{ "0807/0807.0809_arXiv.txt": { "abstract": "Studies of nucleosynthesis in neutrino-driven winds from nascent neutron stars show that the elements from Sr through Ag with mass numbers $A\\sim 88$--110 are produced by charged-particle reactions (CPR) during the $\\alpha$-process in the winds. Accordingly, we have attributed all these elements in stars of low metallicities (${\\rm [Fe/H]}\\lesssim -1.5$) to low-mass and normal supernovae (SNe) from progenitors of $\\sim 8$--$11\\,M_\\odot$ and $\\sim 12$--$25\\,M_\\odot$, respectively, which leave behind neutron stars. Using this rule and attributing all Fe production to normal SNe, we previously developed a phenomenological two-component model, which predicts that ${\\rm [Sr/Fe]}\\geq -0.32$ for all metal-poor stars. The high-resolution data now available on Sr abundances in Galactic halo stars show that there is a great shortfall of Sr relative to Fe in many stars with ${\\rm [Fe/H]}\\lesssim -3$. This is in direct conflict with the above prediction. The same conflict also exists for two other CPR elements Y and Zr. The very low abundances of Sr, Y, and Zr observed in stars with ${\\rm [Fe/H]}\\lesssim -3$ thus require a stellar source that cannot be low-mass or normal SNe. We show that this observation requires a stellar source leaving behind black holes and that hypernovae (HNe) from progenitors of $\\sim 25$--$50\\,M_\\odot$ are the most plausible candidates. Pair-instability SNe from very massive stars of $\\sim 140$--$260\\,M_\\odot$ that leave behind no remnants are not suitable as they are extremely deficient in producing the elements of odd atomic numbers such as Na, Al, K, Sc, V, Mn, and Co relative to the neighboring elements of even atomic numbers, but this extreme odd-even effect is not observed in the elemental abundance patterns of metal-poor stars. If we expand our previous phenomenological two-component model to include three components (low-mass and normal SNe and HNe) and use for example, the observed abundances of Ba, Sr, and Fe to separate the contributions from these components, we find that essentially all of the data are very well described by the new model. This model provides strong constraints on the evolution of [Sr/Fe] with [Ba/Fe] in terms of the allowed domain for these abundance ratios. This model also gives an equally good description of the data when any CPR element besides Sr (e.g., Y or Zr) or any heavy $r$-process element besides Ba (e.g., La) is used. As the stars deficient in Sr, Y, and Zr are dominated by contributions from HNe, they define the self-consistent yield pattern of that hypothecated source. This inferred HN yield pattern for the low-$A$ elements from Na through Zn ($A\\sim 23$--70) including Fe is almost indistinguishable from what we had previously attributed to normal SNe. As HNe are plausible candidates for the first generation of stars and are also known to be ongoing in the present epoch, it is necessary to re-evaluate the extent to which normal SNe are substantial contributors to the Fe inventory of the Galaxy. We conclude that HNe are important contributors to the abundances of the low-$A$ elements over the history of the universe. We estimate that they contributed $\\sim 24\\%$ of the bulk solar Fe inventory while normal SNe contributed only $\\sim 9\\%$ (not the usually assumed $\\sim 33\\%$). This implies a greatly reduced role of normal SNe in the chemical evolution of the low-$A$ elements. ", "introduction": "In this paper we consider that the elements from Sr through Ag in metal-poor stars represent the products of nucleosynthesis in neutrino-driven winds from forming neutron stars. This approach allows us to obtain information on the stellar sources that contributed to the chemical enrichment of the interstellar medium (ISM) in the Galaxy and the intergalactic medium (IGM) at early and recent times. We previously proposed a phenomenological two-component model (\\citealt{qw07}; hereafter QW07) to account for the abundances of heavy elements in metal-poor stars. That model focused on the elements commonly considered to be produced by the generic ``$r$-process''. It specifically attributed all the elements from Sr through Ag in metal-poor stars to the charged-particle reactions (CPR) in the neutrino-driven winds from nascent neutron stars and used this as a diagnostic of the sources for these CPR elements. In contrast, the true $r$-process elements (e.g., Ba and higher atomic numbers) are produced by extensive rapid neutron capture. It was assumed in the two-component model that Fe was only produced by normal supernovae (SNe) from progenitors of $\\sim 12$--$25\\,M_\\odot$, which leave behind neutron stars, and that the heavy $r$-process elements ($r$-elements) with mass numbers $A>130$ were formed in low-mass SNe from progenitors of $\\sim 8$--$11\\,M_\\odot$, which also leave behind neutron stars but produce no Fe. Thus the CPR elements would be produced by both low-mass and normal SNe and the corresponding yields were estimated (QW07). It follows that if these SNe were the only sources, then the presence of Fe should always be associated with that of the CPR elements. A more extensive study of the available observational data shows that some low-metallicity stars have Fe but essentially no Sr. In particular, \\citet{fulbright04} found a star with ${\\rm [Fe/H]}=\\log{\\rm (Fe/H)}-\\log{\\rm (Fe/H)}_\\odot=-2.88$ and $\\log\\epsilon({\\rm Sr})=\\log({\\rm Sr/H})+12<-2.6$ in the dwarf galaxy Draco. From the two-component model we would have estimated $\\log\\epsilon({\\rm Sr})=-0.28$ for this star, which is far above the observational upper limit. These results clearly indicate that if the CPR elements are always produced during the formation of neutron stars, then there must be an additional stellar source contributing Fe that does not leave behind neutron stars, or else the above model for the production of the CPR elements is in error. Utilizing a more extensive data base than QW07 and especially treating the data on stars very deficient in Sr, Y, and Zr relative to Fe at ${\\rm [Fe/H]}<-3$, the present paper will show that a third source in addition to the two sources (low-mass and normal SNe) in the model of QW07 is required to account for the elemental abundances in metal-poor stars. It will be argued that the third source producing Fe but no CPR elements is most likely associated with hypernovae (HNe) from progenitors of $\\sim 25$--$50\\,M_\\odot$, which leave behind black holes instead of neutron stars. It is then shown that essentially all of the stellar data on elemental abundances at ${\\rm [Fe/H]}\\lesssim -1.5$ can be decomposed in terms of three distinct types of sources. This decomposition also identifies a yield pattern for the elements from Na through Zn including Fe that is attributable to HNe. An important conclusion is that this HN yield pattern is almost indistinguishable from what is attributed to normal SNe. Further, the discovery of extremely energetic HNe associated with gamma-ray bursters (e.g., \\citealt{galama,iwa98}) in the present universe requires that contributions from this source must be considered both in early epochs and on to the present. This leads to a reassessment of the contributions from different sources to the Galactic Fe inventory, which shows that ongoing HNe must play an important role and that the usual attribution of $\\sim 1/3$ of the solar Fe inventory to normal SNe is not valid. We aim to present a phenomenological three-component (low-mass and normal SNe and HNe) model for the chemical evolution of the early Galaxy that may provide a quantitative, self-consistent explanation for many of the results from stellar observations. We focus on three groups of elements: the low-$A$ elements from Na through Zn ($A \\sim 23$--70), the CPR elements from Sr through Ag ($A \\sim 88$--110), and the heavy $r$-elements ($A > 130$, Ba and higher atomic numbers). In \\S\\ref{sec-2cm} we give a brief outline of the two-component model of QW07 with low-mass and normal SNe represented by the $H$ and $L$ sources, respectively. In \\S\\ref{sec-data} we present the data on abundances of Sr and Ba as well as Y and La for a large sample of metal-poor stars, and show that the two-component model fails at ${\\rm [Fe/H]}\\lesssim -3$ and that an additional source producing Fe but no Sr or heavier elements is required to account for the data at such low metallicities. This source is identified with HNe. It is then shown that the extended three-component model with HNe, $H$, and $L$ sources gives a good representation of nearly all the data on the CPR elements Sr, Y, and Zr, but leads to the conclusion that the HN yield pattern is indistinguishable from that of the $L$ source for all the low-$A$ elements. Considering that HNe not only represent the first massive stars (Population III stars) but also must continue into the present epoch, we reinterpret the yields attributed to the hypothetical $L$ source as the combined contributions from normal SNe, which we designate as the $L^*$ source, and HNe. In \\S\\ref{sec-3cm} we show that the three-component model with HNe, $H$, and $L^*$ sources gives a very good representation of essentially all the data on the CPR elements Sr, Y, and Zr and further discuss the characteristics of these sources and their roles in the chemical evolution of the universe. We give our conclusions in \\S\\ref{sec-con}. ", "conclusions": "\\label{sec-con} The two-component model of QW07 with the $H$ and $L$ sources provided a good description of the elemental abundances in metal-poor stars of the Galactic halo for $-2.7<{\\rm [Fe/H]}\\lesssim -1.5$. A key ingredient of that model is the attribution of the elements from Sr through Ag in metal-poor stars to the charged-particle reactions in neutrino-driven winds from nascent neutron stars but not to the $r$-process. However, that model cannot explain the great shortfall in the abundances of Sr, Y, and Zr relative to Fe for stars with ${\\rm [Fe/H]}\\lesssim -3$. The observations on these three CPR elements require that there be an early source producing Fe but no Sr or heavier elements. It is shown that if such a third source is assumed, then the data can be well explained by an extended three-component model. From considerations of the abundance patterns of the low-$A$ elements (from Na through Zn), it is concluded that this third source is most likely associated with HNe from massive stars of $\\sim 25$--$50\\,M_\\odot$ that do not leave behind neutron stars. We here consider the third source to be HNe. It is shown that the available data on the evolution of [Sr/Fe] with [Ba/Fe], that of [Y/Fe] with [La/Fe], and that of [Zr/Fe] with [Ba/Fe] are well described by the extended model with HNe, $H$ and $L$ sources, which also provides clear constraints on the abundance ratios that should be seen. It is further shown that the abundance patterns of the low-$A$ elements for HNe and the $L$ sources are not distinguishable. Considering that HNe are observed to be ongoing events in the present universe, we are forced to conclude that the $L$ source, which was assumed to have provided $\\sim 1/3$ of the solar Fe inventory (the rest attributed to SNe Ia), is in fact a combination of normal SNe (from progenitors of $\\sim 12$--$25\\,M_\\odot$), which we define as the $L^*$ source, and HNe. The net Fe contributions from HNe are found to be $\\sim 3$ times larger than those from normal SNe. Using the three-component model with HNe, $H$, and $L^*$ sources, we obtain a very good quantitative description of essentially all the available data. In particular, this model provides strong constraints on the evolution of [Sr/Fe] with [Ba/Fe] in terms of the allowed domain for these abundance ratios. It gives an equally good description of the data when any CPR element besides Sr (e.g., Y or Zr) or any heavy $r$-element besides Ba (e.g., La) is used. The model is also compatible with the non-$s$-process contributions to the solar abundances of all the CPR elements. The anomalous abundance patterns of the low-$A$ elements observed in a small number of stars appear to fit the description of faint SNe (e.g., \\citealt{iwamoto}), which are a rarer type of events from the same progenitor mass range as HNe but with even weaker explosion energies and smaller Fe yields than normal SNe (e.g., \\citealt{nomoto06}). The anomalous abundance patterns observed reflect the fact that faint SNe produce very little of the Fe group elements but an abundant amount of the elements from hydrostatic burning in their outer shells. This gives rise to the extremely high abundances of Na, Mg, and Al relative to Fe observed in the anomalous stars. The quasi-uniform abundance patterns of the elements from Si through Zn in all cases (including the stars with anomalous abundances of Na, Mg, and Al) appear to reflect some robustness in the outcome of explosive burning that may arise from the limited range of conditions required for such nucleosynthesis. In this paper we used the elemental yield patterns for three prototypical model sources to calculate the abundances of an extensive set of elements (relative to hydrogen) for metal-poor stars with ${\\rm [Fe/H]}\\lesssim -1.5$. As the yield patterns adopted for the assumed prototypical sources are taken from the data on two template stars, they must represent the results of stellar nucleosynthesis. The full version of the three-component model appears very successful in calculating the abundances of the elements ranging from Na through Pt in stars with ${\\rm [Fe/H]}\\lesssim -1.5$. In contrast to this phenomenological approach, there are extensive studies of Galactic chemical evolution (GCE) that use the various theoretical results on the absolute yields of metals for different stellar types. These theoretical yields are not calculated from first principles, but are dependent on the parametrization used in the various stellar models. In those GCE studies, the elemental abundances for an individual star are not predicted. Instead, general trends for the elemental abundances are calculated assuming different sources, the rates at which they contribute, and a model of mixing in the ISM for different regions of the Galaxy. These results give a good broad description for typical elemental abundances in the general stellar population at higher metallicities of ${\\rm [Fe/H]}>-1.5$. This is a regime in which the observational data are quite convergent with only limited variability. However, as anticipated by \\citet{gilroy} and supported by the considerable scatter in the abundances of heavy elements observed in stars with ${\\rm [Fe/H]}\\lesssim -1.5$, the chemical composition of the ISM in the early Galaxy was extremely inhomogeneous. For ${\\rm [Fe/H]}<-2$ there are gross discrepancies between the observations and the smoothed model of GCE. In no case does that model give the elemental abundances for an individual star. It is our view that the simple phenomenological model used here permits a clearer distinction between the different stellar sources contributing to the ISM and the IGM at early times. This model also gives specific testable predictions, which can be used to further check its validity. In conclusion, we consider that the general three-component model with HNe, $H$, and $L^*$ sources provides a quantitative and self-consistent description of nearly all the available data on elemental abundances in stars with ${\\rm [Fe/H]}\\lesssim -1.5$. Further, HNe may be not only explosions from the first massive stars (i.e., the Population III stars much sought after by many) that provided a very early and variable inventory to the IGM through ejection of enriched gas from small halos, but also are important ongoing contributors to the chemical evolution of the universe." }, "0807/0807.2384_arXiv.txt": { "abstract": "LS 5039, a possible black hole x-ray binary, was recently observed with Giant Meterwave Radio Telescope. The observed spectrum presented here shows that the spectrum is inverted at the low frequency. When combined with the archival data with orbital phase similar to the present observations, it shows a clear indication of a spectral turnover. The combined data are fitted with a broken power-law and the break frequency signifies a possible spectral turnover of the spectrum around 964 MHz. Truly simultaneous observations in radio wavelength covering a wide range of frequencies are required to fix the spectrum and the spectral turn over which will play a crucial role in developing a deeper understanding of the radio emitting jet in LS 5039. ", "introduction": "LS5039 is a X-ray binary system consisting of a compact object and a companion star of ON6.5V(f) type with mass 22.9M$_{\\odot}$. The mass of the compact object is not accurately known. But recent study shows that the mass of the compact object is 3.7M$_{\\odot}$ (\\cite{cas05}). Although there are lot of debates regading the true nature of LS 5039 (\\cite{mira06}), the dynamical mass estimate indicates that it is a possible black hole x-ray binary candidate. The orbital period of the binary system is 3.9 days (\\cite{cas05}). The binary orbit is highly eccentric with an eccentricity of 0.35 and has an inclination angle 24$^0$.9$\\pm$2$^0$.8 (\\cite{cas05}). LS5039 was proposed to be a very high energy gamma-ray emitter by \\cite{parde00} and was recently seen at TeV $\\gamma$-ray energies by gamma-ray telescope High Energy Stereoscopic System (\\cite{aha06}). The gamma-ray light curve shows an orbital modulation of $\\sim3.9$ days. The gamma-ray flux as well as the spectral hardness vary with the orbital phase. The flux variation with orbital phase can be accounted for by considering the absorption of gamma-rays due to $\\gamma\\gamma$ pair production. This also implies that $\\gamma$-photons are emitted from a region within 1 AU of the compact object (\\cite{aha06}). The observed spectrum is very steep at the superior conjunction and becomes harder as the compact object moves away from it and the spectral index goes below 2.0 as the compact object reaches the inferior conjuntion. But this variation of spectral hardness can not be adequately explained by the gamma-ray absorption via pair production alone. This source was observed in X-ray by $ROSAT$ (\\cite{mot97}), $RXTE$ (\\cite{ribo99}), $ASCA$ (\\cite{mart05}), $BeppoSAX$ (\\cite{reig03}), $Chandra$ (\\cite{bosc05}) and $XMM-Newton$ (\\cite{mart05}) missions during 1996 to 2003. These observations were effectively carried out during different orbital phases of the binary system. All these observations revealed different flux levels and different spectral indices in the X-ray energy band, but no x-ray spectral state transition were reported in these observations. More recently \\cite{bosc05} reported results of $RXTE$ observations when the source was observed consecutively for four days in July 2003. The X-ray flux is found to vary with the orbital phases and it maximizes near periastron passage. This phase dependence of X-ray flux is due to the motion of the compact object accreting matter from the wind of the massive companion while moving in a highly eccentric orbit (\\cite{bosc05}). But the observed anti-correlation between the photon index and the X-ray flux does not fit to the scenario where X-rays are considered to be produced in and around an accretion disc. \\cite{bosc05} argued that the x-ray emission might be due to inverse Compton/synchrotron process in a relativistic jet which might possibly explain the observed anti-correlation between the photon index and the observed flux. But the photon indices reported in these observations are very similar to that generally observed in case of hard state spectrum of black hole x-ray binaries. LS5039 was first observed by \\cite{marti98} in radio with Very Large Array (VLA) and the observation resulted a power-law spectrum with a negative power-law index. This indicates an optically thin synchrotron emission by non-thermal electrons. This observation also revealed a moderate variability in the radio flux. Later, \\cite{ribo99} carried out an observation campaign in radio with VLA at 2.0, 3.6, 6.0 and 20.0 cm wavelengths and with Green Bank Interferometer (GBI) at 3.6 and 13.3 cm wavelengths. The observed spectrum was a power-law with a power-law index of -0.46$\\pm$0.01 supporting the previous observation of \\cite{marti98}. \\cite{pare02} first resolved the source by observing with European VLBI Network (EVN) and the Multi-Element Radio-Linked Interferometer Network (MERLIN) and it was found that LS 5039 consists of an asymmetric two-sided jet with an maximum extension of one side $\\sim$ 1000 AU. Recently \\cite{ribo08} studied the radio morphology of LS5039 using Very Long Baseline Array (VLBA) and it was found that the radio morphology changed into asymmetric from a very symmetric configuration within five days, but the radio emission from the core did not vary appreciably. {\\it But what is most interesting found in these radio observations is the optically thin non-thermal radio spectrum of the source which is in direct contradiction to the generally observed trend of flat or inverted radio spectrum for persistent black hole x-ray binaries in the hard state~(\\cite{fen01,gal03}). Therefore it is important to study LS5039 in the low-frequency radio band.