oiii]/\\hb\\ versus [\\sii]/\\ha\\ diagram. We therefore investigated the possibility that the offset of high-redshift galaxies in the [\\oiii]/\\hb\\ versus [\\nii]/\\ha\\ diagram may result, at least in part, from elevated nitrogen abundances. The composite FIRE spectrum was used to infer the N/O abundance ratio, which we find to be $\\sim$0.4 dex higher in comparison with local galaxies of similar metallicity. We speculate that an elevated nitrogen abundance in high redshift star-forming galaxies may be common, and could be explained by the presence of a substantial population of Wolf-Rayet stars embedded in super star clusters in such galaxies." }, "1402/1402.5932_arXiv.txt": { "abstract": "Astronomy produces extremely large data sets from ground-based telescopes, space missions, and simulation. The volume and complexity of these rich data sets require new approaches and advanced tools to understand the information contained therein. No one can load this data on their own computer, most cannot even keep it at their institution, and worse, no platform exists that allows one to evaluate their models across the whole of the data. Simply having an extremely large volume of data available in one place is not sufficient; one must be able to make valid, rigorous, scientific comparisons across very different data sets from very different instrumentation. We propose a framework to directly address this which has the following components: a model-based computational platform, streamlined access to large volumes of data, and an educational and social platform for both researchers and the public. ", "introduction": "The era of big data is upon us. Existing astronomical surveys like the Sloan Digital Sky Survey (SDSS), with a data volume of $\\sim$15 TB, already stretch the limits of what our analysis methods and software are capable of. Next-generation surveys, such as the Large Synoptic Survey Telescope (LSST), will generate an SDSS every night \\citep{2008arXiv0805.2366I}. How is a research astronomer to cope? The mishmash of mismatched survey outputs from data sets documented to varying degrees with differing catalog conventions, clunky web forms, small departmental computing clusters, and roll-your-own analysis software already wastes enormous quantities of time that could be better spent on actual science: building and testing models, looking for unexplained features, and using data visualization to build physical intuition and motivate new understanding. We present here a new framework that addresses the major challenges of doing astronomy with big data. The framework can be roughly described as having three primary components: a model-based computational platform, streamlined access to large volumes of data, and an educational and social platform for both researchers and the public. It will let theorists compare physical models to data from a wide variety of astronomical surveys without having to understand for each survey the point-spread functions, photometric calibration, or the effect of aperture size choice on photometry. It will allow researchers to write code to run on data as if it were local, without the hassle of managing hundreds of terabytes of local data volume. Our framework will attach quantitative models and their associated fit likelihoods to astronomical objects, letting researchers ask questions like ``How many stars have colors or spectra that look like my model?''~or ``What things in the universe are unexplained by existing models?''~with straightforward database queries. Associating quantitative models, their constituent code, and their detailed descriptions with these objects creates a mechanism to search for and compare with the relevant literature, and it provides a path into the subject for educators and the general public. We will start by identifying current problems in data analysis, offer a high-level overview of the proposed framework, then describe the implementation of the framework in greater detail. Astronomical data sets and models are used as examples to present the ideas in a more concrete manner as this is the primary domain of the authors, but nothing presented here need be specific to astronomy. This framework as offered can be extended such that the data sets and models of many fields may overlap in new and novel ways. While pieces of this framework, such as designing large databases, data visualization, etc.~are certainly being developed in fields such as astrophysics, computer science, and human interface design, the approach described here fully integrates these and other ideas to enable new and novel scientific research and education. ", "conclusions": "We have presented here a framework designed to enable a new way to do science, taking advantage of the large volumes of data now available (and to come) in a way that hasn't been possible before, going beyond existing interfaces to data. The core concepts are the creation and integration of an astronomical data repository, the cross referencing of this data that understands and incorporates instrumental effects, transparent and simple programmatic access to the data, the ability to apply an arbitrary theoretical model on a computing platform, and an integrated education tool for researchers and the public alike. Managing access to increasingly larger data sets is necessary to enable us to ask the important questions. Do we understand the data? How accurate are our models? Do our models hold up with new data? What is in our data that does not fit any of our models? This last question is one of the most compelling that we look forward to addressing with this framework." }, "1402/1402.4808_arXiv.txt": { "abstract": "A recent cross-correlation between the SDSS DR7 White Dwarf Catalog with the Wide-Field Infrared Survey Explorer (\\emph{WISE}) all-sky photometry at 3.4, 4.6, 12, and 22 microns performed by \\cite{debes11} resulted in the discovery of 52 candidate dusty white dwarfs (WDs). The 6$\\arcsec$ \\emph{WISE} beam allows for the possibility that many of the excesses exhibited by these WDs may be due to contamination from a nearby source, however. We present MMT$+$SWIRC $J$- and $H$-band imaging observations (0.5-1.5$\\arcsec$ PSF) of 16 of these candidate dusty WDs and confirm that four have spectral energy distributions (SEDs) consistent with a dusty disk and are not accompanied by a nearby source contaminant. The remaining 12 WDs have contaminated \\emph{WISE} photometry and SEDs inconsistent with a dusty disk when the contaminating sources are not included in the photometry measurements. We find the frequency of disks around single WDs in the \\emph{WISE} $\\cap$ SDSS sample to be 2.6-4.1\\%. One of the four new dusty WDs has a mass of $1.04~M_{\\odot}$ (progenitor mass $5.4~M_{\\odot}$) and its discovery offers the first confirmation that massive WDs (and their massive progenitor stars) host planetary systems. ", "introduction": "A new class of dusty white dwarfs (WDs) is rapidly being populated through infrared excess searches using \\emph{Spitzer}, the Wide-Field Infrared Survey Explorer (\\emph{WISE}; \\citealt{wright10}), and various ground based telescopes such as the IRTF and Gemini (\\citealt{zuckerman87,kilic05,becklin05,vonhippel07,jura07a,farihi10,debes11,barber12,xu12}). These WDs exhibit excess emission in the near- and mid-infrared due to the thermal reprocessing of light by a disk of circumstellar dust. These dust disks are the debris resultant from the tidal disruption of an asteroid that has veered off its orbit and passed within the WD's Roche lobe (\\citealt{debes02, jura03,jura08,bonsor11,debes12}). The origin of these asteroids is dynamical instability, initiated by post-main sequence stellar mass loss, whence a massive planet (if present) will start to gravitationally interact with smaller planetary bodies. After numerous interactions, an asteroid's orbit will become increasingly eccentric and in some cases approach close enough to the WD to be ripped apart by tidal forces. The debris produced by this disruption embodies a circular disk geometry after many subsequent orbits (\\citealt{debes12}). Dust persists around a WD inside the tidal radius of the WD and outside the radius at which the equilibrium temperature is such that the dust will sublimate. Viscous torques cause the sublimated dust to accrete onto WD's surface (\\citealt{rafikov11a,rafikov11b,veras13}). This accretion results in a spectroscopically detectable pollution of the otherwise pristine WD atmosphere. Photospheric abundance analyses of these WDs show that the accreted metals originate from tidally disrupted minor bodies similar in composition to that of bulk Earth (\\citealt{zuckerman07,klein10,klein11,dufour10,dufour12,xu14}). Since at least one planet is required to perturb minor bodies out of their stable orbits (\\citealt{debes12}), photospheric pollution as well as circumstellar debris disks serve as tracers for remnant planetary systems at WDs. Dusty disks surrounding WDs are generally assumed to be geometrically flat and optically thick (\\citealt{jura03,rafikov11a,rafikov11b}). Based on flat disk models, they vary in width from a narrow ring, a few tenths of a solar radius, to a disk filling the entire region interior to the tidal radius of the star and exterior to the radius of dust sublimation. The WDs known to host debris disks typically range in temperature from 9500-24,000 $K$ and in mass from 0.5-0.9 $M_{\\odot}$. With this work, we expand the parameter space occupied by WD+disk systems with the confirmation of the first disk orbiting a WD hotter than 24,000 $K$ and the first disk orbiting a WD more massive than 1 $M_{\\odot}$. We present follow up near-infrared (NIR) $J$- and $H$-band photometry using the 6.5 m MMT with SWIRC of 16 WDs exhibiting a mid-infrared excess indicative of a debris disk detected by \\emph{WISE} (\\citealt{debes11}). We use the higher spatial resolution of SWIRC to confirm the presence of a debris disk or to identify a photometric contamination in the six arcsecond \\emph{WISE} beam. We start in Section \\ref{target} with the selection of our targets. Section \\ref{obs} is a description of the observations and data reduction as well as the debris disk modeling of the detected excesses. Finally, we present our results in Section \\ref{results}. ", "conclusions": "We have confirmed the presence of dust surrounding four WDs with excess flux detected by \\emph{WISE}. These four new dusty WDs, including the two bonafide UKIDSS$+$WISE disks, increases the total number of confirmed WD$+$disk systems from 29 to 35 (\\citealt{farihi09,xu12,hoard13}). The discovery of these four new debris disks enriches the current dusty WD population with one of the coolest (J1507), hottest (J1537), and the most massive (J1234) WDs known to host circumstellar dust. Expanding the parameter space known to be hospitable to circumstellar dust at WD stars will not only guide future infrared excess searches, but also enhance our understanding of the formation and evolution of these remnant planetary systems." }, "1402/1402.6385_arXiv.txt": { "abstract": "\\noindent We present the results of a time series analysis of the long-term radio lightcurves of four blazars: 3C 279, 3C 345, 3C 446, and BL Lacertae. We exploit the data base of the University of Michigan Radio Astronomy Observatory (UMRAO) monitoring program which provides densely sampled lightcurves spanning 32 years in time in three frequency bands located at 4.8, 8, and 14.5\\,GHz. Our sources show mostly flat or inverted (spectral indices $-0.5\\lesssim\\alpha\\lesssim0$) spectra, in agreement with optically thick emission. All lightcurves show strong variability on all time scales. Analyzing the time lags between the lightcurves from different frequency bands, we find that we can distinguish high-peaking flares and low-peaking flares in accord with the classification of \\cite{Valtaoja}. The periodograms (temporal power spectra) of the observed lightcurves are consistent with random-walk powerlaw noise without any indication of (quasi-)periodic variability. The fact that all four sources studied are in agreement with being random-walk noise emitters at radio wavelengths suggests that such behavior is a general property of blazars. ", "introduction": "} \\noindent The strong and complex temporal flux variability of Active Galactic Nuclei (AGN; see, e.g., \\citealt{Beckmann} and references therein for a recent review) provides valuable information on the internal conditions of accretion zones and plasma outflows. Various characteristic variability patterns have been associated with a wide range of physical phenomena, from shocks in continuous (e.g., \\citealt{Marscher}) or discontinuous (e.g., \\citealt{spada2001}) jets to orbiting plasma ``hotspots'' (e.g., \\citealt{abramowicz1991}) or plasma density waves (e.g., \\citealt{kato2000}) in accretion disks. Accordingly, multiple studies have aimed at quantifying the properties of AGN variability on all time scales and throughout the electromagnetic spectrum. At radio wavelengths, variability time scales probed by observations range from tens of minutes (\\citealt{Schodel}, studying the mm/radio lightcurve of M 81*; see also \\citealt{Kim2013} for a discussion of the detectability of intra-day variability) to tens of years (\\citealt{Hovatta2007}, in a statistical analysis of the long-term flux variability of 80 AGN). Of particular interest is the possible presence of quasi-periodic oscillations (QPO) which has been reported by several studies of blazar lightcurves (e.g. \\citealt{Rani2009, Rani2010, Gupta2012}). Fourier transform, period folding, power spectrum, and periodogram methods (cf. \\citealt{priestley1981} for an exhaustive review of time series analysis) have been used extensively for quantifying the statistical properties of AGN variability and for the search for possible QPOs (e.g. \\citealt{Benlloch} for X-ray, \\citealt{Webb} for optical, \\citealt{Fan1999} for near infrared, and \\citealt{Aller} for radio observations). As already noted by \\citet{Press}, power spectra of AGN lightcurves globally follow power laws $A_f{\\propto}f^{-\\beta}$ with $\\beta>0$, corresponding to \\emph{red noise};\\footnote{In the context of time series analysis, the term ``noise'' refers to stochastic emission from a source of radiation, \\emph{not} to measurement errors or instrumental noise.} here $A_f$ denotes the power spectral amplitude as function of sampling frequency $f$. Lightcurves composed of pure Gaussian \\emph{white noise} have flat power spectra ($\\beta=0$). Other important special cases are \\emph{random walk noise} ($\\beta = 2$) -- which is the integral of white noise -- and \\emph{flicker noise} ($\\beta = 1$, ``$1/f$ noise'') as intermediate case between white noise and random walk noise (see also \\citealt{Park2012} for a detailed technical discussion). \\citet{Lawrence1987} found that the power spectrum of the Seyfert galaxy NGC 4051 can be described by flicker noise. Red noise power spectra were also observed by \\cite{Lawrence1993} who used 12 high-quality ``long look'' X-ray lightcurves of AGN. \\begin{deluxetable*}{ccccccccc} \\tablecaption{Properties of our four target blazars \\label{Information}} \\tablehead{ \\colhead{Object} & \\colhead{RA} & \\colhead{DEC} & \\colhead{Type} & \\colhead{Redshift} & \\colhead{$T$ [yr]} & \\colhead{$N_{4.8}$} & \\colhead{$N_{8.0}$} & \\colhead{$N_{14.5}$} } \\startdata 3C 279 & 12:56:11 & $-$05:47:22 & FSRQ & 0.536 & 32.52 & 1086 & 1337 & 1473 \\\\ 3C 345 & 16:42:59 & +39:48:37 & FSRQ & 0.593 & 32.54 & 1323 & 1315 & 1415 \\\\ 3C 446 & 22:25:47 & $-$04:57:01 & BL Lac & 1.404 & 32.12 & 680 & 902 & 1088 \\\\ BL Lac & 22:02:43 & +42:16:40 & BL Lac & 0.069 & 32.54 & 1256 & 1315 & 1755 \\enddata \\tablecomments{J2000 coordinates, source types, and redshifts are taken from the NED. We also give the total monitoring time $T$ (in years) and the numbers $N$ of flux data points for 4.8, 8.0, and 14.5\\,GHz, respectively.} \\end{deluxetable*} A multitude of studies illustrates the difficulties of determining the statistical significance of supposed QPO signals in the power spectra of AGN lightcurves. The analysis of \\cite{Benlloch} concluded that a previously reported quasi-periodic signal in X-ray lightcurves of the Seyfert galaxy Mrk 766 was actually statistically insignificant. \\cite{Uttley} pointed out the importance of sampling effects leading to \\emph{red-noise leaks} and \\emph{aliasing}. \\cite{Vaughan} gives an analytical approach to derive significance levels for peaks in red-noise power spectra. \\cite{Do} demonstrated the power of Monte-Carlo techniques for deriving significance levels by comparing the red-noise power spectra of actual and simulated flux data. The temporal flux variability of AGN can be exploited for elucidating the physical conditions within active galaxies especially at radio frequencies where monitoring observations of hundreds of targets have been conducted over several decades by various observatories. Remarkably, many studies aimed at analyzing long-term AGN radio variability do not take into account the intrinsic red-noise properties of the lightcurves. The incorrect assumption of constant (as function of $f$) significance levels in power spectra (following from the assumption of white-noise dominated lightcurves) has lead to reports of ``characteristic'' time scales which are actually not special at all (cf., e.g., \\citealt{Ciaramella, Hovatta2007, Nieppola}). Blazars, characterized by violent flux variability across the entire electromagnetic spectrum, are a subset of AGN which include BL Lacertae (BL Lac) objects and Flat Spectrum Radio Quasars (FSRQs). In accordance with the standard viewing angle unification scheme of AGN \\citep{Urry}, it is commonly assumed that their observed emission is generated by synchrotron radiation -- dominating from radio to optical frequencies -- and inverse Compton emission -- dominating at frequencies higher than optical -- from relativistic plasma jets (almost) aligned with the line of sight. In order to perform a thorough study of the statistical properties of blazar emission, we analyze the lightcurves of four radio-bright blazars with strong flux variability -- 3C 279, 3C 345, 3C 446, and BL Lac -- provided by the University of Michigan Radio Astronomy Observatory (UMRAO) monitoring program of AGN. The data set comprises data spanning $\\approx$32 years in time and covering three frequency bands located at 4.8\\,GHz, 8.0\\,GHz, and 14.5\\,GHz. ", "conclusions": " \\begin{enumerate} \\item Our sources show mostly flat or inverted ($-0.5\\lesssim\\alpha\\lesssim0$) spectral indices, in agreement with optically thick synchrotron emission. The lightcurves of different frequencies are either simultaneous (within errors) or shifted relative to each other such that the high-frequency emission leads the low-frequency emission by up to $\\approx$1.5 years. We are able to distinguish high-peaking and low-peaking flares according to the classification of \\cite{Valtaoja}. \\item All lightcurves show variability on all time scales. Their periodograms (power spectra) are in agreement with being pure red-noise powerlaw spectra without any indication for (quasi-)periodic signals. When taking into account the sampling patterns via dedicated Monte Carlo simulations, we find that all lightcurves are consistent with being random walk noise signals with powerlaw slopes $\\beta\\approx2$. Given that we find this behavior in all four sources under study, this suggests that random walk noise lightcurves are a general feature of blazars. \\end{enumerate} Our results imply that careful time series analysis of high-quality blazar lightcurves provides information on the source structure even if a target is not resolved spatially. Obviously, it will be necessary to systematically study much larger blazar samples in order to decide if the trends we have uncovered are indeed general." }, "1402/1402.5948_arXiv.txt": { "abstract": "Fourier analysis of the light curve of AC And from the HATNet database reveals the rich frequency structure of this object. Above $30$ components are found down to the amplitude of $3$~mmag. Several of these frequencies are not the linear combinations of the three basic components. We detect period increase in all three components that may lend support to the Pop I classification of this variable. ", "introduction": " ", "conclusions": "" }, "1402/1402.2103_arXiv.txt": { "abstract": "We present new observations of the XZ~Tau system made at high angular resolution (55\\,milliarcsec) with the Karl G.\\ Jansky Very Large Array (VLA) at a wavelength of 7\\,mm. Observations of XZ~Tau made with the VLA in 2004 appeared to show a triple-star system, with XZ~Tau~A resolved into two sources, XZ~Tau~A and XZ~Tau~C. The angular separation of XZ~Tau~A and C (0.09\\,arcsec) suggested a projected orbital separation of around 13\\,AU with a possible orbital period of around 40\\,yr. Our follow-up observations were obtained approximately 8\\,yr later, a fifth of this putative orbital period, and should therefore allow us to constrain the orbital parameters of XZ~Tau~C, and evaluate the possibility that a recent periastron passage of C coincided with the launch of extended optical outflows from XZ~Tau~A. Despite improved sensitivity and resolution, as compared with the 2004 observations, we find no evidence of XZ~Tau~C in our data. Components A and B are detected with a signal-to-noise ratio greater than ten; their orbital motions are consistent with previous studies of the system, although the emission from XZ~Tau~A appears to be weaker. Three possible interpretations are offered: either XZ~Tau~C is transiting XZ~Tau~A, which is broadly consistent with the periastron passage hypothesis, or the emission seen in 2004 was that of a transient, or XZ~Tau~C does not exist. A fourth interpretation, that XZ~Tau~C was ejected from the system, is dismissed due to the lack of angular momentum redistribution in the orbits of XZ~Tau~A and XZ~Tau~B that would result from such an event. Transients are rare but cannot be ruled out in a T~Tauri system known to exhibit variable behaviour. Our observations are insufficient to distinguish between the remaining possibilities, at least not until we obtain further VLA observations at a sufficiently later time. A further non-detection would allow us to reject the transit hypothesis, and the periastron passage of XZ~Tau~C as agent of XZ Tau A's outflows. ", "introduction": "XZ~Tau is a binary system composed of a T~Tauri star, XZ~Tau~A, with a cool companion, XZ~Tau~B, separated by approximately 0.3\\,arcsec, at a distance of approximately 140\\,pc from Earth \\citep{Haas1990,Kenyon1994,Torres2009}. Like many other T~Tauri stars, XZ~Tau~A drives collimated jets and optical outflows \\citep{Mundt1990,Krist1997}. {\\it Hubble Space Telescope} imaging of these outflows shows nebular emission in the shape of an elongated bubble with expansion velocities of around 70\\,$\\mathrm{km\\,s^{-1}}$ \\citep{Krist1999}. The substructure displayed by the bubble suggests its driver is episodic, with the cause attributed to a velocity pulse in the jet of XZ~Tau~A, triggered in the early 1980s \\citep{Krist2008}. These previous studies, particularly that of \\citeauthor{Krist2008}, explored the possibility that the periastron passage of XZ~Tau~B could have caused the outflows (cf.\\ \\citealt{enc_outburst}). However, this would require an eccentric orbit, which is inconsistent with observations of the A/B system, which instead suggest a circular, face-on orbit. Also, the periastron passage of XZ~Tau~B would have occurred in the 1950s, too early to cause the outflow. The periastron passage hypothesis was revived by more recent observations of the XZ~Tau system using the Very Large Array (VLA) by \\citet{Carrasco-Gonzalez2009}. Observations of the 7-mm continuum resolved XZ~Tau~A into two components, with the new component, XZ~Tau~C, separated by around 0.09\\,arcsec (13\\,AU). The non-detection of XZ~Tau~C in the optical waveband was suggestive of a stellar object heavily embedded in a dusty envelope or disk. While a single detection yielded no information on the orbit of component C around A, the existence of XZ~Tau~C increased the likelihood of a close approach to XZ~Tau~A, making it a potential trigger for the outflow. \\citet{Carrasco-Gonzalez2009} speculate that if the orbit of XZ~Tau~C is circular, and the total system mass is $\\approx$1\\,M$_\\odot$, then the orbital period of XZ~Tau~C should be around 40\\,yr. If the A/C system was close to apastron at the epoch of detection (2004), this would place XZ~Tau~C at periastron in the 1980s, as required, and a further ejection from XZ~Tau~A could be expected around 2020. However, the data available to the authors was insufficient to confirm orbital parameters, and as such this explanation for the outflows was tentative. Further observations of XZ~Tau~C at sufficiently later epochs would be required to either confirm or refute the periastron passage model for outflow generation. To this end, we observed the XZ~Tau system at high angular resolution, using the newly upgraded Karl G.\\ Jansky VLA at 7\\,mm, to confirm the existence of XZ~Tau~C and constrain its orbital parameters. The time interval between our observations (2012) and the previous observations (2004) corresponds to approximately one fifth of the potential orbital period of XZ~Tau~C. These observations yield no detection of XZ~Tau~C, and the positions of XZ~Tau~A and B are consistent with the orbital solutions presented by previous studies. This Letter is composed as follows: we describe the observations taken in \\S\\ref{sec:observations}; we discuss the results in \\S\\ref{sec:results}, and we summarise the work in \\S\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} \\noindent We report observations of the XZ~Tau system, recently observed to possess a third component, XZ~Tau~C \\citep{Carrasco-Gonzalez2009}. This third component had been proposed as a potential driver for the outflows generated by XZ~Tau~A, which could have been generated by a periastron passage of this new object in the 1980s. Our new observations were made using the VLA in its most extended configuration, after a period of roughly one fifth of the object's assumed orbital period, and can thereby constrain the orbit of XZ~Tau~C and test the periastron passage theory. However, our observations yield no detection of XZ~Tau~C. XZ~Tau~A and B are both detected with a high degree of confidence, with positions and orbits consistent with those described in the literature. Three potential interpretations of the data are possible. Of these, the most prosaic is that XZ~Tau~C does not exist, and was an artefact of reducing interferometric data, which is consistent with its non-detection in the optical. Alternatively, XZ~Tau~C may have been caught whilst transiting XZ~Tau~A, which is consistent with the orbital requirements for XZ~Tau~C to trigger outflows at periastron passage. This second interpretation would require the hierarchical triple to possess a large mutual inclination, and hence be dynamically unstable, but on timescales that are sufficiently long to remain consistent with the observations. Finally, we cannot rule out the possibility of a transient event local to this T~Tauri system, a system known to be variable. To distinguish between these potential interpretations, determining the nature of XZ~Tau~C, requires a very brief, future observation using the VLA in A~configuration. If the transit hypothesis is true, future observations should be able to detect XZ~Tau~C once it has moved away from XZ~Tau~A. If observations at this time fail to detect XZ~Tau~C, this would be sufficient to reject it as a cause of outflows being generated by XZ~Tau~A." }, "1402/1402.2784_arXiv.txt": { "abstract": "{ We investigate the Euclidean vacuum mode functions of a massive vector field in a spatially open chart of de Sitter spacetime. In the one-bubble open inflationary scenario that naturally predicts a negative spatial curvature after a quantum tunneling, it is known that a light scalar field has the so-called supercurvature mode, i.e. an additional discrete mode which describes fluctuations over scales larger than the spatial curvature scale. If such supercurvature modes exist for a vector field with a sufficiently light mass, then they would decay slower and easily survive the inflationary era. However, the existence of supercurvature mode strongly depends on details of the system. To clarify whether a massive vector field has supercurvature modes, we consider a U(1) gauge field with gauge and conformal invariances spontaneously broken through the Higgs mechanism, and present explicit expressions for the Euclidean vacuum mode functions. We find that, for any values of the vector field mass, there is no supercurvature mode. In the massless limit, the absence of supercurvature modes in the scalar sector stems from the gauge symmetry. } \\begin{document} ", "introduction": "Recent observational data provide a strong support of the existence of extragalactic magnetic fields, in the range of $\\mcO (10^{-14}$-$10^{-20})\\,{\\rm G}$ on Mpc scales~\\cite{Neronov:1900zz,Tavecchio:2010mk,Dermer:2010mm,Huan:2011kp,Dolag:2010ni,Essey:2010nd,Taylor:2011bn,Vovk:2011aa,Takahashi:2011ac,Finke:2013bua}\\,. The generation of the magnetic field in high-redshift galaxies, clusters, and even in empty intergalactic region is still an unresolved problem in cosmology. No promising astrophysical process to generate the sufficient amount of the magnetic field on the large scales are known. As for the inflationary magnetogenesis, though the various mechanism are proposed, the several difficulties such as the strong coupling problem, the backreaction problem and the curvature perturbation problem in some specific models prevent successful production of magnetic field~\\cite{Demozzi:2009fu, Barnaby:2012xt, Fujita:2013qxa}. Actually both upper and lower limits on the inflation energy scale can be derived from these problems in model independent ways and the limits are considerably severe if the extragalactic magnetic fields are stronger than $10^{-16}$G at present~\\cite{Fujita:2012rb,Suyama:2012wh,Fujita:2014sna}. Thus it is known to be very difficult to generate the magnetic field in the context of the inflationary magnetogenesis on the flat Friedmann-Lema\\^itre-Robertson-Walker (FLRW) universe. The superadiabatic growth of the magnetic fields in the open FLRW universe has been discussed in the literatures~\\cite{Barrow:2011ic,Barrow:2012ty}. The authors of these literatures assumed the existence of supercurvature modes of the magnetic field, which describes the fluctuations with the wavelength exceeding the spatial curvature scale. If a supercurvature mode exists, it decays slower than $1/a^2$\\,, where $a$ corresponds to the conventional scale factor of a FLRW universe, and can easily survive the inflationary era. Hence the relatively large amount of the magnetic field on supercurvature scales would remain at late time. However, the existence of supercurvature modes of magnetic fields is non-trivial and should be critically studied. Adamek {\\it et al.}~\\cite{Adamek:2011hi} recently pointed out that the equations of motion of a U(1) gauge field with unbroken conformal and gauge symmetries can be rewritten in the form that is identical to those of massive scalar fields for which there is no supercurvature mode~\\footnote{Rigorously speaking, the proof of the absence of supercurvature modes requires knowledge of not only the equation of motion but also a Klein-Gordon norm and proper boundary conditions. The present paper fills those gaps for the analysis of the massless vector field, although our main focus will be on a massive vector field.}. The purpose of the present paper is to investigate whether supercurvature modes exist for a massive vector field, in both scalar and vector sectors of the physical spectrum. To be specific, we consider a U(1) gauge field with both gauge and conformal symmetries spontaneously broken through the Higgs mechanism. As for the background geometry, we consider a de Sitter spacetime in the open chart. This is relevant to the one-bubble open inflation scenario that naturally predicts the spatially negative curvature universe. While the recent observational data show that the universe is almost exactly flat with accuracy of about $1\\%$\\,, $|1-\\Omega_0 |\\leq 10^{-2}$~\\cite{Ade:2013zuv}\\,, open inflation scenario is attracting a renewed interest in the context of the string landscape scenario~\\cite{Susskind:2003kw,Freivogel:2004rd}. There are a huge number of metastable de Sitter vacua and the tunneling transition generally occurs through the nucleation of a true vacuum bubble in the false vacuum background. Because of the symmetry of the instanton solution, a bubble formed by the Coleman-De Luccia (CDL) instanton~\\cite{Coleman:1977py,Coleman:1980aw} looks like an infinite open universe from the viewpoint of an observer inside. If the universe experienced a sufficiently long inflation after the bubble nucleation, then the universe becomes almost exactly flat and subsequently evolves as a slightly open FLRW universe. This leads to a natural realization of one-bubble open inflation (see e.g. \\cite{Sasaki:1994yt,Yamamoto:1996qq,Tanaka:1997kq,Garriga:1998he,Garriga:1997wz}) and can be tested against observations~\\cite{Yamauchi:2011qq,Sugimura:2012kr,Sugimura:2013cra}. This paper is organized as follows. We first illustrate the background spacetime in section \\ref{sec:Background}. In section \\ref{sec: KG norm}\\,, we expand the U(1) gauge field by harmonic functions and write down the reduced action for the even and odd modes of the U(1) gauge field. In order to investigate the existence/absence of supercurvature modes, we show the quantization conditions for the even and odd modes on a Cauchy surface. With the obtained normalization conditions, we then analyze whether the supercurvature modes, which are normalizable on the Cauchy surface, exist in section \\ref{sec: Mode functions}. In section \\ref{sec: massless limit}\\,, as a consistency check, we explicitly calculate the Wightman function in the decoupling limit by using the (subcurvature) mode functions derived in section \\ref{sec: KG norm}. It is shown that the correct expression for the Euclidean Wightman function is recovered in the decoupling limit without need for any supercurvature modes. Finally, section \\ref{sec:summary} is devoted to a summary and discussions. ", "conclusions": "\\label{sec:summary} In this paper, we have investigated the Euclidean vacuum mode functions of a massive vector field in the spatially open chart of de Sitter spacetime. In order to clarify whether supercurvature modes exist, we have studied the U(1) gauge field with gauge and conformal symmetries spontaneously broken through the Higgs mechanism. We have found that there is no supercurvature mode for both the even and odd parity sectors. This implies that it is difficult to generate the sufficient amount of the magnetic field on large scales by using the superadiabatic growth within the one-bubble open inflation scenario even if the Higgs mechanism spontaneously breaks gauge and conformal invariances. Utilizing the obtained mode functions, we have explicitly computed the Wightman function of the U(1) gauge field in terms of the coordinates in the open chart of the de Sitter spacetime, and have compared it with one obtained by other methods. It was found that the leading-order Wightman function in the decoupling ($e\\to 0$) limit is correctly reproduced by the sum of the products of the subcurvature modes without need for introducing supercurvature modes. In consequence we have verified that the supercurvature mode is not needed as a part of a complete set of mode functions of the U(1) gauge field in the decoupling limit. An interesting observation made in subsection \\ref{sec:exact massless} is that the existence/absence of supercurvature modes can be strongly related to symmetries of the theory. While a massive scalar field with a sufficiently light mass has a supercurvature mode~\\cite{Sasaki:1994yt} that survives the massless limit, a theory of a scalar field with shift symmetry does not allow a physical supercurvature mode. This is because the would-be supercurvature mode does not show up in any correlation functions invariant under the shift symmetry. Furthermore, a vector field with a U(1) gauge symmetry does not have a supercurvature mode even when the vector field is given a mass by the Higgs mechanism and thus absorbs a light scalar degree of freedom (the phase of the complex Higgs field). It may be interesting to investigate supercurvature modes of the vector field when we take metric perturbations into account since gravity has the diffeomorphism symmetry, although in the present paper we take account of the effect of gravity only through a curved background. The evaluation of the metric perturbations is beyond the scope of the present paper and we hope to come back to this issue in a future publication. In this paper, we have assumed several simplifications: (i) the universe during inflationary era after a quantum tunneling is assumed to be well approximated by an exact de Sitter spacetime in the open chart; (ii) the origin of the breaking of the gauge and conformal symmetries and the mass of the vector field is assumed to be the standard Higgs mechanism; (iii) the mass squared $V''_{\\Phi}(|\\Phi|)$ around the minimum of the Higgs potential is assumed to be large enough so that the mass of the vector field can be considered as constant during inflation. Relaxing some of these assumptions would in principle affect details of our results, although generic features that we have found are expected to remain the same. Furthermore, we have neglected the interactions between the tunneling field and the other fields such as the Higgs field. If we take into account such interactions, then spatially localized, bubble-shaped features may appear~\\cite{Sugimura:2012kr}. We hope to come back to these issues in the near future." }, "1402/1402.7168_arXiv.txt": { "abstract": "{The results of a search for eclipsing Am star binaries using photometry from the SuperWASP survey are presented. The light curves of 1742 Am stars fainter than $V$ = 8.0 were analysed for the presences of eclipses. A total of 70 stars were found to exhibit eclipses, with 66 having sufficient observations to enable orbital periods to be determined and 28 of which are newly identified eclipsing systems. Also presented are spectroscopic orbits for 5 of the systems. The number of systems and the period distribution is found to be consistent with that identified in previous radial velocity surveys of `classical' Am stars.} ", "introduction": "\\label{introduction} Amongst the A and F stars there exists a subclass of peculiar stars called the metallic-lined (Am) stars, in which the \\ion{Ca}{II}~K line is considerably weaker than would be expected from the average metallic line type \\citep{1940ApJ....92..256T,1948ApJ...107..107R}. These stars exhibit an apparent underabundance of calcium and scandium, overabundances of iron-group elements, and extreme enhancements of rare-earth elements \\citep{1970PASP...82..781C}. In contrast to normal A-type stars, the Am stars are slowly rotating \\citep{1983aspp.book.....W} with maximum $v \\sin i$ values of $\\sim$100\\,km\\,s$^{-1}$ \\citep{1973ApJ...182..809A}. The abundance anomalies are thought to be due to radiative diffusion of elements within the stable atmospheres of these relatively slowly rotating stars \\citep{1970ApJ...160..641M,1980AJ.....85..589M,1983ApJ...269..239M}. Early spectral studies of Am stars hinted at a high fraction of spectroscopic binaries \\citep{1948ApJ...107..107R}, while the systematic study by \\cite{1961ApJS....6...37A} led to the conclusion that all Am stars are members of spectroscopic binaries. Hence, it was assumed that the slow rotation of Am stars, required for radiative diffusion to occur, was the result of the reduction of rotational velocities due to tidal interaction. While there were spectroscopic orbits for many Am stars \\citep[e.g.][]{2004A&A...424..727P}, only a handful were known to be eclipsing \\citep[e.g.][]{1980ARA&A..18..115P,1991A&ARv...3...91A}. It was this that led \\cite{1990clst.book.....J} to conclude that: \\begin{quote} \\emph{A curious fact is that among the many Am stars known (all of which are binaries) there should be many eclipsing binaries, but surprisingly very few cases are known.} \\end{quote} Comprehensive spectroscopic radial velocity studies of carefully selected Am stars, have found a binary fraction of nearer 60--70\\% \\citep{1985ApJS...59..229A,2007MNRAS.380.1064C}. The period distribution shows that the majority of systems have periods $\\la$\\,50\\,days, consistent with the slow rotation being due to tidal synchronisation or pseudo-synchronization \\citep{1996A&A...313..523B,1997A&A...326..655B}. There are, nonetheless, systems with longer periods, suggesting that these Am stars were formed with low initial rotation velocities \\citep{2004ASPC..318..297N}. The key to the Am phenomenon appears to be slow rotation and not binarity per se. The recent \\citet{2009A&A...498..961R} catalogue lists only 61 Am stars as eclipsing or possibly eclipsing. This represents only 1.4\\% of the Am stars in the catalogue. Given that a large fraction of Am stars are supposed to be in binary systems, this percentage does appear rather low. For example, in a binary with a period of $\\sim$5\\,days (typical of many Am spectroscopic binaries), there is a $\\sim$10\\% probability that the system should be eclipsing. Hence, there is a perhaps somewhat na\\\"{i}ve expectation that there ought to be more eclipsing Am stars. It is this which led us to investigate the number of eclipsing systems that can be found using light curves obtained from the SuperWASP exoplanet transit survey. ", "conclusions": "\\subsection{Expected period distribution of Am binaries} The results from the radial velocity studies of \\cite{1985ApJS...59..229A} and \\cite{2007MNRAS.380.1064C} can be used to predict the number and period distribution of eclipsing Am binaries. The combined sample comprises 151 Am stars, with 61 SB1s and 28 SB2s. This was binned onto 1-day bins, normalized by the total number of stars, to generate a period probability distribution for Am stars. Multiplying by the estimated WASP detection probability ($p_{\\rm overall}$) obtained in Sect.\\,\\ref{DetProb} and by the number of stars in the WASP sample (1742), yields an estimate of the expected period distribution of eclipsing Am stars. In Figure\\,\\ref{dist-predicted} the WASP eclipsing Am star period distribution is compared to that predicted for the two identical stars and the dark companion cases. Since these represent the extrema of the probabilities, we also include the predicted distribution obtained using the ratio of SB1 and SB2 systems from the RV studies. The number of eclipsing Am stars found by WASP does appear to be broadly consistent with the expected number of systems. We recall from Sect.\\,\\ref{introduction} that the fraction of spectroscopic binaries is 60$\\sim$70\\%. Thus, the eclipsing fraction appears to be similar, suggesting a significant fraction of Am stars might be single or have hard to detect companions. The period distribution is, however, slightly different, with a pronounced peak at shorter periods due to the inclusion of close binaries. The $v \\sin i$ distributions of both \\cite{1985ApJS...59..229A} and \\cite{2007MNRAS.380.1064C} are skewed toward lower values than the \\citet{2009A&A...498..961R} sample. These radial velocity studies have preferentially avoided stars with high rotation, which accounts for the excess of short period systems found in the WASP sample. The distribution of Am-type spectroscopic binaries in the \\citet{2009A&A...498..961R} catalogue (some 210 systems) shows a similar short orbital period excess, due to the inclusion of Am stars with a wide range of rotational velocities. \\begin{figure}[h] \\includegraphics[width=\\columnwidth]{dist-predicted.eps} \\caption{Period distribution of eclipsing Am star binaries. The WASP eclipsing Am star distribution is given as solid grey. The predicted period distribution based on the results of spectroscopic binaries are given as dashed-line for the two identical stars and dotted-line for the dark companion case. The thicker solid-line is that predicted based on the ratio of SB1 and SB2 systems.} \\label{dist-predicted} \\end{figure} \\subsection{Mass ratio distribution} \\begin{table*} \\caption{Results from \\textsc{jktebop} fits to light curves (Detached systems only) and approximate stellar parameters, assuming primary is ${T_\\mathrm{eff}}_A = 7500$\\,K. See Text for details.} \\label{JKTEBOP} \\begin{tabular}{llllllllllll}\\hline\\hline Renson &$\\frac{R_A+R_B}{a}$ & $R_B/R_A$ & $i$ & $e$ & $J_B/J_A$ & $P$ & ${T_\\mathrm{eff}}_B$ & $R_A$ & $R_B$ & $q$ & $q$ (literature) \\\\ \\hline 3750& 0.307& 0.111& 89.95& 0.00& 0.00& 1.2199& 2000\\S& 1.63& 0.18& 0.07 & $<$0.003 {\\citep{2010MNRAS.407..507C}}\\\\ 4660& 0.298& 0.910& 87.61& 0.00& 0.84& 2.7807& 7176& 1.97& 1.80& 0.92 & 0.91 {\\citep{2011MNRAS.414.3740S}}\\\\ 6720& 0.344& 1.354& 79.59& 0.00& 0.91& 2.3832& 7333& 1.68& 2.26& 1.07 &\\\\ 7310& 0.161& 0.601& 86.37& 0.00& 0.50& 6.6637& 6322& 2.20& 1.32& 0.71 &\\\\ 7730& 0.156& 3.000& 89.57& 0.24& 0.10& 39.2827& 4218& 2.93& 8.71& 0.84 & 0.96 {\\citep{1988AJ.....96.1040P}}\\\\ 8215& 0.105& 0.940& 86.43& 0.00& 0.44& 8.7959& 6126& 1.38& 1.29& 0.77&\\\\ 9237& 0.248& 1.036& 80.52& 0.00& 0.49& 2.8562& 6288& 1.51& 1.55& 0.81&\\\\ 9318& 0.148& 0.472& 85.89& 0.06& 0.72& 11.1133& 6901& 3.24& 1.52& 0.71&\\\\ 9410& 0.189& 0.634& 87.74& 0.02& 0.30& 16.7873& 5537& 5.07& 3.19& 0.59& 0.81{\\dag} {\\citep{1971AJ.....76..544L}}\\\\ 10016& 0.242& 1.323& 86.98& 0.25& 1.03& 5.4309& 7557& 2.11& 2.81& 1.11&\\\\ 11100& 0.165& 0.397& 85.07& 0.00& 0.36& 4.0375& 5818& 1.79& 0.71& 0.61&\\\\ 11470& 0.228& 0.580& 89.36& 0.00& 0.39& 2.8721& 5931& 1.75& 1.02& 0.66&\\\\ 14850& 0.215& 0.272& 83.02& 0.00& 0.08& 3.9795& 3925& 2.43& 0.66& 0.30&\\\\ 15034& 0.202& 0.469& 83.88& 0.00& 0.54& 7.5392& 6423& 3.39& 1.59& 0.65&\\\\ 15190& 0.174& 0.439& 82.60& 0.00& 0.02& 5.1229& 2952& 1.97& 0.86& 0.21& 0.33{\\dag} {\\citep{2003MNRAS.346..555C}}\\\\ 15445& 0.270& 0.881& 86.32& 0.00& 0.92& 5.7606& 7355& 3.10& 2.72& 0.93&\\\\ 21400& 0.106& 0.513& 86.63& 0.00& 0.04& 7.7729& 3435& 1.49& 0.77& 0.29&\\\\ 22860& 0.136& 0.558& 83.84& 0.28& 2.81& 7.9006& 9712& 2.33& 1.30& 1.17&\\\\ 25020& 0.188& 0.081& 89.10& 0.00& 0.00& 6.6284& 2000\\S& 3.39& 0.28& 0.06&\\\\ 25070& 0.185& 0.500& 85.60& 0.14& 0.91& 6.7149& 7323& 2.85& 1.42& 0.79& 1.00 (this work)\\\\ 29290& 0.314& 0.977& 79.45& 0.00& 0.95& 3.2922& 7406& 2.30& 2.23& 0.98&\\\\ 30090& 0.221& 0.788& 82.26& 0.00& 0.97& 4.3486& 7451& 2.11& 1.67& 0.93&\\\\ 30110& 0.148& 0.162& 84.55& 0.00& 0.09& 4.3145& 4108& 1.92& 0.31& 0.37&\\\\ 30457& 0.137& 0.231& 86.82& 0.00& 0.00& 11.9420& 2000\\S& 3.20& 0.74& 0.08&\\\\ 30650& 0.060& 0.378& 88.38& 0.00& 0.00& 18.1210& 2000\\S& 1.53& 0.58& 0.09&\\\\ 35000& 0.190& 0.835& 82.07& 0.00& 0.00& 10.2861& 2000\\S& 2.69& 2.26& 0.13&\\\\ 36660& 0.096& 0.698& 88.61& 0.41& 0.71& 16.3653& 6892& 2.30& 1.60& 0.81& 0.80 (this work)\\\\ 37220& 0.135& 0.962& 85.57& 0.00& 0.97& 5.7931& 7437& 1.38& 1.33& 0.98&\\\\ 37610& 0.240& 1.455& 85.50& 0.00& 1.28& 3.2387& 7974& 1.38& 2.02& 1.19&\\\\ 38500& 0.173& 0.348& 83.65& 0.00& 0.19& 3.9938& 4945& 1.87& 0.65& 0.47&\\\\ 40350& 0.156& 0.272& 88.22& 0.00& 0.00& 7.0687& 2000\\S& 2.43& 0.66& 0.08&\\\\ 40910& 0.132& 1.114& 87.17& 0.05& 1.05& 11.1163& 7594& 2.03& 2.23& 1.05&\\\\ 42906& 0.062& 0.860& 87.84& 0.00& 0.00& 19.6987& 2000\\S& 1.23& 1.06& 0.12& 0.98 {\\citep{2003MNRAS.346..555C}}\\\\ 44140& 0.309& 1.391& 86.51& 0.02& 0.66& 2.0598& 6763& 1.31& 1.81& 0.96& 0.90 {\\citep{1970ApJ...162..925P}}\\\\ 49380& 0.161& 0.355& 85.06& 0.11& 0.57& 5.0204& 6508& 2.14& 0.76& 0.68& 0.83 (this work)\\\\ 51506& 0.322& 0.119& 80.30& 0.12& 0.17& 1.9196& 4787& 2.69& 0.32& 0.46& 0.12 (this work)\\\\ 56310& 0.285& 1.108& 84.26& 0.00& 0.88& 2.6959& 7258& 1.68& 1.85& 0.99&\\\\ 58170& 0.366& 0.854& 88.24& 0.00& 0.48& 1.6047& 6246& 1.65& 1.42& 0.77& 0.78 {\\citep{1968ApJ...154..191P}}\\\\ 58256& 0.271& 0.378& 81.63& 0.00& 0.41& 2.9673& 5983& 2.50& 0.95& 0.59&\\\\ 59780& 0.228& 1.200& 87.87& 0.33& 0.95& 7.3514& 7412& 2.61& 3.15& 1.06& 0.98 {\\citep{1999AJ....118.1831T}}\\\\ \\hline \\end{tabular} \\tablefoot{\\dag Mass ratio (q) obtained using spectroscopic binary mass function, $f(m)$, and assuming $M_1$ = 1.7\\,$M_\\sun$ and $i$ = 90.\\newline \\S lower-limit on {${T_\\mathrm{eff}}_{B}$} imposed when $J_B/J_A = 0$. } \\end{table*} Without direct determinations of masses from spectroscopic studies, we can only make a rather crude estimate of the mass distribution of the eclipsing systems from their light curves and the \\textsc{jktebop} fits. Since the bolometric correction for late-A stars is small, we can make the approximation that the ratio of bolometric surface brightnesses is given by the WASP bandpass surface brightness ratio ($J_B/J_A$). Thus the effective temperature of the secondary (${T_\\mathrm{eff}}_{B}$) can be obtained from, \\[{T_\\mathrm{eff}}_{B} \\approx {T_\\mathrm{eff}}_{A} \\times (J_B/J_A)^{1/4},\\] where the effective temperature of the primary (${T_\\mathrm{eff}}_{A}$) is assumed to be 7500\\,K. With initial mass estimates of $M_A$ = $M_B$ = 1.7, the known orbital period ($P_{\\rm orb}$) and sum of the radii ($\\frac{R_A+R_B}{a}$) from \\textsc{jktebop}, we determine initial values for $R_A + R_B$. Using the ratio of the radii also from \\textsc{jktebop} ($R_B/R_A$) we can determine individual values for $R_A$ and $R_B$. Next, the \\cite{2010A&ARv..18...67T} relations are used to determine stellar mass estimates, by varying $\\log g$ so that radii obtained from \\cite{2010A&ARv..18...67T} agrees with that expected from the \\textsc{jktebop} analysis. The procedure is iterated until there are no changes in parameters. The results are presented in Table\\,\\ref{JKTEBOP}, along with values for mass ratio ($q$) from spectroscopic analyses in the literature and those determined in Section\\,\\ref{Data_modelling}. The average difference between our estimated $q$ values and the spectroscopic values is $-$0.11, but with an rms scatter of $\\pm$0.25. \\begin{figure}[h] \\includegraphics[width=\\columnwidth]{q-dist.eps} \\caption{Mass ratio distribution for Am binary systems. The solid-line histogram is the distribution based on eclipsing binaries in the current work, while the solid grey histogram is that presented by \\cite{2010A&A...524A..14B} based on the spectroscopic sample. The dashed histogram is the estimated eclipsing binary distribution after making allowance for the WASP detection probabilities.} \\label{q-dist} \\end{figure} As discussed in Sect.~\\ref{ecl_prob}, eccentric systems may not always show two eclipses when both stars are similar. For example Renson\\,42906 (HD\\,151604; V916 Her) is an eccentric ($e = 0.566$) system with mass ratio close to unity \\citep{2007MNRAS.380.1064C}, but the WASP data only shows one eclipse per orbit and value of $q=0.12$. Thus, these systems will appear to have anomalously low $q$ values, further adding to the uncertainty in the mass-ratio distribution. \\cite{2010A&A...524A..14B} concluded that the mass-ratio distribution showed hints of a double-peaked distribution, with peaks at $q \\sim 0.3$ and $q \\sim 1$. The mass-ratio distribution based on the estimated properties from the WASP light curves for detached systems is noticeably different (Fig.\\,\\ref{q-dist}). The estimated WASP mass-ratio distribution shows a broad peak near unity, with a deficit around $q \\simeq 0.3$. However, the WASP detection probability varies with companion size (Sect.