} We observed LS5039 at wavelengths 21 cm, 50 cm and 128 cm (1280 MHz, 614 MHz and 234 MHz respectively) using Giant Meterwave Radio Telescope~(GMRT). In this letter we discuss the results of the observation. In Section 2 we discuss the observation and data reduction, results are discussed in Section 3 and we conclude the paper in Section 4. ", "conclusions": "Here, we report the low frequency radio spectrum of a X-ray binary LS5039 as observed with GMRT. The spectrum is inverted and shows an indication of spectral turnover. With the assumption that the variability of radio emission from the source is not appreciable, present data are plotted with the archival data. The fitting of the data with a broken power-law reveals a possible indication of a spectral turnover at 964 MHz and an inverted spectrum with spectral index 0.749. This estimation of turnover frequency is indeed obtained using the archival VLA data (\\cite{marti98}) with the assumption that the source is not variable. But this estimation may depend on the variability of the source. To have an idea about the sensitivity of the spectral break frequency on flux variability, we introduced a moderate ($\\sim 10\\%$) flux variation to VLA data and fitted the spectrum. The spectral fitting shows that the break frequency $\\nu_b$ does not change appreciably and it is well within the uncertainities of the statistical fit. As LS 5039 is a possible black hole x-ray binary persistently in low/hard x-ray spectral state, the results of the present observation are consistent with the trend of inverted radio spectrum observed for persistent black hole x-ray binary and this is the first source of this kind for which the indication of a spectral turnover is obtained. The spectral turnover can be used to determine the magnetic field in the jet provided the size of the source is known at that frequency. But it is to be noted that the magnetic field depends on the fourth power of the angular size. Therefore, a reliable measurement of magnetic field is possible if and only if the angular size of the source at the turnover frequency is determined with sufficient accuracy. This result will further help to constrain the theoretical models to explain the broadband spectrum from the source. To have a deeper understanding on the radio spectrum and its dependence on the orbital phases it is important to have simultaneous observations over the entire radio band. Nevertheless, we are conducting a long term campaign to study the spectral behaviour at frequencies 234, 614 and 1280 MHz over different orbital phases of LS5039. This will allow us to constrain the low frequency end of the spectrum of LS5039." }, "0807/0807.2451_arXiv.txt": { "abstract": "Emission from Active Galactic Nuclei is known to vary strongly over time over a wide energy band, but the origin of the variability and especially of the inter-band correlations is still not well established. Here we present the results of our X-ray and optical monitoring campaign of the quasar \\mr , covering a period of 2.5 years. The X-ray 2--10 keV flux is remarkably well correlated with the optical B, V and R bands, their fluctuations are almost simultaneous with a delay consistent with 0 days and not larger than 4 days in either direction. The amplitude of variations shows an intriguing behaviour: rapid, large amplitude fluctuations over tens of days in the X-rays have only small counterparts in the optical bands, while the long-term trends over hundreds of days are \\emph{stronger} in the B band than in X-rays. We show that simple reprocessing models, where all the optical variability arises from the variable X-ray heating, cannot simultaneously explain the discrepant variability amplitudes on different time-scales and the short delays between X-ray and optical bands. We interpret the variability and correlations, in the optically-thick accretion disc plus corona scenario, as the result of intrinsic accretion rate variations modulating both X-ray and optical emission, together with reprocessing of X-rays by the accretion disc. ", "introduction": "Optical continuum variability is a universal property of radio-quiet AGN which has been studied for several decades, but its origin is still unclear. The optical continuum emission almost certainly originates from the accretion disc \\citep{koratkarblaes99}, so it is natural to assume that disc variability (e.g. through accretion rate fluctuations) drives the optical variability. Models where the variability is {\\it intrinsic} to the disc, however, suffer from several problems. Firstly, optical variability is seen on time-scales as short as a day, whereas accretion variability time-scales in standard accretion discs \\citep{shakura} should be long in the optically-emitting regions, comparable to the viscous time-scale which can be of the order of a year or more \\citep{trevesetal88}. Secondly, variations are well-correlated in different optical bands, with only short (days) delays between bands, in the sense that longer-wavelength variations lag shorter-wavelength variations \\citep{wandersetal97,collieretal01,cackettetal07}. Since local disc temperature $T$ should decrease with radius $R$ as $T\\propto R^{-3/4}$, shorter wavelengths originate from smaller radii. Therefore, in the case of inward-propagating fluctuations the long-wavelength variations lead, rather than lag, the short-wavelength variations. Also, the resulting delays would be large, comparable to the radial drift time-scale, contrary to the observed day-scale delays. To account for the observed optical properties, \\citet{kroliketal91} suggested that the variability is driven by changes of the X-ray continuum which illuminates and thereby heats the disc, causing optical continuum variations. Since the X-ray emission is likely to be centrally concentrated, blue bands should respond to the X-ray variations first, followed by red, with short delays corresponding to the light-travel time delay between the blue and red-emitting parts of the disc. The predicted optical time-delay $\\tau$ scales with wavelength $\\lambda$ as $\\tau \\propto \\lambda^{4/3}$ \\citep{collieretal99}, and the observed dependencies of optical continuum lags on wavelength are consistent with this prediction \\citep{cackettetal07}. A more stringent test of reprocessing models is to directly compare X-ray and optical variations, using multi-wavelength monitoring campaigns. To date, a handful of these campaigns have been carried out, largely facilitated by the {\\it Rossi X-ray Timing Explorer} ({\\it RXTE}) in conjunction with various ground based observatories. The results have been mixed, with some AGN showing good evidence for correlated variability, as would be expected from reprocessing models (e.g. NGC~4051, \\citealt{petersonetal00,shemmeretal03}; NGC~5548, \\citealt{uttleyetal03}; Mrk~509, \\citealt{marshalletal08}), while others show more complicated but possibly correlated behaviour (e.g. NGC~7469, \\citealt{nandraetal98,nandraetal00}) and one AGN shows no apparent correlation at all (NGC~3516, \\citealt{maozetal02}). The best case for correlated variability to date is NGC~5548, but \\citet{uttleyetal03} note that it is difficult to reconcile the large fractional amplitude of optical variability on time-scales of months-years with reprocessing models, since the X-ray variability shows smaller-amplitude variations than the optical on these long time-scales. One would expect smaller relative variations in optical since the intrinsic emission due to viscous disc heating would dilute any variable reprocessed component. Furthermore, Gaskell (2007) notes that a simple energetics argument rules out reprocessing as the dominant source of optical variability in many AGN (including NGC~5548) which have `big blue bumps' dominating the total luminosity, significantly exceeding the X-ray luminosity available for reprocessing. Detailed reprocessing calculations, produced by \\citet{reprocessing} for the case of NGC~3516 show that simple reprocessing of X-rays on their own could not produce the observed optical variability. Also, in some AGN with simultaneous X-ray and optical monitoring, there is evidence that the optical may {\\it lead} the X-ray variations, which might be expected if at least some of the optical variability is produced by intrinsic accretion fluctuations propagating through the disc \\citep{shemmeretal03,marshalletal08}. Clearly, no single model provides a satisfactory explanation for all the data. It is possible, however, that some combination of reprocessing and intrinsic accretion variations may explain the diverse range of optical/X-ray behaviour which is already observed in only a small sample of AGN. The location of the optical emitting region probably plays a key role in determining the balance of intrinsic versus reprocessed variability. It is governed by the disc temperature, which scales inversely with radius $R$ (in units of the Gravitational radius $R_g=GM/c^2$) but also scales with black hole mass $M_{\\rm BH}$ and accretion rate (as a fraction of the Eddington rate) $\\dot{m}$ as $T\\propto[(\\dot{m}/{M}_{\\rm BH})R^{-3}]^{1/4}$ \\citep{trevesetal88}. Thus over the 3 decade range in black hole mass expected for AGN, we expect the radius corresponding to a given temperature to change by a factor of 10 (at the same fractional accretion rate), perhaps even more for different accretion rates. Since disc temperature governs the radius where peak optical emission is produced, \\citet{uttleyetal03} suggested that a diverse range of optical/X-ray behaviour might result from the range of masses and accretion rates expected in AGN \\citep[see also][]{li}. For example, the AGN with the most massive black holes will have cool discs and very centrally concentrated optical-emitting regions, in terms of gravitational radii. Therefore, the disc variability time-scales will be short, comparable with the X-ray variability time-scales, so intrinsic disc variability as well as reprocessing may contribute significantly to the optical variations. In contrast, AGN with lower-mass black holes will have hotter discs so optical emission will originate from larger radii, where disc variability time-scales are very long compared to the corresponding time-scales in the innermost regions, so that only reprocessing may contribute significantly to the rapid variability we observe. To test composite models of optical variability, it is necessary to carry out combined optical/X-ray monitoring of AGN with widely varying masses and accretion rates. To date, these campaigns have focused on Seyfert galaxies, which typically cover bolometric luminosities ranging from $10^{43}$ to $10^{45}$ erg~s$^{-1}$. Here we report the first optical/X-ray correlation for a more luminous radio-quiet AGN, the quasar \\mr\\, which we have monitored simultaneously in X-ray and optical bands for the past 2.5 years. We describe the data in Section~2 and show the cross correlations in Section~3. In Section~4 we explore reprocessing and propagating fluctuations scenarios to explain the X-ray/optical variability in this source and discuss the implications of our results in Section~5. ", "conclusions": "We have carried out a simultaneous monitoring campaign in X-ray and optical B, V and R bands on the quasar \\mr\\ over two and a half years. All bands are significantly variable and their fluctuations are well correlated. The cross correlation functions show peaks around a lag of zero days, significant over the 99\\% confidence level when compared to uncorrelated simulated data. All the delays between the bands are consistent with 0 days. We used daily sampled light curves in X-ray, B and V to constrain the lags between these bands, obtaining values of $0.6 \\pm 3.1$ days between X-ray and B, $1.3 \\pm 3.3$ day for X-ray vs V and $0.5 \\pm 2.4$ days between B and V lags, where positive values indicate higher energy band leading. The lag between X-ray and R band was determined using the long but more sparsely sampled light curve, obtaining a value of -4.5$\\pm$16.8 days. The long term trends, over hundreds of days, are well matched by all the bands observed and are stronger in the B band than in X-rays, while the other optical bands show smaller-amplitude trends. The optical light curves are not corrected by host galaxy contribution, however, so the decreased amplitude of variability in lower energy bands might be partly due to a constant galactic contribution. On short time-scales, of tens of days, there are large X-ray flares which appear strongly attenuated in the optical bands. The intensively sampled light curves show that the rapid X-ray flares do have optical counterparts, but that these are very weak. Pure reprocessing of X-rays cannot produce both the short and long time-scale optical variability: if the long term optical fluctuations were produced by reprocessing, we would expect similarly large rapid optical fluctuations, unless these could be smoothed by light travel effects. Any geometry of the reprocessor, however, cannot smooth out the fluctuations on time-scales of 20--50 days without introducing time-lags of a similar length, which are not observed. A similar conclusion was reached by \\citet{reprocessing} for the Seyfert galaxy NGC~3516, where the X-ray-UV variability could not be explained by simple reprocessing. We show that reprocessing of the observed X-rays cannot reproduce the optical variability in \\mr, either producing rapid fluctuations that were too large or long term trends that were too small. The short term fluctuations however could be reproduced reasonably well with the model if we limited the light curve lengths to $\\sim 100$ days, which reduces the contribution of long term trends to each light curve segment. Therefore, a moderate amount of reprocessing acting on an already long-term variable optical emission can reproduce the observed behaviour. A simple way to decouple the amplitude of short and long term fluctuations is to have two distinct processes producing the optical variability, e.g.\\ accretion rate fluctuations plus reprocessing. The covering fraction of the disc, together with the ratio of intrinsic disc emission to X-ray heating determine the fraction of optical flux arising from reprocessed X-rays. Therefore, the size of the rapid optical fluctuations depends on these parameters and can be largely independent of the amplitude of the long term trends. To test this possibility, we generated simulated light curves and compared their statistical properties with the observed data. In the model, fluctuations propagate inward through the accretion flow, modulating the optical emission as they travel through the thermally emitting region and finally modulate the X-ray emission, assumed to be located in the centre. This process leads to the long term correlated variability. As the fluctuations are produced at every radius in the flow, on a radially dependent time-scale, the centrally emitted X-rays contain short time-scale fluctuations which are originally absent in the optical light curves. The rapid fluctuations, only present in the X-ray emitting region are then introduced into the optical light curves through thermal reprocessing. Lastly, although propagation times across the accretion disc can be long and produce a large lag between optical and X-ray fluctuations, reprocessing reduces the lag and can even revert it to a value of the order of the light crossing time to the reprocessor. For a black hole mass of $10^9 M_\\odot$ the light crossing time of one gravitational radius is 5000 s, so even for a large source height and inner truncation radius of 50 $R_g$ the crossing time would be 4 days, less for a less massive black hole, which is within the observed limits on the lag. We show using the simulated light curves, that small-amplitude rapid fluctuations in the optical band can shift the CCF peak towards a lag of zero days, even when the long term trends of the optical lead the X-rays by $\\sim$50 days. As inward propagation of the fluctuations and reprocessing have opposite effects on the time lag, the relative amount of optical/UV variability produced by each process determines the sign of the lag. This might be the case in other monitored AGN where the optical/UV can either lead or lag the X-rays. An interesting feature of our fits to the optical variability is that we expect the disc to subtend a small solid angle with respect to the X-ray source, in order to reduce the effects of X-ray reprocessing and produce the correct amplitude of optical rapid fluctuations. This small solid angle is consistent with the small amount of reflection inferred from the broadband X-ray spectrum of MR~2251-178 observed by {\\sl BeppoSAX} \\citep{Orr}, which shows only a weak iron line with an equivalent width of $EW\\sim 70$~eV, and modest reflection continuum (reflected fraction $<0.4$). It is therefore possible that MR~2251-178 represents a source in a state where the thin disc is slightly truncated and/or the X-ray emitting region is at a small height over the disc. Optical/X-ray studies of other AGN should provide further useful constraints on their disc/corona geometries." }, "0807/0807.3457_arXiv.txt": { "abstract": "We use time-evolutions of the linear perturbation equations to study the oscillations of rapidly rotating neutrons stars. Our models account for the buoyancy due to composition gradients and we study, for the first time, the nature of the resultant g-modes in a fast spinning star. We provide detailed comparisons of non-stratified and stratified models. This leads to an improved understanding of the relationship between the inertial modes of a non-stratified star and the g-modes of a stratified system. In particular, we demonstrate that each g-mode becomes rotation-dominated, i.e. approaches a particular inertial mode, as the rotation rate of the star is increased. We also discuss issues relating to the gravitational-wave driven instability of the various classes of oscillation modes. ", "introduction": "The oscillations of rotating neutron stars are of great interest in astrophysics. Recent efforts to understand the various pulsation modes of compact stars have to a large extent been motivated by gravitational-wave astronomy~\\citep{1996PhRvL..77.4134A, 1998MNRAS.299.1059A}. Different modes of oscillation depend on different pieces of internal physics and one may hope to use observations to probe, for example, the composition of the high-density region~\\citep{2008GReGr..40..945F,2007GReGr..39.1323B}. Work in this area, which has a long history \\citep[for a review see e.g.,][]{lrr-1999-2}, gained further momentum recently with the relatively successful matching between observed quasiperiodic oscillations in the tails of magnetar flares and calculated torsional oscillations of the neutron star crust~\\citep{1998ApJ...498L..45D, 2005ApJ...634L.153P, 2007AdSpR..40.1446W,2007MNRAS.374..256S}. The excitement following these, the first ever, observations of likely neutron star vibrations is significant. It is clear that our models need to be improved significantly if we expect to carry out a quantitative astero-seismology programme for neutron stars, but the prospects for improvements look good. After all, we already have a reasonable understanding of the dynamics of the crust region \\citep[see][]{2006CQGra..23.5367C,2007MNRAS.374..256S} as well as the superfluid components in the core of the star~\\citep{lrr-2007-1}. The present work is motivated by the need to consider more realistic models of rapidly rotating stars. Our emphasis will be on the role of the internal stratification associated with composition variations. This leads to the presence of the so-called g-modes~\\citep{1992ApJ...395..240R}, and we want to investigate how these modes are affected by fast rotation (the analogous problem for thermal ocean g-modes of rapidly rotating neutron stars has already been considered, see \\citet*{1996ApJ...460..827B}). In order to determine the rotational effects on the oscillation spectrum we study linear perturbations of axisymmetric stellar configurations, where the Coriolis force and the centrifugal flattening of the star are included in the background. In this way, we go beyond the so-called slow rotation approximation, where the rotation itself is treated perturbatively and the mode-frequencies of non-rotating models are determined by a perturbation expansion with respect to the angular velocity of the star $\\Omega$. We are particularly interested in the relationship between the g-modes of a stratified stellar model and the inertial modes~\\citep{1999ApJ...521..764L}, for which the Coriolis force is the main restoring agent. The inertial modes are important since they may be driven unstable by the emission of gravitational waves. The strongest instability is associated with the so-called r-modes~\\citep{2001IJMPD..10..381A,Andersson:2002ch}. A key issue concerns the amplitude at which an unstable mode saturates. Detailed work has shown that an unstable r-mode saturates due to nonlinear couplings to other inertial modes~\\citep{2003ApJ...591.1129A}. However, the mode saturation has only been considered for slowly rotating, non-stratified models and it is important to ask to what extent these results will change if the models are made more realistic. In stratified neutron stars, the g-modes tend to become rotationally dominated beyond a certain rotation rate~\\citep[see e.g.,][]{2000ApJS..129..353Y, 2000A&A...354...86D} and may then be called inertia-gravity modes~\\citep{1989nos..book.....U}. Effects due to the stratification are thus weakened, which indicates that the r-mode saturation estimates may still be reliable, at least for the fastest spinning systems. Improving and extending a numerical code that has been used to study non-stratified stars~\\citep*{2002MNRAS.334..933J}, we approach the problem via time-evolutions of the equations that govern the linear perturbations of a, potentially rapidly, rotating Newtonian star. This strategy has the advantage that one does not have to deal with the many rotationally coupled multipoles, that tend to make a slow-rotation calculation less tractable. On the other hand, a numerical evolution is not expected to have the precision of a frequency domain calculation. Neither should one expect it to yield the complete spectrum of modes. After all, the simulation results depend on the chosen initial data. In absence of a clear idea of the nature of the various modes one does not know how to excite specific oscillations. One way around this problem is to follow the ``recycling'' strategy developed by~\\citet*{Stergioulas:2003ep} and \\citet*{Dimmelmeier:2005zk}. Another, more pragmatic, approach is to simply study the modes that are excited by ``generic'' initial data. This is the attitude that we adopt here. ", "conclusions": "} \\label{sec:concl} Information concerning the oscillation spectra of realistic neutron star models helps develop our understanding of the physics required to describe these objects. Any spectral property can in principle be attributed to particular physical quantities or configurations of the star. Astero-seismology studies, using either electromagnetic or gravitational signals (or indeed both) may thus help constrain the state of matter at supernuclear densities. Of course, in order to facilitate this, we need to improve our theoretical models and clarify the origin of various stellar pulsation features. In this paper, we have studied the pulsations of stratified and rapidly rotating neutron stars with the aim of understanding the dependence of the composition g-modes on the rotation rate of the star. The stellar pulsations were studied using the linearised Euler and conservation equations on an axisymmetric background. In order to simplify the problem, we used the Cowling approximation where the gravitational potential perturbations are neglected. This approximation generates only a small error in the g-mode frequencies. The system of perturbation equations was evolved in time by a 2D numerical code based on a standard finite differencing scheme. This code was tested against various results available in literature and demonstrated good accuracy. Since both Coriolis and buoyancy forces act on a perturbed fluid element of a rotating and stratified star, the low-frequency modes have a mixed inertia-gravity character. They typically behave as g-modes in the slowly rotating limit, while they assume the properties of the inertial modes when the star rotates rapidly. By comparing the oscillation frequencies and the associated eigenfunctions of barotropic inertial modes and the inertia-gravity modes of non-barotropic stars, we have demonstrated how the two sets of modes are associated at fast rotation rates. Our results show that it would be difficult to deduce the presence of composition gradients from the inertia-gravity mode spectrum of fast rotating stars. However, as the neutron star ages and spins down, the dependence of the inertia-gravity modes on the star's rotation and their deviation from the inertial barotropic modes can, at least in principle, be used to estimate the Brunt-V\\\"{a}is\\\"{a}l\\\"{a} frequency and the degree of stratification. As an extension of this work, we are currently developing a numerical code for studying the dynamics of rapidly rotating superfluid neutron stars. This is an important development since all mature neutron stars are expected to have superfluid components in the core. It is also well known that the associated multi-fluid dynamics leaves an imprint on the stellar oscillation spectrum, see \\citet{2007arXiv0709.0660L} for a recent discussion. We hope to be able to report on the initial results of this investigation soon." }, "0807/0807.4870_arXiv.txt": { "abstract": "To first approximation, a binary system conserves its angular momentum while it evolves to its state of minimum kinetic energy: circular orbit, all spins aligned, and components rotating in synchronism with the orbital motion. The pace at which this final state is achieved depends on the physical processes that are responsible for the dissipation of the tidal kinetic energy. For stars (or planets) with an outer convection zone, the dominant mechanism identified so far is the viscous dissipation acting on the equilibrium tide. For stars with an outer radiation zone, it is the radiative damping operating on the dynamical tide. After a brief presentation of the tides, I shall review these physical processes; I shall discuss the uncertainties of their present treatment, describe the latest developments, and compare the theoretical predictions with the observed properties concerning the orbital circularization of close binaries. ", "introduction": "A fundamental property of closed mechanical systems is that they conserve their total momentum. This is true in particular for binary stars, star-planet(s) systems, whether they possess or not a circumstellar disc, if one can ignore the angular momentum that is carried away by winds and by gravitational waves. Through tidal interaction, kinetic energy and angular momentum are exchanged between the rotation of the components their orbital motion and the disc. In the absence of such a disc, which is the case that we shall consider here, they evolve due to viscous and radiative dissipation to the state of minimum kinetic energy, in which the orbit is circular, the rotation of both stars is synchronized with the orbital motion, and their spin axis are perpendicular to the orbital plane. How rapidly the system tends to that state is determined by the strength of the tidal interaction, and thus by the separation of the two components: the closer the system, the faster its dynamical evolution. But it also depends on the efficiency of the physical processes that are responsible for the dissipation of the kinetic energy. Provided these dissipation processes are well enough understood, the observed properties of a binary system can deliver important information on its evolutionary state, on its past history, and even on the conditions of its formation. The first step is thus to identify these physical processes, and it is surprising that this has not been seriously undertaken until the mid-sixties, while tidal theory as such had already reached a high degree of sophistication, starting with the pioneering work of Darwin (1879). In his classical treatise, Kopal (1959) states from the onset that he is interested only in ``dynamical phenomena which are likely to manifest observable consequences in time intervals of the order of 10 or 100 years, and if so, tidal friction can be safely ignored\". But stars live much longer than that, and this is why we shall consider here changes in the properties of binary systems that occur over their evolutionary time scale, and in particular the circularization of their orbits, which is both easy to observe and easy to interpret. We shall deal mainly with binary stars, although much of what follows may be applied also to star-planet systems. ", "conclusions": "Let me summarize. The two tidal dissipation processes that have received most attention so far are the turbulent friction acting on the equilibrium tide, which was first described in the 60's (Zahn 1966b), and the radiative damping of the dynamical tide, identified in the 70's (Zahn 1975). These processes operate respectively in convection and radiation zones, and they have been quite successful in explaining the observed orbital circularization of binary stars. This is particularly true for the early-type MS binaries, for which we have now at our disposal very large samples gathered during the OGGLE and MACHO campaigns: their transition period is precisely defined and it agrees extremely well with that predicted by the theory of the dynamical tide, which is thus validated. However many of these binaries are circularized well above this transition period, as if they had experienced another, more efficient tidal dissipation mechanism. A very likely explanation for this behavior is that these binaries have undergone several episodes of resonance locking, as was described by Witte and Savonije (1999a, 1999b). On the other hand, the equilibrium tide damped by turbulent dissipation accounts very well for the properties of binaries containing a red giant, as was demonstrated by Verbunt \\& Phinney (1995). It also explains the transition period of about 8 days observed in late-type binaries that are younger than about 1 Gyr: the explanation is that these have been circularized during the PMS phase, when they were much more voluminous and fully convective. The only serious discrepancy today seems to be the behavior of late-type main-sequence binaries older than 1~Gyr, whose transition period increases with age and is higher than that predicted when applying straightforward the theory of the equilibrium tide. Here again one may invoke the dynamical tide with resonance locking in the radiative core of these stars, as was shown by Witte and Savonije (2002). Their mechanism appears thus highly promising, and it ought to be further explored. For instance, one should take into account that the tidal torque is applied primarily to specific regions: the outer convection zone in late-type MS stars and the outermost part of the radiation zone in early-type stars. These regions are synchronized more quickly then the rest of the star, and therefore differential rotation develops in their radiation zone. This increases the radiative damping, since the local tidal frequency tends then to zero as the tidal wave approaches the synchronized region, as we explained in \\S\\ref{grav-exc}. For late-type binaries, a highly interesting alternative is offered by the damping of inertial waves in their convective envelope, which is being explored by Ogilvie and Lin (2007). This process is likely to play an important role also in giant planets (Ogilvie \\& Lin 2004). The difficulty in studying these waves is that they require highly resolved 2D numerical calculations, since the so-called traditional approximation is no longer applicable to render the problem separable. Work is in progress on several other points, and I shall quote only a few. Kumar and Goodman (1996) have studied the enhanced damping of the oscillations triggered in tidal-capture binaries, due to non-linear coupling between the eigenmodes, which is extremely strong in such highly eccentric orbits. Rieutord (2004, and in this volume) is examining the possibility that the so-called elliptic instability may occur in binary stars; this instability is observed in the laboratory when fluid is forced to rotate between boundaries that have a slight ellipticity, and it leads to turbulence. Even the equilibrium tide is being revisited, taking into account the differential rotation of the convection zone (Mathis \\& Zahn, in preparation). Finally, it remains to explain why I made no attempt here to reconcile the theoretical predictions for the synchronization of the binary components with their observed surface rotation. The reason is that in most cases the tidal torque is applied mainly to the outermost part of the star, which is synchronized much more rapidly than the interior; therefore the interpretation of the surface rotation requires to model the transport of angular momentum within the star, and in particular where it proceeds slowest, i.e. in the radiation zones. This is a difficult task that only now begins to be undertaken seriously: for recent accounts on this problem, see the reviews by Talon (2007) and Zahn (2007). I am confident that we will see much progress in solving this problem when the next school will be held on that theme, hopefully in a not too distant future! \\bigskip" }, "0807/0807.1452_arXiv.txt": { "abstract": "{The precision of radial velocity (RV) measurements to detect indirectly planetary companions of nearby stars has improved to enable the discovery of extrasolar planets in the Neptune and Super-Earth mass range. Detections of extremely low mass planets, even as small as 1 Earth mass or below, in short-period orbits now appears conceivable in ongoing RV planet searches. Discoveries of these Earth-like planets by means of ground-based RV programs will help to determine the parameter $\\eta_{\\oplus}$, the frequency of potentially habitable planets around other stars.} {In search of low-mass planetary companions we monitored Proxima Centauri (M5V) as part of our M dwarf program. In the absence of a significant detection, we use these data to demonstrate the general capability of the RV method in finding terrestrial planets. For late M dwarfs the classic liquid surface water habitable zone (HZ) is located close to the star, in which circumstances the RV method is most effective. We want to demonstrate that late M dwarfs are ideal targets for the search of terrestrial planets with the RV technique.} {Using the iodine cell technique we obtained differential RV measurements of Proxima Cen over a time span of 7 years with the UVES spectrograph at the ESO VLT. We determine upper limits to the masses of companions in circular orbits by means of numerical simulations.} {The RV data of Proxima Cen have a total rms scatter of $3.1~{\\rm m\\,s}^{-1}$ and a period search does not reveal any significant signals. In contrast to our earlier results for Barnard's star, the RV results for the active M dwarf Proxima Cen are only weakly correlated with H$_{\\alpha}$ line index measurements. As a result of our companion limit calculations, we find that we successfully recover all test signals with RV amplitudes corresponding to planets with $m \\sin i \\ge 2 - 3~M_\\oplus$ residing inside the HZ of Proxima Cen with a statistical significance of $>99\\%$. Over the same period range, we can recover 50\\% of the test planets with masses of $m \\sin i \\ge 1.5 - 2.5~M_\\oplus$. Based on our simulations, we exclude the presence of any planet in a circular orbit with $m \\sin i \\ge 1~M_{\\rm Neptune}$ at separations of $a \\le 1$~AU.} {} ", "introduction": "Over the next decades we will be able to derive a first estimate of the frequency of stars with a potentially habitable Earth-like planet. This frequency is usually denoted by the parameter $\\eta_{\\oplus}$. CoRoT and Kepler are two space mission that have the capability to detect ``Super-Earths'' and even Earth analogs in short-period orbits (CoRoT) and at 1~AU (Kepler) using the transit method. The astrometry mission SIM Planetquest will achieve the sensitivity to detect Earth-like planets around a sample of nearby stars. Also, ground-based Doppler measurements have attained precision levels that make discoveries of planets with a few Earth masses possible (e.g. Lovis et al.~2006). Already the very first extrasolar planets found around the pulsar PSR 1257+12 (Wolszczan \\& Frail~1992), have such small masses, that they qualify as terrestrial planets. However, their formation is likely to have followed a different path than what is currently envisioned for the formation of terrestrial planets around main sequence stars. Besides its obvious astrobiological implications, and its crucial role for the design and overall costs of any future TPF/Darwin-type mission, the value of $\\eta_{\\oplus}$ will also impact our understanding of the formation of terrestrial planets in general (similar to the effect the extrasolar giant planets have on giant planet formation models.) M dwarfs comprise the majority of stars in the solar neighborhood (e.g. Reid et al.~2004) and once their intrinsic faintness is overcome, they represent attractive targets for high precision Doppler surveys. Due to their lower masses the reflex motion of a planet of a given mass is higher than for a solar mass star. Thus, it is no surprise that so far the lowest mass extrasolar planets detected by the Doppler method are all orbiting M dwarfs (Rivera et al.~2005; Udry et al.~2007). The micro-lensing event reported by Beaulieu et al.~(2006) is also attributed to lensing by a very low-mass planetary companion to, most likely, an M dwarf host. In this paper we present 7 years of high precision RV data for our closest neighbor in space: the M5V star Proxima Centauri. We demonstrate that with the data in hand we could have already detected planets with minimum masses as small as $1 - 2~{\\rm M}_{\\oplus}$. Constraints for giant planetary companions to Proxima Cen using HST Fine Guidance Sensor astrometry were reported by Benedict et al.~(1999) and the companion limits based on less precise RV data from the ESO Coud\\'e Echelle Spectrometer planet search were presented by K\\\"urster et al.