\\,\\ref{det_prob}). Assuming that companions with masses around 0.5\\,{$M_\\sun$} correspond to the transition between the two system scenarios discussed earlier, there would be an increase in the number of low $q$ systems relative to the high $q$ systems. The distribution would become flatter, similar to that found by \\cite{2010A&A...524A..14B}, who noted that a flat mass-ratio distribution also appeared to be a good fit. While there are genuinely low $q$ systems (e.g. Renson 3750), the apparent excess of such systems may not be real, since some of these may be pairs of similar stars with a true period equal to twice the assumed one (as noted with a dagger in Table~\\ref{table:Am Binaries}). Hence, since our mass-ratio estimates are based on photometry alone, radial velocity studies are required to determine spectroscopic mass-ratios of the whole sample, before any firmer conclusions can be drawn." }, "1402/1402.4524.txt": { "abstract": "We present {\\it Herschel}/PACS 100 and 160 $\\mu$m integrated photometry for the 323 galaxies in the {\\it Herschel} Reference Survey (HRS), a K-band-, volume-limited sample of galaxies in the local Universe. Once combined with the {\\it Herschel}/SPIRE observations already available, these data make the HRS the largest representative sample of nearby galaxies with homogeneous coverage across the 100-500 $\\mu$m wavelength range. In this paper, we take advantage of this unique dataset to investigate the properties and shape of the far-infrared/sub-millimeter spectral energy distribution in nearby galaxies. We show that, in the stellar mass range covered by the HRS (8\\ls$\\log(M_{*}/M_{\\odot})$\\ls12), the far-infrared/sub-millimeter colours are inconsistent with a single modified black-body having the same dust emissivity index $\\beta$ for all galaxies. In particular, either $\\beta$ decreases, or multiple temperature components are needed, when moving from metal-rich/gas-poor to metal-poor/gas-rich galaxies. We thus investigate how the dust temperature and mass obtained from a single modified black-body depend on the assumptions made on $\\beta$. We show that, while the correlations between dust temperature, galaxy structure and star formation rate are strongly model dependent, the dust mass scaling relations are much more reliable, and variations of $\\beta$ only change the strength of the observed trends. %Our results suggests that far-infrared/sub-millimeter colours may not always be a good proxy for the average %temperature of the cold dust reservoir in galaxies. ", "introduction": "It is now well established that approximately half of the radiative energy produced by galaxies is absorbed by dust grains and re-emitted in the infrared regime \\citep{hauser01,bosellised,dole06,dale07,burga2013}. Thus, observations in the $\\sim$10-1000 $\\mu$m wavelength range provide us with a unique opportunity not only to quantify half of the bolometric luminosity of galaxies, but also to characterise the properties of cosmic dust. Moreover, since dust grains are crucial for the star formation cycle \\citep{hollenbach71}, such information can give us important insights into the physical processes regulating galaxy evolution (e.g., \\citealp{dunne11}). Unfortunately, despite its paramount importance, we are still missing a complete and coherent picture of dust properties in galaxies across the Hubble sequence, and of the exact role played by grains in regulating star formation \\citep{mckee10}. Indeed, we know very little about the dust composition in galaxies outside our own Local Group \\citep{draine07b,compiegne11} and if/how it is regulated by the physical conditions experienced by grains in the inter-stellar medium (ISM). Hence, our estimates of dust masses in galaxies are still highly uncertain \\citep{fink99,dupac03,gordon10,paradis10,plankbeta}. Luckily, the last decade has seen the start of a golden age for observational far-infrared (FIR) and sub-millimeter (submm) astronomy, providing a new boost to the refinement of theoretical dust models \\citep{meny07,draine07b,hoang2010,compiegne11,steinacker13}. In particular, the {\\it Spitzer} \\citep{spitzer}, and more recently {\\it Herschel} \\citep{pilbratt10} and {\\it Planck} \\citep{planck} space telescopes are finally gathering a wealth of information on the dust emission from thousands of galaxies up to $z\\sim$2. Particularly important for a proper characterisation of dust in galaxies is the radiation emitted at wavelengths \\gs100-200 $\\mu$m. In this regime, the integrated emission from galaxies originates predominantly from dust in thermal equilibrium, heated by the diffuse interstellar radiation field (ISRF), which represents the bulk of the dust mass in a galaxy (e.g., \\citealp{sodrosky89,sauvage92,calzetti95,walterbos,bendo10,boquien11,bendo12b}). Thus, by characterising the dust emission in the \\gs100 $\\mu$m wavelength domain, we have a unique opportunity to provide strong constraints to theoretical models, and to refine our census of the dust budget in galaxies. The first natural step in this direction is to quantify how the shape of the dust spectral energy distribution (SED) varies with galaxy properties across a wide range of morphological type, star formation activity, cold gas mass and metal content. This is necessary to determine if the amount of radiation emitted at each wavelength is simply regulated by the intensity of the ISRF responsible for the dust heating, or whether it retains an imprint of the chemical composition of the grains. Indeed, only after a careful characterisation of the physical parameters regulating the dust SED, will it be possible to properly convert observables into physical quantities such as dust temperatures and dust masses. Many recent works \\citep{gordon10,skibba11,davies11,plankbeta,galametz12,auld13} have shown that, above $\\sim$100 $\\mu$m, the dust SED is very well approximated by a simple modified black-body (but see also \\citealp{bendo12b}): \\begin{equation} F_{\\nu} = \\frac{M_{dust}}{D^{2}} \\kappa_{\\nu_{0}}\\left(\\frac{\\nu}{\\nu_{0}}\\right)^{\\beta} B_{\\nu}(T) \\label{eq1} \\end{equation} where $F_{\\nu}$ is the flux density emitted at the frequency $\\nu$, $\\kappa_{\\nu_{0}}$ is the dust mass absorption coefficient at the frequency $\\nu_{0}$, $\\beta$ gives its variation as a function of frequency, $D$ is the galaxy distance and $B_{\\nu}(T)$ is the Planck function. Mounting evidence is emerging that $\\beta$ is not the same in all galaxies (e.g., \\citealp{remy13}), and may also vary within galaxies (e.g., \\citealp{galametz12,smith12b}). Modified black-bodies are simple models and cannot properly reproduce real dust properties (e.g., \\citealp{draine07b,shetty09b,bernard10}). Several dust components at various temperatures contribute to the total emission along the lines-of-sight. This implies the presence of temperature mixing that can cause variations of the infrared slope, and thus in the apparent emissivity index $\\beta$. Nevertheless, parameterization of the dust SEDs through modified black-body fitting is a powerful tool to help understand variations of dust properties with other galaxy characteristics, especially in case of sparse sampling of the FIR/sub-mm wavelength range (e.g., high-redshift galaxies \\citealp{magdis2011,symenodis2013}). Therefore, it is extremely important to determine in which cases a single modified back-body can be used, and how temperature and dust mass estimates are affected by the assumptions made on $\\beta$. In order to ascertain the dust properties of galaxies in the local Universe, and to provide new constraints to theoretical models, we have carried out the {\\it Herschel} Reference Survey (HRS, \\citealp{HRS}), a {\\it Herschel} guaranteed time project focused on the study of the interplay between dust, gas and star formation in a statistically significant sample of $\\sim$300 galaxies spanning a wide range of morphologies, stellar masses (8\\ls log(M$_{*}$/M$_{\\odot})$\\ls 12), cold gas contents (-3\\ls log(M$_{HI}$/M$_{*}$)\\ls 1), metallicities (8.2\\ls 12+log(O/H) \\ls 8.9), and specific star formation rates (-12 \\ls log(SFR/M$_{*}$)\\ls -9). The combination of {\\it Herschel}/SPIRE \\citep{spire} observations with the multi-wavelength dataset we have been assembling \\citep{ciesla12,hrsgalex,boselli13,hughes13}, has already allowed us to have a first glimpse at how the dust content and shape of the dust SED vary with internal galaxy properties \\citep{boselli10,boselli12,cortese12}. In particular, \\cite{boselli10,boselli12} have shown that the slope of the dust SED in the 200-500 $\\mu$m interval decreases from $\\beta\\sim$2 to $\\beta\\sim$1 when moving from metal-rich to metal-poor galaxies. However, our analyses have so far been limited by the lack of data in the $\\sim$100-200 $\\mu$m wavelength range for the entire sample. Thus, in this paper we present integrated {\\it Herschel}/PACS \\citep{pacs} 100 and 160 $\\mu$m flux densities for all the HRS sample and take advantage of our multiwavelength dataset to perform a first analysis of the properties of the dust SED across our entire sample. Corresponding to the peak of the dust SED, the 100-200 $\\mu$m wavelength interval is crucial not only to properly quantify the shape of the SED, but also to accurately determine the average dust temperature and total dust mass in galaxies. These data make the HRS the largest representative sample of nearby galaxies with homogeneous coverage across the $\\sim$100-500 $\\mu$m wavelength range. In addition to releasing our dataset to the community, our primary goals are 1) to investigate how the shape of the dust SED varies with internal galaxy properties, and 2) to determine whether the integrated dust SED of HRS galaxies can always be reduced to a single modified black-body with a constant value of $\\beta$ and, if not, what are the possible biases introduced by this assumption. The results of SED fitting with the dust models of \\cite{draine07} will be presented in a forthcoming paper (Ciesla et al., submitted.). This paper is organized as follows. In Sect. 2 we describe the {\\it Herschel} observations, data reduction, flux density estimates and comparison with the literature. In Sect. 3 we use the PACS and SPIRE colours to investigate how the shape of the dust SED varies with internal galaxy properties. In Sec. 4, we show how the dust temperature and mass obtained from fitting a single modified black-body to the {\\it Herschel} data depend on the assumptions made on $\\beta$. Finally, the summary and implications of our results are presented in Sec. 5. \\begin{figure*} \\centering \\includegraphics[width=17.5cm, angle=0]{images.epsi} \\caption{Comparison of the quality of our PACS images with the Sloan Digital Sky Survey optical and SPIRE 250 $\\mu$m images. We show three types of objects: an early-type with dust lanes (top row), an unperturbed late-type spiral and an un-detected elliptical and its spiral companion. The size of the PACS and SPIRE beams is shown in the bottom left corner of each panel.} \\label{images} \\end{figure*} ", "conclusions": "In this paper we presented PACS 100 and 160 $\\mu$m integrated photometry for the {\\it Herschel} Reference Survey. We have combined these data with SPIRE observations to investigate how the shape of dust SED varies across the Hubble sequence. Being the largest representative sample of nearby galaxies with homogeneous coverage in the 100-500 $\\mu$m wavelength domain, the HRS is ideal to quantify if and how dust emission varies across the local galaxy population. Our main results are as follows. \\begin{itemize} \\item The shape of the dust SED is not well described by a single modified black-body having just the dust temperature as a free parameter. Instead, there is a clear need to vary the dependence of the dust emissivity ($\\beta$) on wavelength, or to invoke multiple temperature components in order to reproduce the colours observed in our sample. This is particularly important as the HRS does not include very metal-poor dwarf galaxies, for which we already knew that the dust SED is significantly different from the one of metal-rich, massive galaxies \\citep{galliano2005,galliano11,engelbracht2008,galametz2009,remy13}. Our results suggest that the difference in FIR/sub-mm colours between giant and dwarf galaxies \\citep{draine07b} may not be the result of a dramatic transition in dust properties, but just the consequence of the gradual variation that we observe as a function of metal and gas content. \\item The variation in the slope of the dust SED strongly affects dust temperature estimates from single modified black-bodies fits. In particular, the correlations between galaxy properties and dust temperatures strongly depend on the assumptions made on $\\beta$: i.e., trends can disappear or even reverse. Conversely, dust mass estimates are more robust, and variations in $\\beta$ do not produce the same dramatic inversion of some correlations observed for the dust temperature. \\item We confirm that the temperature of a single modified black-body is mainly related to specific star formation rate, while $\\beta$ varies more with the degree of metal enrichment of the ISM. \\end{itemize} The results presented in this paper may appear in contradiction with several recent works showing that the dust SED is very well reproduced by a simple modified black-body with $\\beta=$2 \\citep{davies11,auld13}. However, all these works were focused on massive, metal-rich and relative gas-poor galaxies, for which we also find that a constant value of $\\beta$ provides a good fit to our data. It is when we move to the gas-rich/metal-poor regime that the shape of the SED starts to change \\citep{boselli10,boselli12,remy13}. Our findings overall reinforce the results already presented in \\cite{boselli10,boselli12}. However, it is important to note that the discovery of a clear variation in the shape of the SED across the HRS has only been possible thanks to the large wavelength coverage obtained by combining both PACS and SPIRE data. Indeed, with SPIRE or PACS data only, it would be not only much more difficult to show under which conditions a simple modified black-body approach does not work, but it would also be nearly impossible to quantify how model assumptions can affect the correlation of dust temperature with star formation, galaxy structure and chemical enrichment. %Despite this, {\\it Herschel} colours still represent an ideal tool to quantify %how the dust SED varies with other characteristics of the ISM in galaxies. %{\\bf Indeed, once we take into account that PACS and SPIRE colours are related to different %dust properties, they can be used as proxies for $T$ and $\\beta$ without being %affected by any degeneracy hidden in the SED fitting technique. %Thus, whenever possible, we suggest using colours to carefully test the reliability %of relations obtained as results of a $\\chi^{2}$ minimizations of a single %modified black-body. %% Finally, the need of varying $\\beta$ even for a sample like the HRS has also profound implications %% on our ability to quantify dust masses in galaxies, as a variation in the value %% of the dust emissivity used to normalise the modified black-body is required \\citep{bianchi13}. %% Of course, this automatically affects the use of far-infrared/sub-millimeter %% observations to constrain the amount of cold gas \\citep{james02,eales2010} and/or the value of the %% CO-to-H$_{2}$ conversion factor in nearby \\citep{guelin93,boselligdust,leroy2011} and high-redshift galaxies \\citep{magdis2011}. %% For example, the simple exercise described in the previous section shows that %% dust masses could be easily underestimated by a factor of $\\sim$2, when moving from %% high to low metallicity environments. This would automatically lead to an underestimate %% of the variation of the CO-to-H$_{2}$ conversion factor with metallicity. %% Unfortunately, no dust models are currently able to predict the evolution %% of the dust emissivity with $\\beta$ across the range of metallicities and gas content %% covered by the HRS. This is an issue that, thanks to the complete coverage now available %% for our sample, we plan to investigate in future work. %% In the meantime, we recommend caution when adopting the same dust model for %% quantifying the dust properties and dust mass for samples of galaxies covering %% a wide range in metallicity, cold gas content and/or specific star formation rates." }, "1402/1402.6358_arXiv.txt": { "abstract": "We present a quantitative analysis of the low-resolution ($\\sim$4.5~\\AA) spectra of 12 late-B and early-A blue supergiants (BSGs) in the metal-poor dwarf galaxy NGC~3109. A modified method of analysis is presented which does not require use of the Balmer jump as an independent T$_{eff}$ indicator, as used in previous studies. We determine stellar effective temperatures, gravities, metallicities, reddening, and luminosities, and combine our sample with the early-B type BSGs analyzed by \\citet{E07} to derive the distance to NGC 3109 using the Flux-weighted Gravity-Luminosity Relation (FGLR). Using primarily Fe-group elements, we find an average metallicity of [\\={Z}] = -0.67 $\\pm$ 0.13, and no evidence of a metallicity gradient in the galaxy. Our metallicities are higher than those found by \\citet{E07} based on the oxygen abundances of early-B supergiants ([\\={Z}] = $-$0.93 $\\pm$ 0.07), suggesting a low $\\alpha$/Fe ratio for the galaxy. We adjust the position of NGC 3109 on the BSG-determined galaxy mass-metallicity relation accordingly and compare it to metallicity studies of HII regions in star-forming galaxies. We derive an FGLR distance modulus of 25.55 $\\pm$ 0.09 (1.27 Mpc) that compares well with Cepheid and tip of the red giant branch (TRGB) distances. The FGLR itself is consistent with those found in other galaxies, demonstrating the reliability of this method as a measure of extragalactic distances. ", "introduction": "The extreme brightness of blue supergiants (BSGs), a short post-main sequence evolutionary stage of 12~M$_{\\odot}$ to 40~M$_{\\odot}$ stars, makes it possible to obtain resolved spectra of individual BSGs out to 10 Mpc with current instrumentation. As such, BSGs are ideal tools to obtain crucial information about the chemical composition of nearby galaxies and provide insight to their chemical evolution \\citep{K08,K12}. Often galaxy metallicities are studied through the spectroscopy of HII regions, which has been widely applied to examine radial abundance gradients of spiral galaxies \\citep{VC92, Z94, P04} and the galaxy mass-metallicity relation \\citep{L79, T04, A13}. However, this approach is limited by its reliance on empirical ``strong line'' analysis methods, which have been shown to yield significantly different absolute metallicities depending on what calibration is used \\citep{Ke08, B09}. Even in cases where metallicities can be measured more directly using the weak auroral lines, HII region studies might be affected by systematic uncertainties difficult to assess, such as oxygen depletion on to dust grains and a possible detection bias towards lower abundances at high metallicities \\citep{ZB12}. BSGs thus provide a valuable independent measure of galaxy metallicity. In addition, it has been shown that BSGs can be used as distance indicators through the Flux-weighted Gravity-Luminosity Relation (FGLR, Kudritzki et al. 2003, 2008). This relation correlates stellar gravity and effective temperature, which can be derived from the stellar spectrum, to the absolute bolometric magnitude. The FGLR is advantageous in that it is free of uncertainties caused by interstellar reddening, since the reddening is determined during the spectral analysis. In addition, a potential metallicity dependence of the FGLR, if present, can be accounted for since metallicity is also determined independently by the analysis of each object. This is especially valuable in light of recent efforts to establish the Hubble constant \\emph{H$_{0}$} to an accuracy better than 5\\%, which would greatly constrain cosmological parameters without having to invoke assumptions about the geometry of the universe \\citep{KU12, Riess11}. FGLR distances found for WLM \\citep{U08}, M33 \\citep{U09}, and M81 \\citep{K12} have been found to be consistent with distances determined by other methods, demonstrating the reliability of the method. We conduct a spectroscopic study of 12 late-B and early-A type BSGs in NGC~3109, a Magellanic SBm galaxy at the edge of the Local Group \\citep{dV91,Bergh99}. With M$_{V}$ = -14.9 \\citep{Mcc12}, the galaxy is the most luminous member of the NGC~3109 group, which according to \\citet{Tully06} is the ``nearest distinct structure of multiple galaxies to the Local Group''. Recent work by \\citet{ST13} and \\citet{B13} indicate that the members of this group form a $\\sim$1070 kpc filamentary structure created by tidal interaction or filamentary accretion. Our purpose is two-fold: to determine the FGLR and calculate a corresponding distance to NGC~3109, and to evaluate its average metallicity using multiple Fe-group element species. An analysis of eight early B-type BSGs by Evans et al. (2007, hereafter E07) found the oxygen abundance of NGC~3109 to be approximately 1/10 of solar, a result consistent with HII regions studies using auroral lines \\citep{Le03, Le03b, P07}. If these values reflect the overall metallicity, NGC~3109 will be the lowest metallicity object for which an FGLR has been constructed, allowing us to investigate the metallicity dependence of the relation and compare with stellar evolution theory. Since we fit metal lines of multiple elements, in particular Fe-group elements, our metallicities will more closely resemble the overall stellar metallicities and are not restricted to oxygen as a proxy of stellar metallicity as in the case of the HII region studies and the work of E07. On the other hand, a comparison with the oxygen abundances obtained in these studies will provide insight to the $\\alpha$/Fe abundance ratio, a key parameter in constraining the chemical evolution and star formation history of the galaxy. ", "conclusions": "We analyze low-resolution ($\\sim$4.5~\\AA) spectra of 12 late-B and early-A BSGs in NGC~3109, obtaining the effective temperatures, gravities, metallicities, reddening, and luminosities of these objects. We use a modified method of analysis that does not use the Balmer jump to break the temperature-metallicity degeneracy. Instead, we employ a $\\chi^2$-based approach that takes advantage of the fact that spectral lines from different atomic species and from different excited levels react differently to changes of temperature and metallicity and use this to constrain stellar parameters. A test analysis of SMC spectra is found to produce parameters consistent with high-resolution analyses, attesting to the accuracy of our technique. A disadvantage of this method is that we must make assumptions regarding the He abundances of our objects, which we cannot determine from our analysis alone. Thus we consider two sets of model atmospheres, one assuming a ``normal'' He abundance (based on averages from BSG studies in the MW) and another assuming an enhanced He abundance, and take the average metallicity and FGLR distance as our final results for NGC 3109. Fortunately, the adopted He abundance only has a small impact on the derived metallicities, with the temperatures, gravities, reddening, and luminosities being very similar for both sets of models. From our sample, we find the average Fe-group metallicity of NGC~3109 to be [\\={Z}] = $-$0.63 $\\pm$ 0.13 in the normal He abundance case and [\\={Z}] = $-$0.71 $\\pm$ 0.14 in the enhanced He abundance case, resulting in our adopted metallicity of [\\={Z}] = $-$0.67 $\\pm$ 0.13. Even in the enhanced He abundance case, this result is higher than the oxygen-based metallicity obtained by \\citet{E07}, who find an average value of [\\={Z}] = $-$0.93 $\\pm$ 0.07 based on an analysis of 8 early-B BSGs. This may indicate a sub-solar $\\alpha$/Fe ratio in the galaxy. We adjust the position of NGC~3109 on the BSG-based galaxy mass-metallicity relation presented in \\citet{K12}, and find that the relation compares well (in a qualitative sense) to the recent mass-metallicity relation of \\citet{A13} based on auroral line metallicity measurements of star-forming galaxies in SDSS. Interestingly, the metallicities of the \\citet{A13} relation appear to be systematically higher than those in the BSG relation, an inconsistency which requires further investigation. We combine our results with the BSGs analyzed by \\citet{E07} to determine the Flux-weighted Gravity-Luminosity Relation (FGLR) of NGC~3109. We find the FGLR to be almost identical to those found in other galaxies, demonstrating the consistency of the relation across a wide range of galaxy masses and metallicities. We obtain an FGLR distance modulus of $\\mu$ = 25.55 $\\pm$ 0.09 (1.27 Mpc) that is effectively independent of the adopted stellar He abundances. This result is in good agreement with distances found in Cepheid variable and TRGB studies, serving as an independent confirmation of these values. The consistency between our FGLR distance and the Cepheid distances of \\citet{Pi06} and \\citet{So06} suggests that the Cepheid period-luminosity (PL) relation is not strongly affected by metallicity, since these studies adopt PL slopes derived for higher metallicities ([Z]~=~$-$0.3). This study is an additional example of the great value of blue supergiants as independent distance and metallicity indicators in nearby galaxies." }, "1402/1402.6672_arXiv.txt": { "abstract": "We perform population synthesis simulations for Population III (Pop III) coalescing compact binary which merge within the age of the universe. We found that the typical mass of Pop III binary black holes (BH-BHs) is $\\sim30~\\msun$ so that the inspiral chirp signal of gravitational waves can be detected up to z=0.28 by KAGRA, Adv. LIGO, Adv. Virgo and GEO network. Our simulations suggest that the detection rate of the coalescing Pop III BH-BHs is $140(68)\\ {\\rm events/yr ~(SFR_p/(10^{-2.5}\\msun/yr/Mpc^3))\\cdot Err_{sys}}$ for the flat (Salpeter) initial mass function (IMF), respectively, where $\\rm SFR_{p}$ and $\\rm Err_{sys}$ are the peak value of the Pop III star formation rate and the possible systematic errors due to the assumptions in Pop III population synthesis, respectively. $\\rm Err_{sys}=1$ correspond to conventional parameters for Pop I stars. From the observation of the chirp signal of the coalescing Pop III BH-BHs, we can determine both the mass and the redshift of the binary for the cosmological parameters determined by Planck satellite. Our simulations suggest that the cumulative redshift distribution of the coalescing Pop III BH-BHs depends almost only on the cosmological parameters. We might be able to confirm the existence of Pop III massive stars of mass $\\sim 30~\\msun$ by the detections of gravitational waves if the merger rate of the Pop III massive BH-BHs dominates that of Pop I BH-BHs. ", "introduction": "Gravitational-wave astronomy with KAGRA\\footnote{http://gwcenter.icrr.u-tokyo.ac.jp/en/}, Adv. LIGO\\footnote{http://www.ligo.caltech.edu/}, Adv. Virgo\\footnote{http://www.ego-gw.it/index.aspx/}, and GEO\\footnote{http://www.geo600.org/} will reveal the formation and evolution of binaries through the observed merger rates of compact binaries, such as binary neutron stars (NS-NSs), neutron star -- black hole binaries (NS-BHs), and binary black holes (BH-BHs). For this gravitational wave astronomy, estimates of the merger rate of compact binaries play key roles to develop observational strategy and to translate the observed merger rates into the binary formation and evolution processes. There are two methods to estimate the merger rate of compact binaries. One is to use observational facts such as the observed NS-NSs whose coalescence time due to the emission of gravitational waves is less than the age of the universe. Taking into account the observation time, the sensitivity of the radio telescope, the luminosity function of pulsars and the beaming factor so on, the probability distribution function of the merger rate can be found. For example, \\citet{Kalogera2004b} found that the event rate of the coalescing NS-NSs is in the range from $10^{-5}\\ {\\rm events}\\ {\\rm yr}^{-1}\\ {\\rm galaxy}^{-1}$ to $4\\times 10^{-4}\\ {\\rm events}\\ {\\rm yr}^{-1}\\ {\\rm galaxy}^{-1}$ at the $99\\ \\%$ confidence level (see their Fig.~2)\\footnote{Note here that there are errors in \\citet{Kalogera2004a} so that the rates in \\citet{Kalogera2004b} are the correct ones.}. The merger rate of NS-NSs can be restricted by the rate of the observed Type Ib and Ic supernovae, supposing that the formation of NS-NSs really starts from the massive binary zero age main sequence~(ZAMS) stars. This is because the formation of the second neutron star should occur in association with Type Ib and Ic supernovae in which the H-rich envelope and the He-layer are lost, respectively, otherwise the binary disrupts due to the sudden large mass loss at the supernova explosion\\footnote{If more than half of the total mass is suddenly lost at the supernova explosion, the binary disrupts.}. Under the assumption of the equality of the formation rate to the merger rate, the merger rate of the NS-NSs is limited by the Type Ib and Ic supernova rate of $\\sim 10^{-3}\\ {\\rm events}\\ {\\rm yr}^{-1}\\ {\\rm galaxy}^{-1}$ \\citep{Cappellaro1997, Cappellaro1999,Li2011}. Therefore the maximum rate of $4\\times 10^{-4}\\ {\\rm events}\\ {\\rm yr}^{-1}\\ {\\rm galaxy}^{-1}$ by \\citet{Kalogera2004b} implies that $\\sim 40\\ \\%$ of the Type Ib and Ic supernovae is associated with the formation of NS-NSs with the coalescence time less than the age of the universe. This percentage seems to be too large. We also note here that under the assumption that central engine of short gamma ray bursts are coalescing binary neutron stars, one can use the observations of short gamma ray bursts to estimate the coalescing rate (\\citet{coward12} ~and references cited there). The dynamical interaction in a globular cluster is another route to the formation of NS-NSs since there exists PSR2127+11C, which is contained in a NS-NS system, in the globular cluster M15 \\citep{Prince1991}. The age of the globular cluster is $\\sim 10^{10}\\ {\\rm yr}$ so that all the massive stars ended their life and the young pulsars do not exist. The coalescence time of PSR2127+11C is $\\sim 2 \\times 10^{8}\\ {\\rm yr}$ which is much smaller than the age of the globular cluster so that it was formed most likely by three body interactions such as the collision of neutron star -- white dwarf binary or neutron star -- dwarf star binary with a single neutron star. Since there exists only one NS-NS observed in the globular cluster, it is difficult to estimate the merger rate. Theoretical simulation is the only method at present~\\citep{Grindlay2006, Ivanova2008}. The another method to estimate the merger rate of compact binaries is to use theoretical computation based on the hypothetical assumptions of binary formation and evolution. For NS-BHs and BH-BHs, in particular, there exists no observations so that we can only use theoretical estimates. The merger rates of compact binaries of Population I~(Pop I) stars were estimated by many authors~\\citep{Belczynski2002, Belczynski2007, Belczynski2012}. \\citet{Dominik2012} computed the merger rates for the progenitor stars of metallicity $Z=0.1\\ Z_\\odot$ and found that the number of the coalescing BH-BHs increases compared with that for $Z=Z_\\odot$. \\citet{Dominik2013} adopted a certain model of the star formation rate and the chemical evolution of the metallicity $Z$ to compute the cumulative redshift distribution of the coalescing compact binaries. In this paper, we focus on the compact binary merger originated from Population III stars (Pop III stars) as gravitational wave sources. Pop III stars are the first stars after the Big Bang which are formed from metal-free gas~\\citep{Omukai1998, Bromm2002, Abel2002, Yoshida2008, Greif2012}. The simulations of a rotating primordial gas cloud suggest that the formation of Pop III star binaries and multiple star systems are frequent \\citep{Machida2008a, Stacy2010}. The main differences of our work from \\cite{Dominik2012,Dominik2013} are the following two: (1) we focus on metal-free Pop III stars and (2) consider the star formation history including the transition to metal-enriched stars (see \\S4). The observed merger rate will be the sum of our work and \\citet{Dominik2013}. There are at least three differences between Pop III and Pop I compact binaries. First of all, Pop III stars are more massive than Pop I stars \\citep{McKee2008, Hosokawa2011, Hosokawa2012, Stacy2012} with mass $10-100\\ {\\rm M}_\\odot$ so that Pop III star binaries probably evolve into BH-BHs. Secondly, since the typical formation time of Pop III stars is at $z\\sim 10$, even if the coalescence time is comparable to the age of the universe, they merge at present and contribute to the sources of gravitational waves for KAGRA, Adv. LIGO, Adv. Virgo, and GEO network. Therefore, if Pop III NS-NSs were formed, they might merge at present so that their rate is free from the constraint of the observed NS-NSs as well as Type Ib and Ic supernova rate discussed in the previous paragraphs. Thirdly, Pop III black holes are expected to be more massive than Pop I black holes due to less mass loss so that the resulting gravitational waves are easier to detect, since the detectable distance is proportional to 5/6 power of the chirp mass ($M_{\\rm chirp}$) of a binary defined by $ M_{\\rm chirp} = (M_1M_2)^{3/5}/(M_1 + M_2)^{1/5}$ (Peters 1964 and Peters \\& Mathews 1963, see also Sathyaprakash \\& Schutz(2009)), where $M_1$ and $M_2$ are the mass of each compact object. The idea of Pop III compact binaries as gravitational-wave sources has been considered by \\cite{Belczynski2004}, \\cite{Kulczycki2006} and \\cite{Kowalska2012}. However they considered very massive Pop III stars with mass over hundred solar masses. Recent study shows that the typical mass of Pop III stars is set to 10--100$\\rm{M_{\\odot}}$ by the stellar radiation feedback on the accretion flow~\\citep{Hosokawa2011, Hosokawa2012}. Therefore, in this paper, we calculate $10^6$ Pop III binary evolutions with the mass range of 10--100$\\rm{M_{\\odot}}$ to estimate merger rates and mass distribution of Pop III compact binaries. In order to calculate Pop III binary evolutions, we upgrade Hurley's binary population synthesis code \\citep{Hurley2002} for the Pop I star to Pop III star case. This paper is organized as follows. We describe Pop III single star evolution in \\S \\ref{sec:single}, the method to calculate Pop III binary star evolutions in \\S \\ref{sec:binary}, numerical calculation methods in \\S \\ref{sec:initial condition}. In \\S 3, we present the results of simulations and argue properties of Pop III compact binaries. We compare Pop III compact binary mergers with Pop I compact binary mergers in \\S \\ref{sec:comparison models}. In \\S 4, we describe the Pop III compact binary merger rates. \\S 5 is devoted to the discussions. In Appendix, we show the details of our numerical methods, the comparison of our results with Hurley's ones and the convergence check of our simulations. We adopt the cosmological parameters of $(\\Omega_{\\Lambda}, \\Omega_{\\rm m}) = (0.6825, 0.3175)$ and the Hubble parameter of $H_0 = 67.11\\ {\\rm km}\\ {\\rm s}^{-1}\\ {\\rm Mpc}^{-1}$~\\citep{Planck2013}. Those who are not interested in the details of the methods in numerical simulations can skip \\S 2. ", "conclusions": "In this paper, we do not include the magnetic braking since Pop III stars are formed from the primordial no metal gas. However, there are discussions against this assumption. Among them, the enhancement of the magnetic field during the star formation has been well studied using the numerical simulations \\citep[e.g.,][]{Maki2004,Machida2008b,Machida2013}. According to their results, the turbulent motions driven in the galaxy formation might increase the magnetic field up to $< 10^{-6}$ G \\citep{Schleicher2010,Sur2010,Schober2012,Turk2012}, which is similar value to that in molecular clouds of the galaxy. If this is the case, we should include the effect of magnetic braking, which will give rise $\\rm Err_{sys}\\neq 1$. \\cite{Dominik2012} discussed the metallicity dependence of the compact binary merger by focusing on Pop I stars with metallicities ${\\rm Z_{\\odot}}$ and $0.1\\ {\\rm Z_{\\odot}}$. Here, we compare our results for the coalescing Pop III BH-BHs with those in \\cite{Dominik2012} and discuss the implications of our results~(see Table~\\ref{dominik}). In Table~\\ref{dominik}, the second and third columns show the results of \\cite{Dominik2012} for metallicity $\\rm Z_\\odot$ and $\\rm 0.1\\ Z_\\odot$ stars. Here, Models A and B correspond to the standard case of submodels A and B in \\cite{Dominik2012}. The last column show our results for Pop III binaries, where we take the fiducial parameter values: $\\rm Err_{sys}= 1$ and $\\rm SFR_p=10^{-2.5}\\ \\msun\\ yr^{-1}\\ Mpc^{-3}$. As in Table \\ref{dominik}, \\cite{Dominik2012} suggested that for Pop I stars with $Z = {\\rm Z_{\\odot}}$, the merger rate of the BH-BHs becomes $8.2\\ (1.9)\\times 10^{-8}\\ {\\rm events\\ yr^{-1}\\ Mpc^{-3}}$ in their Model A~(Model B), respectively, while for Pop I stars with $Z = 0.1\\ {\\rm Z_{\\odot}}$, it becomes $7.33\\ (1.36) \\times 10^{-7}\\ {\\rm events\\ yr^{-1}\\ Mpc^{-3}}$ in their Model A~(Model B), respectively. Here we assume that the galaxy with metallicity $Z=0.1\\ {\\rm Z}_\\odot$ has the hypothetical number density $\\sim 10^{-2}\\ \\rm Mpc^{-3}$ in order to compare the results. On the other hand, for our Pop III BH-BHs, the expected merger rate is $2.5\\ (1.2)\\times 10^{-8}\\ {\\rm events\\ yr^{-1}\\ Mpc^{-3}}$ for the flat (Salpeter) IMF, respectively. This means that the merger rate of the Pop III BH-BHs is smaller than that of Pop I. However, for Pop I BH-BHs with metallicity $Z={\\rm Z}_\\odot\\ (0.1\\ {\\rm Z}_\\odot)$, the typical chirp mass is 6.7~(13.2) $\\msun$, while for Pop III ones it is $\\sim 30\\ \\msun$ as we showed in Fig.~7. Since the detection range of merger events increases in proportion to $M_{\\rm chirp}^{5/6}$, the detection rate increases in proportion to $M_{\\rm chirp}^{5/2}$. Therefore, the detectable event rate of the Pop III BH-BHs is 13~(6) times larger for the flat (Salpeter) IMF than that of Pop I with $Z={\\rm Z}_\\odot$ in Model A, while it is 0.26~(0.13) times larger than that of Pop I with $Z=0.1\\ {\\rm Z}_\\odot$, if the galaxy consists of stars with $Z=0.1\\ {\\rm Z}_\\odot$. In Model B, these numbers become 56~(26) for $Z={\\rm Z}_\\odot$ case and 1.4~(0.7) for $Z=0.1\\ {\\rm Z}_\\odot$ case, respectively. Thus, for the fiducial parameters of Pop III binaries, the contribution of Pop III BH-BHs is comparable to or larger than that of Pop I with $Z={\\rm Z}_\\odot$ and $0.1\\ {\\rm Z}_\\odot$, since the major part of a galaxy does not necessarily consist of Pop I stars with $Z=0.1\\ {\\rm Z}_\\odot$. \\begin{table*} \\caption[]{The comparison of the merger rate density of the BH-BHs and typical chirp mass between previous studies and our study. The second and third columns show the results of \\cite{Dominik2012} for metallicity $\\rm Z_\\odot$ and $\\rm 0.1\\ Z_\\odot$ stars. Here, Models A and B correspond to the standard case of submodels A and B in \\cite{Dominik2012}. The last column show our results for Pop III binaries. Here, we take the fiducial parameter values: $\\rm Err_{sys}= 1$ and $\\rm SFR_p=10^{-2.5}\\ \\msun\\ yr^{-1}\\ Mpc^{-3}$.} \\label{dominik} \\begin{tabular}{l c c c} % \\hline & Z$_\\odot$ & 0.1 Z$_\\odot$ & Pop III\\\\%\\midrule \\hline Model A [$10^{-8}\\ {\\rm events\\ yr^{-1}\\ Mpc^{-3}}$] & 8.2 & 73.3 & 2.5 (flat)\\\\ Model B [$10^{-8}\\ {\\rm events\\ yr^{-1}\\ Mpc^{-3}}$] & 1.9 & 13.6 & 1.2 (Salpeter)\\\\ chirp mass [$\\msun$] & 6.7 &13.2 & 30\\\\ \\hline \\end{tabular} \\end{table*} While \\cite{Dominik2012} did not take account of the evolution of the star formation rate, \\cite{Dominik2013} adopted a certain model of the star formation rate~\\citep[Eq.~1 of][]{Dominik2013} and the metallicity $Z$ evolution~\\citep[Eqs. 3 to 5 of][]{Dominik2013} to compute the cumulative redshift distribution of the coalescing compact binaries. They also took into account the lower metal stars such as Pop II and even those with $Z<10^{-4} \\ {\\rm Z}_\\odot$, but not completely metal-free stars, Pop III.1 and Pop III.2 stars. The star formation rate expressed by Eq.~(1) in \\cite{Dominik2013} is completely different from the one shown in Fig.~8 of the present paper. In the latter case, the star formation rate at $z=0$ is zero, while in the former case, it is the present star formation rate of our Galaxy which is not zero. In Fig.~6 of \\cite{Dominik2013}, they show the cumulative merger rate as a function of redshift $z$ for different four models which corresponds to Fig.~10 of the present paper. In our Fig.~10, for the second~(third) generation gravitational wave detectors, $ z\\sim 0.3\\ (3)$ is the detection range, respectively. In Fig.~6 of \\cite{Dominik2013}, information on the detectability is not available, since the chirp mass distribution function is not available. Assuming it is similar to that in \\cite{Dominik2012}, the merger rate for the second and third generation gravitational wave detectors is either higher or lower than Fig.~10 of the present paper taking into account that the chirp mass of Pop I and Pop II BH-BHs is smaller than that for Pop III.1 and Pop III.2 BH-BHs. If the detection rate of the coalescing Pop I and Pop II BH-BHs is lower than that of Pop III, it might be possible to confirm the existence of the massive Pop III stars by detecting gravitational waves from their remnant black hole and identifying the typical chirp mass $\\sim 30\\ \\msun$ and the cumulative redshift distribution which depends almost only on the cosmological parameters. On the other hand, if the detection rate of the coalescing Pop I and Pop II BH-BHs is higher, Pop III BH-BHs contribute only some parts of the gravitational wave events of BH-BHs. If the detected number is $\\sim 10^4$ for the third generation detector like ET, we might select a Pop III BH-BH from its mass and be able to draw the cumulative redshift distribution function to confirm the existence of Pop III stars. In any case, it is needless to say that there are many undetermined parameters and distribution functions for the Pop III population synthesis so that more theoretical study on the evolution and initial conditions of Pop III binaries including the star formation rate is urgent." }, "1402/1402.0959_arXiv.txt": { "abstract": "Neutrino emissions from electron/positron capture on the deuteron and the nucleon-nucleon fusion processes in the surface region of a supernova core are studied. These weak processes are evaluated in the standard nuclear physics approach, which consists of one-nucleon and two-nucleon-exchange currents and nuclear wave functions generated by a high precision nucleon-nucleon potential. In addition to the cross sections for these processes involving the deuteron, we present neutrino emissivities due to these processes calculated for typical profiles of core-collapsed supernovae. These novel neutrino emissivities are compared with the standard neutrino emission mechanisms. We find that the neutrino emissivity due to the electron capture on the deuteron is comparable to that on the proton in the deuteron abundant region. The implications of the new channels involving deuterons for the supernova mechanism are discussed. ", "introduction": "\\label{sec:introduction} The neutrinos play pivotal roles in core-collapse supernovae and their subsequent evolution to neutron stars. Neutrino reactions in the dense matter % of a supernova core are crucial for understanding the supernova explosion mechanism, which is still elusive despite extensive studies over decades. It is therefore essential to identify all neutrino processes, both neutrino-emission and neutrino-absorption processes, that can be important in the supernova environment. Failing to include all the relevant neutrino processes in supernova modeling may have significant consequences for the theoretical understanding of the supernova explosion. The emission of neutrinos acts as a cooling mechanism of the central core, and the addition of any neutrino emission channels that have not been considered so far could affect this cooling mechanism. A portion of the emitted neutrinos are subsequently absorbed by the material behind the shock wave and thereby act as a heating agent. Additional sources of neutrinos enhance this neutrino-heating mechanism and may help the revival of the stalled shock wave and lead to a successful modeling of supernova explosion \\citep{bet90,kot06,jan07}. The neutrinos emitted gradually ($\\sim$20 s) from a nascent neutron star (proto-neutron star) in a supernova explosion, can be detected as supernova neutrinos at terrestrial neutrino detectors, like in the case of SN1987A \\citep{suz94}, and % can be a useful source of information about the neutrino emission mechanisms. Recent calculations have shown that deuterons, tritons and $^{3}$He can appear copiously in the regions between the supernova core and the shockwave \\citep{sumroe08,arc08,hemp12}. These light elements have so far not been included in the tables of equation of state (EOS)~\\citep{lat91,she98a,she98b} that are routinely used in supernova simulations where the nuclear species are limited to the proton, neutron, $^{4}$He and one ``representative heavy nucleus\" that is assumed to simulate the roles of all heavy nuclei. The light elements with mass number $A\\!=\\!2\\,{\\rm and}\\,3$ can be abundant in hot and moderately dense matter ($ < 10^{13}g/cm^{3}$) under nuclear statistical equilibrium~\\citep{sumroe08,arc08,hemp12,furusawa13a} and should be considered in studying the supernova mechanism. They appear in the heating region behind a shockwave, and also in the cooling region at the surface of a proto-neutron star; their appearance gives a new contribution to the neutrino opacity. For example, ~\\citet{sumroe08} showed in a snapshot 150 msec after the bounce that the deuteron mass fraction amounts to about 10\\% % in the neutrino-emitting cooling regions at densities $\\sim$10$^{11}$-10$^{12}$ g/cm$^{3}$. Neutrino processes in this cooling region are essential for determining the flux and spectra of emitted neutrinos, which in turn affect the efficiency of neutrino heating behind the shockwave. Nakamura et al.~\\citep{nakamura09} investigated neutrino absorption on deuterons as an additional heating mechanism on top of the neutrino reactions on nucleons and $^{4}$He, while Arcones et al.~\\citep{arc08} studied neutrino reactions on tritons and $^{3}$He to evaluate their influences on the neutrino spectra at the outer layer of a proto-neutron star. According to Refs.~\\citet{sumroe08,arc08}, the deuteron fraction can be larger than the proton fraction in part of the neutrino-sphere region between the shock wave and the proto-neutron star surface. This indicates that weak-interaction deuteron breakup may play a significant role in neutrino emission processes, possibly altering the conventional understanding of the role of the protons in the neutrino-emission processes as well as the neutronization of matter. An additional reaction to be investigated in this work is deuteron formation in nucleon-nucleon scattering, which also leads to neutrino emission. Although both these neutrino emission processes certainly exist on top of the conventional neutrino-emission processes, they have so far not been considered in supernova simulations. In this article we study neutrino emissions from \\underline{d}euteron \\underline{b}reakup/\\underline{f}ormation processes (DBF for short) in the surface region of a proto-neutron star, where neutrino emissions act as a cooling mechanism. We present the first evaluation of the neutrino emissivities from electron/position-capture on the deuteron (deuteron breakup) and from the nucleon-nucleon weak fusion processes (deuteron formation); see (\\ref{eqn:eled})-(\\ref{eqn:pnd}) below. The neutrino emissivities arising from DBF will be compared with those coming from the ``conventional\" processes; by ``conventional\" processes we mean the neutrino emission processes which have been previously considered in the literature, and which are listed in (\\ref{eqn:elep})-(\\ref{eqn:epem}) below. The neutrino emissivities due to DBF reported here are expected to be useful for numerical simulations of supernova explosion and proto-neutron star cooling. Theoretical treatments of electroweak processes in two-nucleon systems % are well developed. For low-energy neutrino-deuteron reactions, serious efforts to reduce theoretical uncertainties have been made in order to analyze data from the Sudbury Neutrino Observatory~\\citep{nsk01,net02}. One approach is the standard nuclear physics approach (SNPA) that involves nuclear wave functions derived from high-precision phenomenological nuclear potentials, and one-nucleon and two-nucleon electroweak currents. SNPA has been well tested by analyses of photo-reactions, electron scattering, and muon capture on the two-nucleon systems~\\citep{doi90,tamu92,sato95}. Another theoretical approach, effective field theory (EFT) consisting of nucleons and pions, has been developed and applied to low-energy electroweak processes~\\citep{kubo04}. Both methods essentially agree with each other for low-energy electroweak processes in the two-nucleon systems. The $pp$-fusion process, $pp\\to d e^- \\nu_e$, is one of such processes relevant to this work. This reaction has been studied with both SNPA and EFT, and good agreement between the two methods has been found~\\citep{sch98,tsp03}. Another nucleon-nucleon fusion process relevant to this work is neutron-neutron fusion, which was previously studied with EFT~\\citep{ando06}. In the present work we adopt SNPA. This article is arranged as follows. In section \\ref{sec:reaction} we discuss neutrino emissions via DBF relevant to the supernova environment. The theoretical framework for calculating the cross sections for neutrino emission via DBF and the corresponding neutrino emissivities are outlined in section \\ref{sec:formulation}, and the numerical results are presented in sections \\ref{sec:cros} and \\ref{sec:SecEmissivity}. The implications and a summary of these results for the supernova mechanism are discussed in section~\\ref{sec:summary}. % ", "conclusions": "\\label{sec:summary} It was pointed out in Refs.~\\citet{sumroe08} and \\citet{arc08} that, in addition to deuterons, tritons can also have large abundances in high density regions (10$^{11}$--10$^{14}$ g/cm$^{3}$), where the electron fraction $Y_e$ is low and the temperature $T$ is high. This suggests the possible importance of neutrino emissivities involving the triton or ``triton\" (triton-like three-nucleon correlation in dense matter). In the present work, however, we have not considered these effects. Our study here is based on the spherical (1D) configurations of supernovae. It would be interesting to study neutrino emission with the abundance of light elements in 2D/3D profiles. Hydrodynamical instabilities can generate non-spherical distribution of matter in the cooling region around a proto-neutron star and the heating region behind a stalled shock wave. Since the density, temperature and electron fractions can have wider ranges in 2D/3D profiles, there may be regions of high deuteron abundance that cannot be found in the 1D profile. The existence of deuterons in the heating regions can contribute to the additional source of heating as studied by \\citet{nakamura09}. It thus seems interesting to study the possible effects of the neutrino emission and absorption channels involving the deuteron in multi-D supernova explosion simulations. In principle, one must study effects of all neutrino processes by solving the neutrino transfer and hydrodynamics, with detailed information on composition from the equation of state of supernova matter. A study along this line has been recently made~\\citep{furusawa13}, taking into account the neutrino processes involving neutrino absorption on the deuteron and the light element abundance. We now summarize. Neutrino emissions from $e^\\mp$-capture on the deuteron and from deuteron formation in nucleon-nucleon weak-fusion processes have been studied as new neutrino emission mechanisms in supernovae. These weak processes are evaluated with the standard nuclear physics approach, which consists of the one-nucleon impulse current and two-nucleon exchange current and nuclear wave functions derived from high-precision phenomenological $NN$ potentials. It is found that the contribution of the two-nucleon meson-exchange current is only a few \\% for the $e^\\mp$-capture reactions, while it can be as large as the one-nucleon current contribution for the NN fusion reaction at higher energies. The consequences of these new neutrino-emission channels have been examined for representative profiles of core-collapse supernovae at 150 ms after core bounce. The emissivity due to the $e^\\mp$ capture reaction on the deuteron is found to be smaller than that on the free nucleon. Therefore, as Fig.~\\ref{fig:ecap-ratio} indicates, the total neutrino emissivity due to electron capture on protons and deuterons is suppressed when an appreciable amount of protons in a supernova are bound inside deuterons. This results in a smaller neutrino luminosity and the lower efficiency of neutrino heating behind a stalled shock wave. Therefore, this new process may contribute unfavorably towards a successful supernova explosions. It might lead to a slower speed of the deleptonization and, hence, a slower evolution of nascent proto-neutron stars. On the other hand, as seen in Figs.~\\ref{fig:ecap-out} and \\ref{fig:nn-out}, neutrino emission via deuteron formation can be comparable to nucleon-nucleon bremsstrahlung in the outer region. This implies that there might exist situations in which the deuteron-formation weak processes are the main channels for neutrino emission. In the inner core region, where the electrons are highly degenerate (high densities at low temperatures), pair-production via $e^-e^+$ annihilation is suppressed, making nucleon-nucleon bremsstrahlung a main channel to produce $\\nu\\bar{\\nu}$ pairs among the conventional processes\\citep{suz93, bur06}. % Meanwhile, ``deuteron\" formation processes in $NN$ scattering can have large rates for $\\nu_\\mu$ and $\\nu_\\tau$ emissions, a feature that may have significant consequences for the cooling of compact stars. Furthermore, the possible modification of the energy spectra of $\\nu_e$, $\\nu_\\mu$ and $\\nu_\\tau$ due to ``deuteron\" formation may influence supernova nucleosynthesis~\\citep{woo90,yoshi04} and the terrestrial observation of supernova neutrinos \\citep{nakazato13}. On the other hand, the possible increase of the $\\nu_\\mu$ and $\\nu_\\tau$ fluxes due to ``deuteron\" formation hardly affects the heating process behind a shockwave, because these low-energy $\\nu_\\mu$ and $\\nu_\\tau$ interact with stellar matter only through the NC. As explained earlier, the ``deuteron\" here stands for a tensor-correlated $NN$ pair that may persist even in dense nuclear matter. A detailed study of deuteron-like two-nucleon tensor correlation in dense matter seems well warranted, but it is beyond the scope of our present exploratory work." }, "1402/1402.4057_arXiv.txt": { "abstract": "{We describe the late spectral variability and flux evolution of TDF Sw J1644+57, a Tidal Disruption Flare which left the typical potential trend proportional to $t^{-5/3}$ in 2012, maintaining a quiescent flux until nowadays. Sixteen X-ray observations of ESA satellite \\emph{XMM-Newton} have been used in this study, including the one performed on 17$^{\\rm th}$July, 2013. A search for optical emission in BOOTES/CASANDRA database has been performed too. Late X-ray fluxes show that the source flux decline does not follow the expected TDF trend at the time of the last \\emph{XMM-Newton} observation. Moreover, the spectra fitting parameters, in particular the neutral hydrogen column density, N$_{\\rm H}$, and the power-law index, $\\Gamma$, indicate that the source darkening has diminished and that the spectral shape has flattened with time. The disruption of the star could have come to an end. Nevertheless, a quiescent X-ray flux continues. Evidence for a quiescent X-ray flux is presented.} \\resumen{Presentamos la variabilidad espectral tard\\'ia y la evoluci\\'on de flujo del TDF Sw J1644+57, un fen\\'omeno transitorio probablemente producido por el desgarramiento de una estrella por fuerzas de marea, el cual en 2012 abandon\\'o la tendencia caracter\\'istica proporcional a $t^{-5/3}$ para mantener un flujo estable en quietud hasta la actualidad. Para este trabajo se han empleado diecis\\'eis observaciones del sat\\'elite de la ESA \\emph{XMM-Newton}, incluida la \\'ultima observaci\\'on (17 de Julio de 2013). Asimismo, se ha llevado a cabo una b\\'usqueda en la base de datos de BOOTES/CASANDRA a fin de detectar la contrapartida \\'optica. Los flujos tard\\'ios en rayos X muestran que el decaimiento de flujo de la fuente ya no sigue la tendencia esperada para un TDF. Adem\\'as, los par\\'ametros del ajuste espectral, en particular la densidad de columna de hidr\\'ogeno neutro, N$_{\\rm H}$, y el \\'indice de la ley de potencias, $\\Gamma$, indican que el oscurecimiento de la fuente ha disminuido y que los espectros se han aplanado con el tiempo. El desgarramiento de la estrella podr\\'ia haber llegado a su fin. Sin embargo, el flujo de rayos X en quietud contin\\'ua. Hay indicacaciones de emisi\\'on en quietud.