~(1999). The combination of both studies already excluded all companions with (minimum) masses higher than $0.8~{\\rm M}_{\\rm Jupiter}$ for the period range $1$ to $600$ days. ", "conclusions": "\\begin{enumerate} \\item We present 7 years of high precision RV data for our closest neighbor in space, the M5V star Proxima Cen, obtained with UVES + I$_2$ cell at the ESO VLT/UT2. We detect no significant periodicities (except close to the 1-yr peak in the window function) that can be attributed to orbiting companions. \\item Using the same set of spectra we measure an H$_{\\alpha}$ line index to estimate the magnetic activity level of Proxima Cen. These line indices show a large amount of scatter due to flaring activity and are only weakly correlated with the RV results. \\item Based on numerical simulations we demonstrate that we could have already detected all planets with $m \\sin i = 2 - 3~{\\rm M}_\\oplus$ on circular orbits inside the classic habitable zone of Proxima Cen (assuming a stellar mass of $0.12~{\\rm M}_{\\oplus}$). \\end{enumerate}" }, "0807/0807.3985_arXiv.txt": { "abstract": "We present evidence for anomalous microwave emission in the RCW175 \\hii~region. Motivated by 33~GHz $13\\arcmin$ resolution data from the Very Small Array (VSA), we observed RCW175 at 31~GHz with the Cosmic Background Imager (CBI) at a resolution of $4\\arcmin$. The region consists of two distinct components, G29.0-0.6 and G29.1-0.7, which are detected at high signal-to-noise ratio. The integrated flux density is $5.97\\pm0.30$~Jy at 31~GHz, in good agreement with the VSA. The 31~GHz flux density is $3.28\\pm0.38$~Jy ($8.6\\sigma$) above the expected value from optically thin free-free emission based on lower frequency radio data and thermal dust constrained by IRAS and WMAP data. Conventional emission mechanisms such as optically thick emission from ultracompact \\hii~regions cannot easily account for this excess. We interpret the excess as evidence for electric dipole emission from small spinning dust grains, which does provide an adequate fit to the data. ", "introduction": "In recent years there has been mounting observational evidence for a new diffuse component emitting at frequencies $\\approx 10-60$~GHz. The anomalous microwave emission was first detected at 14 and 32~GHz by \\cite{Leitch97}. Since then, a similar picture has emerged both at high latitudes \\citep{Banday03,deOliveira-Costa04,Fernandez-Cerezo06,Hildebrandt07,Bonaldi07} and from individual Galactic sources \\citep{Casassus04,Casassus06,Casassus07,Watson05,Scaife07,Dickinson07}, although negative detections have also been reported \\citep{Dickinson06,Scaife08}. The spectral index between 20 and 40~GHz is $\\alpha \\approx -1.1$ \\citep{Davies06} with some evidence of flattening at $\\sim10-15$~GHz \\citep{Leitch97,deOliveira-Costa04,Hildebrandt07}. The emission appears to be very closely correlated with far-IR data suggesting a dust origin. Various emission mechanisms have been suggested, including hot ($T\\sim10^{6}$~K) free-free \\citep{Leitch97}, flat spectrum synchrotron \\citep{Bennett03b}, spinning dust \\citep{Draine98a,Draine98b} and magnetic dust \\citep{Draine99}. The overall picture is still very unclear and new data covering the range $10-60$~GHz are urgently needed. RCW175 \\citep{Rodgers60} is a diffuse \\hii~region, which consists of a ``medium brightness'' optical filament (G29.1-0.7, S65) $\\sim 7\\arcmin \\times 5\\arcmin$ in extent, and a nearby compact source (G29.0-0.6), which is heavily obscured by dust. Although the filament is clearly seen in high resolution data, the compact counterpart is considerably brighter. The ionization is thought to be provided by a single B-1 II type star, which forms part of a 5-star cluster \\citep{Forbes89,Sharpless59} at a distance of 3.6~kpc. Observations made with the Very Small Array (VSA) at 33~GHz \\citep{Watson03,Dickinson04} as part of a Galactic plane survey (Todorovi\\'c et al., in prep.) indicate that RCW175 is anomalously bright by a factor of $\\approx 2$, when compared with lower frequency data. In this {\\it Letter}, we present accurate Cosmic Background Imager (CBI) $31$~GHz observations of RCW175 and make a comparison with ancillary radio/FIR data. We find that the emission at $31$~GHz is significantly above what is expected from a simple model of free-free and vibrational dust emissions. ", "conclusions": "\\label{sec:conclusions} Using data from the VSA and CBI, we have observed excess emission at $\\sim 30$~GHz from the \\hii~region RCW175. The flux density spectrum indicates that about half of the flux in this region is from optically thin free-free emission leaving about half unaccounted for. An upper limit at 94~GHz from WMAP data constrains the contribution of thermal dust emission. We have discarded optically thick free-free emission from UC~\\hii~regions and GPS sources as possible candidates for this excess; an upper limit at 94~GHz and high resolution radio data rule these out as the primary contributor at $\\sim 30$~GHz. We interpret the excess as electric dipole radiation from small rapidly spinning dust grains as predicted by \\cite{Draine98b}. These models provide a reasonable fit to the data that is consistent both in terms of spectral shape and emissivity. High resolution, multi-frequency data in the range $10-100$~GHz are needed to confirm this result and to investigate the nature of anomalous emission." }, "0807/0807.3382_arXiv.txt": { "abstract": "In this letter, we determine the $\\kappa$-distribution function for a gas in the presence of an external field of force described by a potential U(${\\bf r}$). In the case of a dilute gas, we show that the $\\kappa$-power law distribution including the potential energy factor term can rigorously be deduced in the framework of kinetic theory with basis on the Vlasov equation. Such a result is significant as a preliminary to the discussion on the role of long range interactions in the Kaniadakis thermostatistics and the underlying kinetic theory. ", "introduction": "In this section we discuss briefly the standard case, i.e., the kinetic description of a classical gas under stationary conditions and immersed in a conservative force field, ${\\bf F}=-\\nabla U(r)$. Typical examples are a gas in the earth's gravitational field and ions in an external magnetic field \\cite{Huang,KT}. This kind of problem is important on their own because it permits to understand how the molecular motion is modified by force-fields different from those exerted by the containing vessel or even by the other particles of the gas. As widely known, its distribution function differs from the Maxwellian velocity statistics trough an extra exponential factor involving the potential energy whose general form reads \\begin{equation}\\label{e1} f({\\bf r},v)=n_0\\left({m\\over 2\\pi k_B T} \\right)^{3/2}\\exp \\left(-{{1\\over 2}m v^2 +U({\\bf r})\\over k_B T}\\right), \\end{equation} where $m$ is the mass of the particles, $T$ is the temperature and $n_0$ is the particle number density in the absence of the external force field. In addition, since this distribution function is normalized, it is easy to see that the number density is given by \\begin{equation}\\label{e2} n({\\bf r})=n_0 \\exp \\left[-{{U({\\bf r}) \\over k_B T}}\\right], \\end{equation} where the factor, $\\exp[-{U({\\bf r})/k_B T}]$, which is responsible for the inhomogeneity of $f({\\bf r},v)$, is usually called the Boltzmann factor. Expression ($\\ref{e1}$) follows naturally from an integration of the collisionless Boltzmann's equation \\begin{equation}\\label{e3} \\frac{\\partial{\\it f}}{\\partial{\\it t}} + {\\bf v}\\cdot\\frac{\\partial{\\it f}}{\\partial{\\bf r}} + \\frac{\\bf F}{m}\\cdot\\frac{\\partial{\\it f}}{\\partial{\\bf v}} = 0, \\end{equation} when stationary conditions are adopted along with the assumption that the total distribution can be factored \\begin{equation}\\label{e4} f({\\bf r},v)=f_0(v)\\chi({\\bf r}), \\end{equation} where $f_0(v)$ represents the Maxwell equilibrium distribution function, and $\\chi({\\bf r})$ is a scalar function of ${\\bf r}$. As one may show, after a simple normalization, the resulting expression for $\\chi({\\bf r})$ is exactly the Boltzmann factor for the potential energy of the external field, namely: \\begin{equation}\\label{e2a} \\chi({\\bf r})=\\exp \\left[-{{U({\\bf r}) \\over k_B T}}\\right], \\end{equation} and combining this result with equation (\\ref{e4}) the Boltzmann stationary distribution (\\ref{e1}) is readily obtained. ", "conclusions": "In the last few years, several applications of the $\\kappa$-power law velocity distribution have been done in many disparate branches of physics [8-21]. However, such investigations are usually related with the $\\kappa$-velocity distribution function as given by equation (6). On the other hand, many physical systems involve naturally the presence of a conservative force field as happens for example with ions in a magnetic field. Probably, the most popular problem of a gas in a force-field is the planetary atmosphere. In the standard simplified treatment, the temperature is uniform and the tree-dimensional motion occurs under the action of a constant gravitational field along the $z$-direction. To all this sort of problems, the extended $\\kappa$-distribution deduced here with basis on the Vlasov equation, namely \\begin{eqnarray*}\\label{eq18} f_{\\kappa}({\\bf r},v) = A_\\kappa\\left[\\sqrt{1 +\\kappa^2\\left(-\\frac{m{\\bf v}^2}{2k_BT} - \\frac{U({\\bf r})}{k_BT}\\right)^2} + \\kappa \\left(-\\frac{m{\\bf v}^2}{2k_BT} - \\frac{U({\\bf r})}{k_BT}\\right) \\right]^{1/\\kappa}, \\end{eqnarray*} can be applied, and, might prove to be of extreme wide usefulness. Note also that a giroscopic term may also be added to the above power law distribution. In a rotating frame with constant angular velocity, the whole effect is just to add a Coriolis term $-1/2m\\omega^{2}R^{2}$ to the potential energy, where $R$ is the perpendicular distance from the axis of rotation. In the Newtonian framework, such a term simulates a change in the potential energy due to gravity. Finally, it is worth mentioning that the present consistency between Vlasov equation and Kaniadakis thermostatistics also is valid in the context of Tsallis nonextensive framework \\cite{1}. \\vspace{0.5cm} \\noindent {\\bf Acknowledgments:} The authors are partially supported by the Conselho Nacional de Desenvolvimento Cient\\'{\\i}fico e Tecnol\\'ogico (CNPq - Brazil). JASL is also grateful to FAPESP No. 04/13668-0." }, "0807/0807.1722_arXiv.txt": { "abstract": "% We study the stability and the modes of non -- isothermal coronal loop models with different intensity values of the equilibrium twisted magnetic field.We use an energy principle obtained via non -- equilibrium thermodynamic arguments. The principle is expressed in terms of Hermitian operators and allows to consider together the coupled system of equations: the balance of energy equation and the equation of motion, to obtain modes and eigenmodes in a spectrum ranging from short to long--wavelength disturbances without having to use weak varying approximations of the equilibrium parameters. Long--wavelength perturbations introduce additional difficulties because the inhomogeneous nature of the medium determines disturbances leading to continuous intervals of eigenfrequencies which cannot be considered as purely sinusoidal.We analyze the modification of periods, modes structure and stability when the helicity, the magnetic field strength and the radius of the fluxtube are varied. The efficiency of the damping due to the resonant absorption mechanism is analyzed in a context of modes that can either impulsively release or storage magnetic energy.We find that the onset of the instability is associated to a critical value of the helicity and that the magnetic energy content has a determinant role on the instability of the system with respect to the stabilizing effect of the resonant absorption mechanism. ", "introduction": "\\subsection{Variational principle} A crucial requirement for any theoretical model of coronal structures is to give account of the stability and evolution of far--from--equilibrium states which are responsible of the characteristic rich topology and dynamics of the solar corona. This implies to consider the coupling of thermal and mechanical equations. Different stability analysis of solar structures can be found in the literature, generally restricted to special types of perturbations and specific equilibrium models. These includes, models that consider adiabatic configuration such as the ones analyzed via the classical criterion of \\citet{ber} or those that presuppose static equilibrium and analyze thermal stability. In the application of Bernstein's criterion, the adiabatic assumption implies that the energy balance equation is not required and thus dissipation is impossible. Also the assumption of static models is a strong, and often unjustified, restriction for open systems. In this paper we apply an energy principle to analyze the stability of solar coronal loops when helical modes are present. The principle was obtained in previous papers (Paper I: \\citep{cos1}; \\citep{cos2}; see also \\citep{us0}) using a general procedure of irreversible thermodynamics -based on firmly established thermodynamic laws- that can be understood as an extension of Bernstein's MHD principle to situations far from thermodynamic equilibrium. In Paper I and in \\citep{cos2} we showed how to obtain the variational principle for solar coronal structures from the equations that describe the dynamics of the system. The method consists of obtaining a Lyapunov function, also known as generalized potential, that represents the mathematical expression of the stability conditions. The principle is subject to physically reasonable requirements of hermiticity and antihermiticity over the matrices. For a more detailed presentation see Paper I and the references therein. \\subsection{Solar coronal loops} MHD loop oscillations in the corona are known to be strongly damped, mostly having decaying times of few periods $N_{p}\\approx 2 - 7 \\ periods.$ While thermal conduction, with the contribution of radiative cooling mechanisms, could be the main cause of the damping of pure MHD slow magnetoacoustic mode oscillations they are unimportant for the MHD fast modes. Resonant absorption and phase mixing seem more promising in giving account of the rapid decay (\\citep{gh94}; hereafter HG, \\citep{go02}) of the ideal fast oscillations of these strongly inhomogeneous and structured plasma systems. Inhomogeneous equilibrium distributions of plasma density and temperature varying continuously across the magnetic field led to plasma waves with continuous intervals of eigenfrequencies. The occurrence of the Alfv\\'en ideal MHD continuum in a thin edge layer is derived from the highly anisotropic character of the fast magnetoacoustic waves giving rise to a peak of the amplitudes where the perturbation develops large gradients and the absorption has maxims. However, there is another type of continuum commonly known as slow magnetosonic continuum associated to the inhomogeneity of the equilibrium parameters along the axis of the loop (see Paper I). This inhomogeneities are associated, for example, to changes in the % density concentration at the loop basis. If the magnetic field is twisted the inhomogeneities led to the coupling of Alfv\\'en and slow magnetosonic continuum modes (\\citep{bel}). The resonant absorption mechanism of wave heating consists on the non--dissipative transference of wave energy from the collective line-tied wave with fast discrete eigenvalues (kinetic energy of the fast radial component) to a local resonant mode in the Alfv\\'en continuum, (kinetic energy of the azimuthal component), which is then dissipated in an enhanced manner. Then, the continuum oscillations are converted into heat by dissipative processes; as the medium has large gradients in the Alfv\\'en speed, the oscillations of neighboring field lines become out of phase and shear Alfv\\'en waves lead to enhanced viscous and ohmic dissipation (see \\citep{pr83} for the linear regime and \\citep{nak97} for the nonlinear one). The mode conversion from the collective to the local mode occurs in a time that is non--dissipative and generally much shorter than the second time scale which is related to the dissipative damping of the small--scale perturbations of the local mode in the resonance layer (\\citep{rob00}; \\citep{van04}). The whole temporal pattern description of modes that exhibit a combination of global (discrete line--tied fast eigenmode) and localize (Alfv\\'en continuum mode) behaviour is known as quasi--mode. Moreover, the mixed nature of the modes is not only due to the temporal behaviour but also to the boundary value problem giving rise to a spatial behaviour which is also of a mixed nature, i.e. coronal loops with line--tying constraints cannot support pure waves: Alfv\\'en, slow or fast magnetoacoustic modes. HG studied the mixed spectral description of coronal loops (i.e. the resulting superposition of basic waves which adjust the line--tied condition) without assuming a straight magnetic field and forcing the loop to follow the photospheric velocity perturbations. They found that pure Alfv\\'en and pure slow modes are obtained as singular limiting cases of cluster spectra of Alfv\\'en--fast or slow--fast modes, where the fast components are localized in a photospheric boundary associated to the line--tied condition: the coronal part of the loop acting as a resonant cavity of large Alfv\\'en components and fast components, with a small but rapidly varying amplitude, located in the photospheric boundary layer. They found that heating of coronal loops by resonant absorption is due to the line--tied Alfv\\'en continuum which no longer depends on the poloidal magnetic field and that the corresponding eigenmodes have a global ballooning feature which is characterized by an accumulation point given by the Alfv\\'en frequency. In \\citep{gh93} (hereafter GH), a variational principle, based in Bernstein's principle, was obtained to derive the Alfv\\'en and slow continuum frequencies in a line--tied inhomogeneous cylinder. Stability considerations led them to conclude the global stability of coronal loops. In this paper, following results of Paper I we apply our energy principle to consider the stability and mode structure of loop inhomogeneous coronal models with non--vanishing helicity. Our principle has the advantages that it does not require a WKB approximation and that, as was mentioned, it allows the consideration of the coupling of the thermal and mechanical equations that are necessary to analyze far from equilibrium states. ", "conclusions": "Convective motion of the photosphere is believed to provide the energy that is storage in twisted magnetic coronal fields allowing the presence of long--lived coronal structures until it is released by instabilities (\\citep{raa}; \\citep{vrs}). On the other hand, continuous spectra are generally associated to stability. An accepted conjecture establishes that unstable modes have a discrete spectrum (see \\citep{fre} or \\citep{pri}). There are two types of possible continuous spectra in this problem. The inhomogeneous character of the equilibrium parameters along the loop axis can lead to a continuum that couples to the Alfv\\'en continuum \\citep{bel}; e.g, when the disturbances considered are comparable to the inhomogeneous characteristic wavelength stable eigenvalues can give rise to a continuous spectrum ($L/2$, the equilibrium structure in the $z$ component). This is the case studied in Paper I. On the other hand, GH established, for non vanishing helicity systems, that there is a continuous spectrum associated with the line-tied Alfv\\'en resonance leading to the damping and heating by the resonant absorption mechanism and thus, directly relate to the stability of loops. They also pointed out how to obtain the resonant singular limit $\\omega$, from the class of physically permissible solutions, \\begin{equation} \\omega(r)=\\frac{nB_{z}(r)}{\\int_{-L}^{L}\\sqrt{\\rho(z)}dz}. \\label{11} \\end{equation} This resonance results because of the absence of an explicit dependence on the azimuthal magnetic field component ($B_{\\varphi}$). Thus, in order to understand in which conditions which mechanism can dominate and give account of the different scenarios i.e., the driving of the instability or the damping of mode oscillations, it is critical to gain knowledge about the dynamics and energetic contribution of twisted structures. Yet, the implications of the twisting in theoretical and observational descriptions are poorly known; e.g., there is no clarity about the modification of the dispersion relation and observational data are indirectly inferred. In this paper we focused our attention to describe the changes in periods, stability and mode structure of coronal loops when the helicity, the magnetic field intensity and the radius are varied. For loops with vanishing helicity it is well established that the Alfv\\'en line--tied resonance continuum is responsible of the damping of kink ($m=1$) quasi--modes via the transfer of energy from the radial component into the azimuthal one, i.e., from discrete global modes into the local continuum modes where phase mixing can take place. Still, the twisting of the magnetic field leads to the coupling of MHD cylindrical modes making difficult to provide a classification in terms of the behavior of pure--like modes. In order to calculate modes and frequencies we followed the schematic procedure described in Paper I and in \\citep{gal1}. We used a symbolic manipulation program to integrate the equations. $\\delta^{2}W_{p}$ and the perturbations were expanded in a six dimensional--Fourier basis on the independent coordinate $z$ that adjusts to border conditions, i.e., the four perturbated components (eq.~\\ref{9}) were expanded in a six mode basis to obtain $24$ eigenvalues and eigenvectors for each of the helicity and the magnetic field values. Only the first eighteen eigenvalues were considered (the others are more than two order of magnitude smaller and accumulate at zero; the eigenvectors are also vanishingly small). Thus, a quadratic form for $\\delta^{2}W_{p}$ was obtained and was minimized with the Ritz variational procedure. A matrix discrete eigenvalue problem subject to a normalization constraint was obtained. From the resulting modes and the generalized potential energy (eq.~\\ref{2}): $\\delta^{2}W_{p}\\geq0$ the stability of each mode was determined. The coronal loop parameters used were: $L=10^{10}cm$ (or $L=100Mm$), $T_{b}=10^{4}K$ $T_{t}=10^{6}K$ $n_{e}=10^{8}cm^{-3}$ electron number density $p_{t} =2k_{B}T_{t} \\;$; $\\rho_{t}=m_{p}p_{t}/k_{B}T_{t}$. Frequencies and modes were calculated for two different values of the magnetic field: $B_{0}=10G$ and $B_{0}=100G$, and for three different values of the helicity $b=(3.1 \\ 10^{-8} ; \\ 3.1 \\ 10^{-7}; \\ 1.9 \\ 10^{-6})$ which correspond to the adimentional values: $b_{a}=(2.8; \\ 28; \\ 170)$ with $N_{t}=(0.45; \\ 4.3; \\ 13.7)$, $N_{t}:$ the number of turns over the cylinder length. These helicity values defined as weak, moderated and strong helicity respectively correspond to the classification given in \\citep{asc4} (Chapter 5). The adimentional radius was initially chosen as $R=0.01$. In what follows we summarize the conclusions obtained from the data analysis which are displayed in three tables. Table 1 shows the periods (in minutes) for weak, moderate and strong helicity for two values of the magnetic field intensity ($B_{0}=10G$ and $B_{0}=100G$ (left and right panel respectively). S and U letters indicate the stable--unstable character of the modes. From the table we see that: \\noindent I) Weak helicity modes are stable. This is in accordance with the analytic results by \\citet{rud} who studied nonaxisymmetric oscillations of a thin twisted magnetic tube with fixed ends in a zero-beta plasma. \\noindent II) Higher modes have an accumulation point at zero, indicating the presence of a continuum spectra of stable modes (as in Paper I). Note that, calculus performed via discrete basis, as in our case, give spectra that are necessarily discrete. Thus, an accumulation of discrete eigenvalues suggests a stable continuum spectrum. \\noindent III) The $B_{0}=10G$ case has larger periods than the $B_{0}=100G$ one. For moderate and strong helicity the eigenvalues of the first panel follow a scaling law with that of the second one i.e., they scale with the magnetic field intensity exactly as the Alfv\\'en speed does $P_{10G}\\simeq 10 P_{100G}$. \\noindent IV) As HG and GH, we note a clustering of the spectra associated to the change from real to imaginary eigenvalues (and viceversa). There is a pronounced change (in the spacing of the periods or/and in the stability) from the sixth mode to the seventh mode. This is noted by a double line in Table 1 and related to the importance of the parallel component with respect to the perpendicular component (see Table 2). Up to period number ten real -- imaginary eigenvalues of the first panel ($B_{0}=10G$) correspond to real--imaginary ones of the second panel ($B_{0}=100G$). Also, excepting large order periods $n>10$, when the helicity is increased from weak to moderate the imaginary stable eigenvalues turn to imaginary unstable ones. For $B_{0}=10G$ and weak and moderate helicity cases there are five different groups of periods ($P_{1}-P_{6};P_{7}-P_{10};P_{11}-P_{12};P_{13}-P_{14};P_{15}- P_{18}$) (see also Table 2). The clustering is more difficult to establish i.e., the differences are less pronounced, with increasing magnetic field intensity and larger order periods. \\noindent In order to compare our results with those given by these authors we calculated the expression eq.~\\ref{11} for our modes. We found that, all periods excepting $P_{1}-P_{6}$ weak helicity modes satisfy the relation and thus, they belong to the Alfv\\'en continuum spectrum justifying the scaling law described in III. As HG, we conclude that the change in the real--complex character of the $P_{6}-P_{7}$ eigenvalues is associated to the existence of an accumulation point of the resonant Alfv\\'en continuum, however we find that this change is not necessarily related to a change in the stability as they claimed. Note that all modes with weak helicity are stable (even the imaginary ones); in all the other cases the imaginary character of the eigenvalues is associated to instability. Yet, the continuum stable eigenvalue conjecture is here still valid \\citep{fre}, \\citep{pri}; it applies to a spectrum with an accumulation point in zero; we found stable modes for all the helicity values and for the two magnetic field values with $P_{n>14}$. Note that the analysis of stable modes is still of interest because depending on the relative characteristic times of stable and unstable modes the stable ones could be active and accessible to observations. \\noindent The presence of at least one unstable mode means that the equilibrium state is unstable. Thus, taking into account the whole range of stable modes, we confirm previous results leading to conclude that field configurations with some degree of twisting give a stabilizing effect allowing the storage of magnetic energy \\citep{raa2}, i.e., when the helicity is augmented the stable weak case turns to an unstable one suggesting a critical value. \\bigskip In Paper I, we obtained only one unstable mode classified as slow magnetoacoustic mode due to the almost longitudinal character (parallel to the magnetic field) of the wavevector perturbation and to the fact that the period did not changed with the intensity of the magnetic field, resembling acoustic waves with sound speeds, $v_{s}$, independent of the magnetic field. The characteristic unstable time obtained in Paper I was $\\tau_{u}=36 \\ min$, corresponding to a typical slow magnetoacoustic fundamental period with a characteristic wavelength of the order of the loop length $L/2$. Also, we obtained a continuous set of stable modes classified as fast magnetoacoustic modes due to their large value component orthogonal to the magnetic field and to the fact that the eigenvalues scale with the intensity of the magnetic field as $P_{11G}\\simeq 10 P_{100G}$; thus resembling the dependence of the Alfv\\'en waves $v_{A} \\sim B_{0}$. Table 2 (First Panel) displays the resulting features associated to the relative intensity of the parallel and perpendicular to the field components ($(\\xi_{\\parallel},\\xi_{\\perp},\\xi_{r}) $ is an orthogonal basis) and their classification as slow--like (S) or fast--like (F). The relative phase between the components is also indicated in the table by P (in phase) and IP (inverted phase). Table 2 (Second Panel) also shows the intensity relationship between the cylindrical components. In order to classify the modes and to compare with the slow and fast magnetoacoustic modes obtained in Paper I, we calculated the cylindrical mode components and also the tangential and normal to the field components ($\\xi_{\\parallel}= (Rb\\xi_{\\phi}+\\xi_{z})/\\Delta; \\xi_{\\perp}= (\\xi_{\\phi}-Rb\\xi_{z})/\\Delta$). Our interest in the $\\xi_{\\parallel}$, $\\xi_{z}$, $\\xi_{r}$, $\\xi_{\\perp}$ and $\\xi_{\\phi}$ components resides in that: First, when the helicity is weak, the $\\xi_{\\parallel}$ component is expected to play the slow-mode role of $\\xi_{z}$ in Paper I. Second, the $\\xi_{r}$ component is related to the fast modes and determines the resonant absorption mechanism when uniform cylindrical flux tubes are considered by the transferring of energy to the $\\xi_{\\phi}$ component. When helicity and inhomogeneous distribution of equilibrium parameters are present it is worth investigating the transferring of energy from the $\\xi_{r}$ component to the others. In this case the resonant damping of global oscillations will occur by conversion of kinetic energy of the radial component into kinetic energy of the $\\xi_{\\parallel}$ and $\\xi_{\\perp}$ components; both components forming the plane orthogonal to $\\xi_{r}$, and equal to the plane formed by $\\xi_{\\phi}$ and $\\xi_{z}$. From the analysis of the amplitude of the components of the $P_{1}-P_{6}$ modes with respect to the $P_{7}-P_{18}$ ones in the weak helicity case i.e., real and imaginary eigenvector respectively, we could classify the first ones as slow-like modes because: I) their tangential components $\\xi_{\\parallel}$ are at least an order of magnitude larger than the normal ones $\\xi_{\\perp}$; II) as the helicity is weak $\\xi_{\\parallel}\\approx \\xi_{z}$ and $\\xi_{r}\\rightarrow 0; \\xi_{\\phi}\\rightarrow 0$, the wavevector is almost tangential to the magnetic field; III) they have a larger characteristic time and a shorter characteristic speed than the imaginary eigenvectors. On the contrary, imaginary eigenvalues are associated to large values of the $\\xi_{r}$ component and $\\xi_{\\perp}$ component (due to large values of $\\xi_{\\phi}$ (see Table 2 Second Panel)) , and small values of the $\\xi_{\\parallel}$ and $\\xi_{z}$ components. As in Paper I, when the eigenvalues change form real to imaginary the period strongly diminishes and a change in the type of mode from the slow to fast magnetoacoustic type occurs. In opposition to Paper I where the acoustic mode has the same eigenvalue for both magnetic field intensities, here the modes are affected by the strengthening of the magnetic field leading to an-order-of-magnitude shorter period than in the non--helicity case. The $\\xi_{\\parallel}$ and $\\xi_{\\perp}$ components are in an inverted phase for real eigenvector modes and in phase for imaginary eigenvector modes. For moderate helicity the overall description is similar but all the cases having non vanishing $\\xi_{\\phi}$ component and all the periods in the resonant line-tied continuum. As was mentioned, real--imaginary eigenvalues correspond to stable--unstable behavior. In the strong helicity case, as the weak and moderate ones, we note for $P_{1}-P_{6}$ larger, but comparable, values of the $\\xi_{\\parallel}$ component with respect to the $\\xi_{\\perp}$ component. In this case the two components of the mode are in phase. This relationship between the $\\xi_{\\parallel}$ and $\\xi_{\\perp}$ components of Table 2 (FP), and their associated phases is found again in the modes with $P_{15} - P_{18}$. In spite that these features are associated to the slow magnetoacoustic characterization, Table 2 (SP) shows that as $\\xi_{z}$ is vanishingly small, the strong helicity case cannot be classify as a slow mode. \\bigskip When helicity is present the mixed character of the modes manifests itself making difficult to identify the components that are involved in the damping mechanism. However, taking into account the resonant frequency of eq.~\\ref{11}, we noted that (HG) all the modes, except those with $P_{1}-P_{6}$ periods of the weak helicity case, have resonant frequencies suggesting that resonant absorption in helical modes is associated to modes with significant values of $\\xi_{\\perp}$ component. If this argument is correct we can affirm that the damping mechanism of body helical modes is associated to the transfer of kinetic energy of the radial component into kinetic energy of the $\\xi_{\\perp}$ component which is not only related to the $\\xi_{\\phi}$ cylindrical contribution but also to the $\\xi_{z}$ one by the expression $ \\xi_{\\perp}=(\\xi_{\\phi}-Rb\\xi_{z}) \\Delta$. \\bigskip We also analyzed the change of the period as a function of the radius for different values of the helicity. We found that, for weak helicity, the increasing of the radius leads to a decrease of the periods. This is in accordance with observations, e.g., observed sausage modes are associated with thicker and denser loop structures and lower periods; while in other case (unstable cases) the increasing of the radius leads to an increase of the period. \\bigskip Table 3 -First and Second Panel- shows the variation of the radius $R$ with the twist $bR$ for weak and moderate helicity respectively. \\citep{rud} has conjectured that the line-tying condition at the tube ends should stabilize the tube and has suggested a critical value ($\\sim L b0$ to $\\delta^{2} W_{p}<0$. Figure~\\ref{fig:tres}a and Figure~\\ref{fig:tres}c display the total energy composed by the compressional, radiative, thermal and magnetic energy contributions of $P_{6}$ mode in the weak and moderate case respectively. The same features but for the $P_{7}$ mode are shown in Figure~\\ref{fig:tres}e and Figure~\\ref{fig:tres}g. Figure~\\ref{fig:tres}b and Figure~\\ref{fig:tres}d show the magnetic energy content alone for $P_{6}$ mode in the weak and moderate case respectively. Figure~\\ref{fig:tres}f and Figure~\\ref{fig:tres}h show the magnetic energy content for $P_{7}$ mode and for the weak and moderate case respectively. It can be seen, in this and in all the other cases, that the magnetic energy content has a determinant role on the stability--instability of the system, i.e., the stability changes when the magnetic generalized potential energy changes sign. Thus, a result of this analysis is that the stability of twisted coronal loops is fundamentally determined by the storing of magnetic energy, being the other contributions less significant. Meanwhile, when the helicity is weak or vanishingly small and the magnetic contribution has a stabilizing effect the other non-dominant contributions, as the non-adiabatic ones, can play an important role. This makes possible, for example, the damping of fast excitations due to resonant absorption. Yet, even when one of these contributions is unstable, stable modes could be active for a while if their characteristic periods are shorter than the characteristic time of the instability. This is the case of Paper I, where we obtained a slow mode with an unstable characteristic times of $\\tau\\sim 36 \\ min$ coexisting with stable fast modes with periods about $P\\sim 1 \\ min$; moreover, we showed that the instability can be nonlinearly saturated giving rise to a limit-cycle solutions, i.e., an oscillation between parallel plasma kinetic energy and plasma internal energy where the magnetic energy plays no relevant role. Thus, the contribution to the stable--unstable character of the modes is mostly due to the magnetic energy content and not to other energetic contributions. Note that as the balance energy equation takes into account non--adiabatic contributions, i.e., radiation, heat flow and heat function (with $L=0$ at the equilibrium), the resulting perturbations are not constrained to the force--free condition. So, one result of the analysis is that the pertubation energy contribution is mainly due to magnetic forces. Thus, for these type of twisted magnetic field models, non--adiabatic perturbations (e.g. thermal perturbations) and resonant absorption seem unimportant to guarantee stability; a loop system with weak storage of magnetic energy (low values of the helicity) could be released if the helicity is suddenly increased, e.g., by footpoint motions. Meanwhile all the \"zoo\" of the coronal seismology can be active and accessible to observations. \\begin{figure*} \\includegraphics[width=4.3cm]{P6ewp.eps} \\includegraphics[width=4.3cm]{P6emwp.eps} \\includegraphics[width=4.3cm]{P6emp.eps} \\includegraphics[width=4.3cm]{P6emmp.eps} \\includegraphics[width=4.3cm]{P7ewp.eps} \\includegraphics[width=4.3cm]{P7emwp.eps} \\includegraphics[width=4.3cm]{P7emp.eps} \\includegraphics[width=4.3cm]{P7emmp.eps} \\caption{Energy content of the sixth and seventh mode for $B_{0}=10G$. a) Total potential energy and b) magnetic potential energy respectively for the sixth mode $P_{6}=1.23 \\ min$ and for weak helicity. c) Total potential energy and d) magnetic potential energy respectively for the sixth mode $P_{6}=1.23 \\ min$ and for moderate helicity. e) Total potential energy and f) magnetic potential energy respectively for the seventh mode $P_{7}=0.07 \\ min$ and for weak helicity. g) Total potential energy and h) magnetic potential energy respectively for the sixth mode $P_{7}=0.07 \\ min$ and for moderate helicity.} \\label{fig:tres} \\end{figure*} \\begin{table*} \\begin{tabular}{cccccccc} \\hline $P_{i}$&$weak $&$moderate$&$strong$&$ \\ \\ \\ $& $weak$&$moderate$&$strong$ \\\\ \\hline $P_{1}$&$1.921 \\ S $&$0.209 \\ S$&$0.525 \\ i \\ U$&$ \\ \\ \\ $&$0.159 \\ S$&$0.021 \\ S$&$0.052 \\ i \\ U$\\\\ \\hline $P_{2}$&$1.869 \\ S $&$0.204 \\ S$&$0.450 \\ S$&$ \\ \\ \\ $& $0.158 \\ S$&$0.020 \\ S$&$0.044 \\ S$\\\\ \\hline $P_{3}$&$1.535 \\ S $&$0.169 \\ S$&$0.430 \\ S$&$ \\ \\ \\ $& $0.154 \\ S$&$0.017 \\ S$& $0.042 \\ S$ \\\\ \\hline $P_{4}$&$1.533 \\ S $&$0.168 \\ S$&$0.424 \\ i \\ U$&$ \\ \\ \\ $&$0.153 \\ S$&$0.0167 \\ S$&$0.042 \\ i \\ U$\\\\ \\hline $P_{5}$&$1.306 \\ S $&$0.143 \\ S$&$0.206 \\ i \\ U$& $ \\ \\ \\ $&$0.151 \\ S$&$0.014 \\ S$&$0.020 \\ i \\ U$\\\\ \\hline $P_{6}$&$1.228 \\ S $&$0.135 \\ S$&$0.177 \\ i \\ U$&$ \\ \\ \\ $& $0.15 \\ S $&$0.013 \\ S$&$0.017 \\ i \\ U$\\\\ \\hline \\hline $P_{7}$&$0.068 \\ i \\ S$&$0.070 \\ i \\ U$&$0.125 \\ S$&$ \\ \\ \\ $& $0.0047 \\ i \\ S$&$0.007 \\ i \\ U$&$0.0125 \\ S$\\\\ \\hline $P_{8}$&$0.064 \\ i \\ S$&$0.066 \\ i \\ U$&$0.122 \\ S$&$ \\ \\ \\ $& $0.0046 \\ i \\ S$&$0.006 \\ i \\ U$&$0.012 \\ S$\\\\ \\hline $P_{9}$&$0.042 \\ i \\ S $&$0.044 \\ i \\ U$&$0.101 \\ S$&$ \\ \\ \\ $& $0.0044 \\ i \\ S$&$0.0043 \\ i \\ U$&$0.0101 \\ S$\\\\ \\hline $P_{10}$&$0.041 \\ i \\ S$&$0.043 \\ i \\ U $&$0.100 \\ S$&$ \\ \\ \\ $& $0.0043 \\ i \\ S$&$0.0042 \\ i \\ U$&$0.01 \\ S$\\\\ \\hline $P_{11}$&$0.033 \\ S $&$0.036 \\ S$&$0.989 \\ S$&$ \\ \\ \\ $& $0.0042 \\ i \\ S$& $0.0036 \\ S$&$0.099 \\ S$\\\\ \\hline $P_{12}$&$0.032 \\ S $&$0.035 \\ S$ &$0.096 \\ S$&$ \\ \\ \\ $& $0.0041 \\ i \\ S$&$0.0035 \\ S$ &$0.0096 \\ S$\\\\ \\hline $P_{13}$&$0.030 \\ i \\ S $&$0.031 \\ i \\ U$ &$0.085 \\ S$&$ \\ \\ \\ $&$0.003 \\ S$&$0.0031 \\ i \\ U$ &$0.0086 \\ S $\\\\ \\hline $P_{14}$ & $0.027 \\ i \\ S $&$0.029 \\ i \\ U$&$0.081 \\ S $&$ \\ \\ \\ $&$0.0026 \\ S$&$0.003 \\ i \\ U$&$0.0081 \\ S $\\\\ \\hline $P_{15}$&$0.025 \\ S $&$0.027 \\ S $&$0.077 \\ S $&$ \\ \\ \\ $&$0.0025 \\ S $&$0.0027 \\ S $&$0.0077 \\ S $\\\\ \\hline $P_{16}$&$0.024 \\ S $&$0.026 \\ S $&$0.076 \\ S $&$ \\ \\ \\ $&$0.002 \\ S $&$0.003 \\ S $&$0.0076 \\ S $\\\\ \\hline $P_{17}$&$0.02 \\ S $&$0.02 \\ S $&$0.063 \\ S $&$ \\ \\ \\ $&$0.0024 \\ S $&$0.0021 \\ S $&$0.0063 \\ S $\\\\ \\hline $P_{18}$&$0.018 \\ S $&$0.02 \\ S $&$0.059 \\ S $&$ \\ \\ \\ $&$0.0024 \\ S $&$0.0025 \\ S $&$0.006 \\ S $\\\\ \\hline \\end{tabular} \\caption{\\label{tab:table1} Eighteen first periods associated to stable (S) and unstable (U) eigenvalues (minutes) for A) Left panel: $B_{0}=10G$ with A1) left column: weak helicity, A2) middle column: moderate helicity, A3) right column: strong helicity and B) Right panel: $B_{0}=100G$ with B1, B2, B3 the same as in A. Larger order modes were discarded.