} \\addkeyword{Galaxies: Active} \\addkeyword{ISM: Jets and outflows} \\addkeyword{Stars: Flare} \\addkeyword{X-rays: Galaxies} \\begin{document} ", "introduction": " ", "conclusions": "\\label{sec:Results} \\begin{asparaitem} \\item The N$_{\\rm H}$ has decreased as Figure~\\ref{fig:NH} shows, between day 188 and 549, which indicates that most of the clouds of gas and dust around the disrupted star have dissipated. The final number of 0.056$\\times10^{22}$cm$^{-2}$ is very close to the galacticN$_{\\rm H}$, and implies that the phenomenon is less energetic now. \\item The spectral shapes have flattened with time, as Figure~\\ref{fig:Flattening} reveals when we compare the spectrum of the first observation, performed on day 3, with the last one, taken on day 842. Besides, in late observations the maximum flux occurs at a lower energy. \\begin{figure}[!t] \\includegraphics[width=\\columnwidth]{Intrinsic_Flux} \\caption{Evolution of the intrinsic flux (0.5$-$8.0 keV). \\emph{Swift} observations are drawn as black circles. The most recent \\emph{Chandra} observation is shown as a black square. The fluxes obtained in this work are presented as orange circles. In red, at the bottom right, our latest \\emph{XMM-Newton} observation. Adapted from \\citet{2013ApJ...767..152Z}} \\label{fig:Intrinsic_Flux} \\end{figure} \\item The intrinsic fluxes between 0.5 and 8.0 keV calculated from the sixteen observations of the \\emph{XMM-Newton} Archive, are consistent with previous works about Sw J1644+57. See Figure~\\ref{fig:Intrinsic_Flux}. \\item Between late 2011 and 2012, the source X-ray flux started to decline, leaving the characteristic potential tendency of a TDF in 2012. Consequently, the black hole has probably used up the star that provided dust and gas clouds, or its remains have escaped from the orbit around the hole \\citep{1988Natur.333..523R}. \\item The object behavior has changed from an X-ray emission compatible with a TDF, to a quiescent emission that remains until nowadays and that is higher than Sagittarius A$^{*}$. We ignore if the accretion already existed before \\emph{Swift} triggered. Thus, the source could be the beginning of an Active Galactic Nucleus, AGN, with the jet faced on, i.e., a mini-blazar. Additional X-ray observations are needed to confirm it. \\item We do find no transient optical emission (brighter than 10$^{\\rm th}$ mag) at the CASANDRA all-sky images taken at the different BOOTES robotic astronomical stations world wide in the period 2009-2013 when data is available \\citep{2012ASInC...7..313C}. \\end{asparaitem}" }, "1402/1402.1444_arXiv.txt": { "abstract": "We formulate a model of pulsar spin evolution (braking, inclination angle evolution and radiative precession) taking into account the non-rigidity of neutron star rotation. We discuss two simple limiting cases of this model and show that the evolution of the inclination angle substantially depends on the model of crust-core interaction. The non-rigidity of core rotation accelerates the inclination angle evolution and makes all pulsars evolve to the orthogonal state. The size of the effect depends on the amount of differentially rotating matter and mechanism of its interaction with the rest of the star. Since the rapid evolution of the inclination angle apparently contradicts the observational data, our results may be used as an additional test for the theories of the cores of neutron stars. ", "introduction": "A rotating magnetized neutron star, if it has perfectly spherical shape, can be characterized by two vectors: angular velocity vector $\\vec{\\Omega}$ and magnetic moment vector $\\vec{m}$. During the life of a neutron star the magnitudes of these vectors as well as the inclination angle $\\chi$ formed by these vectors evolve in time. This evolution is caused by electromagnetic torque acting on rotating magnetized star. A neutron star is surrounded by a large magnetosphere. Strictly speaking, the electromagnetic torque can be calculated only by using a self-consistent theory of this magnetosphere which despite the achieved progress is far from complete at present \\cite{Spitkovsky2008}. The problem becomes even more complicated if one wishes to take into account the internal structure of neutron star. Neutron stars are not perfectly rigid. They contain a liquid core. There are two main mechanisms of crust-core interaction: coupling through the magnetic field and viscosity \\cite{Easson1979}. The first mechanism is much more effective and, if it works, then the protons, electrons and normal neutrons in the core can be considered as rigidly rotating with the crust. In contrast, the superfluid neutrons which are believed to be present in neutron star's core \\cite{YakovlevLevenfishShibanov1999} are decoupled from the rest of the matter and interact with it only by weak mutual friction force (vortex-mediated interaction) \\cite{HallVinen1956}. The configurations when the magnetic field does not penetrate the core have been discussed in the literature \\cite{PonsMirallesGeppert2009,GourgouliatosEtAl2013}. The expulsion of magnetic field can be caused, for example, by type-I superconductivity of core protons. Despite the fact that the type-II proton superconductivity is more likely from the point of view of microscopic calculations (see, however, \\cite{BuckleyMetlitskiZhitnitsky2004a}), its implications for the neutron star rotation dynamics are not so clear. The coexistence of proton magnetic fluxoids with neutron superfluidity seems to be inconsistent with the observed long periods of precession in some neutron stars \\cite{Link2006}. In the present paper we consider in detail the two limiting cases of strong and weak coupling between the crust and the core of a neutron star, which correspond to the two physical cases discussed above. ", "conclusions": "One can see that the evolution of the inclination angle dramatically differs for different crust-core interaction models. The rate of the evolution depends on the amount of the matter which rotates differentially. Moreover, all trajectories sharply turn upwards when the action of anomalous torque becomes comparable with the action of normal torque. These features should be included in the studies of population synthesis of pulsars. They can also be used as additional tests for the theories of neutron star interiors." }, "1402/1402.3394_arXiv.txt": { "abstract": "Using data from the SDSS-DR7, including structural measurements from 2D surface brightness fits with GIM2D, we show how the fraction of quiescent galaxies depends on galaxy stellar mass $M_*$, effective radius $R_e$, fraction of $r-$band light in the bulge, $B/T$, and their status as a central or satellite galaxy at $0.010.3$ are excluded. For satellite galaxies, the quiescent fraction is always larger than that of central galaxies, for any combination of $M_*$, $R_e$ and $B/T$. The quenching efficiency is not constant, but reaches a maximum of $\\sim 0.7$ for galaxies with $9<\\log(M_*/M_\\Sun)<9.5$ and $R_e<1$ kpc. This is the same region of parameter space in which the satellite fraction itself reaches its maximum value, suggesting that the transformation from an active central galaxy to a quiescent satellite is associated with a reduction in $R_e$ due to an increase in dominance of a bulge component. ", "introduction": "It is now well known that galaxies in the local Universe can generally be separated into two broad classifications: blue, star-forming, disk galaxies and red, elliptical and lenticular galaxies with little or no recent star formation \\citep[e.g.][]{Strateva,Kauffmann3,Baldry,Balogh2}. The star formation rate (SFR) of blue galaxies is well correlated with their stellar mass (M$_*$), forming the ``main sequence'' or ``star-forming sequence'' in the SFR-M$_*$ plane \\citep{Brinchmann,Salim,Gilbank,MBW+}. Surveys show that the star-forming sequence has persisted out to z $=$ 1 \\citep{Bell,Willmer,Noeske} and possibly z $=$ 2 and further \\citep{Brammer,Taylor2}. Most tellingly, the number density of red, passive galaxies has roughly doubled since z $=$ 1, while the number of blue galaxies has remained nearly constant \\citep{Bell,Faber,Brown,Goncalves,Muzzin}. This motivates the search for processes that cause blue, star-forming galaxies to transition to red, non-star-forming ones. This transition has become known as ``quenching''. It has also been established that the galaxy population varies with environment, in the sense that low-density environments are dominated by less massive, blue star-forming galaxies while denser groups and clusters tend to host massive red elliptical and lenticular galaxies \\citep[e.g.][]{Oemler,DavisGeller,Dressler,Baldry2,Pasquali+10}. Interestingly, the star-forming sequence itself appears to be largely independent of environment, while the fraction of star-forming galaxies decreases dramatically with increasing local density \\citep[e.g.][]{Carter,Balogh2,Wolf,P+08,Peng10}. \\begin{table*} \\begin{tabular}{cccccccccccccc} \\hline VAGC& Yang & Simard & RA & Dec & z&R$_e$&B/T&$M_*$&SFR& $m_r$&Central&C&V$_{\\rm max}$\\\\ ID & ID & Rec. no. & \\multicolumn{2}{c}{(J2000)} && (kpc)& &($10^{10} M_\\Sun$)& ($M_\\Sun/$yr)&(mag)&flag&&($10^{6}$Mpc$^3$)\\\\ \\hline 234667 & 1129345 & 432305 & 161.86430 & 13.889750 & 0.010 & 1.00 & 0.23 & 0.02 & 0.04 & 16.34 & 1 & 0.93 & 0.315\\\\ 204290 & 958798 & 1095350 & 173.42712 & 50.449570 & 0.010 & 1.47 & 0.36 & 0.05 & 0.06 & 15.73 & 1 & 0.88 & 0.761 \\\\ 114459 & 431764 & 897676 & 151.88120 & 57.729069 & 0.010 & 1.83 & 0.01 & 0.006 & 0.04 & 16.95 & 1 & 0.89 & 0.098 \\\\ 256139 & 1219961 & 975660 & 178.49693 & 12.198710 & 0.010 & 1.87 & 0.17 & 0.08 & 0.15 & 15.08 & 1 & 0.86 & 1.875 \\\\ \\hline \\end{tabular} \\caption{A sample of entries in our catalogue, available in the online version. These data are simply matched entries from the published catalogues of \\citet{Blanton2005}, \\citet{Brinchmann}, \\citet{Simard}, and \\citet{YangDR7}. Columns 1-3 provide the ID for simple matching with \\citet{Blanton2005}, \\citet{YangDR7} and \\citet{Simard}, respectively. All catalogues were matched on RA, Dec and redshift (columns 4-6). Column 7 is the circularized half-light radius from the \\citet{Simard} bulge/disk decomposition model, with free index $n$ for the bulge component; column 8 is the corresponding fraction of light in the bulge. Columns 9 and 10 are the stellar mass and SFR from \\citet{Brinchmann}, and column 11 is the $r$-band petrosian magnitude from the original SDSS catalogue. Column 12 is a flag that indicates whether (1) or not (2) a galaxy is the most luminous in its halo according to the \\citet{YangDR7} catalogue; galaxies with a flag of 1 are considered a ``central'' in our analysis. Finally, column 12 is the spatial completeness from \\citet{Blanton2005}, and column 13 is the selection volume that we compute, based on the absolute magnitude and $R_e$ of the galaxy. Additional columns that may be useful, including some that we make use of in the Appendix, are provided in the online table.\\label{tab-catalogue} }. \\end{table*} There is some evidence to show that galaxy properties depend on the mass of their host, dark matter halo, with distinctly different behaviour for the most massive galaxies in the halo (known as centrals, since they are generally expected to lie near the centre of the halo) and the smaller, satellite galaxies that orbit around them \\citep{Hogg,Berlind,Pasquali+10}. Some \\citep{vdB,Peng12,Woo} suggest that the correlations with environment are due to transformations that take place when a galaxy first becomes a satellite. This is natural, as such galaxies are expected to experience diminished cosmological gas accretion \\citep{Balogh1,Kawata}, ram-pressure with the intrahalo gas \\citep{Gunn,Grebel}, and a different type of interaction with neighbouring galaxies \\citep{Moore1,Moore2} and the tidal field \\citep{Read}. The fraction of quiescent galaxies clearly correlates with galaxy stellar mass and this, combined with the apparent ease with which stellar masses can be computed from photometric data, has motivated modelers and observers to consider stellar mass as the fundamental parameter that determines a galaxy's properties \\citep[e.g.][]{Peng10,Woo}. However, various authors \\citep[e.g.][]{Brinchmann,Kauffmann2,Franx} found that galaxy surface density is more predictive of the specific SFR (sSFR) than stellar mass. Others have shown the correlation is best with velocity dispersion, either as measured directly from dynamics \\citep[e.g.][]{SLH,Wake2,Wake1}, or inferred from the virial relations between mass and size \\citep{Franx}. There are good indications that the trends are driven by conditions in the centre of galaxies, either as inferred from S\\'{e}rsic indices derived from single-profile fits to the surface brightness \\citep{Blanton2003c,Driver,Allen,Schiminovich,Bell2,Donofrio+11}, central surface density \\citep[e.g.][]{Cheung,Fang}, or presence of a central bulge \\citep{Martig,Cappellari2}. While it remains controversial which of these characterisations is most fundamental, it seems clear that galaxy structure, and not just stellar mass, is relevant \\citep[e.g.][]{P+13}. If galaxy properties are primarily determined by some aspect of their structure, rather than their stellar mass alone, it reopens the question of how (or if) galaxies of a given {\\it structure} depend on their environment. The first purpose of this paper is to quantify the quenching efficiency of satellite galaxies as a function of their structure. This leads us to take a more general approach than some of the studies listed above, and consider how the quiescent fraction of central galaxies depends on galaxy stellar mass and radius in the local Universe sampled by the Sloan Digital Sky Survey \\citep[SDSS,][]{SDSS}. All calculations use the $\\Lambda$CDM cosmology consistent with the nine year data release from the WMAP mission, with parameters $h$ = 0.693, $\\Omega_M=0.286$, and $\\Omega_\\Lambda=0.714$ \\citep{WMAP}. ", "conclusions": "We have used a sample of 471,554 galaxies at 0.01 $<$ z $<$ 0.2 selected from the SDSS DR7, to examine how the fraction of quiescent galaxies depends upon stellar mass, effective radius, bulge fraction, and whether a galaxy is central or satellite in the group catalogue of \\citet{YangDR7}. We use the GIM2D parametric fits of \\citet{Simard} to determine the circular half-light radius ($R_e$) of all galaxies in the sample, and the specific star formation rates and stellar masses of \\citet{Brinchmann} to classify galaxies as active (on the star--forming main sequence) or quiescent. We make the following observations: \\begin{enumerate} \\item The fraction of quiescent galaxies is a steeper function of M$_*$/R$_e$ and M$_*$/R$_e^2$ than it is of stellar mass alone. This confirms the findings of many other authors \\citep[e.g.][]{Kauffmann2,Franx,Wake2,Wake1,Cheung}. \\item The quiescent fraction of satellite galaxies is always higher than that of central galaxies, except perhaps at the highest masses where almost all galaxies are quenched centrals. Thus a galaxy's status as satellite or central in its halo plays a role in its evolution. The quenching efficiency \\citep{Peng10}, which attempts to quantify the fraction of satellite galaxies that had their star formation quenched as a satellite, is $\\sim 40$ per cent, approximately independent of stellar mass for $M_*>10^9M_\\Sun$. \\item We presented the quiescent fraction of central galaxies as function of both $M_\\ast$ and $R_e$ in Figure~\\ref{fig:qfd}. This representation makes it clear that the fraction does not depend on stellar mass alone, and that over most of the parameter space it appears to depend most strongly on $\\sim$M$_*$/R$_e^{1.5}$. \\item We showed that the quiescent fraction of central galaxies correlates strongly with $B/T$ in a simple way that largely accounts for the more complex dependence on $R_e$. Galaxies with more than 40 per cent of their light in a bulge component are almost all quiescent, and galaxies less massive than $M_*=10^9M_\\Sun$ are almost all forming stars. Otherwise, the quiescent fraction increases monotonically with increasing B/T at fixed $M_*$. \\item The quenching efficiency is highest for galaxies with 10$^{9}<$ M$_*<$ 10$^{10}$ M$_\\Sun$ and 10$^{-0.2}<$ R$_e<$ 10$^{0.2}$ kpc (Figure~\\ref{fig:qeff}), but appears roughly constant above a certain threshold when taken solely as a function of M$_*$, M$_*$/R$_e$, or M$_*$/R$_e^2$. \\item There are structural differences between satellite and central galaxies, such that, at a given stellar mass, satellite galaxies tend to have smaller $R_e$ and more light in a bulge component. The fraction of satellites is greatest (at about 50 per cent) for galaxies with $9.0<\\log({M_*}/M_\\Sun)<10.0$, and $R_e<2$kpc. Even after accounting for this difference by matching central and satellite galaxies to have the same $M_*$, $R_e$, and B/T, the quenching efficiency remains positive at all masses, and reaches a maximum of 70 per cent for galaxies with $\\log(M_*/M_\\Sun)\\sim 9.4$ and $R_e\\sim$1 kpc. \\end{enumerate} For central galaxies, it is appealing to consider a probability of star formation activity ending that depends on their internal structure. In particular the strong correlation between quiescent fraction and $B/T$ points to the dominance of a central bulge component as the driving parameter. We showed that, amongst star--forming central galaxies with $M_*<10^{11}M_\\Sun$, the average sSFR decreases with increasing $M_*/R_e$, qualitatively suporting a picture in which the presence of a bulge reduces the sSFR and increases the probability of being quenched. Alternatively, the data could be explained if the probability of quenching depends only on $M_*$, but is accompanied by a change in structure, with $R_e$ shrinking by a factor $\\sim 2$ due to the increased prominence of a bulge when star formation is terminated. For satellite galaxies, that fact that the quenching efficiency is a maximum for the same combination of $M_*$ and $R_e$ at which the satellite fraction itself peaks almost certainly means that their $R_e$, and perhaps their mass, changes upon quenching." }, "1402/1402.4782_arXiv.txt": { "abstract": "{ The knowledge of atmospheric parameters --~such as temperature, pressure, and humidity~-- is very important for a proper reconstruction of air showers, especially with the fluorescence technique. The Global Data Assimilation System (GDAS) provides altitude-dependent profiles of these state variables of the atmosphere and several more. Every three hours, a new data set on 23~constant pressure level plus an additional surface values is available for the entire globe. These GDAS data are now used in the standard air shower reconstruction of the \\pao. The validity of the data was verified by comparisons with monthly models that were averaged from on-site meteorological radio soundings and weather station measurements obtained at the Observatory in \\mal. Comparisons of reconstructions using the GDAS data and the monthly models are also presented. Since GDAS is a global model, the data can potentially be used for other cosmic and gamma ray detectors. Several studies were already performed or are underway for several locations worldwide. As an example, a study performed in Colorado as part of an Atmospheric \\rnd for a possible future cosmic ray observatory is presented. } ", "introduction": "} A cosmic ray particle entering the atmosphere can initiate an extensive air shower. The secondary shower particles excite nitrogen molecules in the air which emit a characteristic, isotropic emission in the UV range as part of their de-excitation process. The light can then be observed by an optical telescope, typically consisting of a collecting mirror and a camera. To properly reconstruct the properties of such air showers, the atmospheric conditions at the site have to be known in order to correct for Rayleigh scattering effects and to estimate the fluorescence yield of the air shower~\\cite{Abraham:2010atmo}. Height-dependent profiles of temperature, pressure and humidity as well as weather conditions near the ground are relevant. The \\pao~\\cite{Abraham:2004auger} is a cosmic ray detector located near \\mal in the Mendoza province in Argentina. It consists of a Surface Detector (SD) array and five Fluorescence Detector (FD) buildings~\\cite{Abraham:2009fd}. Between 2002 and 2010, atmospheric conditions over the Observatory were measured by intermittent meteorological radio soundings. Additionally, ground-based weather stations measure surface data continuously in order to provide the atmospheric parameters to properly reconstruct the measured air showers. In south-east Colorado, several balloon soundings were performed as part of an atmospheric \\rnd project. The aim of this effort was to study possible enhancements and performance improvements for the \\pao, as well as explore technological advancements for a possible future ground-based observatory. The ground station used for the soundings was a mobile and slightly advanced version of the equipment used in Argentina. The launches were performed at two sites, the Atmospheric Monitoring Telescope (AMT) and the Distant Raman Laser Facility (DRLF)~\\cite{Wiencke:2011icrc}. The sites are about 40\\,km apart and are both equipped with identical weather stations. Performing radio soundings imposes a large burden, both in terms of funds and manpower. We investigated the possibility of using data from the Global Data Assimilation System (GDAS)~\\cite{GDASinformation}, a global atmospheric model, for the site of the \\pao~\\cite{Abreu:2012gdas,Will:2011icrc}. GDAS data are publicly available free of charge via READY (Real-time Environmental Applications and Display sYstem). Each data set contains all the main state variables as a function of altitude. The data gathered in Colorado were also compared to GDAS data in order to evaluate the possibility to use GDAS also in different locations and for a possible future ground-based cosmic ray detector. ", "conclusions": "} The comparison of GDAS data for the site of the \\pao in Argentina with local atmospheric measurements validated the adequate accuracy of the 3-hourly GDAS data. An air shower reconstruction analysis confirmed the applicability of GDAS for Auger reconstructions and simulations, giving improved accuracy when incorporating GDAS data instead of MM. Also, the value of using an atmosphere-dependent fluorescence description has been demonstrated. For the Colorado \\rnd site, the differences between the measured radiosonde data and GDAS are of the same order as in Argentina, further supporting the general validity of GDAS data as an atmospheric description to be used in current and future cosmic ray observatories. \\vspace*{0.5cm} \\footnotesize{{\\bf Acknowledgment:} We would like to thank the organizers of the workshop \\emph{AtmoHEAD: Atmospheric Monitoring for High-Energy Astroparticle Detectors} in Saclay, France, 2013 for the inspiring meeting. Part of these investigations are supported by the Bundesministerium f\\\"ur Bildung und Forschung (BMBF) under contracts 05A08VK1 and 05A11VK1. Furthermore, these studies would not have been possible without the entire Pierre Auger Collaboration and the local staff of the \\pao.{}}" }, "1402/1402.1391_arXiv.txt": { "abstract": "The oscillations in solar like stars are described in terms of the phase shifts of the eigenmodes from simple sine-waves. We discuss model fitting and inversion techniques based on this representation. We analyse the periodic signatures from the HeII ionisation zone and base of the convective envelope of the CoRoT star HD49933. ", "introduction": " ", "conclusions": "" }, "1402/1402.1672_arXiv.txt": { "abstract": "We present an algorithm for the simultaneous measurement of a pulse time-of-arrival (TOA) and dispersion measure (DM) from folded wideband pulsar data. We extend the prescription from Taylor (1992) to accommodate a general two-dimensional template ``portrait'', the alignment of which can be used to measure a pulse phase and DM. We show that there is a dedispersion reference frequency that removes the covariance between these two quantities, and note that the recovered pulse profile scaling amplitudes can provide useful information. We experiment with pulse modeling by using a Gaussian-component scheme that allows for independent component evolution with frequency, a ``fiducial component'', and the inclusion of scattering. We showcase the algorithm using our publicly available code on three years of wideband data from the bright millisecond pulsar J1824-2452A (M28A) from the Green Bank Telescope, and a suite of Monte Carlo analyses validates the algorithm. By using a simple model portrait of M28A we obtain DM trends comparable to those measured by more standard methods, with improved TOA and DM precisions by factors of a few. Measurements from our algorithm will yield precisions at least as good as those from traditional techniques, but is prone to fewer systematic effects and is without ad hoc parameters. A broad application of this new method for dispersion measure tracking with modern large-bandwidth observing systems should improve the timing residuals for pulsar timing array experiments, like the North American Nanohertz Observatory for Gravitational Waves. ", "introduction": "\\label{intro} The practice of pulsar timing attempts to model the rotation of a neutron star by phase-connecting periodic observations of its pulsed, broadband radio signal. The earliest demonstration of long-term timing observations came relatively soon after the discovery of pulsars~\\citep{Roberts&Richards71, Hewish68}. The scientific merits garnered from pulsar timing span astrophysical fields such as planetary science, the interstellar medium, nuclear physics, gravitational wave physics, and are all well-documented (for a review, see eg. Chapter 2,~\\citet{L&K04}). Pulsar timing and its related experiments have carved out a ``sweet spot'' in the radio frequency regime that naturally emerges as a trade-off between the pulsars' steep power-law spectra at the high-frequency end, and the low-frequency drawbacks arising from the pulsed radio signal having to propagate through the ionized interstellar medium (ISM) and the Earth's ionosphere, as well as having to compete with the diffuse background of the galactic synchrotron continuum. The latter has a spectral index in the 1--10~GHz range of $\\approx-2.8$~\\citep{Platania98}. Population studies have shown that pulsars have an average spectral index around -1.4 at gigahertz frequencies~\\citep{Bates13}. The most relevant ISM effect arises from propagation through a homogeneously ionized medium. Interstellar dispersion alters the group-velocity of the radio signal, retarding the arrival of pulses emitted at a radio frequency $\\nu$ by a time $t_{\\textrm{DM}}$ (relative to an infinite frequency signal) according to the cold-plasma dispersion law, \\begin{equation} \\label{DMlaw} t_{\\textrm{DM}} = K \\times \\textrm{DM} \\times \\nu^{-2}, \\end{equation} where $K \\equiv \\frac{e^2}{2\\pi m_ec} = 4.148808(3) \\times 10^{3}$~MHz$^2$~cm$^{3}$~pc$^{-1}$~sec is called the dispersion constant\\footnote{$K$ is a combination of the electron charge $e$, electron mass $m_e$, and speed of light $c$. It is common practice in the pulsar community to adopt the approximation $K^{-1} = 2.41\\times10^{-4} $~MHz$^{-2}$~cm$^{-3}$~pc~sec$^{-1}$~\\citep{L&K04}, which we have used in~\\S\\ref{m28a} and~\\S\\ref{mcs}.}, and DM is the dispersion measure. The dispersion measure is defined as \\begin{equation} \\label{DM} \\textrm{DM} \\equiv \\int\\limits_l n_e\\,dl, \\end{equation} which is the free-electron column density along the path-of-propagation $l$ to the pulsar. The pulse-broadening effect of multi-path propagation through a turbulent, inhomogeneous ISM, known as interstellar scattering, has an even stronger spectral index $\\approx-4$, and becomes increasingly important at lower frequencies for the highest-DM, farthest pulsars~\\citep{L&K04}. Scattering not only broadens the pulsed signal, but delays an intrinsically sharp pulse by an amount roughly proportional to its width, and so is a source of bias in timing measurements. The determination of dispersion measures and effects from scattering have been non-trivial problems concomitant with timing measurements since the beginning~\\citep{Rankin70, Rankin71}. Nearly all observations taken for (high-precision) pulsar timing experiments are taken within the radio window mentioned above, which lies somewhere in the two decades bounded by about 100~MHz and 10~GHz. The middle decade centered around 1500~MHz seems to be the perennial favorite for timing experiments. Recent developments in pulsar instrumentation and computing over the last 5--10 years have enabled more accurate and sensitive timing measurements. Namely, coherent dedispersion, which completely removes the quadratic time-delay due to a known amount of interstellar dispersion~\\citep{Hank&Rick75}, required significant advances in computer technology before becoming feasible in real-time on a wide-bandwidth signal. Historically, observations that implemented coherent dedispersion were limited by computing resources to a bandwidth of order $\\sim$100~MHz or less, which is less than most receiver bandwidths. Thus, if one wanted to cover a large portion of the pulsar spectrum, either for timing, spectral, or interstellar medium purposes, several adjacent receiver bands had to be observed separately, which often meant asynchronous measurements and non-contiguous frequency coverage. The implementation of real-time coherent dedispersion to large, instantaneously observed bandwidths has led to the regime wherein the receiver bandwidth (BW) is a limiting factor. The first generation of GHz-bandwidth, coherent dedispersion instruments has been proliferating in the pulsar community for the past several years, beginning with the Green Bank Ultimate Pulsar Processing Instrument (GUPPI)\\footnote{\\url{www.safe.nrao.edu/wiki/bin/view/CICADA/NGNPP}} outfitted for the 100-m Robert C. Byrd Green Bank Telescope (GBT)\\footnote{The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.}~\\citep{Duplain08}. GUPPI is an FPGA- and GPU-based system capable of real-time coherent dedispersion of an 800~MHz bandwidth. The smearing $\\delta t_{\\textrm{DM}}$ incurred from incorrectly dedispersing a narrow frequency-channel of bandwidth $\\Delta\\nu = \\frac{\\textrm{BW}}{n_{chan}}$ and center frequency $\\nu_c$ by an amount $\\delta{\\textrm{DM}}$ goes as \\begin{equation} \\begin{split} \\label{DMsmear} &\\delta t_{\\textrm{DM}} \\approx \\frac{2K\\ \\delta{\\textrm{DM}}\\ \\Delta\\nu}{\\nu_c^3} \\\\ &\\qquad \\approx 8.3\\ \\Big(\\frac{\\delta{\\textrm{DM}}}{\\textrm{cm}^{-3}\\ \\textrm{pc}}\\Big)\\Big(\\frac{\\Delta\\nu}{1\\ \\textrm{MHz}}\\Big)\\Big(\\frac{\\nu_c}{1\\ \\textrm{GHz}}\\Big)^{-3}\\ \\mu\\textrm{s}. \\end{split} \\end{equation} This equation demonstrates why it was difficult to obtain precise, broadband measurements from millisecond pulsars (MSPs) prior to coherent dedispersion. Incoherent dedispersion shifts individual frequency channels of the data based on the assumed DM without compensating for dispersion within each channel. This means that $\\delta{\\textrm{DM}}$ is equivalent to the full, true DM and so the dispersive smearing $\\delta t_{\\textrm{DM}}$ can easily be a significant fraction of the pulsar's spin period ($P_{s} \\lesssim 10$~ms for MSPs). To mitigate this problem, a large number of filterbank channels is needed, but it comes at the expense of a degraded time resolution due to Nyquist sampling constraints. In turn, poorer time resolution means a poorer timing precision. Moreover, tracking the dispersion measure changes in MSP observations is necessary for minimizing the timing residuals used in gravitational wave searches with a pulsar timing array (PTA)~\\citep{You07}. Equation~\\ref{DMlaw} says that an incorrect DM of $10^{-3}$~cm$^{-3}$~pc at 1500~MHz introduces a delay of $\\sim2 \\mu$s relative to infinite frequency, which is well above the timing quality desired by PTA experiments. As part of the Parkes Pulsar Timing Array\\footnote{\\url{www.atnf.csiro.au/research/pulsar/array/}} project~\\citep{Hobbs13}, \\citet{Keith13} developed a method to correct for inaccurate dispersion measures based on modeling the multi-frequency timing residuals. However, the authors also postulate that more accurate DM variations could be measured from wideband receivers, which ameliorate the difficulties of aligning pulsar data taken with different receivers in different epochs. The desire for very broadband pulsar observations (i.e. with significantly high fractional bandwidths, $\\gtrsim$1) necessitates new, unique receiver designs that can cover much of the frequency range once concatenated from disjoint observations. Wideband receivers and their complimentary, real-time coherent dedispersion backends will quickly facilitate developments in all realms of pulsar astrophysics, including studies of the pulsar spectrum, magnetosphere, and ISM properties. One such instrument, called the Ultra-Broad-Band (UBB) receiver and associated backend\\footnote{\\url{www3.mpifr-bonn.mpg.de/staff/pfreire/BEACON.html}}, has been recently installed at the Effelsberg 100-m Telescope and covers a frequency range from $\\sim$600~--~$\\sim$3000~MHz. However, the current method for making pulse time-of-arrival (TOA) measurements that is used almost ubiquitously in the pulsar timing community does not use all of the information contained in new broadband observations. In summary, the protocol employs frequency-averaged pulse profiles as models of the pulsar's signal for entire receiver bands, which ignores any profile evolution intrinsic to the pulsar or imposed by the ISM. Both intrinsic profile evolution and DM changes are usually taken into account in the timing model for the pulsar's rotation, but there is no modeling of the effects from scattering or scintillation. Arbitrary phase-offsets (known as ``JUMPs'') are introduced to align disparate template profiles that are used to measure TOAs from different frequency bands. Multi-channel TOAs are also parameterized by both a quadratic delay (proportional to the DM) and an arbitrary function to remove residual frequency structure from otherwise unmodeled profile evolution. Additionally, many multi-channel TOAs that are adversely affected by scintillation are often cut from being included in the timing model fit due to their inaccurately determined uncertainties; in effect, one is throwing away portions of the band that do contain much signal. These methods are ad hoc and incomplete in that they were developed as the availability of bandwidth and multi-frequency observations became a ``problem'' (cf. ``the wide-bandwidth problem''~\\citep{L&D13}), and were appropriate when observations covered a narrow bandwidth: phase JUMPs account for profile evolution occurring in frequency gaps that are not observed. It seems natural in the era of wideband receivers --- when frequency evolution {\\it is} observed in the band --- to devise a method for TOA measurement that includes a frequency-dependent model of the average pulse profile. In doing so, it becomes straightforward to include a simultaneous measurement of the dispersion measure. As we will show, a very simple extension to the algorithm that is currently used is a first step in a more comprehensive and necessary description of the received pulsar signal. ", "conclusions": "\\label{conc} In this paper, we have presented a simple method for measuring TOAs in folded pulsar data by using a frequency-dependent model of the pulse profile. This algorithm is a straightforward, yet novel extension of FFTFIT from~\\citet{Taylor92}, but it has some advantages over more standard techniques. These include: \\begin{itemize} \\item{Simultaneous measurement of a phase (TOA) and dispersion measure (DM) across a channelized bandwidth, allowing for easy DM tracking.} \\item{{\\it In situ} accommodation for profile evolution, both intrinsic and extrinsic (eg. scattering), by the use of an arbitrary phase-frequency model (a ``portrait'') of the pulsar's signal.} \\item{Mitigation of scintillation effects by automatic weighting of frequency channels, which appropriately uses the information contained in the data.} \\item{Simplification of the timing procedure by eliminating (where appropriate) the use of phase offsets (i.e. JUMPs), multi-channel TOAs, and ad hoc methods for describing profile evolution, etc.} \\end{itemize} Any arbitrary model can be used, but the choice of model will affect the measured values. We have made our code publicly available online. The demonstration of a simple Gaussian modeling scheme to make these measurements in a 3-year, wideband dataset of the millisecond pulsar M28A shows that we are able to obtain reliable measurements of the dispersion measure, as well as improved TOA and DM precisions by up to a factor of four for the former and two for the latter. The biggest improvements in the parameter precisions and in mitigating profile evolution were seen in the high signal-to-noise 1500~MHz data, which was our best ``wideband data'' for demonstration purposes in the sense that it has the largest fractional bandwidth (and therefore the most obvious effects from interstellar dispersion, scattering, and profile evolution). We note that M28A was chosen for our demonstration precisely because it would show obvious improvements from using our new technique, and so its results may be atypical. Similar improvements in other pulsars will depend on the mitigation of intrinsic and extrinsic profile evolution. It became clear in our comparisons that there is a necessity for quantitative model selection based on more robust two-dimensional portrait modeling, which potentially can lead to the detection of a frequency-dependent DM, or other interesting signals. We probed three typical SNR regimes with Monte Carlo tests and found that the algorithm performs well. Except for the lowest SNR cases, the results from our Monte Carlo analyses has led us to the conclusion that a large number of frequency channels is appropriate for applying this technique. A larger number of channels will provide the highest precision DM measurements and avoids averaging over profile evolution. The proper incorporation of discrete DM measurements with their own heteroscedastic errors (besides the TOAs') into the determination of a timing model (eg. by using \\texttt{tempo}) is not trivial, but a Bayesian approach has been investigated in~\\citet{Lentati13}. Relatedly, measuring DMs from non-simultaneous but temporally proximate multi-frequency data can be an intermediate improvement until larger bandwidths become readily available. This is another avenue of future development, although it comes with the drawback of having correlated TOAs. One important caveat in these Monte Carlo tests is that the model fitted to the simulated data was the true model from which the data were generated. In practice, a Gaussian-component model fitted to real data will not match perfectly (leaving behind non-Gaussian residuals in all but the simplest or low SNR cases) and there will be a much stronger dependence of the measured DM and TOA on the number of channels and the amount of profile evolution. This is another area requiring further testing, but it also suggests to err on the side of more frequency channels. Although general, the algorithm will be most useful when applied to MSPs because of their sensitivities to small dispersion measure changes, as highlighted by Equation~\\ref{DMsmear}, and because of the need to correct for their profile evolution in wideband data to obtain the highest possible timing precisions. For these reasons, we believe this algorithm will provide a natural TOA and DM measurement procedure for campaigns of MSP monitoring, like that of NANOGrav or other PTA experiments. To determine how much is gained in timing precision (and, perhaps, sensitivity to gravitational waves), a direct application of our timing method to the wideband data from a subset of NANOGrav MSPs will be presented in a future paper." }, "1402/1402.3731_arXiv.txt": { "abstract": "LUX, the world's largest dual-phase xenon time-projection chamber, with a fiducial target mass of 118~kg and 10,091 kg-days of exposure thus far, is currently the most sensitive direct dark matter search experiment. The initial null-result limit on the spin-independent WIMP-nucleon scattering cross-section was released in October~2013, with a primary scintillation threshold of 2~phe, roughly 3~\\keVnr~for LUX. The detector has been deployed at the Sanford Underground Research Facility (SURF) in Lead, South Dakota, and is the first experiment to achieve a limit on the WIMP cross-section lower than $10^{-45}$~cm$^{2}$. Here we present a more in-depth discussion of the novel energy scale employed to better understand the nuclear recoil light and charge yields, and of the calibration sources, including the new internal tritium source. We found the LUX data to be in conflict with low-mass WIMP signal interpretations of other results. ", "introduction": "The body of indirect evidence for dark matter is extensive. It includes the best-fit model for explaining the angular power spectrum of the Cosmic Microwave Background temperature anisotropy, gravitational lensing studies, large-scale structure observations and simulations, and galactic rotation curves~\\cite{Blumenthal:1984bp,Davis:1985,Clowe:2006,Hinshaw:2012aka,Ade:2013zuv}. All these point to a significant non-baryonic, cold (heavy and non-relativistic) component of matter in the Universe: $\\sim$85\\% of the matter, or, $\\sim$27\\% of the total mass-energy content of the Universe. The WIMP (Weakly Interacting Massive Particle) is a favored candidate, with many theories (for example, supersymmetry and Kaluza-Klein) providing natural candidate particles. Most direct dark matter searches are therefore geared towards finding WIMPs, which are expected to produce low-energy nuclear recoils (NR) in a detector, while electron recoils (ER) constitute a primary background to be reduced and identified~\\cite{PhysRevD.31.3059, Feng:2010}. The noble element xenon has many favorable properties as a target for direct WIMP detection experiments~\\cite{araujoReview}. For such an element, deposited energy in the material is expressed in three possible channels: excitation, ionization, and heat. The most prominent channel for NR is heat, reducing the amount of energy in the first two significantly, in contrast to ER. Ionization electrons can recombine, or escape. Excitation and recombination together lead to the so-called primary scintillation signal (S1), with S1 from each source indistinguishable. Escaping ionization electrons lead to the secondary scintillation (S2) in the gas region of a two-phase detector. Lastly, noble elements are transparent to their own scintillation light because it originates in decaying molecules (excited dimers, i.e. excimers) rather than in atoms. ", "conclusions": "LUX has the largest exposure of any xenon TPC, as well as the lowest threshold. A new internal ER calibration source, tritium, was successfully implemented, and low-energy NR data agree with Monte Carlo, with the location of the NR band well predicted in terms of absolute charge and light yields for an electric field not studied previously. LUX has achieved the most stringent WIMP limit across a wide range of masses. In spite of assumptions more conservative than have been used in the past for xenon detectors, but which we have shown to agree well with the LUX data, our result is in conflict with low-mass WIMP interpretations of signals seen in DAMA, CoGeNT, CRESST, and CDMS Si. With a 300-day run LUX will probe cross-section versus mass parameter space previously unexplored by any other direct detection experiment, with a significantly improved sensitivity compared to this initial result." }, "1402/1402.0075_arXiv.txt": { "abstract": "We present a new version of a semi-analytic model of cosmological galaxy formation, incorporating a star formation law with a feedback depending on the galaxy-scale mean dust opacity and metallicity, motivated by recent observations of star formation in nearby galaxies and theoretical considerations. This new model is used to investigate the effect of such a feedback on shaping the galaxy luminosity function and its evolution. Star formation activity is significantly suppressed in dwarf galaxies by the new feedback effect, and the faint-end slope of local luminosity functions can be reproduced with a reasonable strength of supernova feedback, which is in contrast to the previous models that require a rather extreme strength of supernova feedback. Our model can also reproduce the early appearance of massive galaxies manifested in the bright-end of high redshift $K$-band luminosity functions. Though some of the previous models also succeeded in reproducing this, they assumed a star formation law depending on the galaxy-scale dynamical time, which is not supported by observations. We argue that the feedback depending on dust opacity (or metal column density) is essential, rather than that simply depending on gas column density, to get these results. ", "introduction": "\\label{sec:intro} The basic picture of galaxy formation and evolution in the cosmological context can be explained in the standard $\\Lambda$ cold dark matter (CDM) cosmology. Particularly, large scale clustering properties and formation and evolution of dark matter halos can reliably be predicted by the theory of gravity. However, in order to obtain the full picture of cosmological galaxy formation, we must solve complicated processes of baryonic physics, such as gas cooling, star formation, feedback, galaxy mergers, and so on. One of the key observables about galaxies that must be explained by the theory of cosmological galaxy formation is the luminosity functions (LFs) and their evolution. Compared with the shape of dark matter halo mass function predicted by the $\\Lambda$CDM cosmology, the observed galaxy LFs have two remarkable features: flatter faint-end slopes and sharp exponential cut-off at the luminous/massive end (see Benson et al. 2003 and references therein), which must be explained by some baryonic processes. A widely accepted solution to achieve a flat faint end is supernova feedback, i.e., energy input into the interstellar medium by supernova explosions to suppress star formation in small galaxies. However, the problem is not yet completely solved at the quantitative level. In fact, unreasonably high efficiency of supernova feedback to remove cold interstellar gas in dwarf galaxies is necessary in many existing theoretical models to reproduce the observed flat faint ends, and such an extreme supernova feedback tends to produce discrepancies with observations other than luminosity function shapes (Nagashima \\& Yoshii 2004, hereafter NY04; Nagashima et al. 2005; Bower et al. 2006, 2012; Guo et al. 2011; Wang, Weinmann \\& Neistein 2012; Mutch, Poole \\& Croton 2013; Puchwein \\& Springel 2013; Hopkins et al. 2013). These results imply that another physical effect may also be taking an important role to produce the observed flat faint end slopes. For the massive end, a popular solution to suppress the formation of too massive galaxies is the feedback by active galactic nuclei (AGNs; e.g., Bower et al. 2006; Croton et al. 2006; Menci et al. 2008; Somerville et al. 2008; Guo et al. 2011). The AGN feedback can also explain the observed trends of the early appearance of massive and quiescent galaxies at high redshifts, and downsizing of star-forming galaxies from high to low redshifts, which are apparently in contradiction with the simple expectation in the $\\Lambda$CDM universe (e.g., Bower et al. 2006; Somerville et al. 2008). However, there are large uncertainties about the physics of AGN feedback both in theoretically and observationally. The current success in explaining the observed trends by this process is based on rather phenomenological modelings including highly uncertain parameters, and further studies are required to confirm the quantitative influence of this process on galaxy evolution. Therefore it is still worth to explore yet other physical effects working to shape galaxy LFs, which is the aim of this paper. It is reasonable to expect that such an effect would be manifested in the scaling laws about star formation efficiency. The relation between the surface densities of star formation rate (SFR) and gas ($\\Sigma_{\\rm SFR}$-$\\Sigma_{\\rm gas}$) has been a subject of intensive research. It is popular to fit this relation by a power-law (so-called Kennicutt-Schmidt law, Kennicutt 1998), but recent observations indicate a cut-off around the total (i.e., {\\hbox{H{\\sc i} }} + H$_2$) gas density of $\\Sigma_{\\rm gas} \\sim 10\\;M_\\odot \\ \\rm pc^{-2}$, under which SFR is suppressed and not well correlated with gas density. This threshold gas density for SFR can be interpreted as a result of less efficient formation of cold molecular gas under the threshold, while the star formation efficiency (SFE) from molecular gas is rather universal in many different environments (Wong \\& Blitz 2002; Kennicutt et al. 2007; Bigiel et al. 2008, 2010; Leroy et al. 2008; Blanc et al. 2009; Heiderman et al. 2010; Lada et al. 2010, Schruba et al. 2011; see Kennicutt \\& Evans 2012 and Schruba 2013 for reviews). A likely physical origin of the suppression of H$_2$ formation under the threshold is radiative feedback by UV photons produced by young massive stars (Schaye 2004; Krumholz, Mckee \\& Tumlinson 2008, 2009; McKee \\& Krumholz 2010; Hopkins et al. 2013). The formation of H$_2$ is driven by collisionally excited metal line cooling and molecule formation on dust grain surfaces, which should be balanced with molecule dissociations by UV photons and grain photoelectric heating, both of which are energetically supplied by UV radiation field. If a region in a galaxy is optically thick to UV radiation field by dust grains, self-shielding of UV radiation would accelerate H$_2$ formation. This implies that the more fundamental threshold about star formation is not the total gas surface density but dust opacity. For a typical dust-to-gas ratio, the observationally indicated threshold in $\\Sigma_{\\rm gas}$ is close to the value at which the effective dust opacity $\\tau_d^{\\rm eff}$ becomes of order unity, where $\\tau_d^{\\rm eff}$ is averaged over wavelength with a weight of the heating radiation energy spectrum (Totani et al. 2011). Therefore it is physically reasonable to expect that a galaxy-scale mean value of $\\tau_d^{\\rm eff}$ has an important role in galaxy formation and evolution. A further observational support to this picture comes from infrared observations. The relations between dust temperature, galaxy size, and infrared luminosity of $\\sim$1,000 nearby star-forming galaxies indicate that almost all of them are in the optically thick regime, and the distribution of dust opacity estimated by gas-phase metal column density suddenly drops around $\\tau_d^{\\rm eff} \\sim 1$, indicating less efficient formation of galaxies at $\\tau_d^{\\rm eff} \\la 1$ (Totani et al. 2011). The purpose of this paper is to investigate the effect of the radiative feedback depend on dust opacity, on cosmological galaxy formation and evolution particularly about the shape of galaxy luminosity functions. The theory of structure formation in the universe predicts that the mean surface density $M/r^2$ of dark halos with mass $M$ and size $r$ nearly scales as $\\propto M^{1/3} (1+z)^2$, indicating higher gas surface density and dust opacity at higher redshifts in more massive objects, and hence more efficient star formation. This may have a favorable effect to explain observations, in a similar way to the feedbacks by supernovae and AGNs. To investigate the effect quantitatively, we use a semi-analytic model (SAM) of cosmological galaxy formation, {\\it the Mitaka model} (NY04). This is a model similar to general SAMs, in which formation and evolution of dark matter halos are solved analytically or calculated by N-body simulations, while complicated baryonic processes are treated phenomenologically (for reviews, see Baugh 2006; Benson 2010). In general, SAMs has many adjustable parameters and the effects of complicated physical processes on the LFs are degenerate (e.g., Neistein \\& Weinmann 2010); therefore a set of best-fit parameters may not be a quantitatively correct description of real galaxy formation. It should be noted that the most important aim of this work is to examine the qualitative effects of the new feedback on luminosity functions. In most of the SAMs, the star formation rate is simply proportional to cold gas mass, and the star formation time scale is modeled as a simple function of the dynamical time scale of galaxy disks or DM halos (e.g., Cole et al. 2000; NY04). Some models (e.g., Kauffmann 1996; Croton et al. 2006; Somerville et al. 2008; Lagos et al. 2011; Wang et al. 2012) incorporated the threshold of gas surface density below which star formation activity is significantly suppressed. In the models of Kauffmann (1996), Croton et al. (2006), and Lagos et al. (2011), they introduced the threshold of gas surface density motivated by the Toomre stability criterion on a galactic scale (Toomre 1964). In this scenario the threshold of gas surface density increases with redshift, and hence the threshold effect should be systematically different in the cosmological context from the threshold by dust opacity considered in this paper. Furthermore, some recent observations indicate that star formation are controlled by the physical state of local interstellar gas, rather than the dynamical state of an entire galaxy (e.g., Leroy et al. 2008; Lada et al. 2010). In other models, such as Somerville et al. (2008), a critical gas surface density threshold for star formation is introduced motivated from the observations of the $\\Sigma_{\\rm SFR}$-$\\Sigma_{\\rm gas}$ relation; however, to our knowledge there are no SAMs that consider a feedback depending on dust surface density rather than gas density. Recently Krumholz \\& Dekel (2012) incorporated a star formation law which depends on gas surface density and gas metallicity, and discussed average evolution of typical galaxies without calculating detailed merger histories of dark halos. The relation between the luminosity function shapes and the dust opacity threshold of star formation has not yet been discussed in previous studies. This paper is organized as follows. In section \\ref{sec:model}, we will describe our model particularly focusing on the modelings of star formation and feedback. In section \\ref{sec:Results}, we show the results of our model, and section \\ref{sec:gasmodel} is devoted for discussion. We will summarize our work in section \\ref{sec:summary}. In this work, the cosmological parameters of $\\Omega_{0}=0.3$, $\\Omega_{\\Lambda}=0.7$, and $H_{0}=70~{\\rm Mpc^{-1}~km~s^{-1}}$ are adopted, and all magnitudes are expressed in the AB system. \\begin{figure*} \\begin{center} \\begin{tabular}{cc} \\includegraphics[width=0.49\\hsize]{g_LF_NY04_z0p0_paper.ps} & \\includegraphics[width=0.49\\hsize]{K_LF_NY04_z0p0_paper.ps} \\\\ \\end{tabular} \\caption{ ({\\it Left}) The local $g$-band LF compared with the NY04 model. The solid and dashed lines represent the NY04 model with strong SN feedback ($\\alpha_{\\rm hot} = 4$) and weak (reasonable) SN feedback ($\\alpha_{\\rm hot} = 2$), respectively. We only show the results of CSF model, since the DSF model gives the almost same results. Filled circles indicate the SDSS $g$-band LF obtained by Blanton et al. (2005), and open circles are the 6dF $b_{\\rm j}$-band LF obtained by Jones et al. (2006). We have transformed the 6dF $b_{\\rm j}$-band LF to match $g$, by subtracting 0.25 mag (Blanton et al. 2005). ({\\it Right}) The same as the left panel but for the local $K$-band LF. Data points are the 6dF galaxy survey (Jones et al. 2006) and 2MASS (Kochanek et al. 2001). } \\label{fig:NY04_LF_z0p0} \\end{center} \\end{figure*} ", "conclusions": "\\label{sec:summary} In this paper, we have considered a new feedback mechanism on star formation depending on galaxy-scale mean optical depth to absorption by dust grains, and examined the effect on galaxy luminosity functions and their cosmological evolution, making use of a semi-analytic model of galaxy formation. The introduction of such feedback process is motivated not only by theoretical considerations but also by recent observations, which indicate that star formation activity is significantly suppressed in galaxies that are transparent to UV radiation. The structure formation theory predicts that the dust-opacity becomes higher in massive objects and at higher redshifts for a fixed dust-to-gas ratio; therefore it is expected that the faint-end of local LFs would be suppressed, which is required for the current galaxy formation models to match the observations. Note that extremely strong supernova feedback was required in the conventional models to reproduce the observed faint end of local LFs. Such feedback process would also accelerate the formation of massive galaxies at high redshifts. We have tested a few models about star formation feedback, and the best fit with observations is found with the model in which star formation is suppressed when the galaxy-scale dust opacity is low and metallicity is higher than a critical value (the evolving $\\varepsilon_{\\min}$ model). The latter condition is introduced phenomenologically, but theoretically motivated by the process of photoelectric heating by dust grains. In this model formation of dwarf galaxies at $z \\sim 0$ is significantly suppressed, and the model successfully reproduces the faint-end slope of local LFs with a physically natural strength of the SN feedback. The new model also succeeded in reproducing the number density of high-$z$ massive galaxies. The early appearance of massive galaxies have been explained by the AGN feedback process; however we have found that the star formation model is also important as well as the AGN feedback. In most of SAMs, star formation time scale is assumed to be proportional to the dynamical time scale of a host DM halo or galaxy disk (e.g., Cole et al. 2000; Bower et al. 2006; NY04). This is essential to explain the early appearance of massive galaxies, because the model with a constant star formation time scale cannot reproduce it even if the AGN feedback is incorporated. However, recent observations suggest that the star formation efficiency is closely related to the gas or dust surface density, rather than the dynamical time scale of an entire galaxy or halo (see section \\ref{sec:intro}). Our new model incorporating the AGN feedback can explain the number density of high-$z$ massive galaxies with the observationally suggested star formation law. The new model is also consistent with the observed cosmic star formation history. We also tested a star formation feedback model depending simply on the gas surface density (the $\\Sigma_{\\rm gas}$ model), rather than the dust opacity, to examine whether the dust opacity is essential or not. Although this model can also reproduce the shape of the local LFs, the difference from the evolving $\\varepsilon_{\\min}$ model appears in the mass function (or $K$-band LF) at high redshifts. The evolving $\\varepsilon_{\\min}$ model predicts more galaxies than the $\\Sigma_{\\rm gas}$ model at the bright end of $K$-band LFs at $z \\sim 2$, which is in better agreement with the observed data. To conclude, we have found that the feedback depending on galaxy-scale dust opacity has significant effects on the cosmological galaxy formation, and has good properties to solve some of the problems found in the previous theoretical models. However, it should also be noted that there are still various uncertainties in our model. For example, we determined the value of star formation efficiency under the dust opacity threshold phenomenologically from fits to the luminosity function data, but these results should be examined in light of theoretical studies of star formation. We assumed that dust mass is simply proportional to the metal mass, but it is not obvious that this proportionality is valid for all galaxies. More observational and theoretical studies on formation/evolution of dust grains are desirable to establish a better star formation modeling for cosmological galaxy formation." }, "1402/1402.0596_arXiv.txt": { "abstract": "{ Recently, there are several reports that the cosmic magnetic fields on Mpc scale in void region is larger than $\\sim 10^{-15}$G with an uncertainty of a few orders from the current blazar observations. On the other hand, in inflationary magnetogenesis models, additional primordial curvature perturbations are inevitably produced from iso-curvature perturbations due to generated electromagnetic fields. We explore such induced curvature perturbations in a model independent way and obtained a severe upper bound for the energy scale of inflation from the observed cosmic magnetic fields and the observed amplitude of the curvature perturbation , as $\\rho_\\inf^{1/4}< 30{\\rm GeV}\\times(B_\\obs/10^{-15}{\\rm G})^{-1}$ where $B_\\obs$ is the strength of the magnetic field at present. Therefore, without a dedicated low energy inflation model or an additional amplification of magnetic fields after inflation, inflationary magnetogenesis on Mpc scale is generally incompatible with CMB observations. } ", "introduction": "It has been known for a long time that galaxies and galactic clusters have their own magnetic fields~\\cite{Wielebinski:2005,Bernet:2008qp,Bonafede:2010xg,Beck:2012}. However, their origin is a big mystery of astronomy and cosmology~\\cite{Giovannini:2003yn,Kandus:2010nw,Durrer:2013pga}. Recently the generation mechanism of the magnetic fields in the universe attracts much attention because there are several reports that magnetic fields are found even in void regions. Such void magnetic fields could be detected by blazar observations~\\cite{Neronov:1900zz, Tavecchio:2010mk, Essey:2010nd, Taylor:2011bn, Takahashi:2013uoa, Finke:2013bua} and it is reported that their strength is larger than $\\sim 10^{-15}$G with an uncertainty of a few orders. On the other hand, the upper bound on primordial magnetic fields could be also obtained from the cosmic microwave background (CMB) and the large scale structure (LSS) observations, and current upper bound is roughly given by $10^{-9}$G (see, e.g. \\cite{Yamazaki:2012pg,Shaw:2010ea}, and references therein) \\footnote{ Ref. \\cite{Kawasaki:2012va} reported an updated constraint on a primordial magnetic field during big bang nucleosynthesis (BBN) as $10^{-6} $G.}. Therefore we know there exist the magnetic fields in the universe with the strength, \\footnote{ The upper bound is irrelevant for magnetic fields which are produced after CMB photons are radiated.} \\begin{equation} 10^{-15}\\G \\lesssim B_\\obs \\lesssim 10^{-9}\\G. \\label{Allowed range} \\end{equation} Nevertheless, their origin is still unknown and no successful quantitative model is established. If the magnetic fields are produced in the primordial universe, they can seed the observed galactic and cluster magnetic field~\\cite{Ryu:2008hi,Pakmor:2013rqa} as well as directly explain the void magnetic fields. As one of the mechanism of generating such cosmic magnetic fields, ``inflationary magnetogenesis\" has been widely discussed. In the context of the inflationary magnetogenesis, large scale magnetic fields, as well as the primordial curvature perturbations, are basically generated from the quantum fluctuations. Although many models of the generation of magnetic fields during inflation are proposed so far~\\cite{Turner:1987bw, Ratra:1991bn, Gasperini:1995dh, Davis:2000zp, Bassett:2000aw, Bamba:2003av,Enqvist:2004yy, Martin:2007ue, Ferreira:2013sqa,Garretson:1992vt,Anber:2006xt}, it is known that these inflationary magnetogenesis models suffer from several problems, namely the strong coupling problem~\\cite{Gasperini:1995dh,Demozzi:2009fu,Fujita:2012rb}, the backreaction problem~\\cite{Bamba:2003av,Demozzi:2009fu,Kanno:2009ei,Fujita:2012rb}, the anisotropy problem~\\cite{Bartolo:2012sd,Thorsrud:2013kya} and the curvature perturbation problem~\\cite{Suyama:2012wh,Barnaby:2012xt,Giovannini:2013rme,Ringeval:2013hfa,Fujita:2013qxa,Nurmi:2013gpa}. In particular, the curvature perturbation problem, where the primordial curvature perturbations which are induced from the generated electromagnetic fields during inflation should not exceed the observed value of CMB experiments, gives strong constraints on inflationary magnetogenesis models. For examples, in our previous paper~\\cite{Fujita:2013qxa}, we have intensively studied the curvature perturbation problem by using a specific model, so-called the kinetic coupling model~\\cite{Ratra:1991bn}, and showed that the allowed strength of the produced magnetic fields is far weaker than the observational lower bound given by eq.~\\eqref{Allowed range}. Ref.~\\cite{Ringeval:2013hfa} have investigated the curvature perturbation problem specifying the time evolution of the magnetic fields during inflation as the power-law of the conformal time and showed limits of the amplitude of the present magnetic fields for the monomial and the hill-top inflation models with several reheating scenarios. Although investigation of the constraint on inflationary magnetogenesis in model dependent ways is important, to discuss whether inflationary magnetogenesis is {\\it really} possible or not, model independent arguments should be also necessary. As for such discussion, in ref.~\\cite{Suyama:2012wh} the authors have put the lower bound on the inflation energy scale $\\rho_{\\rm inf}$ only by requiring the production of magnetic fields with the sufficient strength $B_{\\rm obs} \\sim 10^{-15}$G, but they assumed that the dominant primordial curvature perturbation is generated during the single slow-roll inflation. In ref.~\\cite{Fujita:2012rb}, apart from the curvature perturbation problem, by requiring to escape from the strong coupling and the backreaction problems, the upper bound on $\\rho_{\\rm inf}$ has been % put in model independent ways. In this paper, we consider the curvature perturbation problem of inflationary magnetogenesis % in a model independent way and we do not specify the dominant contribution of the primordial curvature perturbations. That is, our result could be also applied to the case where the dominant primordial curvature perturbation is sourced from a light scalar field other than inflaton. We focus on the existence of the electric fields due to the time evolution of the magnetic fields in the Friedmann-Lemaitre-Robertson-Walker (FLRW) universe and we show that if one requires inflation magnetogenesis is responsible for the generation of the observed magnetic fields and assumes no additional amplification after inflation, the inflation energy scale is constrained by the curvature power spectrum $\\mathcal{P}_\\zeta$ as \\begin{equation} \\mathcal{P}_\\zeta^\\obs > \\mathcal{P}_\\zeta^\\em \\quad\\Rightarrow\\quad \\rho_{\\rm inf}^{1/4} < 30{\\rm GeV}\\times \\left(\\frac{p_B}{1{\\rm Mpc}^{-1}}\\right)^{\\frac{5}{4}} \\left( \\frac{B_{\\rm obs}}{10^{-15}{\\rm G}} \\right)^{-1}, \\label{result intro} \\end{equation} where $\\rho_\\inf$ is the energy scale of inflation, $p_B > 1{\\rm Mpc}^{-1}$ is the peak wave number of the void magnetic field and $B_\\obs$ is the magnetic field strength today. Therefore, our result indicates some tension between inflationary magnetogenesis and phenomenologies in the very early universe, e.g., genesis of the baryon or dark matter, where high energy physics are involved. We also discuss a possible way out of our constraint. If strong magnetic fields are produced without amplifying electric fields, one could avoid our constraint. Such situation is apparently realized in a tree-level analysis of the so-called strong coupling regime of the kinetic coupling model ~\\cite{Ratra:1991bn, Gasperini:1995dh, Bamba:2003av}. However, since the coupling constant becomes huge in the model, a non-perturbative analysis beyond the tree-level is required to make the correct prediction~\\cite{Demozzi:2009fu}. Furthermore, an additional amplification or a non-adiabatic dilution of magnetic fields after inflation can relax our constraint. For example, if the inverse cascade works, the constraint is alleviated~\\cite{Campanelli:2007tc, Saveliev:2013uva}. The rest of paper is organized as follows. In section 2, we briefly review the current lower bound on the cosmic magnetic field from the blazar observations and outline how we constrain inflationary magnetogenesis in a model independent way. In section 3, we derive an expression of curvature perturbations induced by electromagnetic fields during inflation. In section 4, the constraint on inflationary magnetogenesis is obtained. Section 5 is devoted to a summary and discussions. In appendix, we discuss the constraint without the assumption of the instantaneous reheating. ", "conclusions": "In this paper, we show that inflationary magnetogenesis is generally constrained as eq.~\\eqref{final expression} by requiring that the curvature perturbation induced by the electric field during inflation should be smaller than the Planck observation value: $\\mcP_\\zeta^\\obs(k) =2.2\\times10^{-9}$. We emphasize that our argument is model independent as we outlined in sec.~\\ref{Model independent approach}. The main result eq.~\\eqref{final expression} indicates that inflationary magnetogenesis is under pressure in several ways. First, it is known that the reheating (thermalization) energy scale is bounded as $\\rho_{\\rm reh}^{1/4} \\gtrsim 10$MeV in order to achieve a successful BBN~\\cite{Hannestad:2004px}. Therefore even if eq.~\\eqref{final expression} is almost saturated, for example $\\rho_{\\inf}^{1/4}\\sim 10$GeV, the reheating should be quickly completed. Second, the generation of the observed curvature perturbation is in danger. Eq.~\\eqref{final expression} can be translated as \\begin{equation} H_\\inf< 2\\times 10^{-7} {\\rm eV} \\left(\\frac{p_B}{1{\\rm Mpc}^{-1}} \\right)^{5/2} \\left(\\frac{B_\\obs}{10^{-15}\\G} \\right)^{-2}, \\label{H constraint} \\end{equation} where $H_\\inf$ is the Hubble parameter during inflation. For a scalar field to acquire a perturbation during inflation, its mass should be smaller than $H_\\inf$. Thus inflaton field or a spectator field which is responsible to produce $\\mcP_\\zeta^\\obs$ must be extremely light during inflation. During reheating era, however, it has to quickly decay into the standard model particles to cause the BBN properly. Furthermore, in the case of single slow roll inflation, eq.\\eqref{H constraint} and the COBE normalization indicate an extreme slow-roll $\\epsilon< 4\\times 10^{-62}$ which demands a dedicated inflation model. It is interesting to note that eq.\\eqref{H constraint} corresponds to the very small tensor-to-scalar ratio, $r< 7\\times 10^{-61}$. Hence a detection of background gravitational waves in the future excludes inflationary magnetogenesis. \\footnote{See ref.~\\cite{Ferreira2} in which our model-independent constraint is followed up in the light of the BICEP2 result~\\cite{Ade:2014xna}. } Third, in such a low reheating temperature, thermal production of the dark matter or the baryon number seems hopeless. Since $30$GeV is accessible by colliders, effects beyond the standard model have been severely restricted. To realize the dark matter production and baryogenesis, a non-thermal mechanism like the direct decay of inflaton should be considered. In spite of these negative implications, since we have the observational evidence of the magnetic fields in the universe and we are lack of a plausible magnetogenesis model, the inflationary origin of the magnetic field is still an appealing idea. It should be noted that we assume no amplification of the magnetic fields after inflation to derive eq.~\\eqref{final expression}. Thus our result might imply that inflationary magnetogenesis need an additional amplification or a non-adiabatic dilution of magnetic fields after inflation. If the magnetic field generated during inflation is amplified by some mechanism like preheating process~\\cite{Finelli:2000sh} or the inverse cascade~\\cite{Campanelli:2007tc, Saveliev:2013uva}, the constraint is alleviated. Another possible way out from our constraint is to produce a large amplitude of the vector potential before the horizon crossing. It is known that, in the so-called strong coupling regime of the kinetic coupling model, the electric field is not much stronger than the magnetic field and the backreaction and curvature perturbation problems are evaded (if loop effects are neglected)~\\cite{Demozzi:2009fu}. This is because the vector potential $\\mcA_k$ is almost constant on super-horizon ($n\\simeq 0$ in our language). The magnetic field is produced since $\\mcA_k$ has a large amplitude at the horizon crossing due to the small kinetic function. However, as discussed below eq.~\\eqref{def of rho^em}, such model suffers from the strong coupling problem and reliable calculations are difficult to be done. If a large amplitude of a static vector potential is realized without the strong coupling or one can take into account the loop effects in some non-perturbative way, sufficient magnetogenesis might be achieved." }, "1402/1402.7180_arXiv.txt": { "abstract": "The advent of synoptic sky surveys has spurred the development of techniques for real-time classification of astronomical sources in order to ensure timely follow-up with appropriate instruments. Previous work has focused on algorithm selection or improved light curve representations, and naively convert light curves into structured feature sets without regard for the time span or phase of the light curves. In this paper, we highlight the violation of a fundamental machine learning assumption that occurs when archival light curves with long observational time spans are used to train classifiers that are applied to light curves with fewer observations. We propose two solutions to deal with the mismatch in the time spans of training and test light curves. The first is the use of classifier committees where each classifier is trained on light curves of different observational time spans. Only the committee member whose training set matches the test light curve time span is invoked for classification. The second solution uses hierarchical classifiers that are able to predict source types both individually and by sub-group, so that the user can trade-off an earlier, more robust classification with classification granularity. We test both methods using light curves from the MACHO survey, and demonstrate their usefulness in improving performance over similar methods that naively train on all available archival data. ", "introduction": "The advent of next generation optical and radio telescopes such as the Large Synoptic Survey Telescope (LSST) \\cite{tyson_large_2003} and the Square Kilometre Array (SKA) \\cite{taylor2013} will enable massive wide-field surveys for highly variable or transient astronomical sources. Prompt multi-wavelength follow-up of candidate sources will be critical in achieving the scientific goals of these surveys. A global network of potential follow-up instruments is available, but due to limited resources, only high quality candidates will be selected for follow-up. In the regime of big data surveys, manual inspection of data is no longer possible, therefore automated classification methods must form part of the data pipelines to determine whether a detected candidate is an object of interest that requires follow-up. The use of machine learning for source type classification of archival light curve data is well-established \\cite{eyer_automated_2004,debosscher_automated_2007,sarro_automated_2008-1,kim2011,richards2011,blomme_automated_2011}. Some methods focus on specific events or source types to classify \\cite{kim2011}, while others focus on periodic variables \\cite{blomme_automated_2011} or a range of source types \\cite{debosscher_automated_2007,richards2011}. All must contend with the challenges inherent in classifying light curve data including the generation of robust training data \\cite{long_optimizing_2012}, the search for discriminating features \\cite{debosscher_automated_2007,richards2011} and methods that are robust to unevenly sampled data, as well as missing, spurious and sparse observations. Various classification algorithms, such as neural networks, support vector machines and random forests, have been explored. However, algorithmic selection is not the most significant factor in achieving high performance. Current surveys that use machine learned classification in their data processing pipelines include the Palomar Transient Factory (PTF) \\cite{law_palomar_2009} and the Catalina Real-time Survey (CRTS) \\cite{djorgovski_catalina_2011}. PTF employs an automated ``real-bogus'' classification system that identifies true astronomical transient candidates using features extracted from candidates in subtracted images \\cite{bloom_towards_2008, brink_using_2012}. To date, the PTF system performs binary classification (real or not), and primarily uses image features rather than light curve information in its decision-making. CRTS has developed a real-time transient detection pipeline that contends with sparse and heterogeneous data sources \\cite{djorgovski_towards_2011}, self-updates as observations are received, and makes robust decisions on known classes while potentially discovering unknown sources. Whether performing real-time or archival classification, the use of machine learning algorithms in the methods cited above necessitates the creation of a structured data set of features from light curves that are arbitrary in time span and out of phase with each other. After training a classifier on this set, a partially-observed light curve is converted into the same feature characterization and classified. Theoretically the test examples belong to the same population as the training examples, however the fact that the test light curves are only partially-observed implies that the extracted features may effectively belong to a different statistical distribution. This violates a basic premise of all supervised learning algorithms that both training and test examples must belong to the same distribution. A large enough violation of this premise will degrade the performance of the classifier, as it cannot generalize to new test data. To our knowledge, this issue has not been addressed in previous work. This paper discusses how to do classification robustly with light curve data arriving in a stream. This type of algorithm is known as an {\\it online learning algorithm} and it incorporates new data in the classification model as they become available. There are two novel aspects to our work. First, in order to handle the mismatch between the light curve time spans in training and test sets, we developed a method that builds a committee of classifiers each trained with light curves of a set time span. The appropriate committee member that matches the test light in time span is invoked for classification. We show that this method that can outperform a naive method that trains a single classifier using all available archival data. Second, we implemented hierarchical classifiers. A non-hierarchical classifier assigns a single label from a set of discrete labels to a test example, while a hierarchical classifier assigns either a single label or a sub-grouping of discrete labels, where sub-groups are pre-defined by the end user in a tree-like structure. The advantage of a hierarchical classifier is its ability to trade classification granularity against earlier predictions. Given a user-specified desired confidence level, the classifier may not be able to predict down to a single discrete label (e.g., RR Lyrae) but able to predict an informative sub-group of the classes (e.g., variable stars). This enables the end user of the system to receive some information earlier rather than a more confident prediction much later. ", "conclusions": "\\label{s_discussion} In this paper, we explored the feasibility of using light curves for classifying transient and variable sources in an online setting. Although machine-learned classification with light curves has been used in astronomy, classification with a {\\em data stream} (online classification) has not received much attention. As we approach the era of large synoptic surveys, online classification will become increasingly relevant. Transient surveys will benefit from having a classification module as part of the processing pipeline. \\begin{table}[h!] \\centering \\begin{tabular}{ lrrrrr } \\hline & Oracle & \\multicolumn{2}{c}{Naive} & \\multicolumn{2}{c}{Committee }\\\\ & ${\\sim}$2500 days& 160 days & 370 days & 160 days & 370 days \\\\ \\hline \\hline Non-variables & $99\\%$ & $75\\%$ & $96\\%$ & $98\\%$ & $99\\%$ \\\\ Variables & $91\\%$ & $67\\%$ & $80\\%$ & $81\\%$ & $86\\%$ \\\\ \\hline Overall & $97\\%$ & $69\\%$ & $91\\%$ & $93\\%$ & $95\\%$ \\\\ \\hline \\end{tabular} \\caption{Comparison of classification accuracies using different methods} \\label{tab:comp} \\end{table} \\begin{table}[h!] \\centering \\begin{tabular}{ lrrrr } \\hline & \\multicolumn{2}{c}{Hierarch} & \\multicolumn{2}{c}{Hierarch + threshold of 0.7} \\\\ & 160 days & 370 days & 160 days & 370 days \\\\ \\hline \\hline Non-variables & $97\\%$ & $98\\%$ & $98\\%$ (95\\% of srcs) &$99\\%$ (96\\% of srcs) \\\\ Variables\t\t & $82\\%$ & $88\\%$ & $99\\%$ (68\\% of srcs) &$99\\%$ (77\\% of srcs) \\\\ \\hline Overall & $93\\%$ & $95\\%$ & $98\\%$ & $99\\%$ \\\\ \\hline \\end{tabular} \\caption{Comparison of leaf node classification accuracies using hierarchical classification.} \\label{tab:hierarch} \\end{table} An online classifier that consists of a committee of classifiers, each learned from light curves of different time spans that match that of the source to be classified, performs significantly better than a classifier that learns from a fully observed light curve. This is shown in the results summarized in Table \\ref{tab:comp}. However, the performance differs dramatically across source types. This shows that online classification with light curves has potential for some source classes, but will need to be supplemented with other information to be useful for other source classes. We also explored using hierarchical classification with the MACHO dataset. By stopping the classification at the level of the hierarchy where the classification probability is less than the confidence threshold, we can achieve a higher level of accuracy but at a coarser level of classification. The results are summarised in Table \\ref{tab:hierarch}. One area to explore is to evaluate the performance of a hierarchical classifier where the hierarchy structure matches that of the follow-up observation requirements. The result of such a classifier would thus be meaningful and can be used for prioritising follow-up observations. More work can also be done on optimising the classifier at each node by selecting the best feature set and algorithm." }, "1402/1402.2246_arXiv.txt": { "abstract": "{A middle-aged radio-quiet pulsar J0357+3205 was discovered in gamma-rays with \\textit{Fermi} and later in X-rays with \\textit{Chandra} and \\textit{XMM-Newton} observatories. It produces an unusual thermally-emitting pulsar wind nebula observed in X-rays. } { Deep optical observations were obtained to search for the pulsar optical counterpart and its nebula using the Gran Telescopio Canarias (GTC). } { The direct imaging mode in the Sloan $g'$ band was used. Archival X-ray data were reanalysed and compared with the optical data. } { No pulsar optical counterpart was detected down to $g'$~$\\geqslant$ 28\\fm1. No pulsar nebula was either identified in the optical. We confirm early results that the X-ray spectrum of the pulsar consists of a nonthermal power-law component of the pulsar magnetospheric origin dominating at high energies and a soft thermal component from the neutron star surface. Using magnetised partially ionised hydrogen atmosphere models in X-ray spectral fits we found that the thermal component can come from entire surface of the cooling neutron star with a temperature of 36$^{+9}_{-6}$ eV, making it one of the coldest among cooling neutron stars known. The surface temperature agrees with the standard neutron star cooling scenario. The optical upper limit does not put any additional constraints on the thermal component, however it implies a strong spectral break for the nonthermal component between the optical and X-rays as is observed in other middle-aged pulsars. } { The thermal emission from the entire surface of the neutron star likely dominates over the nonthermal emission in the UV range. Observations of the \\psr~in this range are promising to put more stringent constraints on its thermal properties. } ", "introduction": "\\label{sec1} Gamma-ray pulsars are considered as one of the main targets of the \\textit{Fermi} mission. For the five years of activity The Large Area Telescope (LAT) has discovered numerous amounts of such sources previously observed in the radio band. But, apart from the ability to detect many known radio pulsars in $\\gamma$-rays, \\textit{Fermi} LAT also affords the opportunity to discover pulsars independently in so-called blind searches \\citep[cf.][]{Saz09}. These blind searches were quite successful leading to discovery of about three dozens of pulsars in $\\gamma$-rays \\citep[see e.g.,][]{Saz13, Pletsch}. Further multiwavelength investigations of these objects are crucial for unveiling the pulsar emission nature. Because \\textit{Fermi} pulsars are typically nearby and energetic \\citep{Saz13}, they, in particular, appear to be promising targets for studies in X-ray and optical domains. A middle-aged radio-quiet PSR J0357$+$3205 with period $P$~= 444~ms, magnetic field $B$~= 2.3~$\\times$~10$^{12}$~G and characteristic age $P$/2$\\dot{P}$~= 5.4~$\\times$ 10$^{5}$ yr was discovered in one of the \\textit{Fermi} LAT blind frequency searches \\citep{BlindFreqAbdo2009}. The distance to the pulsar of about 500 pc was estimated by \\citet{XrayFirstDeLuca} based on the $\\gamma$-ray ``pseudo-distance'' relation \\citep[see e.g.,][]{sazparkinson2010ApJ}. First X-ray observations of the pulsar field with \\textit{Chandra} had revealed a faint X-ray counterpart of the object with an extended (9 arcmin) X-ray tail \\citep{XrayFirstDeLuca}. Subsequent \\textit{XMM-Newton} observations had shown clearly that emission from the pulsar itself is generally nonthermal with a soft thermal component \\citep{XraySecondDeLuca}. The pulsar field was also observed in the optical and near-infrared bands with 2.5--4 m class telescopes. No counterpart was found down to $V \\ga $ 26\\fm7 \\citep{XrayFirstDeLuca}. To search for an optical counterpart of PSR J0357$+$3205 and/or its tail at a higher sensitivity level, we performed deep optical observations with the 10.4 m GTC. The details of observations and data reduction are described in Sect.~\\ref{sec2}, our results together with reanalysis of the archival X-ray data are presented in Sect.~\\ref{sec3} and are discussed in Sect.~\\ref{sec4}. \\begin{figure*}[t] \\setlength{\\unitlength}{1mm} \\begin{center} \\begin{picture}(145,65)(0,0) \\put (-20,-20) {\\includegraphics[width=100.0mm, clip=]{paper_2.ps}} \\put (80,-12.5) {\\includegraphics[width=80mm, clip=]{small.ps}} \\end{picture} \\end{center} \\vspace{21mm} \\caption{{\\sl Left} panel: GTC/OSIRIS $\\sim$~51\\asec~$\\times$~51\\asec~Sloan $g'$-image fragment of the \\psr~field. The circle shows 3$\\sigma$ X-ray pulsar position uncertainty for the optical observations epoch (see text for details). The 15\\asec~$\\times$~15\\asec~pulsar vicinity within the dashed rectangle is enlarged in the {\\sl right} panel and smoothed with a one pixel Gaussian kernel. The sources discussed in the text are labelled by numbers. } \\label{fig:1} \\end{figure*} ", "conclusions": "\\begin{figure}[t] \\begin{center} \\setlength{\\unitlength}{1mm} \\resizebox{15.5cm}{!}{ \\begin{picture}(50,24.5)(0,0) \\put (-1,0) {\\includegraphics[scale=0.127]{0357_superspec12.eps}} \\end{picture} } \\end{center} \\caption{ Unabsorbed spectrum of PSR J0357$+$3205 in the optical-X-ray range. The GTC $g'$-band flux upper limit is indicated by the bar with an arrow. The \\textit{XMM-Newton} and \\textit{Chandra} data points are shown by grey bars.\\protect\\footnotemark\\ The thick solid line is the best-fit BB+PL model to the X-ray data. The thin dotted line with the dark-grey filled region is the PL component with 90\\% uncertainties. The thick dotted line is the BB hot spot component. The thin/thick dash-dotted and dashed lines are the best-fit NSA and NSMAX components with variable/fixed normalisations. The upper limit on the thermal spectral flux from the entire surface of the NS obtained with the BB+PL+BB model is shown by the light-grey filled region. The extrapolation of the $\\gamma$-ray \\textit{Fermi} spectrum with its uncertainties is shown by the thin solid line with a hatched region. } \\label{fig:spectrum} \\end{figure} From three statistically acceptable X-ray models, BB+PL, NSA+PL, and NSMAX+PL, considered in the previous Section, two latter result in similar parameters of the pulsar thermal emission. However, the NSMAX model is more justified from the physical reasons. Therefore in what follows we omit the NSA+PL model for simplicity. In order to compare the X-ray data and the optical upper limit, the latter must be corrected for the interstellar extinction $A_V$. The standard $A_V$--$N_H$ relation \\citep{predehl1995AsAp} can be used to estimate $A_V$. The BB+PL and NSMAX+PL X-ray spectral fits suggest $N_H$ in a range of (1.0--2.3)~$\\times$ 10$^{21}$ cm$^{-2}$ (Table~\\ref{t:x-fit}). This corresponds to the $A_V$ range of 0.6$-$1.4. However, $A_V$ can hardly exceed the entire Galactic extinction in this direction of 0.8 recently estimated by \\citet{Schlaf}. Therefore, we accept 0.8 as a conservative extinction value for dereddening the optical upper limit. At the same time, the actual $N_H$ value can be larger than 1.4~$\\times$ 10$^{21}$ cm$^{-2}$ which corresponds to this $A_V$. For instance, \\citet{XraySecondDeLuca} estimate entire Galactic $N_H=(2.1\\pm0.2)\\times 10^{21}$~cm$^{-2}$ from the spectral analysis of extra-galactic sources in the pulsar field. Considering the half-thickness of $\\sim$~100 pc for the Galactic gaseous disk responsible for the extinction, the pulsar latitude $b$~= $-$16$^{\\circ}$, and the minimal $A_V=0.6$ ($N_H=10^{21}$~cm$^{-2}$) we obtain the minimum distance to the pulsar of $\\sim$~270 pc. This value is derived assuming the uniform $A_V$ scaling with distance within the disk. The upper limit on the distance of $\\sim$900 pc was estimated by \\citet{XraySecondDeLuca} based on an assumption that the pulsar intrinsic $\\gamma$-ray luminosity cannot exceed its spin-down energy loss rate. The 500 pc distance accepted in Sect. 3.2 is consistent with these limits. At the distance of 500 pc the optical (in the $V$ band) and nonthermal X-ray (in the range of 2--10 keV) luminosities of the pulsar are $L_{V}$~$\\la$ 1.1~$\\times$~10$^{27}$ erg~s$^{-1}$ and $L_{X}$~= 6.0~$\\times$~10$^{29}$ erg~s$^{-1}$. Accounting for the spin-down luminosity $\\dot E= 5.8 \\times 10^{33}$ erg~s$^{-1}$ \\citep{BlindFreqAbdo2009} they yield the optical and X-ray efficiencies of the pulsar $L_{V}/\\dot E$~$\\la$ 10$^{-6.7}$ and $L_{X}/\\dot E$~$\\approx$ 10$^{-4.0}$. These values are compatible with the empirical X-ray luminosity and efficiency {\\sl vs} age dependencies demonstrated by the pulsars detected in the optical and X-rays \\citep[e.g.,][]{zharikov2006AdSpR, zhar2}. This also supports the distance estimate of 500 pc. In Fig.~\\ref{fig:spectrum} we compare the unabsorbed X-ray and $\\gamma$-ray spectra of the pulsar with the optical upper limit of 0.052 $\\mu$Jy derived from the GTC observations and dereddened with $A_V$~= 0.8. \\footnotetext{The data were unfolded and unabsorbed by applying the factor (unfolded unabsorbed model)/(folded absorbed model) in each spectral bin, assuming the best-fit BB+PL model. This procedure is analogous to how an unfolded spectrum is plotted by the \\texttt{Xspec} \\texttt{plot ufspec} command. Note that the data points obtained this way are model-dependent and for different models will follow the respective best-fit lines. } The optical upper limit is two orders of magnitude lower than it would be expected from the extrapolation of the PL spectral component to the optical range. According to Sect.~\\ref{s:x-ray} the PL component is essential to describe the high-energy tail of the pulsar X-ray spectrum. This suggests a spectral break in the PL component between the optical and X-rays, as it is observed for all middle-aged pulsars detected in both domains \\citep{shibanov2006AsAp}. The extrapolation of the $\\gamma$-ray PL spectrum of the pulsar (see Fig.~\\ref{fig:spectrum}) lies well below the optical upper limit. The best-fit NSMAX spectral components with variable and fixed normalisations are shown by thin and thick dashed lines in Fig.~\\ref{fig:spectrum}, respectively. For completeness we also show the NSA spectral components (dash-dotted lines). As seen, to significantly constrain the NS thermal emission in these models one has to go as deep as $\\sim$~30\\fm0 in the optical, which is not feasible with current instrumentation. The BB hot spot spectral component derived from the X-ray data with the BB+PL model (dotted line in the Fig.~\\ref{fig:spectrum}) all the more cannot be currently reached neither in the optical nor in UV. The light-grey region in Fig.~\\ref{fig:spectrum} contains all possibilities for the soft thermal component in the BB+BB+PL model for allowed $R^{\\infty}$ range in accordance with the 99\\% confidence contour of Fig.~\\ref{fig:weiss}. We may conclude that the optical upper limit does not put any additional constraints on thermal emission from the NS surface. However, according to Fig.~\\ref{fig:spectrum}, the entire surface thermal spectral component can be reached in UV. It can also dominate over the PL component there, if the PL component has approximately flat spectral slope from the optical to the UV, as it is observed for other middle-aged pulsars. The latter would be better constrained at longer optical wavelengths, less affected by the interstellar extinction. Therefore, UV observations of J0357$+$3205 would be useful to constrain its surface temperature. There are only few pulsars with thermal emission detected in the UV range namely PSR B0656+14 \\citep{Durant}, PSR B1055$-$52 \\citep{Mignani1055}, PSR J0437$-$4715 \\citep{Karg2004}, and Geminga \\citep{kargaltsev2005ApJ}. In addition, \\citet{Kaplan} reported detection of UV thermal emission from a few isolated neutron stars. In all these cases the UV data on thermal emission were of a great complement to the X-ray data. Accounting for the direction of the pulsar proper motion and the spindown age, we find its likely birth place in the $\\lambda$-Orionis cluster, a 5 Myr active star forming region located in $\\sim$~32\\degs\\ from the pulsar and in $\\sim$~450$\\pm$50 pc from Earth \\citep{mayne2008MNRAS}. Several authors proposed that an expanding molecular ring surrounding the cluster is a supernova remnant left by a Type II supernova explosion of a massive companion of the O-type $\\lambda$~Ori star about 1 Myr ago \\citep[see e.g.,][]{cunha1996AsAp,dolan2002AJ}. Adopting this birth place we independently constrained the pulsar age of 0.2--1.3 Myr, accounting for the pulsar proper motion uncertainties, and the cluster and the pulsar distance ranges. This is consistent with its spindown age of 0.54 Myr. \\label{sec4} \\begin{figure}[t] \\begin{center} \\includegraphics[scale=0.61]{coolcurves7.eps} \\end{center} \\caption{ Measured entire surface temperatures $T^{\\infty}$ as seen by a distant observer (filled circles) for NSs of different age $t$ in comparison with the cooling theory predictions (hatched regions). Dense hatched region correspond to the standard cooling theory, while sparse hatched region shows the minimal cooling theory predictions. The bold star with error-bars show the PSR~J0357$+$3205 surface temperature. } \\label{fig:cool} \\end{figure} Accepting this age range and the NS effective temperature of 36$^{+8}_{-6}$ eV derived from the NSMAX+PL fit we can compare these with the NS cooling theory predictions. The J0357$+$3205 position on the temperature--age plane is shown in Fig.~\\ref{fig:cool} with the bold star with error-bars. The data for other isolated neutron stars (filled circles) are taken from \\citet{shternin2011MNRAS}. It is seen, that J0357$+$3205 is among the coldest cooling NSs known. The dense hatched region shows the range of the NS temperatures that can be obtained by the standard cooling theory where the modified Urca processes are considered as the main neutrino emission mechanism \\citep[e.g.,][]{yakovlev2004ARA}. The J0357$+$3205 position agrees well with the standard cooling theory. However, as seen from Fig.~\\ref{fig:cool} the standard cooling theory is insufficient to reproduce the data on all cooling NSs. Therefore, with sparse hatched region we show the range of cooling curves obtained within the minimal cooling scenario \\citep{gusakov2004AsAp,page2004ApJS,page2009ApJ} which takes into account the presence of the baryon superfluidity inside neutron stars. In this scenario the specific process of the neutrino emission due to a Cooper pair formation cools the star more effectively than the modified Urca process. To date, the parameters of the superfluidity can be plausibly adjusted (sparse hatched region) to fit all the data on the observed NSs temperatures \\citep{gusakov2004AsAp}, including the likely rapidly cooling NS in Cas A \\citep{shternin2011MNRAS}. Obviously, J0357$+$3205 agrees with the minimal cooling scenario as well. At the same time, according to Fig.~\\ref{fig:weiss}, the entire surface temperature in the blackbody spectral model is poorly constrained, taking into account the uncertainties in the NS radius and distance to the pulsar. Thus it is not possible to extract any valuable information from comparison of the BB+BB+PL fit results with the cooling theories. To summarise, our deep optical observations of PSR J0357$+$3205 allowed us to constrain the pulsar nonthermal emission, suggesting a strong spectral break in this emission between the optical and X-rays. Reanalysis of X-ray data allowed us to constrain the NS thermal spectrum and to measure the effective temperature of the NS surface $T^{\\infty}$~= 36$^{+8}_{-6}$ eV. Comparing the optical upper limit with the NS thermal spectrum we conclude that the thermal emission from the entire surface of the NS can be feasibly examined in the UV range and likely dominate there over the nonthermal emission and the emission from pulsar hot spot(s). This makes J0357$+$3205 a promising target for UV observations." }, "1402/1402.5699_arXiv.txt": { "abstract": "Unlike general relativity, the scalar gravitational waves can be excited due to the radial oscillations in scalar-tensor gravity. To examine the scalar gravitational waves in scalar-tensor gravity, we derive the evolution equations of the radial oscillations of neutron stars and determine the specific oscillation frequencies of the matter oscillations and scalar gravitational waves, where we adopt two different numerical approaches, i.e., the mode analysis and direct time evolution. As a result, we observe the spontaneous scalarization even in the radial oscillations. Depending on the background scalar field and coupling constant, the total energy radiated by the scalar gravitational waves dramatically changes, where the specific oscillation frequencies are completely same as the matter oscillations. That is, via the direct observations of scalar gravitational waves, one can not only reveal the gravitational theory, but also extract the radial oscillations of neutron stars. ", "introduction": "\\label{sec:I} Since general relativity has been proposed, many experiments have been performed to verify the gravitational theory. Most of these attempts are done in the weak gravitational field such as our solar system, but nothing indicates the failure of general relativity. On the other hand, since the astronomical observations in the strong gravitational field are very poor, the gravitational theory in such a strong-field regime could be still unconstrained. That is, the gravitational theory to describe the phenomena in the strong-field regime might be different from general relativity, and one might be possible to probe the gravitational theory via the observations of the deviation from general relativity. In practice, up to now there are many suggestions to observationally test the gravitational theory in the strong-field regime \\cite{W1993,W2001,P2008}. The technology is developing more and more, which will enable us to accurately observe the phenomena in the strong-field regime. These coming new observations might be possible to use as the test of gravitational theory. So far, a lot of alternative gravitational theories are proposed. Among them, scalar-tensor theory is the one of the simplest alternative gravitational theories, which must be a natural extension of standard general relativity \\cite{DE1992}. One of the motivations to consider scalar-tensor gravity is that this theory can be obtained in the low energy limit of string and/or other gauge theories. In scalar-tensor gravity, the scalar field plays an essential role in addition to the usual tensor field in general relativity, where the matter field is described by using the effective metric $\\tilde{g}_{\\mu\\nu}$ associated with the scalar and gravitational field, $\\varphi$ and $g_{*\\mu\\nu}$, via the conformal transformation, i.e., $\\tilde{g}_{\\mu\\nu}=A^2(\\varphi)g_{*\\mu\\nu}$. The Brans-Dicke theory \\cite{BD} is the simplest version of scalar-tensor gravity, where $A(\\varphi)$ is defined as $A(\\varphi)=exp(\\alpha\\varphi)$. The coupling parameter $\\alpha$ can be associated with the Brans-Dicke parameter $\\omega_{\\rm BD}$ as $\\alpha^2=1/(2\\omega_{\\rm BD}+3)$, which are constrained through the solar system experiments, i.e., $\\omega_{\\rm BD} \\gsim 40000$ which is corresponding to $\\alpha < 10^{-5}$ \\cite{EF2004}. Within this restriction on the coupling parameter, it is almost impossible to predict a large deviation from general relativity in the strong-field regime. A different functional form of the conformal factor is also suggested by Damour and Esposito-Far\\`{e}se \\cite{DE1993,DE1996}, where $A(\\varphi)\\equiv exp(\\alpha\\varphi + \\beta\\varphi^2/2)$. With this type of coupling, even if $\\alpha$ is almost zero, the relativistic stellar models in scalar-tensor gravity can significantly deviate from the predictions in general relativity. Additionally, they found that the stellar models in scalar-tensor gravity suddenly deviate from those in general relativity for the specific values of coupling parameters, which is referred to as {\\it spontaneous scalarization}. With respect to this phenomenon, Harada systematically examined with the technique of catastrophe theory and found that the spontaneous scalarization can happen for $\\beta\\lsim -4.35$ \\cite{H1998}. Recently, it is found that the spontaneous scalarization are possible for larger value of $\\beta$ in fast rotating relativistic stars \\cite{Daniela2013} and in the neutron star binary system \\cite{BPPL2013,Shibata2013,PBPL2013}. On the other hand, using the observations of pulsar white dwarf binary, Freire {\\it et al}. set a severe constraint on $\\beta$, i.e., $\\beta\\gsim -5$ \\cite{F2012}. Additionally, it is reported that $\\beta$ could be constrained to be larger than $-4.5$, depending on the equation of state \\cite{Shibata2013}. Maybe, although the constraint on $\\beta$ would become severer via the future observations, we focus on the range of $\\beta \\gsim -5$ in this paper. The several attempts to observationally distinguish scalar-tensor gravity from general relativity have been already done in the past by using the redshift in the absorption lines of the X and $\\gamma$ rays emitted from the stellar surface \\cite{DP2003}, the spectrum of the gravitational waves radiated from relativistic stars \\cite{SK2004,SK2005}, and the rotational effect around compact objects \\cite{Sotani2012}. In this paper, we consider the different approach, i.e., the scalar gravitational waves driven by the radial oscillations. In fact, the gravitational waves can not be excited due to the radial oscillations in general relativity. This means that the detection of scalar gravitational waves itself becomes the proof the existence of scalar field. From the observational point of view, if the scalar gravitational waves exist, one could in principle identify the scalar gravitational waves with more than three gravitational wave detectors, because we have only three degrees of polarizations in scalar-tensor gravity, such as two usual tensor gravitational waves and scalar gravitational wave. In fact, we expect that five gravitational wave detectors will be in operation in the future, such as two advanced LIGOs \\cite{aLIGO}, advanced Virgo \\cite{aVirgo}, KAGRA \\cite{KAGRA}, and IndIGO \\cite{IndIGO}. On the other hand, the method how to separate and reconstruct an arbitrary number of polarization modes by using the observational data by multiple interferometric gravitational wave detectors, is also developing, which is a model-independent approach \\cite{HN2013}. It should be notices that the scalar gravitational waves in physical frame are proportional to the cosmological value of scalar field \\cite{BPPL2013,Shibata2013}, whose value must be quite small. That is, if the scalar gravitational waves exist, they might be quite weak and difficult to detect in the current gravitational wave detectors. The radial oscillations of relativistic stars in general relativity have been examined since early times \\cite{C1964,GL1983,GL1992,KR2001} in the context of the stability analysis. Meanwhile, the scalar gravitational waves in scalar-tensor gravity are also examined in the black hole formation due to the dust collapse \\cite{Shibata1994,Scheel1995,H1997} and the test particle around a Kerr black hole \\cite{Saijo1997}. Anyway, this is the first time to calculate the scalar gravitational waves driven by the stellar radial oscillations in scalar-tensor gravity suggested by Damour and Esposito-Far\\`{e}se \\cite{DE1993,DE1996}. For this purpose, we will derive the perturbation equations of radial oscillations and make numerical calculations to examine it. This paper is organized as follows. In the next section, we briefly mention the equilibrium of nonrotating relativistic stars in scalar-tensor gravity. In Sec. \\ref{sec:III}, we derive the perturbation equations describing the radial oscillations of relativistic stars in scalar-tensor gravity. The numerical results are shown in Sec. \\ref{sec:IV}, where the specific frequencies are determined with the mode analysis and the direct time evolution. Then, we make a conclusion in Sec. \\ref{sec:V}. We adopt the geometric units, $c=G_*=1$, where $c$ and $G_*$ denote the speed of light and the gravitational constant, respectively, and use the metric signature is $(-,+,+,+)$. ", "conclusions": "\\label{sec:V} Neutron stars are one of the best candidates to probe the gravitational theory in the strong-field regime. In this paper, we especially focus on the radial oscillations of neutron stars in scalar-tensor gravity to examine the scalar gravitational waves driven by the matter oscillations. In fact, the gravitational waves are not excited due to the radial stellar oscillations in general relativity, while one can expect to observe the scalar gravitational waves due to such oscillations in scalar-tensor gravity. For the calculations of the scalar gravitational waves, we first derive the perturbation equations for radial oscillations in scalar-tensor gravity. From the equation of system of radial oscillations, we find that the matter oscillations depend only on the background scalar field, independently of the scalar gravitational waves. On the other hand, the wave equation of scalar gravitational waves has a source term composed of the matter oscillations. Due to such a specific coupling, we can determine the frequencies of matter radial oscillations by the mode analysis. As a result, we show that the spontaneous scalarization can be observed even in the radial oscillations, which might enable us to find the imprint of gravitational theory with the help of the other observations such as stellar mass and/or compactness. Additionally, to examine the scalar gravitational waves driven by the matter radial oscillations, we directly make a numerical simulation of the evolution equations, where we fix the initial energy of matter oscillations to be $10^{-4}M_\\odot$. Then, we find that the scalar gravitational waves can be excited if the background scalar field exists. We also find that the total energy radiated by the scalar gravitational waves depends strongly on the background scalar field and the coupling constant $\\beta$, where the massive star has a potential to radiate more energy of scalar gravitational waves. Furthermore, we make the fast Fourier transform to see the specific oscillation frequencies of radiated scalar gravitational waves, which are exactly same as the frequencies of matter oscillations. That is, via the observations of scalar gravitational waves, one can extract the frequencies of stellar radial oscillations. This is an advantage in scalar-tensor gravity, because the radial gravitational waves can not be excited in general relativity as mentioned before. We remark that one might have another chance to observe an imprint of radial oscillation in the gravitational waves, if radial oscillations are strongly excited, for example in core-collapse supernovae, and those oscillations make nonlinear coupling with nonradial oscillations, where the oscillations with combination frequencies could be excited \\cite{PSN2007}. If so, one might be possible to measure the background scalar field via such nonlinear coupling. In this paper, as a first step, we neglect the effects of the solid crust layer, magnetic fields, and the exotic matter inside the star, which are also important properties of neutron stars. Such effects could bring us the additional information about the stellar properties \\cite{SKLS2013,Sotani2007,Sotani2011,Yasutake}, which might make observational constraints in the gravitational theory stronger." }, "1402/1402.3240_arXiv.txt": { "abstract": "FU Orionis stars (FUORS) are eruptive pre-main sequence objects thought to represent quasi-periodic or recurring stages of enhanced accretion during the low-mass star-forming process. We characterize the sample of known and candidate FUORS in an homogeneous and consistent way, deriving stellar and circumstellar parameters for each object. We emphasize the analysis in those parameters that are supposed to vary during the FUORS stage. We modeled the SEDs of 24 of the 26 currently known FUORS, using the radiative transfer code of \\citet{whitney2003b}. We compare our models with those obtained by \\citet{robitaille2007} for Taurus class~II and I sources in quiescence periods, by calculating the cumulative distribution of the different parameters. FUORS have more massive disks: we find that $\\sim80\\%$ of the disks in FUORS are more massive than any Taurus class~II and I sources in the sample. Median values for the disk mass accretion rates are $\\sim$ 10$^{-7}$ $\\msun$/yr vs $\\sim$ 10$^{-5}$ $\\msun$/yr for standard YSOs (young stellar objects) and FUORS, respectively. While the distributions of envelope mass accretion rates for class I FUORS and for standard class I objects are similar, FUORS, on average, have higher envelope mass accretion rates than standard class~II and class~I sources. Most FUORS ($\\sim$ 70\\%) have envelope mass accretion rates above $10^{-7}\\,\\msun$/yr. In contrast, 60\\% of the classical YSO sample have accretion rates below this value. Our results support the current scenario in which changes experimented by the circumstellar disk explain the observed properties of these stars. However, the increase in the disk mass accretion rate is smaller than theoretically predicted \\citep{frank1992,hartmann1996a}, though in good agreement with previous determinations. ", "introduction": "FU Orionis stars (FUORS) are a class of variable young stellar objects that show brightness variations of the eruptive type \\citep{herbig1977}. The main feature observed in these variables is a sudden increase in brightness ($ 3-6\\,$mag in the optical), in an elapse of time of a few months. This episode is known as the ``outburst'', after which the object remains bright for years or decades, and then fades in a few centuries back to the pre-outburst stage. The outburst, however, occurs in a different way for each FUORS (see, e.g., \\citealt{hartmann1996a,clarke2005a}). These stars exhibit several indicators of youth, such as the presence of the lithium $ 6707\\,$\\AA~ line in optical spectra, and the association with reflection nebulae and infrared excesses originating from dust grains in circumstellar disks. Moreover, they are spatially and kinematically related to known star-forming regions and in some cases, FUORS have high extinction values in the optical suggesting that they are still embedded in the parent cloud material (see, e.g., \\citealt{hartmann1996a}). FUORS show several properties that strongly suggest the presence of a circumstellar disk, such as broad spectral energy distributions (SEDs, \\citealt{kenyon1988b}), stellar spectral types that become progressively colder with increasing wavelength \\citep{hartmann1985,kenyon1988b}, spectral linewidths that increase with decreasing wavelength \\citep{hartmann1987a,hartmann1987b}, double-peaked line profiles in high-resolution optical and near-infrared (NIR) spectra \\citep{hartmann1985,kenyon1988b} as well as P-Cygni profiles with no evidence for redshifted emission or absorption \\citep{kenyon1988b,hartmann1995}, and finally, deep, broadened infrared CO bands in absorption \\citep{kenyon1988a,reipurth1997b}. Another class of eruptive variables are the so-called EXOR stars, named after EX Lup, the progenitor of the class \\citep{herbig1989,herbig2008}. Their