} \\end{table*} \\begin{table*} \\begin{tabular}{ccccccc} \\hline $P_{i}$&$weak $&$moderate$&$strong \\ \\ \\ \\ $&$weak $&$moderate$&$strong$\\\\ \\hline $P_{1}$&$\\xi_{\\parallel} \\gg \\xi_{\\perp} \\mapsto 0 \\ S; \\ IP $&$\\xi_{\\parallel} > \\xi_{\\perp} \\ S; \\ IP $&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r}\\mapsto 0 $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r} $&$ \\xi_{r} \\leq \\xi_{\\phi}; \\xi_{z} \\mapsto 0 $\\\\ \\hline $P_{2}$&$\\xi_{\\parallel} \\gg \\xi_{\\perp} \\mapsto 0 \\ S; \\ IP $&$\\xi_{\\parallel} > \\xi_{\\perp} \\ S; \\ IP $&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r}\\mapsto 0 $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r} $&$ \\xi_{r} \\leq \\xi_{\\phi}; \\xi_{z} \\mapsto 0 $\\\\ \\hline $P_{3}$&$\\xi_{\\parallel} \\gg \\xi_{\\perp} \\mapsto 0 \\ S; \\ IP $&$\\xi_{\\parallel} > \\xi_{\\perp} \\ S; \\ IP $&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r}\\mapsto 0 $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r} $&$ \\xi_{r} \\leq \\xi_{\\phi}; \\xi_{z} \\mapsto 0 \\ $\\\\ \\hline $P_{4}$&$\\xi_{\\parallel} \\gg \\xi_{\\perp} \\mapsto 0 \\ S; \\ IP $&$\\xi_{\\parallel} > \\xi_{\\perp} \\ S; \\ IP $&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r}\\mapsto 0 $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r} $&$ \\xi_{r} \\leq \\xi_{\\phi}; \\xi_{z} \\mapsto 0 $\\\\ \\hline $P_{5}$&$\\xi_{\\parallel} \\gg \\xi_{\\perp} \\mapsto 0 \\ S; \\ IP $&$\\xi_{\\parallel} > \\xi_{\\perp} \\ S; \\ IP $&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r}\\mapsto 0 $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r} $&$ \\xi_{r} \\leq \\xi_{\\phi}; \\xi_{z} \\mapsto 0 $\\\\ \\hline $P_{6}$&$\\xi_{\\parallel} \\gg \\xi_{\\perp} \\mapsto 0 \\ S; \\ IP $&$\\xi_{\\parallel} > \\xi_{\\perp} \\ S; \\ IP $&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r}\\mapsto 0 $&$ \\xi_{z}\\gg \\xi_{\\phi} \\sim \\xi_{r} $&$ \\xi_{r} \\leq \\xi_{\\phi}; \\xi_{z} \\mapsto 0 $\\\\ \\hline \\hline $P_{7}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel} \\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$ &$\\xi_{\\perp} \\geq \\xi_{\\parallel} \\ IP \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0$&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z} \\mapsto 0 $&$\\xi_{z} > \\xi_{r} > \\xi_{\\phi} $\\\\ \\hline $P_{8}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel}\\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$&$\\xi_{\\perp} \\geq \\xi_{\\parallel} \\ IP \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{z} > \\xi_{r}> \\xi_{\\phi} $ \\\\ \\hline $P_{9}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel}\\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$&$\\xi_{\\perp} \\geq \\xi_{\\parallel} \\ IP \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{z} > \\xi_{r} > \\xi_{\\phi} $\\\\ \\hline $P_{10}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel}\\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$&$\\xi_{\\perp} \\geq \\xi_{\\parallel} \\ IP \\ \\ \\ \\ $& $\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{z}> \\xi_{r} > \\xi_{\\phi} $\\\\ \\hline $P_{11}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel}\\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$&$\\xi_{\\perp} \\geq \\xi_{\\parallel} \\ IP \\ \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi} \\gg \\xi_{z}\\mapsto 0 $&$\\xi_{z} > \\xi_{r} > \\xi_{\\phi} $ \\\\ \\hline $P_{12}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel}\\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ IP \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $ &$ \\xi_{r}> \\xi_{\\phi} > \\xi_{z} $ \\\\ \\hline $P_{13}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel} \\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$&$\\xi_{\\perp} \\geq \\xi_{\\parallel} \\ IP \\ \\ \\ \\ $&$\\xi_{r}\\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{z} > \\xi_{r} > \\xi_{\\phi} $\\\\ \\hline $P_{14}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel}\\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$ &$\\xi_{\\perp} \\geq \\xi_{\\parallel} \\ IP \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{z} > \\xi_{r} > \\xi_{\\phi} $\\\\ \\hline $P_{15}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel} \\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0$&$ \\xi_{r}> \\xi_{\\phi} > \\xi_{z} $ \\\\ \\hline $P_{16}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel} \\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$ \\xi_{r}> \\xi_{\\phi} > \\xi_{z} $ \\\\ \\hline $P_{17}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel} \\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$&$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0$&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$ \\xi_{r}> \\xi_{\\phi} > \\xi_{z} $ \\\\ \\hline $P_{18}$&$\\xi_{\\perp} \\gg \\xi_{\\parallel} \\mapsto 0 \\ F; \\ P $&$\\xi_{\\perp} > \\xi_{\\parallel} \\ F; \\ P$ &$\\xi_{\\parallel} \\geq \\xi_{\\perp} \\ P \\ \\ \\ \\ $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$\\xi_{r} \\sim \\xi_{\\phi}\\gg \\xi_{z}\\mapsto 0 $&$ \\xi_{r}> \\xi_{\\phi} > \\xi_{z} $\\\\ \\hline \\end{tabular} \\caption{\\label{tab:table2} First Panel: Intensity relationship between the tangential and normal to the field components of the eighteen first periods for $B_{0}=10G$ and for weak (first column), moderate (second column) and strong helicity (third column) cases. The (P) indicates in phase and (IP) indicates inverted phase. Second Panel: Intensity relationship between the cylindrical components of the eighteen first periods for $B_{0}=10G$ and for weak (first column), moderate (second column) and strong helicity (third column) cases.} \\end{table*} \\begin{table*} \\begin{tabular}{cccccccc} \\hline $R$&$L $&$R/2L$&$Twist=bR \\ \\ \\ \\ \\ \\ \\ \\ \\ $&$R$&$L $&$R/2L$&$Twist=bR$ \\\\ \\hline $0.01$& $ 9.05 \\ 10^{7} $&$0.005 $&$ 0.028\\ \\ \\ \\ \\ \\ \\ \\ \\ $&$0.01$& $ 8.07 \\ 10^{7} $&$0.005 $&$ 0.28$\\\\ \\hline $0.02$&$ 9.04 \\ 10^{7} $&$ 0.01 $&$ 0.057 \\ \\ \\ \\ \\ \\ \\ \\ \\ $&$0.015$&$ 8.32 \\ 10^{7} $&$ 0.008 $&$ 0.43 $\\\\ \\hline $0.03$&$ 9.02 \\ 10^{7} $&$ 0.015 $&$ 0.085 \\ \\ \\ \\ \\ \\ \\ \\ \\ $&$0.02$&$ 7.86 \\ 10^{7} $&$ 0.011 $&$ 0.57$\\\\ \\hline $0.04 $&$ 8.99 \\ 10^{7} $&$ 0.02 $&$ 0.11\\ \\ \\ \\ \\ \\ \\ \\ \\ $&$0.025 $&$ 7.38 \\ 10^{7} $&$ 0.015 $&$ 0.71$\\\\ \\hline $0.05 $&$ 8.96 \\ 10^{7} $&$ 0.025 $&$ 0.14\\ \\ \\ \\ \\ \\ \\ \\ \\ $&$0.03 $&$ 6.88 \\ 10^{7} $&$ 0.02 $&$ 0.85$\\\\ \\hline $0.06 $&$ 8.9 \\ 10^{7} $&$ 0.03 $&$ 0.17\\ \\ \\ \\ \\ \\ \\ \\ \\ $&$0.04 $&$ 5.97 \\ 10^{7} $&$ 0.03 $&$ 1.13$\\\\ \\hline $0.1 $&$ 8.7 \\ 10^{7} $&$ 0.05 $&$ 0.28\\ \\ \\ \\ \\ \\ \\ \\ \\ $&$0.05 $&$ 5.2 \\ 10^{7} $&$ 0.04 $&$ 1.42$\\\\ \\hline \\hline \\end{tabular} \\caption{\\label{tab:table5} First Panel - Stable case: Variation of the Radius with the Twist for weak helicity $b=0.05$ and $B_{0}=10G$. Second Panel - Unstable case: Variation of the Radius with the Twist for moderate helicity $b=0.5$ and $B_{0}=10G$.} \\end{table*}" }, "0807/0807.0935_arXiv.txt": { "abstract": "We present a study of flux-calibrated low-resolution optical spectroscopy of ten stars belonging to eight systems in the $\\sim 5$ Myr-old $\\epsilon$ Chamaeleontis ($\\epsilon$ Cha) pre-main-sequence (PMS) star cluster. Using synthetic broadband colours, narrow-band continuum, atomic and molecular lines derived from the spectra, we compare the $\\epsilon$ Cha stars to a slightly older PMS cluster, the $\\approx 8$ Myr-old $\\eta$ Cha cluster, and to main-sequence dwarfs. Using synthetic {\\it VRI\\,} colours and other indices, we find that the relationship between broadband colours and spectroscopic temperature indicators for $\\epsilon$ Cha cluster members is indistinguishable from that of Gyr-old dwarfs. This result is identical to that found earlier in $\\eta$ Cha. Gravity-sensitive line indices place the cluster between the giant and dwarf sequences, and there is clear evidence that $\\epsilon$ Cha stars have lower surface gravity than $\\eta$ Cha stars. This result is consistent with $\\epsilon$ Cha being the slightly younger PMS association, a few Myr younger according to the Hertzsprung-Russell (HR) diagram placement of these two clusters and comparison with PMS evolutionary grids. Late M-type $\\epsilon$ Cha cluster members show a $B$-band flux excess of $\\approx 0.2$ mag compared to observations of standard dwarfs, which might be related to enhanced magnetic activity. A similar level of excess $B$-band emission appears to be a ubiquitous feature of low mass members of young stellar populations with ages less than a few hundred Myr, a very similar timescale to the PMS phase of elevated relative X-ray luminosity. ", "introduction": "Analysis of flux-calibrated low-resolution optical spectra is a useful method for investigating the physical properties of stars in comparison with spectra of standard stars with well-defined properties. For nearby stars and stellar associations, such observations are readily obtained even with telescopes of modest aperture (e.g., Bessell 1991; Lyo, Lawson \\& Bessell 2004). Using synthetic broadband colours, narrow-band continuum, atomic and molecular line indices derived from calibrated spectra, we can study temperature, surface gravity and metallicity effects, e.g. Lyo et al. (2004) showed that the relationship between colours and spectroscopic temperature indicators for the $\\approx 8$ Myr-old $\\eta$ Cha cluster was indistinguishable from that for Gyr-old disk dwarfs, although the spectra showed some evidence for higher metallicity and clear evidence for lower gravity. The latter is a consequence of the elevated location of these stars, several mag above the zero main sequence in the HR diagram. The $\\eta$ Cha stars also displayed, like other young stellar populations, a $B$-band flux excess attaining $\\approx$ 0.2 mag for late-M cluster members. By studying PMS clusters of different isochronal ages, we can use spectrophotometric techniques to address various stellar evolutionary issues, e.g. is there evidence in the spectroscopy for differences in the temperature-spectral type sequence as a function of age? Do gravity-dependent features scale with age as stellar low mass stars descend their Hayashi tracks? Is the observed $B$-band excess in late-M PMS stars a general property of young stellar populations, and does it vary as a function of spectral type and age? In this paper, our target group is the young $\\epsilon$ Cha cluster associated with the early-type system $\\epsilon$ Cha AB and the codistant, comoving star HD104237 (DX Cha). HD104237 is the nearest-known Herbig Ae star and forms at least a quintet with low mass companions based upon {\\it Chandra X-ray Observatory} observations, optical/infrared imaging and spectroscopic study, some of which are likely multiple themselves (Feigelson et al. 2003; Grady et al. 2004). HD104237A itself is a spectroscopic binary with a K3 companion in an eccentric 20-d orbit (B\\\"ohm et al. 2004). The group is nearby ($d \\approx 114$ pc), compact in extent ($\\approx 1$ pc) and sparsely populated, containing eight stellar systems in its central region with stars ranging in spectral type from B9 to M5. HR diagram placement of the members ages the group at $3-5$ Myr (Feigelson et al. 2003) when compared to Siess, Dufour \\& Forestini (2000) evolutionary tracks, and $\\sim 6$ Myr (Luhman 2004) when using a mix of Palla \\& Stahler (1999), Baraffe et al. (1998) and Chabrier et al. (2000) tracks. Both the HR diagram comparisons of Feigelson et al. (2003) and Luhman (2004) assign younger ages to the early-type members of $2-3$ Myr. This discrepancy in age with spectral type might be a consequence of nearby low-mass companions elevating the luminosities of the early-type stars, or might serve to highlight differences between model isochrones and observational isochrones, or indicate a genuine age difference between the high-mass and low-mass stars in this cluster. Kinematic study indicates the group is another out-lying population of the Oph-Sco-Cen OB association (Feigelson et al. 2003), following groups and associations such as $\\eta$ Cha, the TW Hya association, and the $\\beta$ Pic moving group. The $\\epsilon$ Cha cluster presents itself then, as a slightly younger analogue of the $\\eta$ Cha star cluster (Mamajek, Lawson \\& Feigelson 1999, 2000), an $\\approx 8$ Myr-old compact, sparse PMS group that was studied using spectrophotometric techniques by Lyo et al. (2004). The ages of PMS clusters derived by HR diagram comparisons using competing evolutionary grids are unreliable; differences can exceed several Myr at $\\sim 10$ Myr (see fig. 8 of Lyo et al. 2004). However, comparison between PMS groups using techniques such as spectrophotometry ought to reveal differences that can be interpreted as age-related trends. With an age difference of only a few Myr, comparison between the $\\epsilon$ Cha and $\\eta$ Cha groups is a demonstration of the sensitivity of such methods to rank in age various PMS populations. ", "conclusions": "Flux-calibrated low-resolution optical spectroscopy is an indispensable tool for investigating the physical properties of stars by comparison with other stellar groups and standard star calibrators. In this paper, we characterized the stellar population of the $\\sim 5$ Myr-old $\\epsilon$ Cha cluster via a number of temperature- and gravity-sensitive spectroscopic indicators, and compared its properties to the slightly older PMS group associated with $\\eta$ Cha, and to those of Gyr-old disk dwarfs. Using synthetic broadband colours, narrow-band continuum, atomic and molecular line indices derived from the spectra, we find that the relationship between the broadband colours and spectroscopic temperature indicators for the $\\epsilon$ Cha cluster stars is indistinguishable from that of the Gyr-old dwarfs. We had previously reached the same conclusion for the slightly-older $\\eta$ Cha cluster (Lyo et al. 2004). However, there is a clear evidence that $\\epsilon$ Cha cluster stars have lower surface gravity than $\\eta$ Cha cluster stars from measurement of the gravity-sensitive $\\lambda\\lambda$8183, 8195 Na I doublet and FeH molecular spectral lines. This result is consistent with the $\\epsilon$ Cha cluster being slightly younger than $\\eta$ Cha, a few Myr younger according to the HR diagram placement of these two clusters and comparison with PMS evolutionary model grids. We also found a $B$-band excess of $\\sim 0.2$ mag in the late M-type cluster members, similar to that found in $\\eta$ Cha and in other PMS populations with ages less than $\\sim 200$ Myr. This result suggests that the blue excess is an ubiquitous property of low-mass young stellar objects, with the most-likely origin being enhanced magnetic activity. The presence of significant excess blue emission appears to closely parallel the phase of high relative X-ray luminosity seen in low-mass $1-100$ Myr-old PMS populations." }, "0807/0807.2271_arXiv.txt": { "abstract": "We report on the discovery of a bright \\lya\\ blob associated with the $z=3$ quasar SDSS~J124020.91$+$145535.6 which is also coincident with strong damped \\lya\\ absorption from a foreground galaxy (a so-called proximate damped \\lya\\ system; PDLA). The one dimensional spectrum acquired by the Sloan Digital Sky Survey (SDSS) shows a broad \\lya\\ emission line with a FWHM $\\simeq 500\\mkms$ and a luminosity of $L_{\\rm Ly\\alpha} = 3.9 \\sci{43} {\\rm erg \\, s^{-1}}$ superposed on the trough of the PDLA. Mechanisms for powering this large \\lya\\ luminosity are discussed. We argue against emission from \\ion{H}{2} regions in the PDLA galaxy since this requires an excessive star-formation rate $\\sim 500\\,\\msol \\rm yr^{-1}$ and would correspond to the largest \\lya\\ luminosity ever measured from a damped \\lya\\ system or starburst galaxy. We use a Monte Carlo radiative transfer simulation to investigate the possibility that the line emission is fluorescent recombination radiation from the PDLA galaxy powered by the ionizing flux of the quasar, but find that the predicted \\lya\\ flux is several orders of magnitude lower than observed. We conclude that the \\lya\\ emission is not associated with the PDLA galaxy at all, but instead is intrinsic to the quasar's host and similar to the extended \\lya\\ ``fuzz'' which is detected around many AGN. PDLAs are natural coronagraphs that block their background quasar at \\lya\\ , and we discuss how systems similar to SDSS~J124020.91$+$145535.6 might be used to image the neutral hydrogen in the PDLA galaxy \\emph{in silhouette} against the screen of extended \\lya\\ emission from the background quasar. ", "introduction": "The ionizing radiation emitted by a luminous quasar can, like a flashlight, illuminate hydrogen clouds in its vicinity, teaching us about the size, kinematic structure, and density of the gas which surrounds it \\citep{rees88,hr01}. This is because recombinations from photoionized hydrogen ultimately produce Ly$\\alpha$ photons -- a fraction of the energy in the quasar's UV continuum is `focused' into fluorescent line radiation and re-emitted, allowing us to study the physical conditions in the emitting gas. Extended \\lya\\ nebulae with luminosities up to $L_{\\rm Ly\\alpha}=5\\sci{42}~{\\rm erg~s^{-1}}$ and angular sizes as large $\\sim 10\\arcsec$ have been observed around many quasars \\citep[e.g.][]{djorgovski85,heckman91a,cjw+06}, but the physical mechanism powering this emission is still unclear. In addition to fluorescent emission powered by photoionization, other mechanisms include emission from material shocked by radio-jets or starburst outflows or \\lya\\ cooling radiation from gravitational collapse. The same set of physical mechanisms have been proposed to explain the analogous \\lya\\ nebulae observed around a subset of luminous radio galaxies \\citep{msd+90,villar07}, as well as the recently discovered \\lya\\ `blobs' \\citep{fmw99,steidel00,matsuda04}. The primary difference between the extended ``fuzz'' around quasars as compared to the nebulae around radio galaxies and \\lya\\ blobs is that a luminous quasar and hence the source of ionizing photons is directly detected in the former but not in the latter. Obscuration and orientation effects, as are often invoked in unified models of AGN, could be responsible for this difference. This explanation is plausible considering that the large energies in the jets $\\gtrsim 10^{60}$\\,erg \\citep[e.g.][]{Miley80} of luminous radio galaxies strongly suggest nuclear activity and evidence for an obscured AGN has been uncovered in several of the \\lya\\ blobs \\citep{csw+04,dbs+05,gsc+07}. In the course of a survey for damped \\lya\\ absorption proximate to high $z$ quasars \\citep[][see also Ellison et al. 2002 and