optical brightness increases by $ 1-4 $\\,mag on time scales of weeks or months, then fading back during a few months to its original state after. During its low activity stage they exhibit T~Tauri-like characteristics, while during the outburst stage they usually display emission in the optical spectrum as well as in the infrared CO bandheads (e.g., \\citealt{aspin2010}). To reproduce the SEDs of FUORS, modelers use dusty disks and infalling envelopes (e.g., \\citealt{hartmann1985,kenyon1988b,calvet1991,hartmann1995,calvet1998,whitney2003a,whitney2003b}). Indeed the presence of a circumstellar disk is essential to explain the FU Orionis phenomenon. The disk is where material coming from the surrounding infalling envelope accumulates, heats-up, and finally destabilizes the structure of the disk itself, causing a thermal \\citep{frank1992, bell1994} and/or a gravitational \\citep{zhu2009,zhu2010,vorobyov2005,vorobyov2006,vorobyov2010} instability that eventually leads to the characteristic outburst. During this episode an increase of the brightness takes place, affecting mainly the optical wavelengths since the excess emission comes from the inner regions of the disk, which are heated by the viscous dissipation released after the instability has triggered an increase in the disk accretion rate. \\citet{frank1992} suggests that the central objects of the FUORS systems alternate between low ($10^{-7}\\,$\\msun/yr) and high ($10^{-4}\\,$\\msun/yr) mass accretion rates. The former corresponds to a low activity, quiescent state, while the latter corresponds to periods of high activity. Furthermore, the transformations undergone by the disk are what cause the observed phenomenon. Similarly, EXOR events are also attributed to thermal instabilities in the inner disks \\citep{aspin2011a}. Currently, 26 FUORS have been identified and classified as class~I or class~II objects according to the shape of their SEDs \\citep{lada1987}. This includes the ``confirmed'' FUORS, for which the sudden increase in brightness has been recorded, and the ``candidate'' FUORS, which share many, if not all, of the properties of bona-fide FUORS but for which an outburst has not been observed. In this paper we model and analyze the SEDs of 24 confirmed and candidate FUORS to determine the physical and geometrical parameters of the star and the disk. In Section 2 we present the sample, describe the adopted model and outline the procedure used in the SED modeling. In Section 3 we analyze the individual sources. Our results are presented in Section 4. Finally in Section 5 we summarize the results and conclusions. ", "conclusions": "\\label{s:summary} In this work we present the modeling of the SEDs of a sample of 24 class~II and class~I FU Orionis stars. These SEDs were constructed from fluxes obtained from the literature (Table~\\ref{t:fluxes}), including Spitzer-IRS infrared spectra in the $5-35\\,\\mu$m range for V1515~Cyg, BBW~76, FU~Ori, V346~Nor and V1057~Cyg, and in the $5 -14\\,\\mu$m range for RNO~1B, RNO~1C, L1551~IRS5 and Par~21 \\citep{green2006,quanz2007b}. For Re~50~N~IRS1, we used an ISO-SWS spectrum in the $5 -15\\,\\mu$m range obtained by \\citet{quanz2007b}. Initially we modeled each source applying the grid of \\citet{robitaille2006}, to later use these models as starting points for a more refined analysis using the code of \\citet{whitney2003b}. The parameters corresponding to the best model fits are given in Tables \\ref{t:resclass2} and \\ref{t:resclass1} for class~II and class~I FUORS, respectively. Figures \\ref{f:v1515cyg} to \\ref{f:v2775} show the corresponding SEDs. For sources V1515~Cyg, BBW~76, PP~13S, V1647~Ori, FU~Ori, V1057~Cyg, Z~CMa, L1551~IRS5, ISO-Cha~192, V2492~Cyg, V1331~Cyg, Par~21, and V2775~Ori we compared our parameters values with those derived by other authors, finding in general a good agreement. For the remaining 11 sources, this is the first time a model of their SED is derived. Figure \\ref{f:cumul} shows the accumulative distribution functions of disk masses, disk accretion rates, envelope accretion rates and stellar temperatures of FUORS in our sample and standard class~II and class~I objects in a quiescence state from \\citet{robitaille2007}. Table \\ref{t:median} gives the median values for both groups. The comparison shows that: \\begin{enumerate} \\item On average FUORS disks are more massive than standard class~II and class~I objects disks. About 80\\% of FUORS disks have masses $> 0.1\\,\\msun$, while standard class~II and class~I objects have disk masses $< 0.06\\,\\msun$. \\item Disks mass accretion rates are higher for FUORS than for classical YSOS. The great majority of FUORS ($\\sim90\\%$) have $\\dot{\\rm M}_{disk}>10^{-6}$\\,M$_{\\odot}$/yr, while $\\sim95\\%$ of the standard class~II and class~I objects have $\\dot{\\rm M}_{disk}<10^{-6}$\\,M$_{\\odot}$/yr. Median disks accretion rates are $\\sim 10^{-5}$\\,M$_{\\odot}$/yr vs $\\sim 10^{-7}$\\,M$_{\\odot}$/yr for FUORS and classical YSOs, respectively. \\item The distributions of envelope accretion rates for class I FUORS and standard class I objects are indistinguisable. Most FUORS ($\\sim70\\%$) have envelope accretion rates $> 10^{-7}\\,\\msun$/yr. Median envelope accretion rates are $\\sim 10^{-6}$\\,M$_{\\odot}$/yr vs $\\sim 10^{-8}$\\,M$_{\\odot}$/yr for FUORS and standard YSOs, respectively. \\item The distribution of stellar temperatures for FUORS and classical YSOs are similar in shape, but the FUORS are shifted $\\sim2000\\,$K to higher temperatures. \\end{enumerate} The cumulative distributions for confirmed and candidate FUORS (see Table~\\ref{t:sample}) show no significant differences, suggesting that most candidate objects, in fact, belong to the FUORS class. We caution, however, on the small number of objects in each class (14 confirmed and 10 candidate FUORS). For the seven objects in our sample, for which we have SEDs both in the outburst and in the quiescence stage (see Tables~\\ref{t:resclass2} and \\ref{t:resclass1}), 2 class II (RNO 1B, and V1647 Ori), and 5 class I (OO Ser, V1313 Cyg, V2492 Cyg, HBC 722, and V2775 Ori), we note that while the disk and stellar parameters show variations, the envelope parameters ($\\dot{M}$, R$ _{\\rm max}$, and M$_{env}$) do not change, suggesting the outbursts are triggered by an instability after a long build-up phase. The current scenario of FUORS events states that the circumstellar disk of a YSO builds up material injected from the envelope until it becomes thermally \\citep{frank1992, bell1994, hartmann1996a} and/or gravitationally \\citep{zhu2009,zhu2010,vorobyov2005,vorobyov2006,vorobyov2010} unstable. In particular, using the model parameters and disk properties listed in Tables~\\ref{t:resclass2} and \\ref{t:resclass1} we calculated the Toomre $ Q $ gravitational stability parameter \\citep{Toomre1964}. If $Q < 1$, the disk is unstable. Most of the models are unstable for $R > 5-20$\\,AU. The only exception is L~1551~IRS~5, which is gravitationally unstable at a larger scale ($R \\gtrsim 50$\\,AU). Nevertheless all disk are unstable well inside the centrifugal radius (see Table \\ref{t:averages}). Consequently, gravitational instabilities may contribute to the outburst eruptions, in addition to thermal instabilities, resulting in an increase of the mass accretion onto the central object. What we have described so far agrees with this picture. However, the disk mass accretion rate $\\dot{\\rm M}_{disk}\\sim 10^{-5}\\,\\msun$/yr we obtain (see Table \\ref{t:median}) is one order of magnitude lower than the $\\sim 10^{-4}\\,\\msun$/yr predicted by the theory \\citep{frank1992,hartmann1996a}. Nevertheless, previous models of individual FUORS objects obtain $\\dot{\\rm M}_{disk}$ values consistent with those presented in this work (see, e.g., \\citealt{pfalzner2008, aspin2008}). Although the average values for the parameters for both groups of FUORS are similar to those theoretically expected, the individual values listed for each object in Tables \\ref{t:resclass2} and \\ref{t:resclass1} differ significantly. This can be in part attributed to the fact that the group of the FU Orionis stars itself is not an homogeneous sample. While they all share a particular set of characteristics, those appear in different ways for each object. For instance, while all FUORS show a sudden brightness increase of several magnitudes, followed by a slow decrease to their previous state, the way the brightness jump develops in time is different for each object. A clear example of this is the great diversity in the light curves of the three prototypes of the class \\citep{hartmann1996a}. It is therefore reasonable to expect that the values of individual parameters of each member of the group will ultimately differ. Lastly, we would like to draw attention to three sources in particular. V1647~Ori is a special case on the FUOR sample since it has been well studied before and after the outburst, having SEDs for both epochs. This makes that source a prime candidate for the study of the FU Orionis event, though it has to be approached with care, since its classification as a FUOR or EXOR is still under debate \\citep{aspin2006,aspin2011b,semkov2012a}. The other two particular sources are V2492~Cyg and HBC~722. These objects are, at the moment of writing, the last two for which an FUORS-like outburst has been observed. They show characteristics proper of bona-fide FUORS, as shown in Sections~\\ref{sec-v2492cyg} and \\ref{sec:hbc722}, however its inclusion in the FUORS class is still not certain. Nevertheless, our SED modeling shows behaviors similar to V1647~Ori, the former newest member of the class. V1647~Ori and V2492~Cyg also show EXOR characteristics, and from our modeling we see that they do not show a large variation between the outburst and the quiescent phases. For example, their disk masses do not change with the outburst (see Tables~\\ref{t:resclass2} and \\ref{t:resclass1}), and the variation of the disk mass accretion rate is lower than for other FUORS. Nevertheless, when compared with other FUORS, the parameter values derived for those sources are still within the range established by the rest of the FUORS sample, and could then be considered FUORS. However, if we had just analyzed only those two sources while taking into consideration that EXOR outbursts are thought to be ``scaled-down'' versions of FUORS outbursts, it is very likely that they would have been considered EXORS. This shows the uncertainty and difficulty of disentangle the two types of outburst episodes. Despite sharing common properties, each FUORS or FUORS candidate has its own peculiarities that are not currently well understood. It is therefore of great interest to study the most extreme objects of the class to reach a full understanding of this period of great activity in circumstellar disks. The work we have presented here is the first compilation of SEDs of the currently known FUORS. Of the 26 currently known FUORS, 2 do not have enough observations as to construct the SEDs, and thus are not analyzed. For 21 of the remaining 24 we compile in one place the observations taken in all wavelengths, producing the most complete SEDs possible so far. For 3 FUORS (AR 6A, AR 6B, V2492 Cyg), SEDs for only a limited range ($\\lambda$ $<$ 20 $\\mu$m) were constructed and thus values for the derived parameters are not fully determined. Finally, for 11 of the 24 FUORS analyzed (V1735 Cyg, V883 Ori, RNO 1B, RNO 1C, AR 6A, AR 6B, V900~Mon, V346~Nor, OO~Ser, RE 50 N IRS 1 and HBC~722) we provide for the first time a complete SED modeling to determine the physical and geometrical parameters of the star$+$disk$+$envelope system. Furthermore, this is the first time all the known FUORS with an observed SED are modeled with the same code at the same time, providing an homogeneous set of results. The data we present here will be of great help for future studies in the field." }, "1402/1402.3595_arXiv.txt": { "abstract": "In order to empirically determine the timescale and environmental dependence of stellar cluster disruption, we have undertaken an analysis of the unprecedented multi-pointing (seven), multi-wavelength (U, B, V, H$\\alpha$, and I) Hubble Space Telescope imaging survey of the nearby, face-on spiral galaxy M83. The images are used to locate stellar clusters and stellar associations throughout the galaxy. Estimation of cluster properties (age, mass, and extinction) was done through a comparison of their spectral energy distributions with simple stellar population models. We constructed the largest catalog of stellar clusters and associations in this galaxy to-date, with $\\sim1800$ sources with masses above $\\sim5000$~\\msun and ages younger than $\\sim300$ Myr. In the present letter, we focus on the age distribution of the resulting clusters and associations. In particular, we explicitly test whether the age distributions are related with the ambient environment. Our results are in excellent agreement with previous studies of age distributions in the centre of the galaxy, which gives us confidence to expand out to search for similarities or differences in the other fields which sample different environments. We find that the age distribution of the clusters inside M83 varies strongly as a function of position within the galaxy, indicating a strong correlation with the galactic environment. If the age distributions are approximated as a power-law of the form $\\dndt \\propto t^{\\zeta}$, we find $\\zeta$ values between $0$ and $-0.62$ ($\\zeta \\sim -0.40$ for the whole galaxy), in good agreement with previous results and theoretical predictions. ", "introduction": "\\label{sec:intro} The age distribution of stellar cluster populations in galaxies is the result of the combined effects of the cluster formation history and cluster disruption. Studies have shown a tight correlation between the maximum intensity of the star formation history and the age distribution of the cluster population in multiple galaxies (e.g. NGC~7252 - \\citet{miller97}; \\citet{schweizes98}, \\citet{chien10}; M82 - \\citet{konstantopoulos09}). More quiescent galaxies, such as the Milky Way (at least in the solar neighbourhood) have a smoother, rather flat, cluster age distribution \\citep[e.g.][]{lamers05,piskunov06}. The age distribution of a star cluster population (\\dndt) is defined as the number of clusters observed within some linear time interval, and, at present, the role of cluster disruption in setting the shape of the age distribution is still an open topic of debate. If cluster disruption has any dependency with environment, we would expect to see the age distribution changing over different locations, e.g. where there are differences in gas density, or in the tidal field. However, if cluster disruption is independent of environment, the age distributions of any location should be similar. We chose the galaxy M83 because it presents such differences in environment, allowing us to test these assumptions \\citep[see e.g.][for differences in gas densities across M83]{lundgren04}. Previous studies \\citep[e.g.][]{fall05} of the Antennae galaxies (a currently starbursting, merging galaxy) have found that the cluster age distribution is quite steep, with \\dndt$ \\sim t^{-1}$, and this has been interpreted as evidence for rapid and strong cluster disruption under the assumption that the cluster formation rate/history has been constant over the past few hundred Myr \\citep{fall05,whitmore07}. However, \\citet[][among others]{bastian09} showed that the steepness of the age distribution can be affected by the star formation rate. The age distribution becomes shallower when corrected for an increasing star formation rate, limiting the role of cluster disruption in shaping the age distribution (assuming that both are studied under the same time interval). Also, based on theoretical studies \\citep[e.g.][and references there in]{elmegreen10,kruijssen11} it was showed that the ambient environment where a cluster resides can drastically affect the steepness of the age distribution, making it disrupt faster when the surface gas density ($\\Sigma_{gas}$) is higher, and live longer when $\\Sigma_{gas}$ is low, indicating a strong relation with environment. \\citet{chandar06} used the publicly available catalogue of clusters compiled by \\citet{hill06} to estimate the age distribution of clusters in the Small Magellanic Cloud (SMC). The authors used a lower mass limit of $10^3$\\msun\\ and ages between $7\\le Log(\\tau /yr) \\le 9$. They found a steep distribution with \\dndt$\\sim t^{-0.85}$ in the age interval mentioned. However, using the same catalogue, \\citet{gieles07} and \\citet{degrijs08} showed that the steepness in the age distribution was caused by the sample of Hill \\& Zaritsky being luminosity-limited. When a mass cut above the completeness was used, the resulting distribution was flat, leading to the conclusion that cluster disruption has not significantly altered the age distribution of SMC clusters. The Large Magellanic Cloud (LMC) hosts a much larger cluster population than the SMC, and has been the subject of numerous studies. \\citet{chandar10} used the \\citet{hunter03} catalogue to study the age distribution, and supplement this catalogue with objects that were avoided by Hunter et al., due to nebular emission. They find a steep age distribution, \\dndt$\\sim t^{-0.8}$. However, \\citet{baumgardt13} and \\citet{degrijs13} combine all publicly available catalogues of clusters in the LMC and do not confirm the Chandar et al. findings. Instead they find much flatter results, again suggesting the cluster disruption has not had a strongly effect on the cluster population. \\citet{silvavilla11} studied the age distributions of clusters in five near-by, face-on spiral galaxies, including M83 (NGC~5236) and NGC~1313. Fitting over the age range of 10 to 500~Myr, they found $\\dndt \\sim t^{-0.25}$ and $t^{-0.65}$ for M83 and NGC~1313, respectively. Their results suggest that the age distribution of star clusters might vary as a function of environment The Panchromatic Hubble Andromeda Treasury survey \\citep[PHAT survey,][]{dalcanton12}, one of the largest studies to-date, has catalogued stellar clusters within M31 \\citep[][to date only $\\sim25$\\% of the survey area has been presented]{johnson12}. Fouesneau et al.~(2014) derived the ages and masses of the clusters in the survey and found a flat \\dndt\\ distribution for ages younger than $\\sim$70 Myr, again indicating the small role of disruption in setting the shape of the age distribution in this galaxy. In the present work we study the age distributions of clusters in the nearby \\citep[$\\sim4.5$~Mpc,][]{thim03}, face-on spiral galaxy M83, based on seven pointings observed with the Wide Field Camera 3 (WFC3) onboard the Hubble Space Telescope (HST). This study aims to complement and extend the previous studies of \\citet[][hereafter C10]{chandar10}, \\citet[][hereafter F12]{fouesneau12} and \\citet[][hereafter B11 and B12, respectively]{bastian11,bastian12}. C10 reported a steep age distribution of clusters and associations in the first of the seven fields analyzed here (F1, see Fig.~1), with $\\dndt \\sim t^{-0.8}$. B12 analyzed the same region (F1) and came to similar conclusions. However, they included another pointing further away from the galaxy centre (F2), and found $\\dndt \\sim t^{-0.5}$. Here, we extend the B11 and B12 analysis by including five further WFC3 pointings of M83. For this study, we try to avoid the problem introduced by a luminosity-limited sample and we will only compare age distributions of mass-limited samples. This paper is organized as follows. In \\S~\\ref{sec:obs} we present the data and techniques used and in \\S~\\ref{sec:results} we show our main results. We discuss the implications of our findings in \\S~\\ref{sec:discussion}. \\begin{figure} \\includegraphics[width=8.5cm]{figure1.eps} \\caption{M83 R-band image \\citep{meurer06} with overlapping lines representing the areas covered with the HST. Taken from NASA/IPAC Extragalactic Database (NED).} \\label{fig:mosaic} \\end{figure} ", "conclusions": "\\label{sec:discussion} \\begin{figure} \\includegraphics[width=8.5cm]{figure3.eps} \\caption{The indices from maximum likelihood fits to each field of the form $\\dndt \\sim t^{\\zeta}$. The symbols represent different choices of the classes of sources used in the fits and different age ranges where the fit is made. Note that the fields are ordered in increasing $\\zeta$ values. Numbers (N) at the top are the number of clusters for the filled circles in the legend. ``Total\" represents the full catalogue. Additionally, we plot the results from the C10 catalogue (with symbols that represent both class 1 and class 2 sources). The error is a combination of the standard deviation of the scatter and the random error in the fit.} \\label{fig:dndt2} \\end{figure} As discussed in \\S~1, the age distribution of clusters is a convolution of the cluster formation history and cluster disruption. \\citet{fall05,whitmore07,chandar10,fall12}, among other authors, have suggested a Universal (or quasi-Universal) age distribution in a form of a steep power-law with an index $\\zeta$ between -0.8 and -1.0. These authors found this in a variety of galaxies, therefore they conclude that this rapid disruption must be independent of the environment. Since it is unlikely that all galaxies would have undergone a rapid increase in the cluster formation rate over the past few hundred Myr, they argue that the steepness of the distributions is caused by rapid cluster disruption. However, if true, there are important discrepancies between this environmental independence and the results from analytic models based on N-body simulations \\citep[e.g.][]{baumgardt03,gieles06}, and it will be require a new framework to understand the evolution of star clusters. It is worth noting that the Universal model was originally developed to explain the stellar cluster population of the merging, starburst galaxy, the Antennae, where it was assumed that the cluster formation rate and history was constant over the past few hundred Myr, contrary to expectations of numerical simulations \\citep[e.g.][]{baumgardt03,bastian09,chien10,kruijssen11}. We have attempted to reduce systematics between the fields, by adopting $(i)$ the same photometric procedure, $(ii)$ the same SED models, and $(iii)$ fitting procedure using identical age bins. Additionally, we have also carried out a maximum likelihood fit to the distributions to avoid binning altogether. Our results from these two different methods are in excellent agreement. The results found for the seven fields in M83 do not agree with the predictions of the Universal model, as we find significantly shallower age distributions, with values of $\\zeta$ between $-0.1$ and $-0.61$. These variations indicate a strong correlation with the local ambient environment. These shallower distributions agree with analyses of the SMC \\citep{gieles07,degrijs08,portegieszwart10}, LMC \\citep{baumgardt13,degrijs13}, M51 \\citep{gieles05}, and NGC~2997 (Ryon et al. in prep.) cluster populations, as well as previous M83 results \\citep[B11; B12;][]{silvavilla11}. Is there a way in which the observations presented here could be made to fit with the Universal scenario? One could argue that the star formation history in M83 has been decreasing over the past 200-300~Myr, which would counteract the observational effects of cluster disruption. In order to change the observed \\dndt\\ distributions to be consistent with the Universal expectations, the star formation rates would have had to have decreased by a factor of 2-3 (Field~1) to 10 (Fields 3 and 7) over the past 100~Myr. However, based on the results presented by \\citet{silvavilla11} and B12, based on resolved stellar populations, such changes are unlikely. Under the assumption that the cluster formation history/rate has not dramatically changed in the time interval we are studying here \\citep[based on resolved stellar population studies, e.g.][]{silvavilla11}, we conclude that the cluster population of M83 has not been dramatically affected by cluster disruption, in contrast to expectations from the Universal scenario, where clusters are disrupted at a high rate independent of their ambient environment \\citep[e.g.][]{chandar06,chandar10,fall12}. On the other hand, the mass and environmentally dependent disruption model \\citep[MDD model,][]{lamers05} predicts that the slope of the age distribution will be strongly tied to the tidal fields and the surface density of giant molecular clouds (GMCs), leading to a strong correlation with the ambient environment. Under the assumption of a (roughly) constant cluster formation history, if the tidal field is strong and the GMC density is high, cluster disruption is expected to be stronger, leading to steeper age distributions. On the contrary, if the tidal field is weak and the GMC density is low, there will be less disruption, and the age distribution will be flatter. We find that the inner fields, where the tidal forces and GMC density are highest \\citep[c.f.][]{bastian12} have steeper age distributions, while in the outer regions they are flatter, in excellent agreement with predictions from the models \\citep[e.g.][]{lamers05,kruijssen11}. We will further quantify the relation between environment and disruption using the M83 cluster dataset in a future study. A study of the resolved stellar population, along with the presentation of the full catalogue and a further study of cluster disruption, will be presented in a future work, Silva-Villa et al. (in preparation). Additionally, we will study the relation between star and cluster formation, and the dependence of environment in Adamo et al. (in preparation)." }, "1402/1402.6186_arXiv.txt": { "abstract": "{ The Fluorescence Detector (FD) of the Pierre Auger Observatory provides a nearly calorimetric measurement of the primary particle energy, since the fluorescence light produced is proportional to the energy dissipated by an Extensive Air Shower (EAS) in the atmosphere. The atmosphere therefore acts as a giant calorimeter, whose properties need to be well known during data taking. Aerosols play a key role in this scenario, since their effect on light transmission is highly variable even on a time scale of one hour, and the corresponding correction to EAS energy can range from a few percent to more than 40$\\%$. For this reason, hourly Vertical Aerosol Optical Depth ($\\rm{\\tau_{aer}(h)}$) profiles are provided for each of the four FD stations. Starting from 2004, up to now 9 years of $\\rm{\\tau_{aer}(h)}$ profiles have been produced using data from the Central Laser Facility (CLF) and the eXtreme Laser Facility (XLF) of the Pierre Auger Observatory. The two laser facilities, the techniques developed to measure $\\rm{\\tau_{aer}(h)}$ profiles using laser data and the results will be discussed.} ", "introduction": "Ultra High Energy Cosmic Rays (UHECR, $\\rm E>10^{18} eV$) entering the atmosphere cannot be directly detected due to their extremely low flux. For this reason, the properties of primary particles (energy, mass composition, direction) are deduced from the study of the cascade of secondary particles (Extensive Air Showers, EAS) that originates in the atmosphere due to the interaction of those primaries with air molecules. The Pierre Auger Observatory is the largest detector of EAS ever built, covering an area of $\\rm 3000$ $\\rm km^2$, located in Argentina in the province of Mendoza. The observatory uses two techniques at the same time : the detection of particles at ground level with the Surface Detector (SD) and the observation of the longitudinal development of the EAS by detecting the fluorescence light emitted with the Fluorescence Detector (FD). The SD array is composed of more than 1600 water Cherenkov detectors, overlooked by 27 fluorescence telescopes grouped in 4 sites located at the array periphery. The observatory was completed in 2008. The FD is designed to perform a nearly calorimetric measurement of the energy of cosmic ray primaries: the detected flux of fluorescence photons, emitted by nitrogen air molecules excited by EAS charged particles, is proportional to the energy deposit per unit slant depth of the traversed atmosphere. Due to the constantly changing properties of the calorimeter (i.e. the atmosphere), in which the light is both produced and through which it is transmitted, an extensive system with several instruments has been set up to perform a continuous monitoring of its properties. In particular, the aerosol attenuation of the fluorescence light, highly variable on a time scale of one hour, needs to be constantly measured during data acquisition. If the aerosol attenuation is not taken into account, the shower energy reconstruction is biased by 8 to 25\\% in the energy range measured by the Pierre Auger Observatory. On average, 20\\% of all showers have an energy correction larger than 20\\%, 7\\% of showers are corrected by more than 30\\% and 3\\% of showers are corrected by more than 40\\% \\cite{bib:SegevPaper}. At the Pierre Auger Observatory, hourly vertical aerosol optical depth profiles have been produced for each FD site from January 2004 to December 2012 for a correct reconstruction of FD events. ", "conclusions": "" }, "1402/1402.6523_arXiv.txt": { "abstract": "{ We propose a supersymmetric scenario in which the small Yukawa couplings for the Dirac neutrino mass term are generated by the spontaneous-breaking of Pecci-Quinn symmetry. In this scenario, a right amount of dark matter relic density can be obtained by either right-handed sneutrino or axino LSP, and a sizable amount of axion dark radiation can be obtained. Interestingly, the decay of right-handed sneutrino NLSP to axino LSP is delayed to around the present epoch, and can leave an observable cosmological background of neutrinos at the energy scale of $\\mathcal{O}(10-100) \\GeV$. } \\citestyle{plain} \\begin{document} ", "introduction": "The standard model (SM) has been extremely successful in describing subatomic world, but many astrophysical and cosmological observations require a theory beyond SM. One of the apparent shortcomings of SM is the lack of the tiny neutrino mass hinted by atmospheric and solar neutrino oscillations \\cite{Fukuda:1998mi,Ahmad:2001an} (also reactor and long-baseline neutrino oscillations \\cite{Eguchi:2002dm,Michael:2006rx}). The mass-squared differences of neutrino mass eigenstates are now known to be \\cite{Fogli:2012ua} \\beq |\\Delta m_{21}^2| \\simeq 7.5 \\times 10^{-5} \\eV^2, \\quad |\\Delta m_{32}^2| \\simeq 2.4 \\times 10^{-3} \\eV^2 \\eeq This implies that at least one neutrino has a mass of at least $0.05 \\eV$. On the other hand, a recent analysis based on data from Planck satellite mission \\cite{planck} and predictions from other phenomena found a consistent picture of $\\Lambda$CDM model with the sum of the active neutrino masses given by \\cite{Battye:2014} \\beq \\sum m_\\nu = 0.320 \\pm 0.081 \\eV \\eeq When one tries to get such small neutrino masses from the Higgs mechanism with right-handed (RH) neutrinos introduced, very tiny Yukawa couplings of $\\mathcal{O}(10^{-13}-10^{-12})$ are required. % It looks quite puzzling to have such small Yukawa couplings. The best known mechanism for this puzzle is the so-called seesaw mechanism \\cite{GellMann:1980vs,Yanagida:1979as,Mohapatra:1979ia} which uses a large Majorana mass term of RH-neutrinos. However one should note that a Majorana particle has never been observed so far and the small Yukawa couplings may have a dynamical origin. Besides phenomenological issues, SM also suffers from an esthetic theoretical issue, strong CP problem \\cite{Kim:1986ax,Peccei:2006as} requiring a tuning of $\\mathcal{O}(10^{-10})$ to match experimental data \\cite{Dress:1976bq,Altarev:1996xs}. Axion \\cite{Peccei:1977hh,Peccei:1977ur,Weinberg:1977ma,Wilczek:1977pj,Kim:1979if,Shifman:1979if,Zhitnitsky:1980tq,Dine:1981rt} from the breaking of a global Abelian symmetry (called $U(1)_{\\rm PQ}$) \\cite{Peccei:1977hh,Peccei:1977ur} provides a very simple and attractive solution to this problem, while becoming a good candidate of cold dark matter. Additionally, as we discuss in this paper, the symmetry breaking field associated with $U(1)_{\\rm PQ}$ may be responsible for the small Dirac neutrino mass term. Meanwhile, low energy supersymmetry (SUSY) is quite attractive because it can provide a fine unification of SM gauge couplings, a natural solution to the hierarchy problem of electroweak Higgs mass and a candidate of dark matter under the assumption of $R$-parity conservation, even though it is facing with \\textit{little hierarchy problem} arising due to the lack of SUSY signature at recent collider experiments \\cite{Feng:2013pwa}. In particular, dark matter might be from the extended non-MSSM sector which is necessary to address various theoretical/phenomenological shortcomings of SM. In this paper, we propose a supersymmetric extension of SM in which the tiny Yukawa couplings of the Dirac neutrino mass term are generated dynamically by Peccei-Quinn field which breaks $U(1)_{\\rm PQ}$ symmetry spontaneously, and discuss its cosmological implications including dark matter, dark radiation and cosmic neutrino flux. This paper is organized as follows. In section~\\ref{sec:model}, our model is described. In section~\\ref{sec:cosmo}, cosmological aspects of scalar fields (particularly, RH-sneutrino and saxion fields) are briefly discussed. In section~\\ref{sec:DM}, we estimate abundances of LSP and NLSP before it decays. In section~\\ref{sec:DR}, the dark radiation contribution of relativistic axions produced in the decay of saxion is estimated. In section~\\ref{sec:neu-flux}, a cosmological neutrino flux produced in the decay of NLSP is estimated. In section~\\ref{sec:con}, our conclusion is provided. ", "conclusions": "\\label{sec:con} In this paper, we proposed a simple supersymmetric extension of the standard model, in which both the MSSM $\\mu$-term and small Yukawa couplings for the tiny Dirac neutrino mass term are simultaneously generated by the intermediate scale vacuum expectation value of PQ-field which breaks $U(1)_{\\rm PQ}$ symmetry. It was shown that a right amount of relic density can be obtained by axino LSP produced in the decay of saxion (flaton responsible for thermal inflation) and/or thermally generated neutralinos. The possibility of right-handed sneutrino LSP which can saturate the observed relic density was also discussed. Additionally, it was shown that in the case of thermal inflation a sizable amount of axion dark radiation, which match to the recent data from Planck satellite mission, can also be obtained in a wide range of parameter space. Interestingly, we found that right-handed sneutrino NLSP decaying to axino LSP can produce a cosmological neutrino flux which may be observable in the near future experiments in the energy range of $\\mathcal{O}(10-100) \\GeV$. This is a very unique possibility of supporting our scenario although there is still a large room of parameter space without any observable signatures." }, "1402/1402.7075_arXiv.txt": { "abstract": "\\noindent {\\it Kepler} has identified over 600 multiplanet systems, many of which have several planets with orbital distances smaller than that of Mercury -- quite different from the Solar System. Because these systems may be difficult to explain in the paradigm of core accretion and disk migration, it has been suggested that they formed in situ within protoplanetary disks with high solid surface densities. The strong connection between giant planet occurrence and stellar metallicity is thought to be linked to enhanced solid surface densities in disks around metal-rich stars, so the presence of a giant planet can be a detectable sign of planet formation in a high solid surface density disk. I formulate quantitative predictions for the frequency of long-period giant planets in these in situ models of planet formation by translating the proposed increase in disk mass into an equivalent metallicity enhancement. I rederive the scaling of giant planet occurrence with metallicity as $P_{\\mathrm{gp}} = 0.05_{-0.02}^{+0.02} \\times 10^{(2.1 \\pm 0.4) [\\mathrm{M/H}]} = 0.08_{-0.03}^{+0.02} \\times 10^{(2.3 \\pm 0.4) [\\mathrm{Fe/H}]}$ and show that there is significant tension between the frequency of giant planets suggested by the minimum mass extrasolar nebula scenario and the observational upper limits. This fact suggests that high-mass disks alone cannot explain the observed properties of the close-in {\\it Kepler} multiplanet systems and that migration is still a necessary contributor to their formation. More speculatively, I combine the metallicity scaling of giant planet occurrence with recently published small planet occurrence rates to estimate the number of Solar System analogs in the Galaxy. I find that in the Milky Way there are perhaps $4 \\times 10^{6}$ true Solar System analogs with an FGK star hosting both a terrestrial planet in the habitable zone and a long-period giant planet companion. ", "introduction": "{\\it Kepler} has discovered many multiplanet systems with several planets with orbital periods $P < 50$ days.\\footnote{See for example \\citet{bor11a,bor11b}, \\citet{bat13}, and \\citet{bur14}.} Indeed, 40\\% of solar-type stars in the {\\it Kepler} field have at least one planet with $P < 50$ days \\citep[e.g.,][]{fre13}. Even though these systems differ from the Solar System, their apparent ubiquity suggests that they may represent a frequent outcome of planet formation. In the traditional minimum mass solar nebula (MMSN) scenario, there is probably insufficient solid material in protoplanetary disks to form the {\\it Kepler} multiplanet systems where they are observed today \\citep{wei77,hay81}. Instead, formation further out in the parent protoplanetary disk combined with subsequent inward migration has been suggested as one possible formation channel for this class of system \\citep[e.g.,][]{ali06}. The apparent excess of planets just outside of mean-motion resonances may also support the formation then inward migration scenario \\citep[e.g.,][]{lis11,fab12}. However, the rate and even direction of migration is known to sensitively depend on the unknown thermodynamic state of the disk \\citep[e.g.,][]{paa10,kle12}. Alternative models of in situ formation in disks with solid surface densities enhanced beyond the MMSN expectation have also been suggested to explain the ubiquity of close-in multiple systems. In the minimum mass extrasolar nebula (MMEN) scenario of \\citet{chi13}, the {\\it Kepler} multiplanet systems formed in protoplanetary disks that were about six times more massive than those envisioned in the MMSN scenario. In that picture, the more massive MMEN disks describe the typical protoplanetary disk in the Galaxy, while the less massive MMSN disk is the outlier. I illustrate the MMEN in Figure~\\ref{fig01}. On the other hand, \\citet{han12} invoke the rapid inward migration of planetesimals into the inner regions of the disk. The enhanced solid surface density of the inner disk then naturally leads to the in situ formation of planetary systems closely resembling those observed by {\\it Kepler} \\citep{han13}. Both models of in situ planet formation described above provide useful, fresh looks at planet formation. As I will show, both models are also amenable to quantitative tests. At face value, both models make qualitative predictions for the formation of long-period giant planets. All else being equal, a disk with higher solid surface density in the giant planet forming region has a better chance of forming a giant planet than does a disk with lower solid surface density in the giant planet forming region \\citep[e.g.,][]{lis09}. Consequently, \\citet{chi13} suggested that the enhanced solid surface density in the MMEN scenario should lead to the efficient formation of giant planets outside of 1 AU. In contrast, the concentration of a significant amount of a disk's solid material in the inner disk as suggested by \\citet{han12} should lead to inefficient formation of giant planets outside of 1 AU. Unfortunately, it is not currently possible to directly assess with either the transit or radial velocity technique the frequency of long-period giant planets in the observed {\\it Kepler} multiple systems themselves. However, it is possible to indirectly infer the frequency of long-period giant planets in at least two ways. First, the frequency can be characterized by proxy, a technique in which objects in the solar neighborhood that can be studied in detail stand in for the more distant {\\it Kepler} objects. In this case, the close-in multiple systems of low-mass planets discovered in the solar neighborhood with the radial velocity technique are a proxy for the more distant {\\it Kepler} multiple systems. One can use the published completeness limits of the radial velocity surveys to establish upper limits on the frequency of long-period giant planets in those systems, then compare that upper limit to quantitative predictions of the in situ models of planet formation. Second, it well known that giant planet host stars are preferentially metal enriched \\citep[e.g.,][]{san04,fis05}. Therefore, the absence of a metallicity effect in solar-type hosts of {\\it Kepler} multiple systems can be used to determine an upper limit on the frequency of giant planets in these systems. In this paper, I compare statistical upper limits on the frequency of long-period giant planets in the {\\it Kepler} multiple systems with quantitative predictions of the in situ models of planet formation. I find that there is significant tension between the derived upper limits and the expectation from the MMEN scenario, though current samples are not large enough to constrain the \\citet{han12} scenario. I describe my sample selection in Section 2, I detail my statistical analyses in Section 3, I discuss the results and implications in Section 4, and I summarize my findings in Section 5. ", "conclusions": "{\\it Kepler} multiple planet systems with several planets orbiting with periods $P < 50$ days may be difficult to explain in the traditional core accretion and Type I migration paradigm. In response, two in situ models of plant formation were proposed: the minimum mass extrasolar nebula (MMEN) scenario of \\citet{chi13} and the planetesimal migration scenario of \\citet{han12}. Both models make predictions for the occurrence rate of long-period giant planets. In the MMEN scenario, they are ubiquitous. In the planetesimal migration scenario, they are less common than in the field population. I find that the prediction from the MMEN scenario for the occurrence of long-period giant planets in multiple low-mass systems discovered with the radial velocity technique fails at 3$\\sigma$. The lack of metallicity enhancement in the hosts of multiple small-planet systems discovered by {\\it Kepler} provides an independent test of the MMEN scenario, which it also fails at 3$\\sigma$. I am unable to constrain the planetesimal accretion scenario with the current sample. As a result, migration is still a necessary step in the formation of systems of close-in low-mass planets. I also rederived the scaling of giant planet occurrence on metallicity, and I find that $P_{\\mathrm{gp}} = 0.05_{-0.02}^{+0.02} \\times 10^{(2.1 \\pm 0.4) [\\mathrm{M/H}]} = 0.08_{-0.03}^{+0.02} \\times 10^{(2.3 \\pm 0.4) [\\mathrm{Fe/H}]}$. I used these relations to calculate the frequency of ``solar systems\" in the Galaxy, where a solar system is defined as a single FGK star orbited by a planetary system with at least one small planet interior to the orbit of a giant planet. The presence of a giant planet exterior to an Earth-size planet may be necessary to prevent frequent comet impacts from inhibiting the evolution of life on an otherwise habitable planet. I find that in the solar neighborhood, about 0.7\\% of solar-type stars have a ``solar system\" consisting of both a Neptune-size planet with $P < 50$ days and a protective long-period companion. Intriguingly, I find that true Solar System analogs with both a terrestrial planet in the habitable zone and a long-period giant planet companion to protect it occur around only 0.06\\% of solar-type stars. There are perhaps $4 \\times 10^{6}$ such systems in the Galaxy." }, "1402/1402.0180_arXiv.txt": { "abstract": "The nonbaryonic dark matter of the Universe can consist of new stable charged species, bound in heavy neutral \"atoms\" by ordinary Coulomb interaction. Stable $\\bar U$ (anti-$U$)quarks of 4th generation, bound in stable colorless ($\\bar U \\bar U \\bar U $) clusters, are captured by the primordial helium, produced in Big Bang Nucleosynthesis, thus forming neutral \"atoms\" of O-helium (OHe), a specific nuclear interacting dark matter that can provide solution for the puzzles of direct dark matter searches. However, the existence of the 4th generation quarks and leptons should influence the production and decay rates of Higgs boson and is ruled out by the experimental results of the Higgs boson searches at the LHC, if the Higgs boson coupling to 4th generation fermions with is not suppressed. Here we argue that the difference between the three known quark-lepton families and the 4th family can naturally lead to suppression of this coupling, relating the accelerator test for such a composite dark matter scenario to the detailed study of the production and modes of decay of the 125.5 GeV boson, discovered at the LHC. ", "introduction": "The cosmological dark matter can consist of dark atoms, in which new stable charged particles are bound by ordinary Coulomb interaction (see \\cite{Levels,Levels1,mpla,DMRev,4QGC} for review and references). In order to avoid anomalous isotopes overproduction, stable particles with charge -1 (and corresponding antiparticles as tera-particles \\cite{Glashow}) should be absent \\cite{Fargion:2005xz}, so that stable negatively charged particles should have charge -2 only. There exist several types of particle models, in which heavy stable -2 charged species, $O^{--}$, are predicted: \\begin{itemize} \\item[(a)] AC-leptons, predicted in the extension of standard model, based on the approach of almost-commutative geometry \\cite{Khlopov:2006dk,5,FKS,bookAC}. \\item[(b)] Technileptons and anti-technibaryons in the framework of walking technicolor models (WTC) \\cite{KK,Sannino:2004qp,Hong:2004td,Dietrich:2005jn,Dietrich:2005wk,Gudnason:2006ug,Gudnason:2006yj}. \\item[(c)] and, finally, stable \"heavy quark clusters\" $\\bar U \\bar U \\bar U$ formed by anti-$U$ quark of 4th generation \\cite{Khlopov:2006dk,Q,I,lom,KPS06,Belotsky:2008se}. \\end{itemize} All these models also predict corresponding +2 charge particles. If these positively charged particles remain free in the early Universe, they can recombine with ordinary electrons in anomalous helium, which is strongly constrained in the terrestrial matter. Therefore cosmological scenario should provide a mechanism, which suppresses anomalous helium. There are two possibilities, requiring two different mechanisms of such suppression: \\begin{itemize} \\item[(i)] The abundance of anomalous helium in the Galaxy may be significant, but in the terrestrial matter there exists a recombination mechanism suppressing this abundance below experimental upper limits \\cite{Khlopov:2006dk,FKS}. The existence of a new strict U(1) gauge symmetry, causing new Coulomb like long range interaction between charged dark matter particles, is crucial for this mechanism. Therefore the existence of dark radiation in the form of hidden photons is inevitable in this approach. \\item[(ii)] Free positively charged particles are already suppressed in the early Universe and the abundance of anomalous helium in the Galaxy is negligible \\cite{mpla,DMRev,I}. \\end{itemize} These two possibilities correspond to two different cosmological scenarios of dark atoms. The first one is realized in the scenario with AC leptons, forming neutral AC atoms \\cite{FKS}. The second assumes charge asymmetric case with the excess of $O^{--}$, which form atom-like states with primordial helium \\cite{mpla,DMRev,I}. If new stable species belong to non-trivial representations of electroweak SU(2) group, sphaleron transitions at high temperatures can provide the relationship between baryon asymmetry and excess of -2 charge stable species, as it was demonstrated in the case of WTC \\cite{KK,KK2,unesco,iwara}. After it is formed in the Standard Big Bang Nucleosynthesis (SBBN), $^4He$ screens the $O^{--}$ charged particles in composite $(^4He^{++}O^{--})$ {\\it O-helium} ``atoms'' \\cite{I}. In all the models of O-helium, $O^{--}$ behaves either as lepton or as specific \"heavy quark cluster\" with strongly suppressed hadronic interaction. Therefore O-helium interaction with matter is determined by nuclear interaction of $He$. These neutral primordial nuclear interacting objects can explain the modern dark matter density and represent a nontrivial form of strongly interacting dark matter \\cite{McGuire:2001qj,McGuire1,McGuire2,Starkman,Starkman2,Starkman3,Starkman4,Starkman5,Starkman6}. The cosmological scenario of O-helium Universe allows to explain many results of experimental searches for dark matter \\cite{mpla}. Such scenario is insensitive to the properties of $O^{--}$, since the main features of OHe dark atoms are determined by their nuclear interacting helium shell. It challenges experimental probes for the new stable charged particles at accelerators and such probes strongly depend on the nature of $O^{--}$. Stable $-2$ charge states ($O^{--}$) can be elementary like AC-leptons or technileptons, or look like elementary as technibaryons. The latter, composed of techniquarks, reveal their structure at much higher energy scale and should be produced at the LHC as elementary species. They can also be composite like \"heavy quark clusters\" $\\bar U \\bar U \\bar U$ formed by anti-$U$ quark in the models of stable fourth generation \\cite{Q,I}. In the context of composite dark matter scenario accelerator probe for new stable quark-lepton generation acquires the meaning of critical test for the existence of charged constituents of cosmological dark matter. The signature for double charged AC leptons and techniparticles is unique and distinctive what has already allowed to obtain the lower bound on their mass of 430 GeV in the ATLAS experiment \\cite{2qatlas}. Test for composite $O^{--}$ at the LHC can be only indirect: through the search for heavy hadrons, composed of single $U$ or $\\bar U$ and light quarks (similar to R-hadrons) \\cite{4QGC}, or by virtual effects of 4th generation fermions in the processes with known particles. Here we study a possibility for experimental probe of the hypothesis of stable 4th generation in the studies of 125.5 GeV Higgs boson, discovered in the ATLAS \\cite{atlas,atlas1} and the CMS experiments \\cite{cms1,cms} at the LHC. The results of these studies \\cite{atlas,atals1,cms1,cms} indicate that the number of the detected events, being the production cross section times the decay rate of Higgs boson to two-photon channel, is consistent with the prediction of the Standard model. On the other hand, as it was first revealed in \\cite{nuHiggs} the existence of 4th generation leads to enhancement of the main mechanism of Higgs boson production in $pp$ collisions, what puts constraints on the effect of 4th generation particles and practically excludes the possibility of their full strength coupling to 125.5 GeV Higgs boson. In the model of stable 4th generation the difference of these fermions from the quarks and leptons of the three known families is related to some new conserved charge (which can be even a gauge charge) that protects the stability the lightest quarks and leptons ($U$-quark and the 4th neutrino $N$). The experimental lower limits on the new quarks and leptons make these particles to be heavier than the three light families, what can be explained by the existence of an additional mechanism of their mass generation, e.g. in the framework of multi-Higgs models. It can naturally lead to suppression of the coupling of 4th generation fermions to the 125.5 GeV Higgs boson, discovered at the LHC. Here we explore a possibility to make the 4th generation hypothesis consistent with the experimental data on the two gamma decays of Higgs boson, what opens the door to the indirect probes of the charged constituents of composite dark matter in the detailed studies of production and modes of decay of the 125.5 GeV Higgs boson. ", "conclusions": "The cosmological dark matter can be formed by stable heavy double charged particles bound in neutral OHe dark atoms with primordial He nuclei by ordinary Coulomb interaction. This scenario sheds new light on the nature of dark matter and offers nontrivial solution for the puzzles of direct dark matter searches. It can be realized in the model of stable 4th generation and challenges for its experimental probe at accelerators. In the context of this scenario search for effects of new heavy quarks and leptons acquires the meaning of direct experimental probe for charged constituents of dark atoms of dark matter. The $O^{--}$ constituents of OHe in the model of stable 4th generation are \"heavy antiquark clusters\" $\\bar U \\bar U \\bar U$. Production of such clusters (and their antiparticles) at accelerators is virtually impossible. Therefore experimental test of the hypothesis of stable 4th generation is reduced to the search for stable or metastable $U$ hadrons, containing only single heavy quark or antiquark, or to the studies of virtual effects of 4th generation quarks and leptons in the processes with known particles. The discovery of the 125.5 GeV Higgs boson at the LHC opens the new room for such indirect test of the model of stable 4th generation. The number of detected events of decays of this boson to the two-photon channel is consistent within the experimental errors with the prediction of the Standard model, putting severe constraints on the contribution of new quarks and leptons. On the other hand, the existence of heavy quarks of the 4th generation should lead to enhancement of the gluon fusion mechanism of Higgs boson production, which is its dominant production mechanism in $pp$ collisions. Therefore, to be compatible with the experimental data the model of the stable 4th generation should involve a mechanism of suppression of new quark and lepton couplings to the 125.5 GeV Higgs boson. Taking into account the difference of the 4th generation from the three known families of quarks and leptons and in particular the lower limits on the masses of new quarks and leptons, one can assume some additional mechanisms of their mass generation, what can lead to suppression of their couplings to the 125.5 GeV Higgs boson. In the present work we have shown that, indeed, the suppression in the Higgs boson couplings to 4th generation quarks and lepton the can make compatible the existence of this generation with the experimental data on the two-photon decays of the 125.5 GeV Higgs boson. We consider this result as the first step in the thorough investigation of the predictions of the model of stable 4th generation for the whole set of decay channels of Higgs boson. The confrontation of these predictions with the detailed experimental study of the 125.5 GeV Higgs boson will provide the complete test for the composite dark matter scenarios based on the model of the stable 4th generation." }, "1402/1402.2466_arXiv.txt": { "abstract": "We study the time evolution of the impact probability for synthetic but realistic impacting and close approaching asteroids detected in a simulated all-sky survey. We use the impact probability to calculate the impact warning time ($t_{warn}$) as the time interval between when an object reaches a Palermo Scale value of $-2$ and when it impacts Earth. A simple argument shows that $t_{warn} \\propto D^x$ with the exponent in the range $[1.0,1.5]$ and our derived value was $x=1.3 \\pm 0.1$ . The low-precision astrometry from the single simulated all-sky survey could require many days or weeks to establish an imminent impact for asteroids $>100\\meter$ diameter that are discovered far from Earth. Most close approaching asteroids are quickly identified as not being impactors but a size-dependent percentage, even for those $>50\\meter$ diameter, have a persistent impact probability of $>10^{-6}$ on the day of closest approach. Thus, a single all-sky survey can be of tremendous value in identifying Earth impacting and close approaching asteroids in advance of their closest approach but it can not solve the problem on its own: high-precision astrometry from other optical or radar systems is necessary to rapidly establish an object as an impactor or close approacher. We show that the parallax afforded by surveying the sky from two sites is only of benefit for a small fraction of the smallest objects detected within a couple days before impact: probably not enough to justify the increased operating costs of a 2-site survey. Finally, the survey cadence within a fixed time span is relatively unimportant to the impact probability calculation. We tested three different reasonable cadences and found that one provided $\\sim10\\times$ higher (better) value for the impact probability on the discovery night for the smallest ($10\\meter$ diameter) objects but the consequences on the overall impact probability calculation is negligible. ", "introduction": "The past quarter century has witnessed an exponential increase in the number of known near-Earth objects\\footnote{NEOs are asteroids or comets that have a perihelion distance $<1.3\\au$} (NEO) accompanied by a concomitant improvement in the ability to calculate their impact probabilities with Earth. The job of identifying the largest and most hazardous NEOs, those larger than about $1\\km$ diameter, has mostly been accomplished and asteroid surveys are now focussing on the individually less hazardous but far more numerous smaller asteroids. The large asteroids can be detected at great distances years to centuries in advance of their impact but the smaller asteroids may only be detected on their final approach, if at all, since about 40\\% of them must approach from the direction of the Sun in daylight sky. This work quantifies how the impact probability and warning time evolve in the impact apparition for the smaller asteroids as a function of their size, time after discovery, and observing cadence. In particular, we examine whether the parallax afforded by observations at nearly the same time from two independent observatories provides leverage in improving the impact probability calculation or increasing the impact warning time. The Catalina Sky Survey \\citep[CSS,\\ ][]{Larson1998} and \\PSone\\ system \\citep[Pan-STARRS; \\eg][]{Kaiser2002,Hodapp2004} currently dominate the field of NEO discovery --- almost 90\\% of all NEOs and about 75\\% of all potentially hazardous objects\\footnote{PHOs are NEOs that have a minimum orbital intersection distance \\citep[\\eg][]{Gronchi2005} with Earth of $<0.05\\au$ and absolute magnitude $H<22$} (PHO) were discovered by these two surveys in calendar years 2012 and 2013. The known population of NEOs larger than $1\\km$ diameter is $>90$\\% complete \\citep{Mainzer2011c} so the discovery rate of NEOs in this size range has decreased by about a factor of 6 from a peak of 93 in the year 2000 to about 15/year in the last two years. Despite the success of the surveys in the past few decades it remains the case that the most likely warning time for an impact is {\\it zero} --- contemporary surveys are unlikely to detect smaller but still dangerous asteroids because they do not survey the entire sky deeply or regularly enough to identify the next impactor. The surveys are further limited by the simple fact that ground-based facilities can not survey during the day and about 40\\% of all impactors will approach from the direction of the Sun. These problems were spectacularly highlighted by the Chelyabinsk impact on the morning of 15 February 2013 \\citep[\\eg][]{Brown2013,Borovicka2013} --- with absolutely no warning a $\\sim17\\meter$ diameter object blew up in the atmosphere with an energy equivalent to about 500\\,kilotons of TNT, damaging buildings 50\\,km away in the city of Chelyabinsk and injuring about 1,500 people. The impact risk associated with the unknown objects $>1\\km$ diameter is now comparable to the impact risk with the much more numerous but individually less destructive objects with diameters $<1\\km$. The new balance in the impact risk, along with the realization that smaller impacts may be more numerous but less destructive than anticipated a decade ago \\citep[\\eg][]{Brown2013}, has contributed to an increased interest and funding for the NEO survey programs in recent years. \\eg\\ NASA's NEO Observations (NEOO) program office now solicits\\footnote{ROSES 2011 NEOO solicitation section C.9.1.1.} proposals for surveys that `provide capability to detect the subset of 90\\% of PHOs down to 140 meters in size' The smaller NEOs are more difficult to detect than the larger ones, will be detected ({\\it if} they are detected) closer to Earth, and consequently have shorter observational arcs. The limited time range of the set of detections can make it difficult to identify real impactors even during the apparition in which the impact will take place. This was not the case for the few-meter diameter asteroids \\TC \\citep[\\eg][]{Jenniskens2009} and 2014~AA\\footnote{Minor Planet Electronic Circular 2014-A02}, the only natural objects to be discovered before striking Earth. The very smallest objects will be discovered so close to Earth that, if individual detections of the object can be associated with one another as a `tracklet' \\citep{Denneau2013}, the non-linear motion of the detections on the sky-plane due to topocentric parallax can provide enough leverage in the orbit solution to predict an impact. The observable characteristics of NEOs that will impact Earth can be quite different from those of other NEOs \\citep[\\eg][]{Farnocchia2012,Veres2009,Chesley2004}. For instance, their observable steady-state distribution on the sky-plane is a function of their size and time before impact. Decades before impact they tend to be concentrated in `sweet spots' near the ecliptic and within about $120\\arcdeg$ of the Sun. As the time until impact decreases from weeks to days they spread out over most of the sky but there are still concentrations in the direction looking towards and away from the Sun. An object on its `death plunge' must be moving directly towards Earth in a geocentric reference frame so that about a week before impact its apparent rate of motion may be small --- likely mimicking the rate of motion of much more distant and totally harmless asteroids and perhaps not triggering followup that would allow an impact probability calculation. The techniques employed for the impact probability calculation have evolved dramatically over the past few decades with the realization that asteroid impacts have shaped the Moon's surface and influenced the evolution of life on Earth. Indeed, it was only 34 years ago that \\citet{Alvarez1980} proposed that the KT extinction was the result of an asteroid impact and, even though \\citet{Opik1952} stated that ``Over a dozen meteor craters are at present known on the earth's surface'', it was only in 1960 that \\citet{Chao1960} found strong physical evidence that Meteor Crater in Arizona, USA, was formed in an impact event. \\citet{Opik1952}'s estimated collision rates using the `Theory of Probabilities' for the entire NEO population were surprisingly good given that only six NEOs were known at the time. His collision probability formulae formed the basis of much of the impact collision work in the next decades \\citep[\\eg][]{Bottke1996,Kessler1981} but were eventually supplanted by new numerical techniques \\citep{Milani2002}. The two primary operational asteroid impact warning systems, the Jet Propulsion Laboratory's Sentry system and the NEODyS CLOMON2 system, calculate the collision probability by generating synthetic `Virtual Asteroids' (VA) on orbits that are consistent with the known set of observations and propagating all of them into the future with a N-body integrator to search for impacts \\citep{Milani2005}. These impact warning systems are based on a geometric sampling technique for which the identification of the Virtual Impactors (VI) is performed on the line-of-variation \\citep[LOV,\\ ][]{Milani2005b} thus avoiding the poor efficiency inherent to the Monte Carlo methods, especially when the collision probability is small. Impact predictions are extremely sensistive to the orbit accuracy which depends on many factors but the primary drivers are the length of the observational arc and the astrometric accuracy \\citep[\\eg\\ ][]{Desmars2013}. The longer the arc and the better the astrometry the more accurate the orbit. The latter effect is best illustrated by radar detection of asteroids that provide exquisite range and range-rate information thereby dramatically improving the impact probability accuracy and/or extending the time frame during which the impact probability can be calculated \\citep{Ostro2002}. Impactors can be either direct or resonant \\citep{Milani1999}. Direct impactors collide with Earth during their first known encounter and must be discovered far away to have a large warning time. The warning time for small impactors can be significantly less than one orbital period because they have to be close to Earth to be detected. On the other hand, resonant impactors experience intervening Earth encounters before collision. The intervening encounters are the main source of non-linearity in the dynamics and usually prevent a conclusive assessment of the impact threat but provide additional observational opportunities to detect and constrain their orbits and the impact threat. In this work we focus on the evolution of the collision probability with time for a single survey and concentrate on direct Earth impactors that are detected in the apparition during which the impact occurs. The smaller the asteroid the more likely this scenario as the likelihood that small asteroids will be detected in earlier apparitions is 1) small and, even if they are detected, 2) it is unlikely that they will be recoverable in future apparitions because of the large uncertainties in their ephemeris based on the short observational arcs in the discovery apparition. Thus, we concentrate on collision probability evolution with time for $300\\meter$, $100\\meter$, $50\\meter$ and $10\\meter$ diameter impactors. We also explore whether the collision probability calculation benefits from simultaneous or nearly-simultaneous parallax measurements from two observatories. The heliocentric motion of the impactor and Earth as well as the topocentric rotation of the observer about the geocenter produce `parallax' between successive observations of the same object even from the same site. For very close objects that will impact within days of discovery there may be benefits from the two-site scenario --- especially in rapidly identifying the object as an impactor. Finally, we measure the single-system impact warning time as a function of impactor diameter. We expect the warning time to be longer for larger objects but the exact relationship between diameter and warning time is not intuitively obvious. The larger objects are discovered at greater distances where their rate of motion is similar to the much more distant main belt objects and the impact probability will be much smaller. If the impact probability is too small it may not cross the threshold to flag the object as an imminent impactor. ", "conclusions": "We have performed a simulation of a single all-sky asteroid survey to study the time evolution of the calculated impact probability for both real and false impactors. We also studied the utility of using two observatories at different locations to perform the survey to take advantage of the parallactic displacement in the detections of the same object. As expected, the impact probability for impactors typically increases monotonically with time after discovery and is larger at the time of discovery for small objects that are detected closer to Earth and with less time to impact. We found that the impact warning time, the time interval between when the impact probability reaches -2 on the Palermo Scale and when the impact takes place, increases with diameter according to $t_{warn} \\propto D^{1.3}$ and developed a simple mathematical argument that the exponent should be in the range $[1.0,1.5]$. Close approaching asteroids can almost always be unambiguously identified as non-impactors but a small percentage will have a non-zero impact probability even on the day of (false) impact. \\ie\\ the simulated survey on its own is unable to eliminate the impact risk. The fraction of objects for which a persistent impact risk exists at the time of impact increases with decreasing diameter of the object because the small objects have smaller observational arc lengths and concomitantly less precise orbit elements. The combination of the PHO size-frequency distribution with the probability of detecting false impactors suggests that the single all-sky system alone will generate 100s of potential impactors that must be ruled out with other followup facilities. The calculated impact probability can take surprisingly long to reach $\\sim100$\\% with just the results from a single low-precision astrometric survey. The impact probability may reach 100\\% only a few days before impact even for $300\\meter$ diameter objects detected a month in advance and imaged nightly thereafter . Our simulations suggest that a 2-site survey is unnecessary, at least in terms of the incremental benefit in improving the impact probability calculation. The parallax afforded by this scenario only improves the impact probability calculation for a small fraction of the smallest asteroids detected shortly before impact. The 2-site survey offers many different cadence options and some can provide more efficient impact probability calculations than others. The derived impact probability was $\\sim10\\times$ higher (\\ie\\ better) on the discovery night using the `full-shift' cadence compared to the other two cadences. This suggests that a real survey that implements the 2-site scenario should carefully test different cadences to select one that maximizes the efficiency and accuracy of the impact probability on the discovery night. The effect of survey cadence on the impact probability calculation is negligible on successive nights." }, "1402/1402.6842_arXiv.txt": { "abstract": "We present a sample of normal type~Ia supernovae from the Nearby Supernova Factory dataset with spectrophotometry at sufficiently late phases to estimate the ejected mass using the bolometric light curve. We measure \\nickel\\ masses from the peak bolometric luminosity, then compare the luminosity in the \\cobalt-decay tail to the expected rate of radioactive energy release from ejecta of a given mass. We infer the ejected mass in a Bayesian context using a semi-analytic model of the ejecta, incorporating constraints from contemporary numerical models as priors on the density structure and distribution of \\nickel\\ throughout the ejecta. We find a strong correlation between ejected mass and light curve decline rate, and consequently \\nickel\\ mass, with ejected masses in our data ranging from 0.9--1.4~\\Msol. Most fast-declining \\mbox{(SALT2 $x_1 < -1$)} normal SNe~Ia have significantly sub-Chandrasekhar ejected masses in our fiducial analysis. ", "introduction": "Type Ia supernovae (SNe~Ia) have been used for well over a decade as precision luminosity distance indicators, leading to the discovery of the universe's accelerated expansion \\citep{riess98,scp99} \\revised{which has been measured in contemporary studies with increasing precision} \\citep{hicken09,kessler09,sullivan11b,suzuki12}. SN~Ia luminosities can be measured to an accuracy of $\\sim 0.15$~mag using correlations between the luminosity, colour, and light curve width \\citep{riess96,tripp98,phillips99,goldhaber01}, and many recent and ongoing studies have sought to further reduce this dispersion by looking for new correlations between SN~Ia luminosities and their spectroscopic properties \\citep{sjb09,wang09,csp10,fk10}. The spectra of SNe~Ia show no hydrogen, no helium, and strong intermediate-mass element signatures; they are generally understood to be thermonuclear explosions of carbon/oxygen white dwarfs in binary systems. The absence of a detectable shock breakout in the early light curve of the nearby SN~Ia~2011fe \\citep{nugent11,bloom12} provides direct evidence that the progenitor primary must be a compact object such as a white dwarf. However, many variables remain which can affect the explosion, including the evolutionary state of the white dwarf progenitor's binary companion, the circumstellar environment, the explosion trigger, and the progress of nuclear burning in the explosion. The low luminosities, small radii, and relatively clean environments of white dwarfs make SN~Ia progenitor systems notoriously hard to constrain. Uncovering the nature of SN~Ia progenitor systems and explosions is therefore an interesting puzzle in its own right. From a cosmological viewpoint, if two or more SN~Ia progenitor channels exist which have slightly different peak luminosities or luminosity standardization relations, and their relative rates evolve with redshift, the resulting shift in the mean luminosity could mimic a time-varying dark energy equation of state \\citep{linder06}. The two main competing SN~Ia progenitor scenarios are the \\emph{single-degenerate} scenario \\citep{wi73}, in which a carbon/oxygen white dwarf slowly accretes mass from a non-degenerate companion until exploding near the Chandrasekhar mass, and the \\emph{double-degenerate} scenario \\citep{it84}, in which two white dwarfs collide or merge. The classical formulations of these scenarios assume the primary white dwarf must explode near the Chandrasekhar limit; however, in the \\emph{sub-Chandrasekhar double-detonation} variant, a sub-Chandrasekhar-mass white dwarf can be made to explode by the detonation of a layer of helium on its surface, accreted from the binary companion \\citep{ww94,sim10,fink10,kromer10,sim12}. Distinguishing which of these models accounts for the majority of spectroscopically ``normal'' \\citep{bfn93}, hence cosmologically useful, SNe~Ia has been a very active subject of current research \\citep[for a recent review see][]{wh12}. Binary population synthesis models of the Chandrasekhar-mass single-degenerate and double-degenerate channels often have trouble producing enough SNe~Ia to reproduce the observed rate \\revised{\\citep[but see][]{hp04,ruiter11}}; this is one of the main motivations for investigating sub-Chandrasekhar models \\citep{vkcj10}. The mass of the progenitor is a fundamental physical variable with power to differentiate between different progenitor scenarios. While Chandrasekhar-mass delayed detonations have been historically favored, viable super-Chandrasekhar-mass evolution pathways and explosion models have been proposed for both single-degenerate \\citep{justham11,hachisu11,rds12} and double-degenerate \\citep{pakmor10,pakmor11,pakmor12} SN~Ia progenitors, and sub-Chandrasekhar-mass models must necessarily involve a different explosion trigger than any of these. The white dwarf progenitor is totally disrupted in theoretical models of normal SNe~Ia, although a bound remnant may remain in some models which try to reproduce underluminous, peculiar events such as SN~2002cx \\citep{kromer13}. For normal SNe~Ia, then, measuring the progenitor mass reduces to measuring the ejected mass. Nebular-phase spectra can be used to estimate the mass of iron-peak elements in the ejecta \\citep[e.g.][]{mazzali07}, but only the closest SNe~Ia are bright enough to yield high-quality spectra in nebular phase $\\sim 1$~year after explosion, which limits the number of SNe on which this technique can be used. \\citet{stritz06} used SN~Ia quasi-bolometric light curves ($UBVRI$) in early nebular phase (50--100~days after $B$-band maximum light) to estimate the ejected mass, as follows: The mass of \\nickel, \\revised{the radioactive decay of which} powers the near-maximum light curve of normal SNe~Ia, can be inferred from the bolometric luminosity at maximum light \\citep{arnett82}. The decay of \\cobalt, itself a decay product of \\nickel, powers the post-maximum light curve. At sufficiently late times, the shape of the bolometric light curve is sensitive to the degree of trapping of gamma rays from \\cobalt\\ decay \\citep{jeffery99}; greater ejected masses provide greater optical depth to Compton scattering, and hence higher luminosity, for a given phase and \\nickel\\ mass. \\citet{scalzo10,scalzo12} refined this method by including more accurate near-infrared (NIR) corrections and a set of prior constraints on model inputs from contemporary explosion models, using it to estimate the masses of several candidate super-Chandrasekhar-mass SNe~Ia; they found ejected masses of $2.30^{+0.27}_{-0.24}$~\\Msol\\ for the superluminous SN~Ia~2007if and $1.79^{+0.28}_{-0.21}$~\\Msol\\ for the spectroscopically 1991T-like \\snf{080723-012}, interpreting them as double-degenerate explosions powered entirely by radioactive decay. In the current work, we use this method as implemented in \\citet{scalzo12} on a set of \\emph{normal} SNe~Ia, attempting to quantify the distribution of progenitor mass scales in the context of different progenitor scenarios. Our supernova discoveries, our sample selection, and the provenance of our data are described in \\S\\ref{sec:observations}. Our \\revised{method for constructing} full $UBVRIYJHK$ (3300--\\revised{23900}~\\AA) bolometric light curves for 19 spectroscopically normal SNe~Ia, (including NIR corrections for the $YJHK$ flux which we do not observe), are presented in \\S\\ref{sec:analysis}. We briefly review the assumptions of our ejected mass reconstruction method in \\S\\ref{sec:modeling}, and present the reconstructed masses for our 19 SNe. We also present \\revised{ejected mass and \\nickel\\ mass reconstructions based on} synthetic observables from a series of contemporary explosion models. In \\S\\ref{sec:discussion} we examine correlations between \\revised{ejected mass} and other quantities, such as photospheric light curve fit parameters (decline rate and colour) and \\nickel\\ \\revised{mass}. We summarize and conclude in \\S\\ref{sec:conclusions}. ", "conclusions": "" }, "1402/1402.3460_arXiv.txt": { "abstract": "The IceCube Neutrino Observatory is a large Cherenkov detector instrumenting $1\\,\\mathrm{km}^3$ of Antarctic ice. The detector can be used to search for signatures of particle physics beyond the Standard Model. Here, we describe the search for non-relativistic, magnetic monopoles as remnants of the GUT (\\textbf{G}rand \\textbf{U}nified \\textbf{T}heory) era shortly after the Big Bang. Depending on the underlying gauge group these monopoles may catalyze the decay of nucleons via the Rubakov-Callan effect with a cross section suggested to be in the range of $10^{-27}\\,\\mathrm{cm^2}$ to $10^{-21}\\,\\mathrm{cm^2}$. In IceCube, the Cherenkov light from nucleon decays along the monopole trajectory would produce a characteristic hit pattern. This paper presents the results of an analysis of first data taken from May 2011 until May 2012 with a dedicated slow-particle trigger for DeepCore, a subdetector of IceCube. A second analysis provides better sensitivity for the brightest non-relativistic monopoles using data taken from May 2009 until May 2010. In both analyses no monopole signal was observed. For catalysis cross sections of $10^{-22}\\,(10^{-24})\\,\\mathrm{cm^2}$ the flux of non-relativistic GUT monopoles is constrained up to a level of $\\Phi_{90} \\le 10^{-18}\\,(10^{-17})\\,\\mathrm{cm^{-2}s^{-1}sr^{-1}}$ at a 90\\% confidence level, which is three orders of magnitude below the Parker bound. The limits assume a dominant decay of the proton into a positron and a neutral pion. These results improve the current best experimental limits by one to two orders of magnitude, for a wide range of assumed speeds and catalysis cross sections. ", "introduction": "\\label{sec:intro} Magnetic monopoles are particles carrying a quantized magnetic charge and are predicted in various theories. In classical electrodynamics, their existence would symmetrize Maxwell's equations with respect to the sources of the electromagnetic field. Quantum mechanically, the existence of magnetic monopoles implies that both electric charge and the hypothetical magnetic charge, are quantized, given that the associated electromagnetic fields still satisfy Maxwell's equations \\cite{Dirac:1931}. The resulting magnetic elementary charge, called the Dirac charge $g_{\\mathrm{D}}$, is \\begin{equation} g_{\\mathrm{D}}=\\frac{e}{2\\alpha}, \\end{equation} where $e$ is the electric elementary charge and $\\alpha$ is the fine structure constant. In Grand Unified Theories (GUTs) \\cite{Georgi:1974FirstGUT} magnetic monopoles appear as stable, finite energy solutions of the field equations \\cite{tHooft:1974MonopoleSolution,Polyakov:1974MonopoleSolution}. The predicted masses range from $10^5\\,\\mathrm{GeV}$ to $10^{17}\\,\\mathrm{GeV}$ \\cite{Georgi:1974_Interaction_GUT,Daniel:1979_SU5Monopoles,Lazarides,Kephart,Wick:CosmicFluxMM} and their magnetic charges are integer multiples of the Dirac charge $\tg_{\\mathrm{D}} $. The lower part of the mass range up to $\\sim 10^{13}\\,\\mathrm{GeV}$ refers to intermediate mass monopoles (IMMs) which arise from intermediate stages of symmetry breaking below the GUT scale. In contrast the superheavy monopoles with masses at the GUT scale may have been created during the phase transition associated with the spontaneous breakdown of the unified gauge symmetry in the early universe at $\\sim 10^{-36}\\,\\mathrm{s}$ after the Big Bang \\cite{Kibble}. The monopole mass and charge depend on the underlying gauge group, the symmetry breaking hierarchy, and the type and temperature of the phase transition in a particular GUT. Since magnetic monopoles are stable, they should still be present in cosmic rays. The number density today depends on the existence of an inflationary epoch and on the time of creation, which could be before, during or after this epoch \\cite{Preskill:MM}. Since then, monopoles have been accelerated by large-scale cosmic magnetic fields. The kinetic-energy gain by passing through a magnetic field $B$ is given by \\begin{equation} E_{\\mathrm{kin}}=g\\int\\limits_{\\mathrm{path}} \\vec{B} \\cdot d\\vec{l}, \\end{equation} where $g =n\\cdot g_{\\mathrm{D}} $ is the magnetic charge. The maximum kinetic energy of a magnetic monopole due to acceleration in cosmic magnetic fields is rather uncertain but can reach $\\sim 10^{14}\\,\\mathrm{GeV}$ \\cite{Wick:CosmicFluxMM}. Therefore, monopoles with masses at, or above, this energy scale should be non-relativistic. Based on the propagation of magnetic monopoles in the Galactic magnetic field an upper bound on the monopole flux can be calculated, assuming the Galactic magnetic field does not decrease faster than it can be regenerated. This assumption constrains the monopole flux to be less than $10^{-15}\\,\\mathrm{cm}^{-2}\\,\\mathrm{s}^{-1}\\,\\mathrm{sr}^{-1}$, which is called the Parker Bound \\cite{Turner:ParkerBound,Groom:SearchSupermassiveMM}. Taking into account the fields during galaxy formation, the limit was extended by Adams et al. to be less than $10^{-16}\\,\\mathrm{cm}^{-2}\\,\\mathrm{s}^{-1}\\,\\mathrm{sr}^{-1}$ for monopoles with masses below $10^{17}\\,\\mathrm{GeV}$ \\cite{ExtensionParkerBound}. Many experiments have searched for relic magnetic monopoles, but there is no experimental proof for their existence. The current best limits for magnetic monopoles constrain their flux to a level of $\\sim 10^{-16}-10^{-18}\\,\\mathrm{cm}^{-2}\\,\\mathrm{s}^{-1}\\,\\mathrm{sr}^{-1}$ depending on the monopole speed and interaction mechanism \\cite{FinalMACRO,MACRO:SlowMonopoles,icecube:relMPs,IceCube:ICRC2013_relMPs}. Consequently, searches for magnetic monopoles require very large detectors. The IceCube Neutrino Observatory currently is the world's largest neutrino detector. The primary goal is the detection of Cherenkov light from electrically-charged secondary particles produced when high-energy astrophysical neutrinos interact in the surrounding matter \\cite{IceCubeSensitivity}. However, IceCube can also be used to search for magnetic monopoles. Depending on their speed monopoles have different signatures in IceCube. Relativistic monopoles with a speed above the Cherenkov threshold, e.g. $\\beta \\approx 0.76$ in ice, can be detected by the Cherenkov light they directly produce \\cite{Tompkins:CherenkovEmission}. Non-relativistic monopoles that catalyze the decay of nucleons in the detector medium can, in contrast, be detected by the Cherenkov light from electrically charged secondary particles produced in subsequent nucleon decays along the monopole trajectory (Sec. \\ref{subsec:Rubakov-Callan_Effect}). Therefore, different analysis strategies are needed in order to cover both detection channels. This paper presents the results of a search for non-relativistic magnetic monopoles which would catalyze the proton decays via the Rubakov-Callan effect in IceCube. ", "conclusions": "Data taken from May 2011 until May 2012 with a dedicated slow-particle trigger and for the brightest monopoles data taken from May 2009 until May 2010 with standard-IceCube-triggers were analysed. The analysis, which is based on data of the slow-particle trigger, was developed by using simulated monopole events and experimental data to estimate background properties. For this first analysis of such a signal in IceCube a robust approach based on a single final selection criterion and the comparison between the number of expected background events and observed experimental events is chosen. Using experimental data, the number of expected background events can be estimated to $n_{\\mathrm{b}} = 3.2^{+1.8}_{-1.1}$. The IC-59 analysis based on standard-IceCube-triggers is sensitive only for bright monopoles with $\\sigma_{\\mathrm{cat}} > 1.7 \\cdot 10^{-23}\\,\\mathrm{cm^2}$. The analysis used Boosted Decision Trees (BDT) to discriminate between monopole signal and background. The expected number of background events is derived from a fit of the BDT scores tails with an exponential function for each ($\\beta$,$\\lambda_{\\mathrm{cat}}$). The number of observed events after unblinding is $1$ for an expected background of $4.8_{-0.6}^{+0.7}$. This event contains multiple coincident muons, which renders it compatible with a background event. The obtained flux limits for $\\beta=10^{-2}$ and $\\lambda_{\\mathrm{cat}} =0.01$\\,m, $0.001$\\,m from the IC-59 analysis are better than the ones from the IC-86/DeepCore analysis because of the bigger effective area. For $\\beta=10^{-3}$ the limits are comparable since the standard IceCube triggers are less sensitive to the monopole signal in comparison to the dedicated slow-particle trigger. In both analyses no monopole signal has been observed. Thus, the limits on the flux of non-relativistic magnetic monopoles -- catalysing the proton decay -- are improved by about more than one order of magnitude in comparison to MACRO \\cite{MACRO:SlowMonopoles} for most of the investigated parameter space and reach down to about three orders of magnitude below the Parker limit. Since May 2012 the dedicated slow-particle trigger has been updated to the full IceCube detector. From this upgrade, we expect an improvement in sensitivity by roughly an order of magnitude \\cite{IceCube:ICRC2013_slowMPs}. This gain is supplemented by improvements of the data selection which have been developed after completion of this analysis. Examples are the implementation of a Kalman-filter based HLC hit selection, which improves the angular and speed reconstruction, and the implementation of an event selection based on a Boosted Decision Tree \\cite{Zierke:Masterarbeit}." }, "1402/1402.4130_arXiv.txt": { "abstract": "The recently discovered dwarf galaxy Perseus I appears to be associated with the dominant plane of non-satellite galaxies in the Local Group (LG). We predict its velocity dispersion and those of the other isolated dSphs Cetus and Tucana to be 6.5, 8.2, and $5.5\\,\\mathrm{km\\,s}^{-1}$, respectively. The NGC 3109 association, including the recently discovered dwarf galaxy Leo P, aligns with the dwarf galaxy structures in the LG such that all known nearby non-satellite galaxies in the northern Galactic hemisphere lie in a common thin plane (rms height 53 kpc; diameter 1.2 Mpc). This plane has an orientation similar to the preferred orbital plane of the Milky Way (MW) satellites in the vast polar structure. Five of seven of these northern galaxies were identified as possible backsplash objects, even though only about one is expected from cosmological simulations. This may pose a problem, or instead the search for local backsplash galaxies might be identifying ancient tidal dwarf galaxies expelled in a past major galaxy encounter. The NGC 3109 association supports the notion that material preferentially falls towards the MW from the Galactic south and recedes towards the north, as if the MW were moving through a stream of dwarf galaxies. ", "introduction": "The Milky Way (MW) is surrounded by a vast polar structure (VPOS) of satellite objects including the satellite galaxies, young halo globular clusters and several stellar and gaseous streams \\citep*{1976MNRAS.174..695L,2012MNRAS.423.1109P}. The proper motions of the 11 classical satellite galaxies reveal that these almost exclusively co-orbit in this VPOS, which allowed us to predict the proper motions of the remaining satellite galaxies \\citep{2013MNRAS.435.2116P}. \\citet*{2005A&A...431..517K} have first identified this planar alignment as being inconsistent with cosmological simulations based on the cold dark matter paradigm with a cosmological constant, $\\Lambda$CDM. This finding subsequently triggered an ongoing debate on whether such structures can be reconciled with cosmological expectations \\citep[e.g.][]{2008ApJ...686L..61D,2008MNRAS.385.1365L,2009ApJ...697..269M,2009MNRAS.399..550L,2011MNRAS.415.2607D,2012MNRAS.424...80P,2013MNRAS.429.1502W,2013MNRAS.435.2116P} or rather points at a different origin such as the formation of phase-space correlated tidal dwarf galaxies \\citep[TDGs, e.g.][]{2005PASJ...57..429S,2007MNRAS.376..387M,2010ApJ...725L..24Y,2011A&A...532A.118P,2012MNRAS.427.1769F,2013MNRAS.431.3543H,2013MNRAS.429.1858D}. \\citet{2013Natur.493...62I} and \\citet{2013ApJ...766..120C} have recently discovered a similar 'Great Plane of Andromeda' (GPoA), a co-orbiting alignment consisting of about half of the satellite galaxies of the Andromeda galaxy (M31), the other major galaxy in the Local Group (LG). Motivated by this discovery that satellite galaxies appear to preferentially live in phase-space correlated structures, \\citet{2013MNRAS.435.1928P} set out to search for similar structures on a LG scale. They have discovered that all but one of the 15 LG dwarf galaxies more distant than 300\\,kpc from the MW and M31 are confined to two narrow (short-to-long axis ratios of $\\approx 0.1$) and highly symmetric planes, termed LGP1 and LGP2. LGP1 is the dominant plane both by number of objects (about nine), and alignment with additional features, such as the Magellanic Stream which traces the positions and line-of-sight (LOS) velocities of the LGP1 plane members in the southern Galactic hemisphere. Given that the number of known dwarf galaxies in the LG more distant than 300\\,kpc from both major galaxies is still low, each additional detection poses a chance to test the existence of the planar LG structures and to potentially refine our understanding of these structures. Such an opportunity is now provided by the recent discovery of the dwarf spheroidal (dSph) galaxy Perseus I at a distance of 374\\,kpc from M31 \\citep{2013arXiv1310.4170M}. In the following, we test whether it can be considered to be associated to either LGP1 or LGP2. In addition, we predict the velocity dispersion of Perseus I and two other non-satellite dSphs as expected in Modified Newtonian Dynamics \\citep[MOND,][]{1983ApJ...270..365M,2012LRR....15...10F}. Similar predictions have been made for other dwarf galaxies in M31's vicinity \\citep{2013ApJ...766...22M} and these have successfully passed the test of observations \\citep{2013ApJ...775..139M}. Unfortunately, no similar predictions are possible in the $\\Lambda$CDM framework. Another recently discovered nearby dwarf galaxy, Leo P \\citep{2013AJ....146...15G,2013AJ....145..149R}, has lead \\citet{2013arXiv1310.6365B} to re-investigate the NGC 3109 association, a group of dwarf galaxies at a distance of about 1.3--1.4\\,Mpc from the MW that consists of NGC 3109, Antlia, Sextans A and Sextans B \\citep{1999ApJ...517L..97V,2006AJ....132..729T}. They realised that Leo P aligns with the four other members of the association in a very narrow, linear structure. As the NGC 3109 association is very close to the LG and has a linear extent of 1.2\\,Mpc, similar to its distance from the MW, we will investigate its orientation in the context of the LG planes of non-satellite dwarf galaxies. This reveals an intriguing alignment with the other three nearby non-satellite galaxies in the northern hemisphere of the MW and leads us to discuss suggested origins for the NGC 3109 association in light of the geometry of the LG. The paper is structured as follows. In Sect. \\ref{sect:perseus} we determine whether Perseus I is associated with one of the dwarf galaxy planes in the LG. In Sect. \\ref{sect:MOND} we predict the velocity dispersion of the distant dSphs in the LG, Perseus I, Cetus and Tucana. In Sect. \\ref{sect:NGC3109association} we determine the orientation of the NGC 3109 association in the same coordinate system used in \\citet{2013MNRAS.435.1928P}, discuss possible origins for the alignment and conclude that the association is likely part of the LG dwarf galaxy structures. In Sect. \\ref{sect:backsplash} we discuss how the search for cosmological backsplash galaxies in the LG might give rise to two additional small-scale problems of cosmology and how it could falsely identify TDGs as backsplash objects. In Sect. \\ref{sect:scheme} we present a sketch of the LG dwarf galaxy structures and their preferred direction of motion and discuss open questions and limitations in Sect. \\ref{sect:openquestions}. Finally, we summarize our results in Sect. \\ref{sect:conclusion}. ", "conclusions": "\\label{sect:conclusion} We have investigated how the recently discovered dSph galaxy Perseus I and the NGC 3109 association extended with the recently discovered dwarf galaxy Leo P are related to, and might fit in with, the dwarf galaxy structures present in the LG. Our work has shown that the NGC 3109 association cannot necessarily be interpreted as an independent group of galaxies, but might be related to the LG dwarf galaxy population and as such might provide important constraints on attempts to model the whole LG. The main results of our analysis are: \\begin{enumerate} \\item Perseus I is consistent with being part of the LGP1$^{\\mathrm{mod}}$, the dominant plane of non-satellite galaxies in the LG, at least if Andromeda XVI is associated with the GPoA. \\item In the context of MOND, we have predicted Perseus I's velocity dispersion to be $\\sigma = 6.5^{+1.2}_{-1.0} \\pm 1.1\\,\\mathrm{km\\,s}^{-1}$. The corresponding prediction for Cetus ($\\sigma = 8.2^{+1.5}_{-1.3} \\pm 0.4\\,\\mathrm{km\\,s}^{-1}$) is in much better agreement with the more recent observational value \\citep[$\\sigma = 8.3 \\pm 1.0\\,\\mathrm{km\\,s}^{-1}$:][]{2014arXiv1401.1208K} than with the previous measurement \\citep[$\\sigma = 17 \\pm 2.0\\,\\mathrm{km\\,s}^{-1}$:][]{2007MNRAS.375.1364L}. The prediction for Tucana ($\\sigma = 5.5^{+1.0}_{-0.9} \\pm 0.4\\,\\mathrm{km\\,s}^{-1}$) is in conflict with the available measurement \\citep[$\\sigma = 15.8^{+4.1}_{-3.1}\\,\\mathrm{km\\,s}^{-1}$:][]{2009A&A...499..121F}. We note that the observations of \\citet{2009A&A...499..121F} lack the spectral resolution to resolve the predicted velocity dispersion. Unfortunately, no similar predictions are possible in a $\\Lambda$CDM context. \\item The orientation of the NGC 3109 association consisting of the dwarf galaxies NGC 3109, Antlia, Sextans A, Sextans B and Leo P has been determined in the same coordinate system used to review the planes of co-orbiting satellite galaxies around the MW and M31 and the symmetric larger-scale dwarf galaxy structure in the LG \\citep{2013MNRAS.435.1928P}. The association aligns with the Supergalactic Plane, is almost perpendicular to our line-of-sight and parallel but offset by 300-500\\,kpc to LGP1$^{\\mathrm{mod}}$. \\item The members of the NGC 3109 association have large receding velocities which indicate that they have been close to the MW in the past, possibly at the same time about 7--9\\,Gyr ago. This is consistent with their orbits passing within $\\approx 25$\\,kpc of the MW suggested by \\citet{2013MNRAS.tmp.2500S} and the identification as likely backsplash galaxies by \\citet{2012MNRAS.426.1808T}. Together with the association's extremely narrow extent and perpendicular orientation this argues against the association tracing a thin and cold cosmological filament. The timing is consistent with independent timing estimates for several suggested major galaxy encounter scenarios in the LG, during which phase-space correlated populations of TDGs could have formed that would today give rise to the observed dwarf galaxy structures \\citep{2012MNRAS.423.1109P,2013MNRAS.431.3543H,2013A&A...557L...3Z}. \\item The association aligns with the other three distant ($> 300$\\,kpc) LG galaxies in the northern hemisphere of the MW in a narrow plane (RMS height of $\\Delta\\,=\\,53$\\,kpc and axis ratios of $c/a\\,=\\,0.10$\\ and $b/a\\,=\\,0.76$). This ``Great Northern Plane'' passes through the center of the LG, is inclined to LGP1$^{\\mathrm{mod}}$\\ by only $27^\\circ$ and to the GPoA by $40^\\circ$ and is consistent with being aligned with the preferred orbital plane of the MW satellites in the VPOS. \\item Five out of seven (6 of 8 if the later discovered Leo P would be included) of the galaxies in the Great Northern Plane have been identified as likely backsplash objects by \\citet{2012MNRAS.426.1808T}, and most of the remaining five backsplash candidates in the southern Galactic hemisphere are also situated close to the same plane. As only a small fraction of sub-haloes in simulations are identified as backsplash objects the finding of a majority of such galaxies in one hemisphere is extremely unlikely ($\\approx 0.1$\\ per cent) and might constitute an ``overabundant backsplash problem'' for $\\Lambda$CDM. It would mean that more galaxies are receding in one direction from the MW than are being accreted onto the MW from that direction. A natural explanation for this would be the local formation of galaxies as TDGs, which, if expelled to large distances, have very similar orbital properties like cosmological backsplash galaxies. That the backsplash candidates preferentially lie in a common plane is also consistent with an interpretation as phase-space correlated TDGs. \\item LG galaxies are found to be preferentially infalling in the Galactic south and receding in the Galactic north, which possibly indicates that the MW is moving through a stream of dwarf galaxies. This would be in qualitative agreement with the M31-merger scenario by \\citet{2010ApJ...725..542H}, in which our Galaxy passes through the tidal debris expelled in a past merger forming M31 \\citep[see also][]{2010ApJ...725L..24Y,2012MNRAS.427.1769F,2013MNRAS.431.3543H}. However, the TDG origin would be in M31, such that the galaxies move away from M31 sufficiently fast which would imply that they have a tangential velocity component relative to the MW, rendering the timing estimate of Sect. \\ref{sect:timing} somewhat useless. The scenario should however be testable on LG scales if proper motion measurements of the distant dwarf galaxies could be obtained. \\end{enumerate}" }, "1402/1402.3474_arXiv.txt": { "abstract": "{To deepen our understanding of the role of small-scale magnetic fields in active regions (ARs) and in the quiet Sun (QS) on the solar irradiance, it is fundamental to investigate the physical processes underlying their continuum brightness. Previous results showed that magnetic elements in the QS reach larger continuum intensities than in ARs at disk center, but left this difference unexplained.} {We use Hinode/SP disk center data to study the influence of the local amount of magnetic flux on the vigour of the convective flows and the continuum intensity contrasts.} {The apparent (i.e. averaged over a pixel) longitudinal field strength and line-of-sight (LOS) plasma velocity were retrieved by means of Milne-Eddington inversions (VFISV code). We analyzed a series of boxes taken over AR plages and the QS, to determine how the continuum intensity contrast of magnetic elements, the amplitude of the vertical flows and the box-averaged contrast were affected by the mean longitudinal field strength in the box (which scales with the total unsigned flux in the box).} {Both the continuum brightness of the magnetic elements and the dispersion of the LOS velocities anti-correlate with the mean longitudinal field strength. This can be attributed to the ``magnetic patches'' (here defined as areas where the longitudinal field strength is above 100 G) carrying most of the flux in the boxes. There the velocity amplitude and the spatial scale of convection are reduced. Due to this hampered convective transport, these patches appear darker than their surroundings. Consequently, the average brightness of a box decreases as the the patches occupy a larger fraction of it and the amount of embedded flux thereby increases.} {Our results suggest that as the magnetic flux increases locally (e.g. from weak network to strong plage), the heating of the magnetic elements is reduced by the intermediate of a more suppressed convective energy transport within the larger and stronger magnetic patches. This, together with the known presence of larger magnetic features, could explain the previously found lower contrasts of the brightest magnetic elements in ARs compared to the QS. The inhibition of convection also affects the average continuum brightness of a photospheric region, so that at disk center, an area of photosphere in strong network or plage appears darker than a purely quiet one. This is qualitatively consistent with the predictions of 3D MHD simulations.} ", "introduction": "\\label{sec_intro} Variations of the Total Solar Irradiance (TSI) on timescales from days to the solar cycle can be reasonably well reproduced based on the evolving distribution of solar surface magnetic fields \\citep[which is the basis of e.g. the SATIRE reconstructions,][]{Fligge00, Solanki02, Krivova03, Wenzler06}. Although the TSI reconstructions proved very successful \\citep[with a correlation of up to 0.98 with the measured irradiance,][]{Ball11}, they are based on an empirical relationship between the emergent intensity of a pixel and the measured magnetic field (via MDI magnetograms) whose simple linearity lacks physical basis; its justification relies on its ability to reproduce the disk-average irradiance and the convenience of using a single model atmosphere for all faculae \\citep[for more details see][]{Fligge00}. In particular, the spectral radiance of a given pixel is assumed to be \\emph{uniquely} determined by the magnetic signal at that pixel, thus neglecting its magnetic environment and location on the Sun. The surface magnetic field outside Sunspots is mainly distributed in active region (AR) plages and in the quiet Sun (QS) network outlining supergranular cells. There the magnetic flux is often concentrated into features traditionally described in terms of flux tubes, with field strengths of the order of kG \\citep[e.g.][]{Frazier72, Stenflo73, Rabin92a, Ruedi92} and a spectrum of sizes \\citep[see][for a review]{Solanki_rev06}. At the lower end of this size spectrum are found the so-called ``magnetic elements'' \\citep[see][]{Schuessler_rev92, Solanki_rev93}, which have been spatially resolved recently by the balloon-borne SUNRISE ImaX observations at an angular resolution of $\\sim 0\\dotarsec14$ \\citep[][]{Lagg10}. As a result of being hotter than their surroundings at equal optical depth \\citep{Schuessler_rev92}, they often appear brighter than the average quiet photosphere, i.e. they have a positive ``contrast''. This is particularly pronounced when observing in the core of spectral lines and in molecular bands \\citep{Chapman68, Sheeley69, Muller84}, but also holds at continuum wavelengths or broader wavelength bands, even at disk center, if the magnetic elements are sufficiently resolved \\citep[see][the two last citations refering to radiation-MHD simulations]{Muller83b, Foukal84, Lawrence88, Riethmueller10, Schuessler88, Voegler03}. Note that the contrast of magnetic features further increases from the disk center to the limb in a way that is still a matter of debate \\citep{Steiner07} and shall be treated in a forthcoming paper. Therefore, taking into account the larger area coverage of magnetic elements compared to sunspots, their radiance overcompensates the solar brightness deficit due to sunspots at activity maximum \\citep{Froehlich00}. It is known that ARs typically harbour more magnetic flux per unit area than the network and thus a higher number density of magnetic elements. The larger available flux in ARs also leads to the formation of larger and darker magnetic features such as pores and micropores, the latter being small (sub-arcsec) magnetic features that are darker than the QS continuum \\citep[][]{Beckers68, Tarbell77}. The inner part of supergranulation cells, the so-called ``internetwork'' (IN), also contains small magnetic elements but in lower number density \\citep[][]{Muller83a, Lites02, Dominguez03, deWijn05, deWijn08}. Finally, the solar photosphere contains weaker equipartition fields everywhere \\citep[detected mostly in the IN by, e.g.][]{Lin95, Solanki96, Lites02, Khomenko03}, but the latter have been estimated to bring negligible contributions to the TSI variations \\citep{Schnerr10}. To determine how these different components of solar surface magnetism contribute to the solar irradiance, one should investigate the intensity-magnetic field relation within ARs, the QS network and even the IN. In the 1990's, \\citet{Title89, Topka92, Lawrence93} and \\citet{Topka97} performed a series of ground-based studies of continuum intensity vs. magnetogram signal in ARs and in the QS at disk center, and found that the QS continuum intensity contrast reaches larger values than in ARs at equal magnetogram signal. This early result was confirmed by our recent study using Hinode/SP data at a constant and higher spatial resolution than the earlier studies \\citep[][hereafter Paper I]{Kobel11a}. Note that in our study, the contrast of magnetic elements could be interpreted as a measure of their intrinsic brightness by virtue of the comparable contrast references used in ARs and in the QS. We also found that the bright magnetic elements in ARs and in the QS share similar filling factors (along with similar kG field strengths and inclinations close to vertical). Assuming that, at Hinode's resolution and at disk center, the filling factor reflects the size of the unresolved magnetic elements (at the height of line formation), this questions the conventional interpretation that the brightness of magnetic features is \\emph{primarily} dictated by their size \\citep[][see also the Introduction of Paper I]{Spruit81}. As discussed in Paper I, the other factor that could possibly influence the brightness of magnetic elements (besides their size) is the surrounding convective energy transport, because it determines the energy available to be radiated into the flux tubes through their ``hot walls'' \\citep[][]{Spruit76}. Since AR plages typically have a larger mean field strength than the QS network, the Lorentz force should inhibit the convective flows more strongly in ARs. The granulation indeed appears ``abnormal'' in plages \\citep[e.g.][]{Dunn73}, such that its spatial scale is reduced compared to the QS, as are both the vertical and horizontal components of the flow field \\citep{Schmidt88, Title89, Title92, Keller91}. 3D MHD simulations also disclose a similar behaviour as the mean field increases \\citep{Voegler05b}. Furthermore, these carefully set up simulations (with constant inflowing entropy density at the bottom boundary) demonstrated that the convective energy transport was increasingly inhibited to the point of altering the \\emph{vertical} radiative energy output from the boxes, thus providing an explanation for magnetic elements appearing darker in more active environments at disk center. The work presented herein is an attempt to explain the effect of the local amount of magnetic flux on the continuum brightness of magnetic elements and on the local average continuum brightness of the solar surface at disk center. Inspired by MHD simulations, we carried out a ``local box analysis'' that considers a series of small square fields of view taken from different regions of QS and ARs (see Sect. \\ref{sec_boxes}) located at disk center, observed with the spectropolarimeter instrument onboard Hinode. In Sect. \\ref{sec_results}, we examine the relation between the contrast and the longitudinal field strength inside these boxes. In particular, we assess the influence of the mean longitudinal field strength in the box on the intensity contrast of the magnetic elements, on the vigour of the vertical convective flows, and on the contrast averaged over the box. One reason to restrict our study to disk center is that in this way we can interpret the Doppler velocities retrieved by Milne-Eddington inversions from the data (see Sect.