Russell et al. 2006]{phh08}, we discovered an extremely luminous \\lya\\ blob coincident with both the redshift of a $z=3$ proximate damped \\lya\\ system (PDLA)\\footnote{ A PDLA is defined as an absorber with $\\mnhi \\ge 2\\sci{20} \\cm{-2}$ located within $\\delta v = 3000 \\mkms$ of its background quasar} and its background quasar. Although intervening DLAs ($\\delta v > 3000 \\mkms$) rarely show \\lya\\ emission \\citep{scb+89,mff04,kwy+06}, PDLAs appear to preferentially exhibit \\lya\\ emission superimposed on their \\lya\\ absorption trough \\citep{mw93,mwf98,eyh+02} which is $\\sim 5$ times brighter than the few intervening detections. Although based on poor statistics and heterogeneous samples, this putative discrepancy suggests that the \\lya\\ emission associated with PDLAs might be powered by its background quasar. In what follows, we analyze the basic characteristics of this absorber and discuss several physical scenarios for the origin of this \\lya\\ blob. Finally, we emphasize several novel applications that exploit this unique configuration. Throughout this paper we use the best fit WMAP3 cosmological model of \\citet{wmap03}, with $\\Omega_m = 0.240$, $\\Omega_\\Lambda =0.76$, $h=0.73$. % ", "conclusions": "(1) the gas is located within approximately 1\\,Mpc of the quasar; (2) the majority of PDLA systems lying within 1\\,Mpc of the quasar should also show significant \\ion{N}{5} absorption. We explore the fluorescence scenario further by performing the following numerical simulation. The PDLA system is modeled as a singular isothermal sphere within a halo of $10^{11}\\msol$, for which 5\\% of the mass is assumed to be in gas. The system is placed between the observer and the quasar with a distance 300\\,kpc to the quasar such that the PDLA's backside is illuminated by the quasar. We use the code developed in \\citet[][see also Cantalupo et al. 2005]{Juna08} to perform the self-shielding calculation, solving for the ionization structure of the gas, the recombination rate, and the resulting \\lya\\ luminosity. The radiative transfer of the fluorescent \\lya\\ photons is solved with a Monte Carlo method \\citep{zheng02,Juna08}. In turn, we produce a map of the expected \\lya\\ surface-brightness as shown in Figure~\\ref{fig:theory}. We have extracted several one-dimensional spectra from this surface-brightness image to illustrate the flux level and the expected emission line profiles. Although the one-dimensional spectrum through the $3''$ SDSS fiber aperture has roughly the observed line width, it has an integrated flux that is several orders of magnitude lower than observed. The low flux results from the fact that we are observing the illuminated galaxy in a `new moon' configuration; the optically thick regions which give rise to the \\lya\\ emission also serve to shield the radiation from our vantage point. The observed flux is instead dominated by the outer regions of fiber which cover optically thin regions of gas in the halo surrounding the galaxy (aperture A2 in Figure~\\ref{fig:theory}). Based on this calculation, we conclude that the observed \\lya\\ emission in \\name\\ is unlikely to be fluorescent recombination radiation from the PDLA galaxy. {\\bf Extended \\lya\\ ``Fuzz'' from the Quasar Host Halo:} We consider the most plausible hypothesis to be that the \\lya\\ emission is not associated with the PDLA galaxy at all, but instead is intrinsic to the quasar's host galaxy. Indeed, \\lya\\ emission has been detected as an extended `fuzz' around many quasars \\citep[e.g.][]{djorgovski85,heckman91a,cjw+06}, and with luminosities comparable to or greater than the $L_{\\rm Ly\\alpha} = 3.9\\sci{43} {\\rm erg \\, s^{-1}}$ measured for \\name. The velocity widths of this extended emission can be as large as $1000-1500$~\\kms consistent with our observed kinematic profile. The physical mechanism responsible for this emission could be fluorescent recombination radiation from gas in the quasar host halo illuminated by quasar's large ionizing flux \\citep{rees88,heckman91a,hr01,weidinger04}. It is important to explain the distinction between the fluorescent \\lya\\ fuzz as it occurs in the quasar host halo to the fluorescence from the DLA galaxy which was considered in the previous section. For the former case the emission is presumed to be coming from small dense self-shielding clouds which permeate the quasar halo but have a small $\\sim 0.5\\%$ covering factor \\citep{msd+90,heckman91a} deduced from the fraction of the ionizing continuum which is being absorbed \\citep{heckman91a}. In the latter case, where a galaxy is illuminated, the emission comes from a large, kpc-scale self-shielding cloud and a large fraction $\\eta \\simeq 60\\%$ of the impingent ionizing continuum is converted into \\lya\\ recombination photons. But the mechanism powering the extended emission line `fuzz' could be among other mechanisms put forth to explain the extended \\lya\\ halos around quasars \\lya\\ blobs, and high-redshift radio galaxies, such as gas shock-heated by a large scale outflow or cooling radiation from gravitational infall \\citep[e.g.][ and references therein]{matsuda04,ctf+06}. In the halo fuzz scenario, the DLA-galaxy is in the foreground of the extended \\lya\\ emitting region of the quasar host, and should hence absorb the fraction of the \\lya\\ emission which it covers. This configuration also offers a reasonable explanation for why PDLAs might preferentially exhibit \\lya\\ emission relative to intervening DLAs: the presence of a PDLA occults the bright quasar making it much easier to detect faint extended emission from the quasar halo." }, "0807/0807.1272_arXiv.txt": { "abstract": " ", "introduction": "\\label{s:intro} Planetary systems are not just made up of \\textit{planets}, but are also composed of numerous small bodies ranging from asteroids and comets as large as 1000 km down to sub-$\\mu$m-sized dust grains. In the solar system the asteroids and comets are confined to relatively narrow rings known as the asteroid belt and the Kuiper belt (see chapters by Nakamura and Jewitt). These belts are the source of the majority of the smaller objects seen in the solar system, since such objects are inevitably created in collisions between objects within the belts (see chapter by Michel). Sublimation of comets as they are heated on approach to the Sun is another source of dust in the solar system. It is known that extrasolar systems also host belts of planetesimals (a generic name for comets and asteroids) that are similar to our own asteroid belt and Kuiper belt. These were first discovered using far-IR observations of nearby stars, which showed excess emission above that expected to come from the stellar photosphere \\cite{auma84}. This emission comes from dust that is heated by the star and which re-radiates that energy in the thermal infrared, at temperatures between 40-200 K, depending on the distance of the dust from the star. The lifetime of the dust is inferred to be short compared with the age of the star, and so it is concluded that the dust cannot be a remnant of the proto-planetary disk that formed with the star (see chapter by Takeuchi), rather it must originate in planetesimal belts much in the same way that dust is created in the solar system \\cite{bp93}. Over 300 main sequence stars are now known with this type of excess emission \\cite{mb98,su06,bryd06}, and such objects are either known as Vega-like (after the first star discovered to have this excess), or as debris disks. Statistical studies have shown that $\\sim 15$\\% of normal main sequence stars have debris disks, although it should be stressed that the disks which can be detected with current technology have greater quantities of dust than is currently present in the solar system by a factor of at least 10 \\cite{gwhd04}. Nevertheless this indicates that debris disks are common, more common in fact that extrasolar planets which are found around $\\sim 6$\\% of stars \\cite{fv05}. Studying these disks provides a unique insight into the structure of the planetary systems of other stars. Indeed, the nearest and brightest debris disks can be imaged, and such studies have provided the first images of nearby planetary systems. These images reveal the distribution of dust in the systems, which can in turn be used to infer the distribution of parent planetesimals, and also the architecture of the underlying planetary system. However, to do so requires an understanding of both the mechanism by which dust is produced in planetesimal belts and its consequent dynamical evolution, as well as of the dynamical interaction between planets and planetesimals and between planets and dust. This chapter reviews our knowledge of debris disks from observations (\\S \\ref{s:obs}) and describes a simple model for planetesimal belt evolution which explains what we see (\\S \\ref{s:mod}), as well as how the detailed interaction between planets and planetesimals imposes structure on that planetesimal belt (\\S \\ref{s:plpl}), and how those perturbations translate into structures seen in the dust distribution (\\S \\ref{s:pldust}). Conclusions, including what has been learned about the planetary systems of nearby stars from studying these disks, are given in \\S \\ref{s:conc}. ", "conclusions": "\\label{s:conc} This chapter has considered the types of structures seen in the dusty debris disks of nearby stars (\\S \\ref{s:obs}) and how those structures can be used to determine the layout of their planetary systems, in terms of the distributions of both planetesimals and planets. The text has dwelled on the successes of the models at explaining the observed structures, because this illustrates the elements that are essential to any debris disk model if the observations are to be successfully explained (\\S \\ref{s:mod}), and because we are confident that we understand how a planet would perturb a planetesimal belt in an idealised system comprised of just one planet (\\S \\ref{s:plpl}) and to some extent how to extrapolate that to consider how the planet would affect the observed dust disk (\\S \\ref{s:pldust}). To summarise what we have learned: \\textbf{(i)} the axisymmetric structure of debris disks can mostly be explained by a model in which dust is created in collisions in a narrow planetesimal belt and is subsequently acted on by radiation forces; \\textbf{(ii)} the asymmetric structure of debris disks can mostly be explained by secular and resonant gravitational perturbations from unseen planets acting on the planetesimal belt and dust derived from it. Knowing the radial location of the planetesimal belts is important because this demonstrates where in a protoplanetary disk grain growth must have continued to km-sized planetesimals \\cite{wd02}, and by analogy with the solar system there is reason to believe that the location of the planetesimal belts tells us indirectly the whereabouts of unseen planets, although it is worth bearing in mind that there may be alternative explanations for gaps in the planetesimal distribution related to the physics of the protoplanetary disk. Nevertheless, it appears that where we have the capability to look for detailed disk structure, there is good correspondence between the asymmetric structures observed with those expected if there are planets in these systems. The modelling is also sufficiently advanced that the disk structure can be used to infer information on the properties of the perturbing planets (such as the planet's mass, orbit and even evolutionary history). The planet properties which have been inferred in this way are particularly exciting when compared with those of exoplanets discovered using the radial velocity and transit techniques. Figure \\ref{fig:exopl} shows how the debris disk planets are similar to Uranus and Neptune in the solar system, occupying a unique region of parameter space. This is possible because the large size of debris disks means that the planets perturbing them are most often at large orbital radii, and it is easy for planets as small as Neptune to impose structure on a debris disk. There is also the tantalising possibility that in the future debris disk structures can be used to identify planets analogous to the Earth and Venus in extrasolar systems. \\begin{figure} \\centering \\begin{tabular}{c} \\includegraphics[height=8.3cm]{wyattfig12.ps} \\end{tabular} \\caption{Distribution of planet masses and semimajor axes. Solar system planets are plotted as open circles, and those known from radial velocity and transit studies with a plus (taken from the list on http://exoplanets.eu dated 24 May 2007). The shaded region shows the current limits of radial velocity surveys for sun-like stars. Debris disk planets inferred from disk structure (all awaiting confirmation) are shown with filled circles. References for the plotted planet parameters are: HR4796 \\cite{wdtf99}, $\\epsilon$ Eridani \\cite{ogmt00}, Vega \\cite{wyat03}, HD141569 \\cite{wyat05b}, $\\eta$ Corvi \\cite{wgdc05}, Fomalhaut \\cite{quil06}, $\\beta$ Pictoris \\cite{fkl07}, although it should be noted that these parameters, particularly planet mass, are often poorly constrained.} \\label{fig:exopl} \\end{figure} However, while it is incontrovertible that if there are planets present they would impose structure on a disk, the question of whether we have already seen these structures in extrasolar systems is still a matter for debate. In many cases the presence of an unseen planet is the only explanation for the observed structures, but that does not mean that it has to be the right explanation. The problem is that it is hard to confirm that the planets are there, since they lie beyond the reach of radial velocity studies (see Fig.~\\ref{fig:exopl}). Direct imaging could detect planets at this distance if they were a few times Jupiter mass \\cite{mhw03}, but not if they are Neptune mass. Thus the onus is on the models to make other testable predictions, and some of these have already been made (such as the orbital motion of the clumpy structures, and the disk structures expected to be seen at different wavelengths) and will be tested in the coming years. If these planets are confirmed, their addition onto plots like that shown in Figure \\ref{fig:exopl} will be invaluable for constraining planet formation models \\cite{il04}. It is also important to remember that this theory cannot yet predict the quantities of small grains we would expect to see in any given disk. There are too many uncertainties regarding the dust production mechanisms, and it is possible that these processes may differ among stars with, e.g., different dust compositions. Applying dynamical models of the kind presented in \\S \\ref{s:mod} to a greater number of resolved disk observations will help to understand these differences. However, there is still the possibility that the problem is more fundamental in a way which is best illustrated by the archetypal debris disk Vega. The observed mass loss rate from $\\beta$ meteoroids in this system is $2M_\\oplus$/Myr, which indicates that this must be a transient, rather than a steady state, component \\cite{su05}. It is thus possible that the small grain population in debris disks is inherently stochastic, perhaps influenced by input from recent massive collisions \\cite{tele05}. Fortunately it appears that the large grain component of the majority of debris disks is evolving in steady state \\cite{wssr07} and so can be understood within the framework described in this chapter, and the same is likely also true for the small grain component (it is just the relative quantities of the different components that is less certain). However, the possibility must be considered that in some systems the observed dust is transient in such a way that its origin will require a significant overhaul to the models presented here. For example, there are a few cases of sun-like stars surrounded by hot dust (e.g., \\S \\ref{ss:axi}) which cannot be maintained by steady state production in a massive asteroid given the age of the stars \\cite{wsgb07}. It is not clear what the origin of the transient event producing the dust is. However, it is known that the quantity of planetesimals in the inner solar system has had a stochastic component, notably involving a large influx $\\sim 700$ Myr after the solar system formed in an event known as the late heavy bombardment, the origin of which is thought to have been a dynamical instability in the architecture of the giant planets \\cite{gltm05}. So perhaps these systems are telling us about the more complex dynamics of their planetary systems. Given the complexity of planetary systems it seems inevitable that the models presented in this chapter are just the start of a very exciting exploration of the dynamics of extrasolar planetary systems." }, "0807/0807.1758_arXiv.txt": { "abstract": "We investigate the formation of carbon-enhanced metal-poor (CEMP) stars via the scenario of mass transfer from a carbon-rich asymptotic giant branch (AGB) primary to a low-mass companion in a binary system. We explore the extent to which material accreted from a companion star becomes mixed with that of the recipient, focusing on the effects of thermohaline mixing and gravitational settling. We have created a new set of asymptotic giant branch models in order to determine what the composition of material being accreted in these systems will be. We then model a range of CEMP systems by evolving a grid of models of low-mass stars, varying the amount of material accreted by the star (to mimic systems with different separations) and also the composition of the accreted material (to mimic accretion from primaries of different mass). We find that with thermohaline mixing alone, the accreted material can become mixed with between 16 and 88\\% of the pristine stellar material of the accretor, depending on the mass accreted and the composition of the material. If we include the effects of gravitational settling, we find that thermohaline mixing can be inhibited and, in the case that only a small quantity of material is accreted, can be suppressed almost completely. ", "introduction": "Carbon-enhanced, metal-poor (CEMP) stars are defined as stars with [C/Fe]\\footnote{[A/B] = $\\log (N_\\mathrm{A}/N_\\mathrm{B}) - \\log (N_\\mathrm{A}/N_\\mathrm{B})_\\odot$}$>$+1.0 \\citep{2005ARA&A..43..531B}, with [Fe/H]$<-2$ in most cases. These objects appear with increasing frequency at low metallicity \\citep{2006ApJ...652L..37L}. The study of CEMP stars is being used to probe conditions in the early universe. For example, CEMP stars have been used to infer the initial mass function in the early Galaxy \\citep[e.g.][]{2005ApJ...625..833L}. Chemical abundance studies have revealed that the majority of the CEMPs are rich in $s$-process elements like barium \\citep{2003IAUJD..15E..19A}, forming the so-called CEMP-s group. Recent survey work has detected a binary companion in around 68\\% of these CEMP-s stars and this is consistent with them all being in binary systems \\citep{2005ApJ...625..825L}. Binary systems provide a natural explanation for these objects, which are of too low a luminosity to have been able to produce their own carbon. The primary of the system was an asymptotic giant branch (AGB) star which became carbon-rich through the action of third dredge-up\\footnote{Third dredge-up occurs when the convective envelope of an AGB star deepens after a thermal pulse and material that has experienced nuclear burning is brought to the surface.