\\ref{sec_scans_inversions}) in terms of vertical flows. These results are then discussed in Sect. \\ref{sec_disc}, with special emphasis on the role of the magnetically-suppressed convection on the contrasts. ", "conclusions": "\\label{sec_concl} Previous studies found that at disk center, the continuum intensity contrast in the QS network reached higher values than in ARs \\citep{Title89, Topka92, Lawrence93} and \\citep{Topka97}. This was confirmed by the first paper of this series \\citep[][Paper I]{Kobel11a} using Hinode/SP data, where we established that the relation between continuum brightness (at 630.2 nm) and longitudinal field strength $B_{\\rm app,los}$ (retrieved by Milne-Eddington inversions) exhibits a higher peak in the QS network than in ARs. Since the magnetic elements producing the peak share similar magnetic filling factors, we argued that this brightness difference between the QS and ARs is unlikely to be explained solely by a difference in the size of the brightest magnetic elements (although the presence of larger features in ARs certainly plays a role in determining their brightness more generally, see below). In this paper, we tested whether the contrast of the magnetic elements at disk center could be affected by an altered convective transport due to the local concentration of magnetic flux. With this aim, we extracted from Hinode/SP scans a series of ``boxes'' containing different amounts of magnetic flux, covering areas from AR plages to weak QS network. We found that the contrast of the brightest magnetic elements (i.e. the peak of the contrast vs. $B_{\\rm app,los}$) continuously decreases with the mean longitudinal field strength $\\left< B_{\\rm app,los} \\right>$ in the boxes (Fig. \\ref{fig_CVlos_vs_avgB}a). Since the $\\left< B_{\\rm app,los} \\right>$ is generally larger in ARs than in the QS, this can be taken as a generalization of the earlier finding that magnetic elements in the network are brighter than in ARs. Hence, as the amount of flux (and thereby $\\left< B_{\\rm app,los} \\right>$) increases in a local area, the observed contrast of the magnetic elements associated to the contrast peak decreases due to a combination of two effects: (i) The known presence of larger (and thus darker) magnetic features, even if pores are removed. These can either provide pixels. Such larger magnetic features reduce the contrast of nearby magnetic elements through the telescope point spread function. Yet this effect is likely to play only a minor role in the decrease of the magnetic element contrasts with $\\left< B_{\\rm app,los} \\right>$ (see discussion in Sect. \\ref{sec_disccontrast}). (ii) The disturbance of convective energy transport in the surrounding of the magnetic features. This is supported by the anti-correlation between the fluctuation of the longitudinal velocities outside the magnetic features and $\\left< B_{\\rm app,los} \\right>$ (Fig. \\ref{fig_CVlos_vs_avgB}b) and the lower brightness of the abnormal granulation surrounding magnetic elements compared to weakly magnetized areas (Fig. \\ref{fig_avgC_vs_avgB_patches}b). A closer inspection of the strong network and plage boxes revealed that the magnetic flux in these boxes is mostly carried by ``magnetic patches'' that are well delimited by $B_{\\rm app,los} > 100$ G (Fig. \\ref{fig_Phi_fA_vs_avgB}a). The flux density in these patches is such that the vertical convective velocities inside them (around magnetic features) are reduced significantly relative to their surroundings (the rms($v_{\\rm los}$) is about 100 m s$^{-1}$ lower), this reduction being more important as the mean field of the patches increases (cf. Fig. \\ref{fig_rmsVlos_vs_avgB_patches}b). These reduced velocities probably limit the heat upflow and consequently lower the brightness of the granulation in the patches compared to the granulation in the weakly magnetized areas (Fig. \\ref{fig_avgC_vs_avgB_patches}b). The reduced spatial scale of the (abnormal) granulation inside the patches probably also contributes to decreasing the efficiency of the convective energy transport. Even with their magnetic elements, the patches are generally darker than their surroundings (Fig. \\ref{fig_avgC_vs_avgB_patches}a), and become all the more dark when $\\left< B_{\\rm app,los} \\right>$ increases in the box and the patches occupy a larger part of it (Fig. \\ref{fig_Phi_fA_vs_avgB}b). We interpret the patches as having the following influence on the brightness of the magnetic elements: \\emph{the more extended the magnetic patch is (independently of the box size) and the larger its mean field, the larger is the inhibition of convection, resulting in cooler granules and smaller heat influx to the ``hot walls'' of the magnetic elements which become cooler and less bright (at least at disk center)}. Since the patches are darker than their surrounding and occupy a larger part of the box as $\\left< B_{\\rm app,los} \\right>$ increases, the average brightness of a box at disk center also decreases with increasing $\\left< B_{\\rm app,los} \\right>$ (by about 1.2\\% from weak network to plage boxes). Previous 3D MHD simulations of \\citet{Voegler05b} already predicted a reduction of the vertical bolometric intensity of boxes with mean field strength $\\left< B\\right> > 200$ G. The qualitative agreement between our results and these simulations (for $\\left< B\\right> > 200$ G) supports their use to model the different component atmospheres used in irradiance reconstructions \\citep[see, e.g.,][]{Unruh09}. Up to now, the effect of the inhibition of convection on the photospheric contrast has not been taken into account in irradiance reconstruction models. These models only use the photometric area of magnetic features \\citep{Penza03, Fontenla05} or their ``filling factor'' within a pixel \\citep[in the case of SATIRE, see][for reviews]{Solanki05, Krivova10} as the time-dependent variables modulating the intensity output given by semi-empirical 1D model atmospheres. Our results suggest that the reconstructions could be possibly enhanced by modulating the emitted intensity (e.g. for a given magnetic filling factor) according to the surrounding magnetic flux in a local (e.g. in a $20\\arsec \\times 20\\arsec$) area. However, the finding that the inhibition of convection in more active areas affects the brightness of magnetic elements has been tested here for the disk center only. In the next article of this series, we shall investigate this effect away from the disk center by comparing center-to-limb variation of the continuum contrast of magnetic elements in the QS and in ARs separately." }, "1402/1402.4853_arXiv.txt": { "abstract": "X--ray photons coming from an X--ray point source not only arrive at the detector directly, but also can be strongly forward-scattered by the interstellar dust along the line of sight (LOS), leading to a detectable diffuse halo around the X--ray point source. The geometry of small angle X--ray scattering is straightforward, namely, the scattered photons travel longer paths and thus arrive later than the unscattered ones; thus the delay time of \\xshalo\\ photons can reveal information of the distances of the interstellar dust and the point source. Here we present a study of the \\xshalo\\ around \\src , which is one of the so--called supergiant fast X--ray transients. \\src\\ underwent a striking outburst when observed with \\chd\\ on 2004 July 3, providing a near $\\delta$--function lightcurve. We find that the X--ray scattered halo around \\src\\ is produced by two interstellar dust clouds along the LOS. The one which is closer to the observer gives the \\xshalo\\ at larger observational angles; whereas the farther one, which is in the vicinity of the point source, explains the halo with a smaller angular size. By comparing the observational angle of the scattered halo photons with that predicted by different dust grain models, we are able to determine the normalized dust distance. With the delay times of the scattered halo photons, we can determine the point source distance, given a dust grain model. Alternatively we can discriminate between the dust grain models, if the point source distance is known independently. ", "introduction": "Immersed in the interstellar medium (ISM), the interstellar dust grains not only absorb X--ray photons but also scatter X--ray photons. \\citet{ove65} first predicted the presence of \\xshalo s around X--ray sources. Nonetheless due to the limitation of the angular resolution of X--ray imaging telescopes, it was not until two decades later, \\citet{rol83} first observed the diffuse \\xshalo\\ around GX 339$-$4 with the Imaging Proportional Counter (IPC) instrument onboard \\ein . The theory of small angle X--ray scattering has since been refined by a number of authors \\citep[e.g.,][etc]{mau86,mat91,smi98}. Meanwhile, X--ray scattering phenomena around both X--ray point sources and extended supernova remnants (SNRs) were found with \\ros\\ \\citep[e.g.,][]{pre95}, \\chd\\ \\citep[e.g.,][]{smi02}, \\xmm\\ and \\swi /XRT \\citep[e.g.,][]{tie10}, etc. Likewise, albeit it had already been pointed out by \\citet{tru73} that the time delay effect in the \\xshalo s can be used to determine the distances of variable X--ray point sources, it was nearly three decades later then came its long overdue application. \\citet{pre00} roughly estimated the distance of Cyg X$-$3, which was observed with the Advanced CCD Imaging Spectroscopy (ACIS) instrument onboard \\chd . Thanks to the fine angular resolution of \\chd , \\citet{tho09},~\\citet{lin09}, and~\\citet{xia11} used the time delay effect in the \\xshalo s to determine the distances of Cen X$-$3, Cyg X$-$3, and Cyg X$-$1, respectively. Additionally, \\citet{vau04} first expanded the scope onto $\\gamma$--ray bursts (GRBs), in which case an even simpler geometry was involved. The evolving \\xsring s around GRB~031203 \\citep{vau04, tie06}, GRB~050724 \\citep{vau06}, GRB~050713A \\citep{tie06}, GRB~061019 and GRB~070129 \\citep{via07} were detected with both \\xmm\\ and \\swi /XRT. It is indisputable that the characteristics of the X--ray dust scattered halos (e.g. the radial profile, etc) depend upon the properties of interstellar dust grains, including chemical composition, dust size distribution, dust spatial distribution, and sometimes the morphology and the alignment of elongated dust grains might play important roles. So far, various types of interstellar dust grain models have been established based on different observational results such as interstellar extinction, diffuse infrared emission feature and so forth \\citep[see][for a review]{lia03}. Among them, the three types of models provided by \\citet[][hereafter MRN]{mat77}, \\citet[][hereafter WD01]{wei01}, \\citet[][hereafter ZDA]{zub04} are most widely used. In this work we analyze the archival \\chd\\ data of \\src , which was first discovered with \\intg\\ on 2003~September~17 \\citep{sun03} as a galactic high mass X--ray binary (HMXB). With several subsequent observations carried out with \\xmm\\ \\citep{gri04}, \\chd\\ \\citep{iza05}, \\swi /XRT \\citep{sid09}, and \\suz\\ \\citep{ram09}, remarkable outbursts with a duration time scale of hours were spotted. Along with the confirmation of an O9Ib blue supergiant donor in the binary system \\citep{pel06}, \\src\\ was confirmed as one of the so--called supergiant fast X--ray transients \\citep[SFXTs][]{sgu05,smi06,neg06}. SFXT, a subclass of HMXB, is characterised by the presence of a supergiant companion and significant outbursts lasting typically a few hours. Typically, the peak luminosity of a flare can be about a factor of 10$^{3}$--10$^{5}$ times the fainter quiescent X--ray luminosity. In Section~\\ref{sct:data}, we analyze the \\chd\\ ACIS-S data of \\src , focusing on timing (Section~\\ref{sct:timing}) and imaging analysis (Section~\\ref{sct:imaging}). Then, we derive the time lags of the \\xshalo\\ photons via cross correlation method in Section~\\ref{sct:ccf}. In Section~\\ref{sct:theta}, we model the deviation of the arithmetic mean of the observation angle from the mid-value of the angular distance of the annular region. We subsequently present the distance measurement for interstellar dust clouds along the line of sight (LOS) and the point source (Section ~\\ref{sct:dist}). A dynamical distance measurement for \\src\\, obtained from a Galactic Center molecular clouds survey, presented in Section \\ref{sct:clouds_dist}, is consistent with the estimated point source distance. In Section~\\ref{sct:caveats}, we briefly discuss the feasibility of a promising application of the relationship between interstellar dust grain models and the average observational angles for annular halos. Finally, we summarize our results in Section \\ref{sct:summary}. ", "conclusions": "\\label{sct:discussion} \\subsection{Kinematic distance measurements of molecular clouds} \\label{sct:clouds_dist} For $d\\sim3.6$~kpc \\citep{rah08}, the distances for the closer and farther dusts are $\\sim1.8$~kpc and $\\sim3.4$~kpc away from us, respectively. In this subsection we try to find kinematic distance measurements of the molecular clouds along the LOS toward \\src\\, for comparison with our geometrical distances of the dust slabs. A rough estimate of the radial velocity of the molecular clouds where the dust slab is embedded can be made via \\citep{rdu09} \\begin{equation} r=R_\\odot \\sin l \\frac{V(r)}{v_{\\rm los}+V_\\odot \\sin l}, \\label{v_los} \\end{equation} where $r$ is the distance of the molecular cloud to the Galactic Center (GC), $R_{\\odot}$ is the distance of the Sun to GC, {\\it l} is the galactic longitude of the LOS, $V_{\\odot}$ is the rotation velocity of the Sun, $V(r)$ is the rotation velocity of the molecular cloud, and $v_{\\rm los}$ is the projection of $V(r)$ to the LOS, also known as the radial velocity. Assuming $R_{\\odot}=8.5$~kpc, $V_{\\odot}=220$~km s$^{-1}$ and a flat rotation curve (i.e. $V(r)=220$~km s$^{-1}$), we derive $v_{\\rm los}$ of $\\sim3.1$~km s$^{-1}$ and $\\sim7.8$~km s$^{-1}$ for the closer and farther molecular cloud, respectively. In fact, \\citet{dah98} reported an averaged $v_{\\rm los}\\sim5$~km s$^{-1}$ for $^{12}$CO(1-0) emission line for the Southern Clump 2 region, $l\\in(2\\fdg7, 3\\fdg5)$, $b\\in(0\\fdg15, 0\\fdg35)$, which is roughly consistent with our result of the farther dust. \\subsection{Issues with observational angles} \\label{sct:caveats} In Section~\\ref{sct:theta}, we have shown that given the normalized distance of a dust slab, different interstellar dust models predict different average observational angles for scattered photons within certain annular regions. This offers the advantage of estimating the parameter $x$ from data, thus breaking the degeneracy between $x$ and $d$. We tested several interstellar dust grain models with the data of the farther dust, for which the predictions of these models become quite different, because the scattering angle $\\theta_{\\rm sca}=\\theta_{\\rm obs}/(1-x)$ is quite large for the farther dust. It is therefore possible to distinguish among different interstellar dust models, as demonstrated above. However in the calculations of the scattering differential cross section, Gaussian approximation is used for the form factor in the RG approximation. \\citet{smi98} pointed out that the Gaussian approximation leads to deviations at large scattering angle ($\\gtrsim200~{\\rm arcsec}$) and large dust grain size. In our case, i.e., $x>0.95$ and $\\theta_{\\rm obs}\\sim10~{\\rm arcsec}$ for the farther dust slab, we have $\\theta_{\\rm sca}>200$~arcsec. Moreover, the upper limits of the grain size are greater than 0.3~micron, and even reach to 0.9~micron. Therefore the Gaussian approximation may cause considerable inaccuracies to the model predictions. Alternatively, one can turn to Mie theory \\citep{vhu57}, which is more accurate for $E\\sim1~$keV and/or large scattering angles, but numerically more difficult to carry out the calculations. In this work, we analyzed the \\xshalo\\ around \\src\\ with cross correlation method. The main results are summarized as follows: \\begin{enumerate} \\item From the cross correlation functions between the streak lightcurve (used as a proxy for the point source lightcurve) and the lightcurves of the annular halos, we conclude that there are at two interstellar dust clouds along the LOS toward \\src . \\item By comparing the observational angle of the scattered halo photons with that predicted by different dust grain models, the normalized dust distance can be determined independent of the distance of the point source. \\item Given the point source distance of $\\sim3.6$~kpc, the closer dust, which is $\\sim 1.8$~kpc away from us, is responsible for \\xshalo s at larger observational angles ($\\gtrsim12.5$ arcsec). The farther dust, which is quite close to the point source ($\\sim3.4$~kpc away from us), explains the X--ray scattered halos at smaller angular distances ($\\lesssim12.5$ arcsec). \\item We determined the model-dependent point source distances, which are compared with that yielded by IR observations. We find that among the 18 tested dust grain models (MRN, WD01, ZDAs and XLNW), the four dust grain models COMP-AC-S/B, COMP-NC-S/FG are better, COMP-GR-S/FG, COMP-AC-FG, COMP-NC-B are also acceptable, but the rest dust grain models fail to obtain consistent source distance. \\end{enumerate} Similar to the GRBs, the transient nature of SFXTs can, in principle, be used to precisely determine the geometrical distances of interstellar dust and the point source by taking advantage of the time delay effect of the small angle X--ray scattering phenomena. However, we have to face some practical difficulties, such as: 1) the angular resolutions of the space telescopes are relatively poor; 2) the effective area is small and thus the photon counts are relatively low; or 3) for observations of \\xshalo\\ caused by dust slab in the vicinity of the point source, the time lags can be quite large, but no observations with sufficiently long effective exposure times are available. The effective exposure time refers to the exposure time for observing the \\xshalo . For instance, in our work, since the outburst occurs $\\sim$~10~ks after the beginning of this observation, the effective exposure time for observing the \\xshalo\\ here is only $\\sim$~9 ks. With the fine angular resolution, \\chd\\ has the capability to observe \\xshalo s around SFXTs especially at smaller angular distance, although the collecting area of ACIS is small. Unfortunately, insofar as the archival \\chd\\ data, only \\src\\ (ObsID 4550) allows us to study the \\xshalo . In terms of other observations of SFXTs, either the exposure time is only several kiloseconds (e.g. XTE J1739$-302$), or no flaring activity was caught (e.g., for IGR J19410$-$0951). Therefore, we suggest that in the future more long term follow up observations of the outbursts of SFXTs shall be made with \\chd\\ to study the \\xshalo\\ and thus the interstellar dust models in further details." }, "1402/1402.1192_arXiv.txt": { "abstract": "We have determined spectroscopic orbits for five single-lined spectroscopic binaries, HD 100167, HD 135991, HD 140667, HD 158222, HD 217924. Their periods range from 60.6 to 2403 days and the eccentricities, from 0.20 to 0.84. Our spectral classes for the stars confirm that they are of solar type, F9 to G5, and all are dwarfs. Their [Fe/H] abundances, determined spectroscopically, are close to the solar value and on average are 0.12 greater than abundances from a photometric calibration. Four of the five stars are rotating faster than their predicted pseudosynchronous rotational velocities. ", "introduction": "Nearly a decade ago a group of solar-type stars with $V$ magnitudes between 5.9 and 8.1 was selected to expand a photometric program that was designed to detect long-term spot cycles in dwarfs similar to the Sun \\citep{h99}. Although most of these moderately bright stars had radial velocities published in the literature, many did not. To remove binaries from the photometric sample, we obtained spectra and measured radial velocities. Our spectroscopic observations soon identified five single-lined binaries, HD~100167, HD~135991, HD~140667, HD~158222, and HD~217924, which we continued to observe. The basic properties of these binaries are given in Table~\\ref{tbl-basic}. During the course of our extended spectroscopic observing campaign, others have also identified these five stars as binaries. From our observations we determine spectroscopic orbits and spectral and luminosity classes, measure iron abundances and $v$~sin~$i$ values, and then briefly discuss the systems. ", "conclusions": "" }, "1402/1402.5778_arXiv.txt": { "abstract": "{Transition disks are protoplanetary disks with inner depleted dust cavities that are excellent candidates for investigating the dust evolution when there is a pressure bump. A pressure bump at the outer edge of the cavity allows dust grains from the outer regions to stop their rapid inward migration towards the star and to efficiently grow to millimetre sizes. Dynamical interactions with planet(s) have been one of the most exciting theories to explain the clearing of the inner disk.} {We look for evidence of millimetre dust particles in transition disks by measuring their spectral index $\\alpha_{\\mathrm{mm}}$ with new and available photometric data. We investigate the influence of the size of the dust depleted cavity on the disk integrated millimetre spectral index.} {We present the 3-millimetre (100~GHz) photometric observations carried out with the Plateau de Bure interferometer of four transition disks: LkH$\\alpha$~330, UX~Tau~A, LRLL~31, and LRLL~67. We used the available values of their fluxes at 345~GHz to calculate their spectral index, as well as the spectral index for a sample of twenty transition disks. We compared the observations with two kinds of models. In the first set of models, we considered coagulation and fragmentation of dust in a disk in which a cavity is formed by a massive planet located at different positions. The second set of models assumes disks with truncated inner parts at different radii and with power-law dust-size distributions, where the maximum size of grains is calculated considering turbulence as the source of destructive collisions.} {We show that the integrated spectral index is higher for transition disks (TD) than for regular protoplanetary disks (PD) with mean values of $\\bar{\\alpha}_{\\mathrm{mm}}^{\\mathrm{TD}}=2.70\\pm0.13$ and $\\bar{\\alpha}_{\\mathrm{mm}}^{\\mathrm{PD}}=2.20\\pm0.07$ respectively. For transition disks, the probability that the measured spectral index is positively correlated with the cavity radius is 95\\%. High angular resolution imaging of transition disks is needed to distinguish between the dust trapping scenario and the truncated disk case.} {} ", "introduction": "\\label{introduction} Observations by infrared telescopes done with e.g. Spitzer and Herschel and by millimetre observations with radio-interferometers such as the Submillimetre Array (SMA), the Very Large Array (VLA) or the Combined Array for Research in Millimeter-wave Astronomy (CARMA) have revealed different types of structures in circumstellar disks surrounding young stars. Transition disks show a lack of mid-infrared radiation in the spectral energy distribution (SED), implying an absence of warm dust in the inner disk \\citep[e.g.][]{calvet2005, pietu06, espaillat10}. Some of these cavities have been resolved by continuum observations \\citep[see][for a review]{williams11}. The recent discoveries of companion candidates inside transition disks, such as HD142527 \\citep{biller2012} and HD169142 \\citep{quanz2013}, promote the idea that holes in transition disks are the result of a brown dwarf, planet, or multiple planets interacting with the disk. However, none of these companion candidates have been confirmed, and alternative explanations have been suggested to explain the observations \\citep{olofsson2013}. Hydrodynamical simulations of planet-disk interactions have been explored for several years \\citep[e.g.][]{lin79, paardekooper04}. The presence of a massive planet in a circumstellar disk not only affects the surrounding gas, but also the dust distribution in the disk. A planet of $\\gtrsim$~1~$M_{\\mathrm{Jup}}$ opens a gap in a viscous disk and dust can flow through a gap, depending on the particle size and the shape of the gap \\citep{rice06, pinilla12a, zhu2012, zhu2013}. The overall dust distribution is unknown, and observations at different wavelengths with high sensitivity and resolution are crucial for understanding how dust evolves under disk clearing environments. Maps of sub-millimetre and microwave emission from the diffuse interstellar dust have shown that the typical values for the dust opacity index $\\beta_{\\rm{mm}}$ are $\\sim~1.7-2.0$ \\citep[see e.g.][]{Finkbeiner1999}. Hence, if the dust in protoplanetary disks is similar to the interstellar medium (ISM) dust, $\\beta_{\\rm{mm}}$ is expected to be similar. However, changes in the opacity between the ISM and denser regions have been found \\citep[e.g.][]{henning1995}. Dust evolution in protoplanetary disks, and especially the process of grain growth to millimetre sizes in the midplane of disks, has been confirmed by observations \\citep[see reviews by][]{natta2007, testi2014}. The slope of the SED, known as the spectral index $\\alpha_{\\mathrm{mm}}$, in the millimetre regime can be interpreted in terms of grains size, via $\\alpha_{\\mathrm{mm}} \\approx \\beta_{\\rm{mm}} +2$ \\citep[see e.g.][]{draine2006}, and low values of $\\alpha_{\\mathrm{mm}}$ ($\\lesssim 3.5$) correspond to big particles ($\\gtrsim$~mm). Sub- and millimetre observations of classical protoplanetary disks have shown that the spectral index reaches low values $\\alpha_{\\mathrm{mm}}\\lesssim 3.5$ in Taurus, Ophiuchus, Orion, Chamaeleon, Lupus, etc., star forming regions \\citep[e.g.][]{testi2001, Testi2003, rodmann2006, lommen2007, lommen2009, lommen2010, Ricci2010a, Ricci2010b, Ricci2011, ubach2012}. Alternative explanations for the low spectral indices in disks have been addressed by different authors. One possible explanation is that the emission or part of the emission may come from unresolved regions where the dust is optically thick at millimetre wavelengths \\citep{testi2001, Testi2003}, but it has been shown that this is unlikely to be the general explanation for the low values of the opacity index measured in large samples of disks \\citep{Ricci2012b}. In addition, the chemical composition \\citep{Semenov2003}, porosity \\citep{henning1996, stognienko1996}, and geometry of the dust particles can affect the opacity index. However, their impact is not significant, for reasonable values of the properties of the dust grains \\citep[see e.g.][]{draine2006}. Thus, the most likely explanation is that dust grains have grown to mm-sizes \\citep[see in addition][]{calvet2002, natta2004a, natta2007}. One of the most critical challenges of planetesimal formation is to understand how the rapid inward migration of dust grains can stop. The theory of radial drift and observations at mm-wavelength are in strong disagreement \\citep[e.g.][]{brauer07}. The confirmation of long-lived presence of mm grains with observations made theoreticians to look for the missing piece of this rapid inward-drift puzzle. Particle trapping in pressure maxima (i.e.\\ locations where $\\nabla p=0$ and $\\nabla^2p<0$) is a result of the dynamic phenomenon that, due to friction with the gas, particles tend to drift in the direction of higher gas pressure, i.e. up the pressure gradient \\citep[e.g.][]{Klahr1997}. Besides the inward drift reduction in pressure bumps, dust grains do not experience the potentially damaging high-velocity collisions due to relative radial and azimuthal drift anymore, thus reducing the fragmentation problem as well. Pressure bumps may themselves be caused by planets that have formed by a yet unknown mechanism. A possible first planet embryo can be explained by the growth of a seed (cm-sized object) by sticking mutual collisions, including mass transfer and bouncing effects as possible outcomes of the collisions \\citep{windmark2012}. Once a massive planet is formed and opens a gap, a pressure bump is expected in the radial direction at the outer edge of the gap due to the depletion in the gas surface density. This pressure bump becomes a ``particle trap'', a region where particles accumulate and grow. This trap is therefore a good planet incubator for the next generation of planets. Particle trapping creates a ring-like emission at mm-wavelengths, whose radial position and structure depend on the disk viscosity, mass, and location of the planet \\citep{pinilla12a}. In addition, the non-axisymmetric emission of transition disks revealed by sub-mm observations \\citep{brown09} and recently confirmed with ALMA and CARMA observations \\citep{casassus13, vandermarel2013, isella2013, fukagawa2013, perez2014} may also be a result of particle trapping that lead to strong asymmetric maps \\citep{birnstiel2013, ataiee2013, lyra2013}. Spatially resolved, multi-frequency observations will provide crucial information about the dust density distribution of transition disks. Radial variations in the dust opacity have been revealed for individual disks \\citep[e.g.][]{perez2012, trotta2013}; however, a statistical study cannot have been done yet. In this paper, we show that in the case of transition disks, a change in the cavity radius affects the spatially integrated spectral index, a quantity already available for a significant number of transition disks. We present 3mm-observations of four transition disks: LkH$\\alpha$~330, UX~Tau~A, LRLL~31, and LRLL~67 carried out with Plateau de Bure Interferometer (PdBI). The observations and data reduction of our four targets are presented in Sect.~\\ref{observations}, together with the information on additional transition disks. Section~\\ref{results} presents the integrated spectral index of our sample with the data collected from the literature. In this section a comparison between the spectral indices of transition and regular protoplanetary disks is presented, as well as a possible correlation between the cavity radii and the spectral index. In section~\\ref{section_discussion} we compare the observations to two different kinds of models: first of all, dust evolution models based on planet-related cavities, where a pressure bump is formed at the outer edge of the gap, which traps millimetre particles; second, models with a power law size distribution of particles with an artificial cut at different radii to imitate dust cavities. Finally, in Sect.~\\ref{conclusion} we give the main conclusions of this work. \\begin{table*} \\label{table1} \\centering \\tabcolsep=0.10cm \\begin{tabular}{c||cccccccccccccc} \\\\ \\textbf{{\\tiny Name}} & {\\tiny Ra} & {\\tiny Dec} & {\\tiny SpT} & {\\tiny $T_{\\mathrm{eff}}$}&{\\tiny d}&{\\tiny $L_\\star$}&{\\tiny $R_\\star$}&{\\tiny $M_\\star$}&{\\tiny $\\dot{M}$}&{\\tiny $R_{\\mathrm{cav}}$} & {\\tiny $F_{\\mathrm{0.8mm}}$(mJy)}&{\\tiny $F_{\\mathrm{3.0mm}} $(mJy)}& {\\tiny $F_{\\mathrm{1.0mm}} $}&$\\alpha_{0.88-3\\mathrm{mm}}$\\\\ &{\\tiny [J2000]}&{\\tiny [J2000]}&&{\\tiny (K)}&{\\tiny (pc)}&{\\tiny ($L_\\odot$)}&{\\tiny ($R_\\odot$)}&{\\tiny ($M_\\odot$)}&{\\tiny ($M_\\odot~$yr$^{-1}$)}&{\\tiny (AU)}& {\\tiny$\\pm \\sigma$(mJy/beam)} & {\\tiny$\\pm \\sigma$(mJy/beam)} &{\\tiny (mJy)} &\\\\ \\hline \\hline \\\\ {\\tiny LkH$\\alpha$~330}&{\\tiny03 45 48.29}&{\\tiny +32 24 11.9} &{\\tiny G3}&{\\tiny5830}&{\\tiny 250}& {\\tiny 15}&{\\tiny 3.75}&{\\tiny 2.2}&{\\tiny 2$.0 \\times 10^{-9}$}&{\\tiny 68$^{R}$}& {\\tiny 210$\\pm$2.1}& {\\tiny 3.9$\\pm$0.12}& {\\tiny 138.6$\\pm$21.1}& {\\tiny 3.25$\\pm$0.40}\\\\ \\\\ {\\tiny UX~Tau~A}&{\\tiny 04 30 04.00}&{\\tiny +18 13 49.3} & {\\tiny G8}&{\\tiny5520}&{\\tiny 140}& {\\tiny 3.5}&{\\tiny 2.05}&{\\tiny 1.5}&{\\tiny $1.0\\times10^{-8}$}&{\\tiny 25$^{R}$}& {\\tiny 150$\\pm$1.5}& {\\tiny 10.1$\\pm$0.33}& {\\tiny 113.2$\\pm$17.2}& {\\tiny 2.20$\\pm$0.40} \\\\ \\\\ {\\tiny LRLL~31}&{\\tiny 03 44 18.00}&{\\tiny +32 04 57.0}&{\\tiny G6}&{\\tiny5700}&{\\tiny 315}& {\\tiny 5.0}&{\\tiny 2.3}&{\\tiny 1.6}&{\\tiny $1.4\\times10^{-8}$}&{\\tiny 14$^{S}$}& {\\tiny 62$\\pm$6.0}& {\\tiny 2.88$\\pm$0.08}& {\\tiny 45.0$\\pm$8.1}& {\\tiny 2.50$\\pm$0.44} \\\\ \\\\ {\\tiny LRLL~67} &{\\tiny 03 44 38.00} &{\\tiny +32 03 29.0} &{\\tiny M0.75}&{\\tiny3720}&{\\tiny 315}& {\\tiny 0.5}&{\\tiny 1.8}&{\\tiny 0.5}&{\\tiny $1.0\\times 10^{-10}$}&{\\tiny 10$^{S}$}& {\\tiny 25$\\pm$11}& {\\tiny 1.21$\\pm$0.09}& {\\tiny 18.2$\\pm$8.5}& {\\tiny 2.47$\\pm$0.92} \\\\ \\\\ \\hline \\end{tabular} \\caption{{\\bf Observed targets.} Column~1: Name of the disks observed with PdBI. Column~2 and~3: Right ascension and declination of the targets. Column~4: Spectral type. Column~5: Effective temperature of the central star. Column~6: Estimated distance to the target. Column~7, 8 and 9: Stellar luminosity, radius and mass respectively. Column~10: Accretion rate. Column~11: Radial cavity extension ($R$ and $S$ implies that the cavity is either resolved or derived from SED modelling, respectively). Column~12: Flux at~345~GHz. The data is from \\cite{andrews11} for LkH$\\alpha$~330 and UX~Tau~A, and from \\cite{espaillat12} for LRLL~31 and LRLL~67 and references therein. Column~13: Flux at~100~GHz from the PdBI observations. Column~14: Calculated flux at~300GHz ($\\sim~1$mm) and the error assuming an additional systematic uncertainty from calibration of 15\\%. Column~15: Integrated spectral index calculated with the observed fluxes (columns 13 and 14).} \\end{table*} ", "conclusions": "\\label{conclusion} We have reported 3mm observations carried out with PdBI of four transition disks LkH$\\alpha$~330, UX-Tau~A, LRLL~31, and LRLL~67. Using previous observations of these targets at 880~$\\mu$m, we calculated the spectral index, and found that for these sources the values for $\\alpha_{\\mathrm{mm}}$ are lower than typical values of the ISM dust. However, the disk-integrated spectral index $\\alpha_{\\mathrm{mm}}$ is close to the value for ISM-dust ($\\sim$~3.7) for LkH$\\alpha$~330 ($\\alpha_{\\mathrm{mm}}=3.25\\pm{0.40}$), which is also the target with the largest inner hole \\citep[$\\sim$~68~AU,][]{andrews11b}. The sources UX-Tau~A and LRLL~31 have smaller cavity radii of 25~AU and 14~AU, respectively, and lower integrated spectral indices, $\\alpha_{\\mathrm{mm}}=2.20\\pm{0.40}$ for UX-Tau~A and $\\alpha_{\\mathrm{mm}}=2.50\\pm{0.44}$ for LRLL~31. Owing to the large uncertainties on the millimetre flux of LRLL~67, there is no conclusive correlation between the integrated spectral index and its cavity radius. Recent observations at different wavelengths have shown that the opacity may change radially throughout the disk \\citep{isella2010a, Guilloteau2011, perez2012, trotta2013}. Studies that combine dust evolution and planet-disk interaction for transition disks \\citep{pinilla12a} indicate that the dust size distribution may present strong spatial variations. In this picture, the inner part is depleted of mm grains leading to radial variations of the dust opacity. Thus, the opacity index has higher values in the inner regions than the outer regions, in contrast to observations of classical protoplanetary disks. From models that combine hydrodynamical simulations and dust evolution models for the planet-related cavities, a linear relation between the disk integrated spectral index and the cavity radius is inferred, because the millimetre emission is dominated by the dust at the outer edge of the cavity, i.e. in the pressure bump. This suggests that the spectral index is higher for larger cavities. Since our own sample is too small to draw statistically significant conclusions, we compiled all the known millimetre fluxes from transition disk sources for which these data exist at least for two separated millimetre wavelengths. Together with our observations, we collecteded a total of twenty transition disks for which the cavity radii are known either by SED modelling or spatially resolved observations, and we calculate the disk integrated spectral indices. We found that the integrated spectral index is higher for transition disks than for regular protoplanetary disks. Moreover, the probability of a positive trend between the spectral index and cavity size is $\\sim$95\\% (Fig.~\\ref{alpha_flux_Rcav2}). Two models are considered in this work to explain this positive correlation: first of all, coagulation and fragmentation of dust in a disk whose cavity is formed by a massive planet located at different positions; second, a disk with different truncation radii and with a power-law dust-size distribution where turbulence sets the maximum grain size. Both models can explain the trend, and multi-wavelength observations with high angular resolution are needed to distinguish between the two cases. Observations with new telescope arrays such as ALMA, EVLA, and NOEMA will provide enough sensitivity and angular resolution to detect the small cavities of many transition disks and constrain the dust opacity distribution of those disks in more detail. Together with scattered imaging, these observations will help us understand the correlation between the spectral index and disk properties, such as the cavity size, hence, the environment where the planets are born." }, "1402/1402.2644_arXiv.txt": { "abstract": "Very Long Baseline Interferometry (VLBI) allows for high-resolution and high-sensitivity observations of relativistic jets, that can reveal periodicities of several years in their structure. We perform an analysis of long-term VLBI data of the quasar S5 1928+738 in terms of a geometric model of a helical structure projected onto the plane of the sky. We monitor the direction of the jet axis through its inclination and position angles. We decompose the variation of the inclination of the inner $2$ milliarcseconds of the jet of S5 1928+738 into a periodic term with amplitude of $\\sim 0.89^{\\circ }$ and a linear decreasing trend with rate of $\\sim 0.05^{\\circ }\\text{yr}^{-1}$. We also decompose the variation of the position angle into a periodic term with amplitude of $\\sim 3.39^{\\circ }$ and a linear increasing trend with rate of $\\sim 0.24^{\\circ }\\text{yr}^{-1}$. We interpret the periodic components as arising from the orbital motion of a binary black hole inspiraling at the jet base and derive corrected values of the mass ratio and separation from the accumulated 18 years of VLBI data. Then we identify the linear trends in the variations as due to the slow reorientation of the spin of the jet emitter black hole induced by the spin-orbit precession and we determine the precession period $T_{\\mathrm{SO}}=4852 \\pm 646$ yr of the more massive black hole, acting as the jet emitter. Our study provides indications, for the first time from VLBI jet kinematics, for the spinning nature of the jet-emitting black hole. ", "introduction": "The existence of Supermassive Black Holes (SMBH) at the centre of galaxies, combined with the important role of galaxy mergers in the hierarchical galaxy formation models and the long time-scale of the SMBH merging process suggest that SMBHs often exist in pairs \\citep[][]{Blandford1986iau,SearleZinn1978,Komossa2006}. The binary passes through three evolutionary stages during the galaxy merger \\citep{Begelman1980}. At first the stellar halo of each galaxy interacts with the central black hole of the other galaxy via dynamical friction \\citep[see e.g.][]{BinneyTremaine1987}. Gradually the black holes sink into a common gravitational potential well, and a bound, widely separated SMBH binary forms \\citep[a detailed review of massive black hole binary evolution can be found in][]{MerrittMilos2005}. After this, a transition stage follows. The separation then decreases due to dynamical friction and beyond a transition radius the dissipation of energy and angular momentum becomes dominated by gravitational radiation. General relativistic effects during this inspiral stage are described in a post-Newtonian (PN) framework by expanding the field equations in terms of the PN parameter $\\varepsilon =Gmc^{-2}r^{-1}$ (where $m$ is the total mass, $r$ is the separation, $G$ is the gravitational constant, and $c$ is the speed of light). PN techniques are involved in the description of merging binary-dynamics if $0.001<\\varepsilon<0.1$, while in the plunge and the ring-down stages numerical calculations are required to characterize the final SMBH. The spins of the black holes generate corrections to the dynamics at $1.5$PN orders through the spin-orbit coupling, at $2$PN through the spin-spin coupling and at even higher orders through various post-Newtonian corrections to these couplings. The complicated PN dynamics in the presence of the spins is described in detail in e.g. \\citet{Barker1975,Barker1979}, \\citet{Kidder1995}, \\citet{Gergely2010b,Gergely2010a}. When the spin and orbital momentum vectors are not aligned, a spin-orbit type precession will cause the spin vectors to slowly rotate about the orbital angular momentum vector. SMBHs are believed to be the main engine of the activity in Active Galactic Nuclei (AGN) \\citep[for a review see e.g.][]{Begelman1984}. Apart from efforts of identifying binary black hole candidates in kpc-scale separated binaries with spatially resolved AGN \\citep[e.g.][]{Komossa2003}, indirect approaches have been also employed for the cases when gravitational radiation already dominates the merger of a sub-pc-scale separated binary. It has been estimated from the jet power, that at least one merger with orbital period of the order of one year should be detected in future blind surveys (at a nominal sky coverage of $ 10^{4}$ deg$^{2}$ with $0.5$ mJy sensitivity) through the electromagnetic (EM) signature of the SMBH binary on the jet \\citep{Shaughnessy2011}, manifesting itself as an increase in the jet power when the binary inspirals toward its barycentre. Relativistic bulk motion in the jet of AGN causes Doppler boosting of the synchrotron radiation of the relativistic particles. Very Long Baseline Interferometry (VLBI) observations reveal superluminally outward moving objects in a number of relativistic jets. Such superluminal motions are only apparent, indicating projection effects. These sources often show extreme variability throughout the EM spectrum. The radio variability can be explained in terms of flaring events following the ejection of jet-components or of a wiggling jet, implying a variable Doppler factor. AGN variability is however mostly non-periodic, hence detecting periodic behaviour of the radio light curve and precessing jets would strongly indicate a SMBH binary at the jet base \\citep[e.g. OJ287,][]{Villata1998}. While a plausible explanation for the precession of the jet is the existence of a merging SMBH binary at its base, alternative models allowing for a single black hole and external causes for the jet precession have been proposed \\citep[e.g. see the Discussion in][and references therein]{Britzen2010}. The precession of a jet originating in a SMBH binary system was discussed in e.g. \\citet[][]{Roos1993,Romero2000,Britzen2001,Lobanov2005,Valtonen2012,Caproni2012}. Several models have been developed to investigate the interaction between a SMBH binary and its astrophysical environment, revealing the presence of the binary in the AGN \\citep[e.g.][]{MacFadyen2008,Tanaka2013}. Recently, the helical modulation of the large-scale radio jet of J1502+1115 was attributed to the effect of the tight pair in a triple supermassive black hole system \\citep{Deane2014}. Observational data have already been employed to identify precessing jet candidates. AGN observables allowed placing constraints on some of the SMBH binary parameters, like the total mass, the orbital period and the separation of the binary, e.g. Mrk 501, \\citet{Villata1999}; 3C 273, \\citet{Romero2000}; BL Lac, \\citet{Stirling2003}; 3C 120, \\citet{Caproni2004}; 3C 345, \\citet{Lobanov2005}; S5 1803+784, \\citet{Roland2008}; NGC 4151, \\citet{Bon2012}. The effect of the spin was not included in the calculations of these works, and the binary motion was approximated by pure Keplerian orbits. The extreme Kerr-limit of the rotating SMBH is characterized by the value $1$ of the dimensionless spin parameter $\\chi$ (the spin $S$ can be calculated as $S= G c^{-1} m^2 \\chi$). However, as the accretion disk radiates and some of that radiation is trapped by the black hole, the extreme limit is actually not reached and rather the so-called canonical spin limit of $\\chi _{\\mathrm{can}}\\simeq 0.998$ applies \\citep{Thorne1974}. A closed magnetic field line topology further connects the black hole horizon to the accretion disk, field lines acting as anchor chains, which further reduces the spin to $\\chi \\simeq 0.89$ \\citep{Kovacs2011}. Nevertheless this value is still extremely high, and as the presence of the jet signals a rich black hole environment, supporting the assumption of accretion, it is plausible to assume that at least the jet emitter black hole spins fast, and as such, the spin effects in the binary dynamics are non-negligible. The spin-powered jets are believed to be aligned with the black hole spin axis \\citep{Blandford1977}, therefore precessing jets can reveal the presence of the spin. In this paper we investigate the source S5 1928+738. This is a core-dominated quasar at a redshift $z=0.302$, its luminosity distance is $D_{\\mathrm{L}}=1620$ Mpc, and at this distance the spatial resolution is $4.6$ pc/milliarcsecond \\citep[where the cosmological parameters are $\\Omega _{\\mathrm{m}}=0.314$, $\\Omega _{\\lambda}=0.686$, $H_{0}=67.4$ km s$^{-1}$ Mpc$^{-1}$,][]{Planck2013}. The source is a well studied member of a complete sample of the extragalactic flat spectrum radio sources taken from the S5 strong source survey \\citep{Kuhr1981}. The VLA map at a wavelength of $20$ cm \\citep{Johnston1987} reveals a structure comprising two lobes and a bright radio-core. Zooming into the jet base a one-sided core-jet structure appears on the milliarcsecond (mas) scale. The maximal apparent velocity detected in the jet, $8.1c$ \\citep[][]{Lister2013} suggests high-speed bulk motion that is nearly head on, the mean jet direction is pointing close to the observer line of sight (LOS). This feature makes the jet a strongly Doppler-boosted one. An analysis of the arcsecond- and the mas-scale jet structure has been published by \\citet{Hummel1992} and the jet structure has been found to exhibit wiggles on both scales. A moving sine wave was fitted to the temporal motion of the components in the jet. \\citet{Roos1993} proposed that the host galaxy of S5 1928+738 harbours a binary SMBH at its centre, inducing the periodic structures detected in the jet. A total mass of $10^{8}$ M$_{\\odot} $ was adopted, which would account for the bolometric luminosity on the basis of the Eddington limit. The orbital period was calculated to be $\\sim 2.9$ years, the orbital separation $\\sim 0.005$ pc and the mass ratio $>0.1$. It has also been noted that if the wiggles in the mas-scale jet of S5 1928+738 are due to the orbital motion of a massive binary, the mean direction of the jet should obey a Lense-Thirring precession \\citep{LenseThirring1918} about the orbital angular momentum of the binary, with a period of the order of $10^{3}$ yr. \\citet{Murphy2009} used the VLBI Space Observatory Programme data to further constrain the binary parameters by fitting their equations to jet component data from $8$ epochs of observation performed at 5 GHz. The derived parameters are however affected by their assumption of the jet being emitted by the fictitious reduced mass particle. The data collected on S5 1928+738 within the framework of the Monitoring Of Jets in Active galactic nuclei with VLBA Experiments (MOJAVE) survey span almost twenty years \\citep{Lister2009}. Careful analysis of this dataset allows us to detect and quantitatively analyse for the first time the spin-orbit precession of the jet emitter SMBH, manifesting itself as a small change over time in the direction of the jet axis. The spatial-length of the jet based on the MOJAVE-VLBI maps at $15$ GHz and assuming an inclination of $7\\degr$ is approximately $560$ pc. The paper is organised as follows. In Section 2 we summarize the observational techniques which can reveal periodicities in the extragalactic jets observed at radio wavelengths. More specifically we present methods based on the investigation of the AGN jet morphology, AGN jet kinematics and AGN jet flux density variability. In Section 3 we present the analysis of the archival MOJAVE data of S5 1928+738, investigating the temporal evolution of the jet structure. In Section 4, we describe a geometrical model for the three-dimensional (3D) helical jet structure. By projecting the spatial jet onto the plane of the sky and employing the maximal apparent velocity seen in the jet (which provides information about the inclination angle), we constrain the model by fitting it to the MOJAVE observations on the jet. Based on the jet modelling and on the independent total mass estimate, we give the binary parameters in Section 5. In Section 6 we discuss our findings and give our conclusions. ", "conclusions": "In this paper we investigated the radio jet of S5 1928+738 based on calibrated data of the MOJAVE survey with focus on the perturbed jet ejection. Our analysis basically confirmed the model of \\citet{Roos1993} which in turn was based on the jet analysis by \\citet{Hummel1992} of a hidden SMBH binary at the jet base, causing the observed wiggling of the jet. The improvement upon this model, advanced in the present paper is the inclusion of the spin of the jet-producing black hole. We developed a detailed geometrical model to describe the inner $2$ mas of the jet by using the $43$ GHz VLBA-map of \\citet{Lister2000}. From that we obtained the inclination and position angle of the jet axis and the intrinsic half-opening angle of the conical helix. Then we fitted the geometric model to the $15$ GHz VLBA data of almost twenty years. Three features of the jet have been revealed; (i) a conical helix shape, (ii) a periodical change in the direction of its symmetry axis, (iii) a slow additional reorientation in the average direction of the jet. Our spinning binary black hole model naturally explained the simultaneous presence of properties (ii) and (iii), and (i) was attributed to the presence of helical kink instabilities. Jet precession (iii) can be induced through the Bardeen-Petterson interaction between a viscous accretion disk and a spinning black hole \\citep{BardeenPetterson1975}, if the disk is misaligned compared to the equatorial plane of the black hole. % However such a scenario cannot explain both (ii) and (iii). In an alternative scenario, the influence of the immediate environment could cause the wiggling of the jet. However, to explain the observed periodicities would require properly fine-tuned structures. So far there is no evidence for such periodic distribution of dense material close to the jet. We adopted the total mass of the binary $m=8.13 \\times 10^{8}M_{\\odot }$ determined by \\citet{Woo2002} from the black hole mass-AGN continuum luminosity scaling relation (this value is different by a factor of $8.13$ from the one adopted by \\citealt{Roos1993}). The helical jet model with periodic jet axis and VLBI data implied the orbital period $T=4.78\\pm 0.14$ yr (this is a factor of $1.59$ larger than the value given by \\citealt{Roos1993}). With these values we calculated the binary separation as $r=0.0128\\pm 0.0003$ pc. These parameters imply that the SMBH binary is in the inspiral phase, but far from coalescence, with post-Newtonian parameter $\\varepsilon \\approx 0.003$. Long-term monitoring of the radio jet allowed us to identify the linear trend in the evolution of the inclination and position angles of the jet axis, interpreted as arising from the spin-orbit precession of the jet emitter SMBH. The mass ratio most likely falls into the range $\\nu \\in \\left[ 0.21:1/3 \\right] $. The spin-orbit precession period was identified as $T_{\\mathrm{SO}}=4852 \\pm 646$ yr and the gravitational lifetime emerged as $T_{\\mathrm{merger}}=(1.44 \\pm 0.19)\\times 10^6$ yr. Although we cannot rule out that other models could explain the observed jet structure, we showed that the VLBI data of S5 1928+738, extending over almost twenty years is consistent with the model of a spinning binary black hole lying at the jet base, where the larger black hole has a spin detectable through its spin-orbit precession. Measurements of slow increase in the average flux density of the jet in the newest three epochs (2013.34, 2013.58 and 2013.96) further support the model, as such an increase is predicted by spin-orbit precession. Our study thus provides indications, for the first time from VLBI jet kinematics, for the spinning nature of the jet-emitting black hole. As the \\-MOJAVE survey is still ongoing, further data on S5 1928+738 may better constrain the parameters of the model. With a significantly increased amount of data in principle it would be possible to monitor higher order post-Newtonian effects too. Beyond spin induced precession, such an analysis could also reveal the magnitude of the spin, unavailable at the accuracy of the present analysis." }, "1402/1402.2702_arXiv.txt": { "abstract": "We present a detailed comparison between the photometric properties of the bulges of two simulated galaxies and those of a uniform sample of observed galaxies. This analysis shows that the simulated galaxies have bulges with realistic surface brightnesses for their sizes and magnitude. These two field disc galaxies have rotational velocities $\\sim$ 100 km/s and were integrated to a redshift of zero in a fully cosmological $\\Lambda$ cold dark matter context as part of high-resolution smoothed particle hydrodynamic simulations. We performed bulge-disc decompositions of the galaxies using artificial observations, in order to conduct a fair comparison to observations. We also dynamically decomposed the galaxies and compared the star formation histories of the bulges to those of the entire galaxies. These star formation histories showed that the bulges were primarily formed before {\\em z} = 1 and during periods of rapid star formation. Both galaxies have large amounts of early star formation, which is likely related to the relatively high bulge-to-disc ratios also measured for them. Unlike almost all previous cosmological simulations, the realistically concentrated bulges of these galaxies do not lead to unphysically high rotational velocities, causing them to naturally lie along the observed Tully--Fisher relation. ", "introduction": "Computational astronomers have made great strides in reproducing the observed disc structure of spiral galaxies in $\\Lambda$ cold dark matter simulations, as studies of the Tully--Fisher relation \\citep{Robertson04, Governato07, Stinson10, Piontek11}, angular-momentum content \\citep{Scannapieco09}, and disc sizes \\citep{Brooks11} have shown. Historically, though, the corresponding bulges of simulated galaxies were much too large and concentrated compared to observed bulges \\citep{Bullock01, vandenBosch01b, Binney01, vdb02, donghia06, dn07,Stinson10, scannapieco11}. This difficulty in reproducing the bulges of disc galaxies represents the more-persistent aspect of the `angular-momentum catastrophe' \\citep{Navarro94}. In the past several years, only a handful of simulations have been able to combine the resolution necessary to resolve bulges in cosmological simulations with star formation and feedback models capable of producing galaxies with realistic mass bulges. These simulations have variously succeeded by limiting the star formation efficiency \\citep{Agertz10}, pre-heating the gas through early stellar feedback \\citep{Stinson13}, or increasing the strength of supernova-driven (SN-driven) outflows, either by scaling a kinetic wind model \\citep{Okamoto12, Vogelsberger13} or concentrating the energy through a high star formation threshold \\citep{Guedes11, Brook11a}. Similar strong outflows have been shown to reduce the amount of low-angular momentum material \\citep{Pontzen11}, resulting in bulgeless or cored dwarf galaxies \\citep{Oh11, Governato12,Teyssier13} and disc galaxies with realistic, rising rotation curves \\citep{Christensen12a, AnglesAlcazar13}. % Despite these improvements, no single method has been shown to reproduce the observed properties of bulges in a range of galaxies. As such, the masses and concentration of bulges remain as firm limits on the formation of galaxies. In particular, the observed scaling relations between bulge surface density, magnitude, and size encapsulate much of the observational constraints. Given the importance of reproducing bulge properties, it is critical to analyse them in an observationally motivated fashion. Here we undertake a careful photometric analysis of the bulges of two simulated galaxies and compare them to an observed sample analysed in an identical fashion. These galaxies were previously shown to have flat rotation curves and a reduced mass of low-angular momentum material as the result of feedback-driven outflows \\citep[][hereafter C12a]{Christensen12a}. In this Letter, we show that they also have realistically concentrated bulges. ", "conclusions": "\\label{sec:res5} In this Letter, we have taken an observationally motivated approach to analyzing the bulges of two simulated galaxies, previously shown to have rising rotation curves (C12a). We photometrically decomposed H-band images into bulge and disc components and compared the bulge properties to an observational sample of disc and elliptical galaxies. We determined that these simulations had bulges of the appropriate surface brightness for their magnitude and size. This success indicates that the centres of the simulated galaxies have appropriate stellar distributions and that they are not overly concentrated. Their relatively high bulge-to-total ratios, however, remain a concern and demonstrate a need for a further reduction in the central stellar mass. In particular, reducing the amount of early star formation could both lower the bulge mass and result in more realistic SFHs. We compared the SFHs of the kinematically selected bulges to those of the entire galaxies. We found that the bulges formed primarily during the first half of the galaxies' lifetimes and that bulge star formation occurred during periods of peak star formation, possibly driven by mergers. Of the two galaxies, the SFHs and photometric properties of one were most consistent with the properties of classical bulges, while the other shared characteristics with both classical and pseudo-bulges. Finally, we verified the global structure of the simulated galaxies by comparing them to the observed baryonic and stellar Tully--Fisher relations." }, "1402/1402.1910_arXiv.txt": { "abstract": "Given an approximately centered image of a spiral galaxy, we describe an entirely automated method that finds, centers, and sizes the galaxy and then automatically extracts structural information about the spiral arms. For each arm segment found, we list the pixels in that segment and perform a least-squares fit of a logarithmic spiral arc to the pixels in the segment. The algorithm takes about 1 minute per galaxy, and can easily be scaled using parallelism. We have run it on all $\\sim$644,000 Sloan objects classified as ``galaxy'' and large enough to observe some structure. Our algorithm is stable in the sense that the statistics across a large sample of galaxies vary smoothly based on algorithmic parameters, although results for individual galaxies can sometimes vary in a non-smooth but easily understood manner. We find a very good correlation between our quantitative description of spiral structure and the qualitative description provided by humans via Galaxy Zoo. In addition, we find that pitch angle often varies significantly segment-to-segment in a single spiral galaxy, making it difficult to define ``the'' pitch angle for a single galaxy. Finally, we point out how complex arm structure (even of long arms) can lead to ambiguity in defining what an ``arm'' is, leading us to prefer the term ``arm segments''. ", "introduction": "\\label{sec:Intro} \\vspace{-3mm} The Hubble Ultra Deep Field represents about 1/13,000,000 of the celestial sphere and contains about 10,000 galaxies, suggesting the entire sky contains upwards of $10^{11}$ galaxies. Gaining quantitative structural information for this number of galaxies will require automated methods. For spiral galaxies, existing methods for visual classification \\cite{Hubble1936,DeVaucouleurs1959} are either subjective or non-quantitative, while currently available semi-automated methods \\cite{DeSouza2004,Simard1998,Peng2002a,Peng2010,BenDavis2012,Ma2001,Au2006,Ripley1990,Shamir2011}. are either too simplistic or require significant human input. \\begin{figure*}[t] \\plotsix{2-autoCrop.eps}{3-contrastEnhanced.eps}{4-orientationField_32x32-from-64x64.eps}{5-clustersAndArcs_arcEnhanced.eps}{6-mergedClustersAndArcs_arcEnhanced.eps}{7-arcOverlayOnAutoCrop_arcEnhanced.eps} $\\begin{array}{cccccc} \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\, ({\\bf a}) \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; &\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\, ({\\bf b}) \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; &\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; ({\\bf c}) \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; &\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; ({\\bf d}) \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; &\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; ({\\bf e}) \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; &\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; ({\\bf f}) \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; \\end{array}$ \\caption{Steps in describing a spiral galaxy image. {\\bf a)} The centered and de-projected image. {\\bf b)} Contrast-enhanced image. {\\bf c)} Orientation field (at reduced resolution for pedagogical reasons). {\\bf d)} Initial arm segments found via Hierarchical Agglomerative Clustering (HAC) of nearby pixels with similar orientations and consistent logarithmic spiral shape, and the associated logarithmic spiral arcs fitted to these clusters. {\\bf e)} Final pixel clusters (and associated arcs) found by merging compatible arcs. {\\bf f)} Final arcs superimposed on image (a). Red arcs wind S-wise, cyan Z-wise. } \\label{fig:parse} \\end{figure*} ", "conclusions": "\\vspace{-3mm} We have run our code on every item in the Sloan Digital Sky Survey (SDSS) that is classified as ``galaxy''. Unfortunately, SDSS does not distinguish between spiral and non-spiral galaxies. We are currently working on a machine learning algorithm that uses the output of our code to distinguish between spiral and non-spiral galaxies. Preliminary results are encouraging (agreeing approximately 90--95\\% with Galaxy Zoo humans), and will be presented in a future paper. Once we can separate out spiral galaxies, further studies will be performed concerning how spiral structure correlates with other variables such as color, redshift, local environment, etc. Code and data are available from WBH upon request. \\vspace*{-3mm}" }, "1402/1402.3915_arXiv.txt": { "abstract": "In observations of flare-heated electrons in the solar corona, a longstanding problem is the unexplained prolonged lifetime of the electrons compared to their transit time across the source. This suggests confinement. Recent particle-in-cell (PIC) simulations, which explored the transport of pre-accelerated hot electrons through ambient cold plasma, showed that the formation of a highly localized electrostatic potential drop, in the form of a double layer (DL), significantly inhibited the transport of hot electrons (T.C. Li, J.F. Drake, and M. Swisdak, 2012, ApJ, 757, 20). The effectiveness of confinement by a DL is linked to the strength of the DL as defined by its potential drop. In this work, we investigate the scaling of the DL strength with the hot electron temperature by PIC simulations, and find a linear scaling. We demonstrate that the strength is limited by the formation of parallel shocks. Based on this, we analytically determine the maximum DL strength, and find also a linear scaling with the hot electron temperature. The DL strength obtained from the analytic calculation is comparable to that from the simulations. At the maximum strength, the DL is capable of confining a significant fraction of hot electrons in the source. ", "introduction": "INTRODUCTION} Coronal electrons can be accelerated to very high energies, over 100 keV \\citep{Lin03, Krucker10, Krucker07,Tomczak09}, during flares and coronal mass ejections. Acceleration is believed to occur high in the corona \\citep{Fletcher11}. Hard and soft X-rays emitted by the accelerated electrons are observed in both the corona and chromosphere. Emission from the chromosphere is usually brighter because of the higher density, which allows for more efficient interaction with the X-ray-producing energetic electrons. Improved detection sensitivity, however, has led to more measurements of X-ray sources in the corona. These sources are situated at or distinctly above the top of the soft X-ray flaring loops. A comprehensive analysis suggests that above-the-looptop sources are a common feature of all flares \\citep{Petrosian02}. The transport of energetic electrons from the corona to the chromosphere is crucial to understanding energy release in flares. Transport effects can modify the energy distribution of the propagating electrons and hence the observed X-ray spectra, affecting the interpretation of acceleration models. There is evidence for both free-propagation and confinement of energetic electrons. Evidence for free-propagation of energetic electrons, i.e., no interaction with the ambient plasma, was reported in previous time-of-flight measurements of hard X-ray emission \\citep{Asch95,Aschwanden96,AschSchwartz95}. A systematic time delay between lower-energy hard X-rays with respect to the higher energies was measured from a large sample (over 600) of solar flares. The time-of-flight delays indicate lower-energy electrons arrive at the chromosphere after those with higher energy as they freely stream down the flare loop from the corona. On the other hand, observations of above-the-looptop sources reveal that the lifetime of the energetic electrons is two orders of magnitude longer than their free-streaming transit time across the source \\citep{Masuda94, Krucker10, Krucker07}. This requires confinement of the electrons in the source region. Recently, a systematic study of solar flares with both looptop and footpoint emissions found that the number of electrons required to explain observations is 2-8 times higher at the looptop than at the footpoint \\citep{Simoes13}. This suggests electron accumulation at the looptop. Another systematic study reported that the difference in spectral index between non-thermal coronal and footpoint emission in some events is considerably greater than 2 - which is expected for the simple case that the same electron population, which produces looptop emission through thin target bremsstrahlung because of the lower density in the corona, free-streams to the footpoints to emit through thick target bremsstrahlung due to the higher density in the chromosphere \\citep{Battaglia06}. This requires a filter effect in the propagation preferentially reducing the distribution at lower energies. Such a filter can be an electric field. An important question on electron transport is why some electrons appear to freely propagate while some are filtered or even trapped. The physics of this subject, however, remains poorly understood. The transport of electron heat flux has previously been modeled as anomalous conduction \\citep{Manheimer77, Tsytovich71} in which anomalous resistivity arises from electron scattering by turbulent wave fields. Turbulence is excited by instabilities that result from the interaction between energetic electrons and ambient plasma. In a similar model, it was suggested that ion-acoustic turbulence driven by a return current in response to hot electron streaming and an induced polarization electric field would drive a potential that would suppress electron transport \\citep{Levin93}. An anomalous conduction front that moved along the flare loop at the head of an expanding hot electron source was considered as a means to confine hot electrons for the production of hard X-rays \\citep{Smith79}. The model, based on a one-dimensional (1D) one-fluid code, however, did not resolve ion inertial lengths, let alone capture processes occurring at electron scales. Later, 1D electrostatic particle-in-cell (PIC) simulation that resolved the shortest electron scale, the Debye length, did not reveal a conduction front \\citep{McKean90}. Recently, the existence of a thermal front, literally defined as a region that links plasmas in thermal nonequilibrium and sustains the temperature difference for longer than the electron free-streaming time, was studied in 1D electrostatic Vlasov simulations \\citep{Arber09}. The formation of a temperature difference propagating at a speed comparable to the ion acoustic speed was observed and the behavior was identified as a conduction front. The physics of the responsible mechanism was not identified or investigated. We suggest, on the basis of the simulations and analysis in this paper, that the propagating front is an ion acoustic shock and the temperature difference to be a result of shock heating due to its extremely sharp transition. Other observed non-propagating temperature jumps that were associated with a potential jump were likely DLs, as mentioned by the authors. More recently, the transport of coronal energetic electrons was studied by electromagnetic 2D PIC simulations \\citep{Li12} (hereafter as LDS). It was shown that transport suppression began as the electrons propagated away from the acceleration site. The suppression was caused by the formation of a DL and associated potential barrier that reflected electrons back to the source region. The DL reduced the electron heat flux by nearly 50\\%. A DL is a localized region that sustains a potential drop in collisionless plasmas \\citep{Block78,Raadu88}. The potential drop comes from a large-amplitude electrostatic electric field sandwiched between two adjacent layers of equal and opposite charges. The structure is globally neutral, but quasi-neutrality is locally violated within it, which occurs at scales of $\\sim$10 Debye lengths $\\lambda_{De}$. An ideal DL is a monopolar electric field, but more generally, a DL can be bipolar. Therefore, instead of a monotonic drop in the potential $\\phi$, there can be dips or bumps at the low or high potential sides, but an overall potential drop $\\phi_{DL}$ across the structure. $\\phi_{DL}$ is the measure of the strength of a DL. A DL reflects particles if $\\phi_{DL}$ is greater than their kinetic energy. In LDS, it is the reflection of energetic electrons by a DL that reduces their transport from the source. Numerical simulations with a kinetic model have been used to study DL dynamics since they can both resolve kinetic scales down to $\\lambda_{De}$, where DLs occur, and probe nonlinear phenomena. Methods used to generate DLs include a current and an applied potential. In LDS and this work, the DL forms as a result of strong currents driven by an imposed field-aligned temperature jump in the initial state. Earlier 1D PIC simulations of systems driven by a subthermal electron current showed that ion trapping in the potential dip of a (bipolar) DL was associated with decay of a DL \\citep{Barnes85}. The trapping of ions slowed down the movement of the DL structure as the trapped ions added an inertial drag to it \\citep{Chanteur83}. It was argued that such slowing down led to the decay of the DL \\citep{Barnes85}. On the other hand, 1D PIC simulations with an applied potential reported that a DL decayed as a result of the propagation of solitons, which developed from the DL structure, across the DL width \\citep{Sato81}. Only very limited work exists on the saturation of DLs. The degree of transport suppression depends on the maximum strength of the DL at saturation. In LDS, the generation mechanism of a DL was studied and identified as the Buneman instability involving the ions and return current electrons. LDS showed that the potential increased with increasing hot electron temperature, but the functional dependence on this temperature was not studied. In this work, we investigate the saturation mechanism of the DL with simulations and analytic modeling. We carry out a series of simulations with different initial values of the hot electron temperature and show that the DL potential at saturation scales linearly with the hot electron temperature. The DL saturates when its potential jump is large enough to accelerate ions above the local sound speed. The result is a parallel ion acoustic shock that stabilizes the Buneman instability and therefore saturates the DL. We demonstrate that the shock formation criterion predicts a maximum DL strength that is proportional to the hot electron temperature in agreement with simulations. At its maximum strength, the DL is observed to reflect, and hence contain, a significant fraction of electrons in the source. In addition, we observe that an anisotropy with $T_\\perp > T_\\parallel$ develops in the hot electrons that pass through a DL. Such an anisotropic (pancake) distribution favors further hot electron trapping in the presence of a magnetic mirror. Since the magnetic geometry of flaring loops resembles a mirror configuration, we explore combining a magnetic mirror with the DL. We demonstrate that the combination significantly enhances confinement. An outline of the work is as follows: the setup of the simulations and parameters used for flare settings are described in Section \\ref{sim}; in Section \\ref{result}, results from the simulations and the shock model are presented, and we provide evidence for the model and determine the saturation amplitude of a DL; in Section \\ref{mu_anis}, we combine the DL with a magnetic mirror and show that further suppression of transport is obtained; further discussion of our results is given in Section \\ref{dis}; and we summarize in Section \\ref{con}. ", "conclusions": "" }, "1402/1402.1584_arXiv.txt": { "abstract": "{Gravity modifies the spectral features of young brown dwarfs (BDs). A proper characterization of these objects is crucial for the identification of the least massive, and latest-type objects in star-forming regions, and to explain the origin(s) of the peculiar spectro-photometric properties of young directly imaged extrasolar planets and BD companions.} {We obtained medium-resolution (R$\\sim$1500-1700) near-infrared (1.1-2.5 $\\mu$m) spectra of seven young M9.5-L3 dwarfs classified at optical wavelengths. We aim to empirically confirm the low surface gravity of the objects in the near-infrared. We also test whether self-consistent atmospheric models correctly represent the formation and the settling of dust clouds in the atmosphere of young late-M and L dwarfs.} {We used ISAAC (Infrared Spectrometer And Array Camera) at VLT (Very Large Telescope) to obtain the spectra of the targets. We compared them to those of mature and young BDs, and young late-type companions to nearby stars with known ages, in order to identify and study gravity-sensitive features. We computed spectral indices weakly sensitive to the surface gravity to derive near-infrared spectral types. Finally, we found the best fit between each spectrum and synthetic spectra from the BT-Settl 2010 and 2013 atmospheric models. Using the best fit, we derived the atmospheric parameters of the objects and identify which spectral characteristics the models do not reproduce.} {We confirmed that our objects are young BDs and we found near-infrared spectral types in agreement with the ones determined at optical wavelengths. The spectrum of the $\\mathrm{L2{\\gamma}}$ dwarf \\object{2MASSJ232252.99-615127.5} reproduces well the spectrum of the planetary mass companion \\object{1RXS J160929.1-210524b}. BT-Settl models fit the spectra and the 1-5 $\\mu$m spectral energy distribution of the L0-L3 dwarfs for temperatures between 1600-2000 K. But the models fail to reproduce the shape of the H~band, and the near-infrared slope of some of our targets. This fact, and the best fit solutions found with super-solar metallicity are indicative of a lack of dust, in particular at high altitude, in the cloud models.} {The modeling of the vertical mixing and of the grain growth will be revised in the next version of the BT-Settl models. These revisions may suppress the remaining non-reproducibilities. Our spectra provide additional templates for the characterization of the numerous young L-type companions that will be detected in the coming years by planet imaging instruments such as VLT/SPHERE, Gemini/GPI, Subaru/SCexAO, and LBTI/LMIRCam.} ", "introduction": "Since the discovery of the first substellar objects (\\citealt{1995Natur.378..463N,Rebolo1995}), large infrared (IR) surveys have unearthed hundreds of brown dwarfs (BDs) in the field and in star forming regions. The spectral energy distributions of BDs peak in the infrared. Their spectra are dominated by broad, overlapping condensate and molecular absorption features \\citep[see][]{Kirkpatrick}. The strength of these features depends on a combination of photospheric temperature, gas, pressure and dust properties, which in turn are related to the effective temperature ($\\mathrm{T_{eff}}$), surface gravity ($g$), and metallicity of the objects. Spectroscopic studies of mature ($\\gg 400$ Myr) field BDs led to the definition of the spectral classes `L' \\citep{1999ApJ...519..802K, 1999AJ....118.2466M} for objects dominated by H$_{2}$O, FeH, and CO absorption bands in the near-infrared (NIR), and `T' \\citep{2002ApJ...564..421B, 2002ApJ...564..466G} for objects which exhibit strong CH$_{4}$+H$_{2}$O bands and collision-induced absorption (CIA) due to H$_{2}$ at these wavelengths. The change in the spectral morphology, used for the classification, is mostly tied to changes in $\\mathrm{T_{eff}}$ \\citep[e.g.,][]{2009A&A...503..639T}. Substellar objects contract and cool down with time \\citep{1997ApJ...491..856B, BCAH98}. Therefore, young objects can have larger radii and higher luminosities than older and more massive objects despite an identical effective temperature. The lower surface gravity of young objects ($g \\varpropto M/R^{2}$) can be directly accessed by observation and be used to break the degeneracy. Low surface gravity results in peculiar spectral characteristics such as the triangular H band shape in the near-infrared \\citep{Zapatero-Osorio2000, Lucas}, and reduced alkali lines in the optical and near-infrared \\citep[e.g.,][]{McGovern, Cruz}. These modifications have nonetheless mostly been investigated for objects earlier than $\\sim$L4 found in young clusters and star forming regions \\citep[e.g.,][]{Luhman1997, 1999ApJ...525..466L, Lucas, 2004ApJ...617..565L, 2004AJ....127..449M, Luhman2004, McGovern, Lodieu, Weights, 2012A&A...539A.151A}. Advances in high contrast and high resolution imaging in the near-infrared (1-5 $\\mu$m; NIR) led to the discovery of late-type companions to young nearby stars straddling the planet/BD boundary \\citep[e.g.,][]{2004A&A...425L..29C, 2005A&A...438L..29C, 2008ApJ...689L.153L, 2008Sci...322.1348M, 2010Natur.468.1080M, 2010Sci...329...57L, 2011ApJ...726..113I, 2013A&A...553L...5D, 2013ApJ...772L..15R, 2013ApJ...774...11K, 2013arXiv1310.4825C}. Most of these objects have estimated effective temperatures similar to those of L dwarfs. The low to medium-resolution (R$\\leq$5000) spectra and infrared photometry of these objects exhibit peculiar features \\citep[red pseudo-continua, triangular H~band shape, lack of methane absorption, reduced K I and Na I lines; ][]{2008ApJ...689L.153L, Bonnefoy2010, Patience, 2011ApJ...729..139W, Barman, Bonnefoy2013, 2013ApJ...768...24O}. These peculiarities are likely directly related to the expected low surface gravity of the objects. Nevertheless, the currently limited sample of young late-type objects with high S/N (signal-to-noise) spectra make the establishment of a proper empirical classification scheme challenging for these objects. \\citet{Cruz} have identified a population of young bright and nearby L dwarfs isolated in the field. They developed a classification scheme in the optical. This scheme has been extended in the NIR for young L dwarfs by \\citet{Bonnefoy2013} and \\citet{Allers2013}. Several of these peculiar L-type dwarfs share similarities with the spectra of young companions at these wavelengths \\citep{Bonnefoy2010, Bonnefoy2013, Allers2013, 2013AJ....145....2F, 2013arXiv1310.0457L}. Getting more NIR spectra of young isolated objects are therefore needed to consolidate the classification scheme and to identify further analogues to directly imaged exoplanets. Atmospheric models allow us to disentangle the effect of varying $T_{eff}$, log~\\textit{g}, and (metallicity) \\textit{M/H} on the spectral features. Below $T_{eff}$ $\\sim$~2600~K, models predict that clouds of iron and silicate grains begin to form, changing the opacity (\\citealt{Lunine}, \\citealt{Tsuji}, \\citealt{Burrows_Sharp}, \\citealt{Lodders}, \\citealt{Marley2000}, \\citealt{Marley2001}, \\citealt{Allard2001}). The formation and the gravitational sedimentation of these dust clouds are influenced by the surface gravity. Dust cloud formation is expected to be more efficient at low gravity because the atmosphere is more extended, and the gas cooler. Low gravity tends to make the convection and the resulting mixing more efficient as well. Modeling of the spectro-photometric data on young L and early-T type companions with parametrized models (\\citealt{Marley_Ackerman}, \\citealt{Burrows}) has revealed anomalously thick clouds (\\citealt{Barman2011a}, \\citealt{Barman2011b}, \\citealt{Skemer}, \\citealt{Marley2013}). These peculiar cloud properties may explain why some companions are \"underluminous\" in some bands (\\citealt{Skemer}, \\citealt{Marley2012}). Self-consistent atmospheric models, such as the BT-Settl models \\citep{Allard} and the Drift-PHOENIX models, \\citep{Helling} use cloud models where the dust properties do not require defining any additional free parameters other than log \\textit{g}, $T_{eff}$, \\textit{M/H}. Synthetic spectra for a specific set of atmospheric parameters can be compared to empirical spectra. These models are just starting to be tested on spectra of young late-type objects \\citep[companions and free-floating objects;][]{Bonnefoy2010, 2011A&A...529A..44W, 2012A&A...540A..85P, Bonnefoy2013}. We used them on spectra of young M5.5-L0 dwarfs in \\citet{Bonnefoy2013} to reveal a drop of the effective temperature at the M/L transition. We suggested that this drop could be induced by an improper handling of the formation of dust clouds at the M/L transition in the models. The test however could not be extended to later spectral types and lower effective temperatures due to the lack of a consistent sample of young objects in the L dwarf regime at that time. In this paper, we present a homogeneous set of seven medium-resolution (R$\\sim$1500-1700) spectra of M9.5-L3 dwarfs, all classified at optical wavelengths. Our sample is composed of the M9.5 object \\object{DENIS-P J124514.1-442907} (also called TWA 29; hereafter DENIS J1245) a member of TW-Hydrae (5-10~Myr), and the L0 dwarf \\object{Cha J1305-7739} \\citep[][hereafter Cha 1305]{Jayawardhana}, one of the least massive objects of the Chameleon II cluster. We also present the spectra of five L dwarfs with features indicative of low surface gravity (L$\\gamma$ dwarfs) in the optical, identified by \\cite{Cruz}. These objects are the two L0$\\gamma$ dwarfs \\object{EROS~J0032-4405} \\citep[][hereafter EROS~J0032]{Goldman} and \\object{2MASS J22134491-2136079} \\citep[][hereafter 2M2213]{Cruz}, the L2$\\gamma$ dwarf \\object{2MASSJ232252.99-615127.5} \\citep[][hereafter 2M2322]{Cruz}, and the two L3$\\gamma$ dwarfs \\object{2MASS J212650.40-814029.3} \\citep[][2M2126]{Cruz} and \\object{2MASSJ220813.63+292121.5} \\citep[][hereafter 2M2208]{Cruz}. We aim to use the spectra to confirm the low surface gravities of the objects in the near-infrared and to test the ability of the BT-Settl models to correctly handle the formation and gravitational settling of dust under reduced surface gravity conditions. These spectra enrich the scarce sample of empirical near-infrared medium-resolution spectra of young late-type objects beyond the M-L transition, especially for spectral type L3. We describe in Section \\ref{observations} our observations and the associated data reduction. We present in Section \\ref{empirical_analysis} an empirical analysis of the spectral features in order to derive near-infrared spectral types and confirm the young age of our targets. In Section \\ref{Spectralsynthesis} we describe the comparison of the atmospheric models to the observed spectra. We discuss these comparisons and derive updated target properties in Section \\ref{discussion}. ", "conclusions": "We obtained and analyzed seven VLT/ISAAC medium-resolution (R$\\sim$1500-1700) spectra of M9.5-L3 dwarfs classified at optical wavelengths and showing indications of low surface gravity. We built an age-sequence of M9.5 objects that allow us to pinpoint age-sensitive and gravity-sensitive features at medium-resolving powers. The comparison of our spectra to those of young reference brown dwarfs and companions, and of mature field dwarfs confirm that our objects have peculiar features in the near-infrared indicative of low surface gravities and young ages. We also confirm the youth of our objects by calculating the equivalent widths of their KI lines and comparing these values per spectral type with the values obtained for young reference brown dwarfs and companions and mature field dwarfs. We derived near-infrared spectral types based on dedicated spectral indices. These spectral types are in agreement with the optical classification, and confirm the coherence of the classification method. The analysis revealed that the L2${\\gamma}$ object \\object{2MASS~J2322} provides a good match to the spectrum of the young planetary mass companion \\object{1RXS J160929.1-210524b}. The spectra and SEDs of the objects can be reproduced by the 2010 and 2013 BT-Settl atmospheric models. The 2013 release of the models fits simultaneously the spectra and the SED for the same temperatures at all wavelengths. L0-L3$\\gamma$ dwarfs have nearly equal temperatures around 1800 K. Nevertheless, we identify that: \\begin{itemize} \\item the 2010 models do not reproduce the 1.1-2.5 $\\mu$m spectral slope of some L2-L3 objects. \\item the H~band shape is not well reproduced by the BT-Settl 2013 models at solar metallicity. The problem disappears when new, but not as well-tested, models at super-solar metallicity are used, but these models remain mostly untested. \\end{itemize} Currently, all these discrepancies point out a lack of dust in the cloud models. The next version of the BT-Settl models will modify the treatment of the vertical mixing and of grain growth processes. These new models are expected to produce thicker clouds, and may solve the issues revealed by the ISAAC spectra. The spectra of the objects will help to confirm the membership of photometrically-selected candidates in star-forming regions. Within the next few years, surveys on the next generation of planet imaging instruments such as SPHERE (Spectro-Polarimetric High-contrast Exoplanet REsearch) at VLT, GPI (Gemini Planet Imager) at Gemini South, ScEXAO (Subaru Coronagraphic Extreme AO Project) at Subaru, and LMIRCam (Large Binocular Telescope mid-infrared camera) at LBT should provide a sample of a few dozen young companions. Several planets similar to $\\beta$ Pictoris b should be unearthed and fall in the same temperature range as our objects. Therefore, our spectra will serve as precious benchmarks for the characterization of the physical and atmospheric properties of these companions." }, "1402/1402.6192_arXiv.txt": { "abstract": "The gamma-ray-detected blazar 3C~454.3 exhibits dramatic flux and polarization variations in the optical and near-infrared bands. In December 2010, the object emitted a very bright outburst. We monitored it for approximately four years (including the 2010 outburst) by optical and near-infrared photopolarimetry. During the 2010 outburst, the object emitted two rapid, redder brightenings, at which the polarization degrees (PDs) in both bands increased significantly and the bands exhibited a frequency-dependent polarization. The observed frequency-dependent polarization leads us to propose that the polarization vector is composed of two vectors. Therefore, we separate the observed polarization vectors into short and long-term components that we attribute to the emissions of the rapid brightenings and the outburst that varied the timescale of days and months, respectively. The estimated PD of the short-term component is greater than the maximum observed PD and is close to the theoretical maximum PD. We constrain the bulk Lorentz factors and inclination angles between the jet axis and the line of sight from the estimated PDs. In this case, the inclination angle of the emitting region of short-term component from the first rapid brightening should be equal to 90$^{\\circ}$, because the estimated PD of the short-term component was approximately equal to the theoretical maximum PD. Thus, the Doppler factor at the emitting region of the first rapid brightening should be equal to the bulk Lorentz factor. ", "introduction": "Blazars are a type of active galactic nuclei with relativistic jets that are widely believed to be viewed at small angles to the line of sight. Blazars frequently show violent variations in flux and polarization, which vary on timescales ranging from minutes \\citep{Aharonian07,Sasada08} to years \\citep{Sillanpaa96}. On the timescale of months, blazars may turn out to be over 10 times brighter than in their quiescent state. These brightening phenomena are often called ``outburst''. The blazar 3C~454.3 with a redshift $z=0.859$ \\citep{Jackson91} is one of the most famous blazar because it has emitted several large-amplitude outbursts. In 2005, it emitted a dramatic outburst that covered the range from radio to hard X-ray bands \\citep{Fuhrmann06,Pian06,Giommi06,Villata07}. The maximum brightness in the optical band reached $R=12.0$. Many authors reported that prominent outbursts occurred from 3C~454.3 in 2007, 2008 and 2009 and covered the range from radio to gamma-ray bands. Gamma-ray emission from the 2007 outburst was detected by the AGILE satellite \\citep{Vercellone08,Donnarumma09,Vercellone10}, and brightenings in other wavelengths were also detected \\citep{Villata08,Raiteri08a,Raiteri08b,Villata09a,Sasada10}. \\citet{Ghisellini07} and \\citet{Sikora08} proposed two possible explanations for the origin of seed photons for inverse Compton scattering emission in the GeV gamma-ray band. A large-amplitude gamma-ray brightening in the 2008 outburst was detected by the {\\it Fermi Gamma-ray Space Telescope} \\citep{Bonning09a,Abdo09}. Based on intensive monitoring, it was determined that the flux variations in the optical bands lagged no more than one day behind those of the GeV gamma-ray band \\citep{Vercellone09a,Donnarumma09,Bonning09a,Vercellone10}. In 2009, an outburst from 3C~454.3 occurred over all wavelengths: in the gamma-ray band it was detected by {\\it Fermi} and AGILE \\citep{Striani09a,Striani09b,Escande09,Striani10,Pacciani10,Ackermann10}, in the X-ray band it was detected by {\\it INTEGRAL} \\citep{Vercellone09b}, {\\it Swift}/XRT \\citep{Sakamoto09}, and {\\it Swift}/BAT \\citep{Krimm09}, in the optical bands it was detected by many groups \\citep{Villata09b,Bonning09b,Sasada09,Raiteri11,Sasada12}, and in the radio bands it was detected by \\citet{Raiteri11} and \\citet{Jorstad13}. In September 2010, a prominent outburst from 3C~454.3 was detected in the GeV gamma-ray band by {\\it Fermi} and AGILE \\citep{Abdo11,Vercellone11}. The peak flux reached 85$\\times$10$^{-6}$ photons~cm$^{-2}$~s$^{-1}$ in December 2010. The optical continuum and emission-line fluxes increased simultaneously in this outburst \\citep{Vercellone11,Leon-Tavares13}, and the X-ray and radio fluxes also increased \\citep{Wehrle12,Jorstad13}. In the present study, we report the results of using the Kanata telescope to monitor emission from 3C~454.3 using optical and near-infrared (NIR) photopolarimetry from July 2007 to January 2011: a period that includes the 2010 outburst. In December 2010, the polarization degree (PD) increased in the rapid brightenings of the 2010 outburst and the polarization difference between the optical and NIR bands became frequency dependent (this phenomenon is called ``frequency-dependent polarization'' or FDP). The temporal variations of polarization angle (PA) indicate that no rotation event occurred during this outburst. The paper is organized as follows: In section~2, we present the methods used for observation and analysis. In section~3, we report the results of the photometric and polarimetry observation. In section 4, we discuss the origin of the observed FDP and emission regions of the observed rapid brightenings and outburst. The conclusion is given in section 5. ", "conclusions": "We monitored blazar 3C~454.3 in the optical and near-infrared bands for approximately four years starting from July 2007, and we detected two types of variations in 2010: a large-amplitude outburst occurring on a timescale of $\\sim$months and extraordinary rapid brightenings occurring on a timescale of $\\sim$days. These brightenings had three features: (1) the $V-J$ bands reddened, (2) the degree of polarization (PD) increased in the $V$ and $J$ bands, and (3) the polarization was frequency dependent (FDP). Based on these results, we suggest that the observed polarization vectors can be decomposed into two components, namely long- and short-term components. We separate the short-term component, estimating the long-term component for the rapid brightenings from the pre- and post-brightening data. The estimated PD of the short-term component for the rapid brightenings was higher than the observed maximum PD and was close to the theoretical maximum PD. This result indicates that the short-term emitting regions responsible for the rapid brightenings are different from the long-term emitting region that radiated the outburst. The blazar FDP gives us important constraints on $\\Gamma$, $\\Phi$, and $\\delta$. \\\\ \\\\ This work was supported by a Grant-in-Aid for JSPS Fellows." }, "1402/1402.1856_arXiv.txt": { "abstract": "We show that the local type non-Gaussianity in a class of curvaton models is suppressed, i.e. the non-linearity parameters $f_{\\rm NL}$ and those related with higher order statistics can be at most $O(1)$, even if the curvaton energy density is subdominant at the decay. This situation is naturally realized in a very simple curvaton potential with quadratic term plus quartic term. ", "introduction": " ", "conclusions": "" }, "1402/1402.3581_arXiv.txt": { "abstract": "Mergers of compact stellar remnants are prime targets for the LIGO/Virgo gravitational wave detectors. The gravitational wave signals from these merger events can be used to study the mass and spin distribution of stellar remnants, and provide information about black hole horizons and the material properties of neutron stars. However, it has been suggested that degeneracies in the way that the star's mass and spin are imprinted in the waveforms may make it impossible to distinguish between black holes and neutron stars. Here we show that the precession of the orbital plane due to spin-orbit coupling breaks the mass-spin degeneracy, and allows us to distinguish between standard neutron stars and alternative possibilities, such as black holes or exotic neutron stars with large masses and spins. ", "introduction": "Compact stellar remnant mergers are the main targets of gravitational wave (GW) detectors such as advanced LIGO (aLIGO)~\\citep{Harry:2010zz} and advanced Virgo (Adv)~\\citep{Acernese:2007zze}, with predicted rates between a few and a few hundred per year at full design sensitivity~\\citep{Abadie:2010cf}. These systems take tens of minutes to sweep through the sensitive band of the detectors, entering the band at $\\sim10$Hz, and terminating in the kHz range with a violent merger lasting just a few milliseconds. The final stages of the inspiral and merger proceed differently for black holes (BHs) and neutron stars (NSs), and in principle, this should allow us to identify the make-up of the system from the GW signal alone. However, the number of GW cycles in the signal and the aLIGO/AdV sensitivity fall off rapidly with increasing frequency, meaning that there is very little information past $\\sim 500$ Hz (less than 2\\% of the SNR). Probes of BH physics and the equation of state of NSs will likely require multiple detections~\\citep{DelPozzo:2011pg, DelPozzo:2013ala}. An electromagnetic counterpart to the GW signal, such as a short-hard gamma-ray burst or an associated kilonova/macronova emission~\\citep{Metzger:2011bv}, would indicate that at least one of the bodies was a NS, but beaming effects or the luminosity of the signal may make detecting a counterpart difficult for the majority of mergers~\\citep{Abadie:2010cf,Aasi:2013wya}. Absent a counterpart, we must rely on the early inspiral to extract information about the make-up of the binary, which poses a challenge since finite size effects are completely negligible during inspiral~\\citep{Read:2009yp}. All we have to go on to decide the composition of the binary are the values of the masses and spins inferred from the inspiral signal. General arguments based on stability and causality limit the mass and spin of NSs to the range $M\\in[0.1, 3.2 ]M_\\odot$ for the mass and $\\chi \\in[0,0.7]$ for the dimensionless spin magnitude,~$\\chi \\equiv |\\vec{S}|/M^{2}$, where $\\vec{S}$ is the spin angular momentum~\\citep{Rhoades:1974fn,Lattimer:2006xb,Yagi:2014bxa}. Realistic equations of state yield a tighter mass range $M\\in[1.0, 2.5] M_\\odot$. The observed range of masses and spins is somewhat tighter~\\citep{Lattimer:2006xb,Ozel:2012ax}: $M\\in[1.0, 2.0]M_\\odot$, $\\chi \\in[0,0.3]$. The old NSs that merge are expected to have spun down by magnetic breaking to the point where the maximum spin is much lower, $\\chi \\lesssim 0.05$, than in the general NS population~\\citep{Mandel:2009nx}. Furthermore, the standard isolated NSNS binary formation scenario ensures that after every common envelope phase (that tends to align the spins) follows a supernovae kick that misaligns the spins (unless the kick is in the orbital plane, though there is evidence that this is not the case~\\citep{Kaplan:2008qm}). Thus, we adopt the definition that {\\em normal} NSs seen by aLIGO/AdV have $M\\in[1, 2.5]M_\\odot$ and $\\chi \\leq 0.05$, and term NSs with larger masses or spins {\\em exotic}. Einstein's theory of gravity allows BHs to have spin in the range $\\chi \\in[0,1]$ with any mass. X-ray observations have identified stellar remnant BHs with $M\\in [3.6, 36]M_\\odot$ and $\\chi \\in[0,1]$. There is currently some debate as to the existence of a mass gap between NSs and BHs~\\citep{Ozel:2010su,Farr:2010tu,Belczynski:2011bn}, but for the purpose of determining whether a normal NS could be misidentified as a BH or an exotic NS, the existence of a gap is moot. The early inspiral phase of a compact binary merger can be modeled analytically by expanding Einstein's equations in powers of the ratio of the orbital velocity to the speed of light, the so-called \\emph{post-Newtonian} (PN) approximation~\\citep{Blanchet:2013haa}. This ratio is small during the inspiral, with $v/c$ of $7\\%$ when the system enters the detector sensitivity band, reaching roughly $40\\%-60\\%$ by contact~\\citep{Bernuzzi:2014kca}. The PN approximation becomes less accurate as the system evolves through the band, eventually breaking down at the end of the inspiral phase. As all forms of energy couple to gravity, both the masses and spins leave an imprint on the binary orbit and the GWs emitted. The coupling between spin and orbital angular momentum can strongly affect the orbital trajectory and the GWs emitted in the inspiral phase. The PN approximation can be used to construct a model of the GWs emitted during inspiral. The combination of such a GW model with a model for the instrument response yields templates for the signals as seen by the detector. Subtracting the model from the data produces a residual, and demanding that the residual is consistent with a model for the instrument noise defines a likelihood function. From this function and our prior knowledge we can derive a posterior distribution for the model parameters that are consistent with the observed data. It often happens that there are strong correlations between these parameters, limiting our ability to measure each parameter individually. Recent work~\\citep{Hannam:2013uu} has suggested that the correlation between mass and spin~\\citep{Cutler:1992tc,Cutler:1994ys} may make it impossible to distinguish between a NSNS binary and a NSBH or a BHBH binary. This result hinges on a simplified waveform model that assumes that the spin and orbital angular momenta are perfectly aligned, and thus, spin-orbit induced precession~\\citep{springerlink:10.1007/BF00756587,Bohe:2012mr} is absent. However, we have no reason to expect the spin and orbital angular momenta to be aligned in stellar remnant binaries. Indeed, the NS binaries observed at much longer orbital periods are far from aligned and are precessing~\\citep{2002ApJ...576..942W,2005ApJ...624..906H,2008Sci...321..104B}. It has been hypothesized~\\citep{Hannam:2013uu,Baird:2012cu} that spin precession would not significantly alter the conclusions drawn using spin-aligned waveforms. We have tested this hypothesis and found, as first suggested by~\\citet{Cutler:1992tc}, that spin precession adds additional richness to the signals that almost completely breaks the mass-spin degeneracy, producing an order-of-magnitude improvement in the extraction of the individual masses and spins, which allows us to distinguish between NSs and BHs. We show that normal NS binaries will not be mistaken for BHs or exotic NSs, but we cannot rule out the possibility that some exotic NSs or low mass/low spin BHs may be misidentified as normal NSs. ", "conclusions": "We showed that the inclusion of spin-precession in waveform templates breaks the degeneracy between the system's individual masses and spins, and allows us to distinguish between NSNS binaries and low-mass, small-spin NSBH or BHBH binaries. Moreover, even for signals with modest SNR, we can distinguish between ``normal'' and ``exotic'' NSs. These results open the door to population studies with the first GW detections, as well as coincident studies between the electromagnetic detection of short gamma-ray bursts and GWs. Indeed, if such a coincident observation is made, being able to identify the source from purely GW observations as a NS binary, a mixed binary or a BH binary would prove invaluable. The results presented here are subject to several assumptions. First, the noise is assumed to be stationary and Gaussian, while in reality this may not be the case. Proper noise modeling along the lines described in~\\citet{Littenberg:2014oda,Cornish:2014kda} will help to restore performance to levels close to the ideal case. Second, calibration errors and non-stationary drifts in the noise spectrum should be marginalized over in a full analysis, but these mostly impact the amplitude parameters, and only have a small impact on the spin measurement. Third, the waveform model inaccuracies do not affect our estimates of the statistical errors at leading order~\\citep{Chatziioannou:2014bma}, so our conclusions will apply to more accurate waveform models. \\newpage \\emph" }, "1402/1402.3062_arXiv.txt": { "abstract": "{The commonly used extinction laws of Cardelli et al. (1989) have limitations that, among other issues, hamper the determination of the effective temperatures of O and early B stars from optical and NIR photometry.} {We aim to develop a new family of extinction laws for 30 Doradus, check their general applicability within that region and elsewhere, and apply them to test the feasibility of using optical and NIR photometry to determine the effective temperature of OB stars.} {We use spectroscopy and NIR photometry from the VLT-FLAMES Tarantula Survey and optical photometry from HST/WFC3 of 30 Doradus and we analyze them with the software code CHORIZOS using different assumptions, such as the family of extinction laws.} {We derive a new family of optical and NIR extinction laws for 30 Doradus and confirm its applicability to extinguished Galactic O-type systems. We conclude that by using the new extinction laws it is possible to measure the effective temperatures of OB stars with moderate uncertainties and only a small bias, at least up to $\\ebv \\sim 1.5$ mag.} {} ", "introduction": "Astronomy is entering a time when massive photometric surveys allow us to obtain information about a very large number of objects. Projects such as Gaia and the Large Synoptic Survey Telescope (LSST) will reinforce this trend in the next decade. The main goal of these surveys is to measure the intrinsic properties of these objects, such as the effective temperature, luminosity, and metallicity of stars; the mass, age, and metallicity of stellar clusters; or the redshift and type of galaxies. These surveys include not only large numbers of targets, but also detailed calibration mechanisms that lead to (internal) precisions and (external) accuracies at the level of one hundredth of a magnitude. In other words, we have not only data in large quantities but also with high quality in the form of random and systematic errors that are significantly lower than what was typical twenty years ago. This high photometric quality is also extended to space missions such as the Hubble Space Telescope (HST) and is due to the stability of the space environment and the resources devoted to ensure the uniformity of the data. Despite such high quality, there is (and always will be) one obstacle for the derivation of the intrinsic properties of astronomical objects: extinction. Every observation has to be corrected for the presence of dust between the target and the observer and that can be (and in many cases is) the main limitation. In the 1980s the great success of the International Ultraviolet Explorer (IUE) satellite prompted a revived interest in the subject of extinction that culminated with the groundbreaking work of \\citet[hereafter CCM]{Cardetal89}. that paper provided for the first time a family of extinction laws that extended from the IR to the UV while simultaneously characterizing the type of extinction with a single parameter, \\rv\\ (see \\citealt{Maiz13b} for a discussion on the name and the precise nature of the parameter). These two characteristics made the CCM laws a resounding success and the paper one of the most cited in astronomy in the last quarter of a century. Despite their unquestioned relevance, different studies in the last two decades have revealed several issues with some aspects of the CCM laws: \\begin{itemize} \\item the use of band-integrated [$E(B-V)$ and $R_V$] quantities to define the amount and type of extinction instead of their monochromatic equivalents (\\ebv\\, and \\rv, respectively)\\footnote{It cannot be emphasized enough that using $R_V$ to parameterize an extinction law is a serious mistake. $R_V\\equiv A_V/E(B-V)$ depends not only on the extinction law but also on the amount of extinction and the input SED. The reader is referred to Figure 3 of \\citet{Maiz13b} to quantify the effect. The parameter called $R_V$ in CCM is not really that (in the sense that an extinction law with a given value of that parameter does not yield that value of $A_V/E(B-V)$ for an arbitrary amount of extinction and an arbitrary SED), but a monochromatic value. In addition, this type of effect in broad-band photometry has been known at least since \\citet{Blan57} but appears to be overlooked by a significant fraction of the astronomical community.}; \\item the validity of a fixed extinction law in the NIR; \\item the functional form used in the optical; \\item the reality of the correlation between \\rv\\ and UV extinction; \\item the applicability of the laws beyond the \\ebv\\ and \\rv\\ values of the sample used to derive them; \\item the photometric calibration of the filters. \\end{itemize} These issues are discussed in \\citet{Maiz13b}, where the reader is referred for details, and they are the reasons that prompted us to attempt an improvement of the CCM laws, concentrating on the correction for extinction for photometric data, their most commonly used application. This paper is part of a series on the VLT-FLAMES Tarantula Survey. The reader is referred to the first paper, \\citet{Evanetal11a}, for details on the project. Within the series, this paper on the optical and NIR extinction law in 30 Doradus and its application to the determination of effective temperatures (\\teff) is part of a subseries on extinction and the ISM. The subseries started with the work of \\citet{vanLetal13} on diffuse interstellar bands and neutral sodium and will continue with another paper on the spatial distribution of extinction in 30 Doradus (Ma{\\'\\i}z Apell\\'aniz et al. in preparation). We start by describing the spectroscopic and photometric data in this paper. We then perform different experiments with the data by processing them with CHORIZOS \\citep{Maiz04c}. The results are discussed and possible future work is described. The paper ends with three appendixes on (a) the detailed changes introduced by the new laws, (b) CHORIZOS and the spectral energy distributions (SEDs) used for this paper, and (c) the extinction along a sightline with more than one type of dust. ", "conclusions": "" }, "1402/1402.1251_arXiv.txt": { "abstract": "{In this paper I provide a general framework based on $\\delta N$ formalism to study the features of unavoidable higher dimensional non-renormalizable K\\\"ahler operators for ${\\cal N}=1$ supergravity (SUGRA) during primordial inflation from the combined constraint on non-Gaussianity, sound speed and CMB dipolar asymmetry as obtained from the recent Planck data. In particular I study the nonlinear evolution of cosmological perturbations on large scales which enables us to compute the curvature perturbation, $\\zeta$, without solving the exact perturbed field equations. Further I compute the non-Gaussian parameters $f_{NL}$ , $\\tau_{NL}$ and $g_{NL}$ for local type of non-Gaussianities and CMB dipolar asymmetry parameter, $ A_{CMB}$, using the $\\delta N$ formalism for a generic class of sub-Planckian models induced by the Hubble-induced corrections for a minimal supersymmetric D-flat direction where inflation occurs at the point of inflection within the visible sector. Hence by using multi parameter scan I constrain the non-minimal couplings appearing in non-renormalizable K\\\"ahler operators within, ${\\cal O}(1)$, for the speed of sound, $0.02\\leq c_s\\leq 1$, and tensor to scalar, $10^{-22} \\leq r_{\\star} \\leq 0.12$. Finally applying all of these constraints I will fix the lower as well as the upper bound of the non-Gaussian parameters within, ${\\cal O}(1-5)\\leq f_{NL}\\leq 8.5$, ${\\cal O}(75-150)\\leq\\tau_{NL}\\leq 2800$ and ${\\cal O}(17.4-34.7)\\leq g_{NL}\\leq 648.2$, and CMB dipolar asymmetry parameter within the range, $0.05\\leq A_{CMB}\\leq 0.09$.} \\begin{document} ", "introduction": "The primordial inflationary paradigm is a very rich idea to explain various aspects of the early universe, which creates the perturbations and the matter. For recent developments see Refs.~\\cite{Mazumdar:2010sa,Mazumdar:2011zd}. Usually inflation prefers slow-rolling of a single scalar field on a flat potential, which has unique predictions for the Cosmic Microwave Background (CMB) observables. The induced cosmological perturbations are generically random Gaussian in nature with a small tilt and running in the primordial spectrum which indicates that inflation must come to an end in our patch of the universe. But a big issue may crop up in model discrimination and also in the removal of the degeneracy of cosmological parameters obtained from CMB observations \\cite{Spergel:2006hy,Ade:2013uln,Ade:2013zuv,Ade:2013ydc}. Non-Gaussianity has emerged as a powerful observational tool to discriminate between different classes of inflationary models \\cite{Maldacena:2002vr,Bartolo:2004if,Komatsu:2009kd,Chen:2010xka,Komatsu:2010hc}. The Planck data show no significant evidence in favour of primordial non-Gaussianity, the current limits \\cite{Ade:2013ydc} are yet to achieve the high statistical accuracy expected from the single-field inflationary models and for this opportunities are galore for the detection of large non-Gaussianity from various types of inflationary models. To achieve this goal, apart from the huge success of cosmological linear perturbation theory, the general focus of the theoretical physicists has now shifted towards the study of nonlinear evolution of cosmological perturbations. Typically any types of nonlinearities are expected to be small; but, that can be estimated via non-Gaussian n-point correlations of cosmological perturbations. The so-called ``$\\delta N$ formalism'' (where $N$ being the number of e-foldings) \\cite{Starobinsky:1982,Salopek:1990,Sasaki:1995aw,Wands:2000dp,Lyth:2004gb,Lyth:2005fi,Mazumdar:2012jj,Sugiyama:2012tj} is a well accepted tool for computing non-linear evolution of cosmological perturbations on large scales ($k<