} and transferred material on to the low-mass secondary (most likely via a stellar wind). The primary became a white dwarf and has long since faded from view, with the carbon-rich secondary now being the only visible component of the system. It has commonly been assumed that the accreted material remains on the surface of the secondary until the star ascends the giant branch, at which point the deepening of the convective envelope (referred to as first dredge-up because material that has experienced CN-cycling in the stellar interior is brought to the surface) mixes the material with the interior of the star. However, the transferred material should become mixed with the interior of the accreting star via the process of thermohaline mixing \\citep[see e.g.][]{2007A&A...464L..57S, 2004MNRAS.355.1182C,2003NewA....8...23B}. This occurs when the mean molecular weight of the stellar gas increases toward the surface. A gas element displaced downwards and compressed will be hotter than its surroundings. It will therefore lose heat, become denser and continue to sink. This leads to mixing on thermal timescales until the molecular weight difference is eliminated \\citep{1980A&A....91..175K}. \\citet{2007A&A...464L..57S} showed that the inclusion of thermohaline mixing could result in the accreted material being mixed throughout 90\\% of the star. Recent work has questioned the efficiency of thermohaline mixing in carbon-enhanced metal-poor stars. Using a sample of barium-rich CEMP stars, \\citet{2008ApJ...678.1351A} showed that the distribution of [C/H] values in turn-off stars (i.e. those stars that have reached the end of their main-sequence lives, are still of low-luminosity and have yet to become giants) was different from that in giants suggesting that significant mixing only happened at first dredge-up. A similar point was made by \\citet{2007arXiv0709.4240D} using the data of \\citet{2006ApJ...652L..37L}. These authors showed that the turn-off stars and giants were consistent with coming from the same distribution if first dredge-up resulted in the [C/H] value (and also the [N/H] value) being reduced by around 0.4 dex. They find that this result is consistent with having an accreted layer of material mixed to an average depth of about 0.2\\ms\\ (or alternatively having an accreted layer of 0.2\\ms\\ that remains unmixed). Neither of these scenarios is consistent with the extensive mixing found by \\citet{2007A&A...464L..57S}. A possible source for reduced thermohaline mixing efficiency has been suggested by \\citet{2008ApJ...677..556T}. These authors suggest that the action of gravitational settling will alter the composition gradient of the accreting star near its surface. Helium will settle from the surface, reducing the mean molecular weight at the surface but leading to an increase in the layers beneath. This produces a small region in which the mean molecular weight, $\\mu$, decreases outwards toward the stellar surface -- a situation which is stable to thermohaline mixing. This stabilising composition gradient (a so-called `$\\mu$-barrier') can inhibit the process of thermohaline mixing. This paper extends the work of \\citet{2007A&A...464L..57S}, examining the effect of varying the composition of the accreted material (mimicking accretion from different masses of companion) and the amount of material that is accreted. We also investigate the effect that gravitational settling has on the extent to which material is mixed. ", "conclusions": "We have modelled the accretion of AGB material on to low-mass stars in order to determine what chemical signatures may be observed in CEMP stars. We have examined three specific cases: canonical evolution including only convective mixing, the inclusion of thermohaline mixing and the inclusion of both thermohaline mixing and gravitational settling. We find that thermohaline mixing can lead to accreted material being mixed with between 16 and 88\\% of the accreting star, depending on the mass and composition of the accreted material. When gravitational settling is included, thermohaline mixing is severely inhibited when only a small amount of material (around a few $10^{-3}$\\ms) is accreted because of the presence of a $\\mu$-barrier formed by the settling of helium. This barrier is less and less effective as the amount of accreted material is increased. It is also less effective for more massive companions as the barrier has had less time to be established before accretion on to the secondary occurs. Models without thermohaline mixing produce turn-off objects with [C/H] values that are too high to match observations. Unless a substantial quantity of material has been accreted, these models also suffer from too much dilution at first dredge-up. Models with thermohaline mixing cannot reproduce the highest [C/H] values observed and predict the existence of low [C/H] objects which are not observed. In order to reproduce the highest [C/H] values, which seem to require the accretion of a large quantity of material from a companion, some alternative mechanism to suppress thermohaline convection is required. The inclusion of gravitational settling can solve the problem of the low [C/H] objects at low luminosity as the $\\mu$-barrier prevents mixing when the amount of accreted material is small. However, the inclusion of this physics presents another serious problem namely that carbon will settle from the surface during the main sequence, making it extremely difficult to form turn-off stars with [C/H] values of -1--0. There also appears to be a problem with the abundance predictions of the AGB models. They do not predict substantial enhancements of both carbon {\\it and} nitrogen at the same time (except in a very narrow mass range). Further work needs to be done to find a mechanism capable of producing the observed abundance patterns." }, "0807/0807.4382_arXiv.txt": { "abstract": "We re-analyse all of the archive observations of the Ophiuchus dark cloud L1688 that were carried out with the submillimetre common-user bolometer array (SCUBA) at the James Clerk Maxwell Telescope (JCMT). For the first time we put together all of the data that were taken of this cloud at different times to make a deeper map at 850$\\mu$m than has ever previously been published. Using this new, deeper map we extract the pre-stellar cores from the data. We use updated values for the distance to the cloud complex, and also for the internal temperatures of the pre-stellar cores to generate an updated core mass function (CMF). This updated CMF is consistent with previous results in so far as they went, but our deeper map gives an improved completeness limit of 0.1M$_{\\sun}$ (0.16 Jy), which enables us to show that a turnover exists in the low-mass regime of the CMF. The L1688 CMF shows the same form as the stellar IMF and can be mapped onto the stellar IMF, showing that the IMF is determined at the prestellar core stage. We compare L1688 with the Orion star-forming region and find that the turnover in the L1688 CMF occurs at a mass roughly a factor of two lower than the CMF turnover in Orion. This suggests that the position of the CMF turnover may be a function of environment. ", "introduction": "Star formation in molecular clouds occurs within prestellar cores, which are gravitationally bound cores within the clouds \\citep{1994MNRAS.268.276W,1996A&A.314.625A,1999MNRAS.305.143W, 2002Sci...295...76W, 2007prpl.conf...17D, 2007prpl.conf...33W}. A number of observations have shown that the core mass function (CMF) of prestellar cores appears to mimic \\citep[Motte Andre \\& Neri 1998, hereafter MAN98,][]{1998ApJ...508L..91T, 2000ApJ...545..327J, 2001A&A...372L..41M, 2001ApJ...559..307J, 2002Sci...295...82K, 2002ApJ...575..950O, 2006ApJ...639..259J} the stellar initial mass function \\citep[IMF;][]{1955ApJ...121..161S}. However, the comparison between the core mass function and stellar IMF has not often been accurately probed at lower masses. It is more difficult to study this part of the mass domain, but recent results have shown that the CMF exhibits a turnover at lower masses in a manner similar to the IMF \\citep{2007MNRAS.374.1413N}. The Ophiuchus star-forming region is located at a distance of 139~pc \\citep{2008AN....329...10M} and is a site of low-mass star formation \\citep{1983ApJ...274..698W}. The region consists of two main clouds, L1688 and L1689, which both have extended streamers leading out to lengths of around 10 pc \\citep{1989ApJ...338..902L}. Specifically, it is the more massive of the two clouds, L1688, which is studied in this paper, and which is generally known as the Ophiuchus main cloud. Very high star formation rates have been measured here, with 14--40\\% of the molecular gas being converted into stars \\citep{1977AJ.....82..198V}. The Ophiuchus cloud has been observed in many wavelengths from the visible to the submilimetre \\citep[e.g. MAN98, ][]{1983ApJ...269..182M, 1989ApJ...340..823W, 1992ApJ...401..667A, 1992ApJ...395..516G, 1997ApJS..112..109B, 2000ApJ...545..327J, 2001MNRAS.323.1025J, 2004ApJ...611L..45J, 2008ApJS..175..277D}. Because of this, the properties of the cloud are very well known and it is therefore a good place to probe low-mass star formation. The Ophiuchus cloud is the nearest example of `clustered' star formation (MAN98). This is important to study because most stars form in clustered environments \\citep{1993prpl.conf..429Z}. Ophiuchus may also be the nearest example of triggered star formation in action \\citep{1989ApJ...338..902L,2006MNRAS.368.1833N}, making it a prime candidate for study. \\begin{figure*} \\includegraphics[angle=0,width=0.85\\textwidth]{./RhoOphWholeLabels2.eps} \\caption{Greyscale image and isophotal contour map of the SCUBA 850$\\mu$m continuum scan-map data of the Ophiuchus dark cloud L1688. Signal-to-noise contours at 5$\\sigma$ and 10$\\sigma$ are shown in black; 25$\\sigma$ and 100$\\sigma$ contours shown in white. 1$\\sigma$ noise levels vary from 15 to 40 mJy/beam (see text for details).} \\label{whole-map} \\end{figure*} \\begin{figure} \\includegraphics[angle=0,width=0.40\\textwidth]{./RhoOphWholeJiggleLabels.eps} \\caption{Greyscale image and isophotal contour map of the SCUBA 850$\\mu$m continuum jiggle map data of the Ophiuchus dark cloud L1688. Contours at 0.25, 0.5 and 1.0 Jy/beam are shown in black; 2.0 and 4.0 Jy/beam contours are shown in white. 1$\\sigma$ noise levels vary from 10 to 180 mJy/beam.} \\label{jiggle-map} \\end{figure} In this study we have combined all of the high signal-to-noise SCUBA wide-field scan-map data and narrow-field jiggle-map data taken of L1688, and re-reduced it to produce the deepest submillimetre map of this cloud ever made. L1688 is the region of the Ophiuchus cloud defined by \\citet{1989ApJ...338..902L} and outlined in their Figure~1a (it is marked out by a solid, 5K contour). Of the original regions covered at 1.3mm by MAN98 (Oph-A, -B, -C, -D, -E and -F) only one, Oph-D, is not included here, as it is not part of the central region of L1688. A newly discovered region, which we name Oph-J, is discussed. Two smaller regions, Oph-H and Oph-I are discussed by \\citet{2004ApJ...611L..45J} but are not included in this study. We produce a CMF and investigate the low-mass end of the CMF of the cloud. We compare this with previous findings and also with the Orion molecular cloud \\citep{2007MNRAS.374.1413N}. ", "conclusions": "In this paper we have re-analysed the SCUBA archive data for L1688, incorporating all available high signal-to-noise scan-map and jiggle-map data. An updated form of the CMF in the L1688 cloud complex has been presented using updated values for the distance to this region as well as new estimates for the temperatures of the cores. We have shown that the CMF for L1688 is consistent with a three part power-law with slopes the same as seen in the stellar IMF. The higher-mass end of the CMF declines as a power law which is consistent with other studies of L1688 \\citep[MAN98;][]{2000ApJ...545..327J, 2006A&A...447..609S} as well as studies of Orion \\citep{2007MNRAS.374.1413N,2001A&A...372L..41M,2001ApJ...559..307J,2006ApJ...639..259J}. Hence, the results are mostly in agreement with those found in earlier studies. However, our deeper maps have allowed the discovery of a turnover in the CMF at 0.7$M_{\\sun}$ which shows that the core mass function continues to mimic the stellar initial mass function to low masses. This agreement is indicative that the stellar IMF is determined at the prestellar core stage. It has been shown that the relationship between the CMF and IMF is not necessarily a simple 1:1 translation in the mass axis. Consistency can also be achieved using a fully multiple star model." }, "0807/0807.3794_arXiv.txt": { "abstract": "We have carried out a multi-site campaign to measure oscillations in the F5 star Procyon~A. We obtained high-precision velocity observations over more than three weeks with eleven telescopes, with almost continuous coverage for the central ten days. This represents the most extensive campaign so far organized on any solar-type oscillator. We describe in detail the methods we used for processing and combining the data. These involved calculating weights for the velocity time series from the measurement uncertainties and adjusting them in order to minimize the noise level of the combined data. The time series of velocities for Procyon shows the clear signature of oscillations, with a plateau of excess power that is centred at 0.9\\,mHz and is broader than has been seen for other stars. The mean amplitude of the radial modes is $38.1\\pm1.3$\\,\\cms{} (2.0 times solar), which is consistent with previous detections from the ground and by the WIRE spacecraft, and also with the upper limit set by the MOST spacecraft. The variation of the amplitude during the observing campaign allows us to estimate the mode lifetime to be $1.5_{-0.8}^{+1.9}$\\,d. We also find a slow variation in the radial velocity of Procyon, with good agreement between different telescopes. These variations are remarkably similar to those seen in the Sun, and we interpret them as being due to rotational modulation from active regions on the stellar surface. The variations appear to have a period of about 10 days, which presumably equals the stellar rotation period or, perhaps, half of it. The amount of power in these slow variations indicates that the fractional area of Procyon covered by active regions is slightly higher than for the Sun. ", "introduction": "Measuring solar-like oscillations in main-sequence and subgiant stars requires high-precision observations -- either with spectroscopy or photometry -- combined with coverage that is as continuous as possible. Most of the results have come from high-precision Doppler measurements using ground-based spectrographs, while measurements from spacecraft have also been reported (see \\citealt{B+K2007c} and \\citealt{AChDC2008} for recent summaries). Procyon has long been a favourite target for oscillation searches. At least eight separate velocity studies have reported an excess in the power spectrum, beginning with that by \\citet{BGN91}, which was the first report of a solar-like power excess in another star. For the most recent examples, see \\citet{MLA2004}, \\citet{ECB2004}, \\citet{BMM2004} and \\citet{LKB2007}. These studies agreed on the location of the excess power (around 0.5--1.5\\,mHz) but they disagreed on the individual oscillation frequencies. However, a consensus has emerged that the large separation (the frequency separation between consecutive overtone modes of a given angular degree) is about 55\\,\\muHz. Evidence for this value was first given by \\citet{MMM98} and the first clear detection was made by \\citet{MSL99}. Controversy was generated when photometric observations obtained with the MOST satellite failed to reveal evidence for oscillations \\citep{MKG2004,GKR2007,BAB2008}. However, \\citet{BKB2005} argued that the MOST non-detection was consistent with the ground-based data. Meanwhile, \\citet{R+RC2005} suggested that the signature of oscillations is indeed present in the MOST data at a low level (see also \\citealt{Mar08}). Using space-based photometry with the WIRE satellite, \\citet{BKB2005b} extracted parameters for the stellar granulation and found evidence for an excess due to oscillations. All published velocity observations of Procyon have been made from a single site, with the exception of two-site observations by \\citet{MLA2004}. Here we describe a multi-site campaign on Procyon carried out in 2007 January, which was the most extensive velocity campaign so far organized on any solar-type oscillator. The only other comparable effort to measure oscillations in this type of star was the multi-site photometric campaign of the open cluster M67 \\citep{GBK93}. ", "conclusions": "We have presented multi-site velocity observations of Procyon that we obtained with eleven telescopes over more than three weeks. Combining data that spans a range of precisions and sampling rates presents a significant challenge. When calculating the power spectrum, it is important to use weights that are based on the measurement uncertainties, otherwise the result is dominated by the noisiest data. We have described in detail our methods for adjusting the weights in order to minimize the noise level in the final power spectrum. Our velocity measurements show the clear signature of oscillations. The power spectrum shows an excess in a plateau that is centred at 0.9\\,mHz and is broader than has been seen for other solar-type stars. The mean amplitude of the radial modes is $38.1\\pm1.3$\\,\\cms{} ($2.04 \\pm 0.10$ times solar), which is consistent with previous detections from the ground and by the WIRE spacecraft, and also with the upper limit set by the MOST spacecraft. The variation of the amplitude during the observing campaign allowed us to estimate the mode lifetime to be $1.5_{-0.8}^{+1.9}$\\,d. We also found a slow variation in the radial velocity of Procyon, with good agreement between different telescopes. These variations are remarkably similar to those seen in the Sun, and we interpret them as being due to rotational modulation from active regions on the stellar surface. The variations appear to have a period of about 10 days, which presumably equals the stellar rotation period or, perhaps, half of it. The amount of power in these slow variations indicates that the fractional area of Procyon covered by active regions is slightly higher than for the Sun. The excellent coverage of the observations and the high signal-to-noise should allow us to produce a good set of oscillation frequencies for Procyon. This analysis will be presented in subsequent papers." }, "0807/0807.3331_arXiv.txt": { "abstract": "{The nature of Type Ia supernova progenitors is still unclear. The outstanding characteristic of the single-degenerate scenario is that it contains hydrogen in the binary companion of the exploding white dwarf star, which, if mixed into the ejecta of the supernova in large amounts may lead to conflicts with the observations thus ruling out the scenario.}{We investigate the effect of the impact of Type Ia supernova ejecta on a main sequence companion star of the progenitor system. With a series of simulations we investigate how different parameters of this system affect the amount of hydrogen stripped from the companion by the impact.}{The stellar evolution code GARSTEC is used to set up the structure of the companion stars mimicking the effect of a binary evolution phase. The impact itself is simulated with the smoothed particle hydrodynamics code GADGET2.}{We reproduce and confirm the results of earlier grid-based hydrodynamical simulation. Parameter studies of the progenitor system are extended to include the results of recent binary evolution studies. The more compact structure of the companion star found here significantly reduces the stripped hydrogen mass.}{The low hydrogen masses resulting from a more realistic companion structure are consistent with current observational constraints. Therefore, the single-degenerate scenario remains a valid possibility for Type Ia supernova progenitors. These new results are not a numerical effect, but the outcome of different initial conditions.} ", "introduction": "\\label{sec:introduction} While the progenitors for Type Ib/c and Type II supernovae are known, Type Ia supernovae (SNe~Ia) still elude an identification of their progenitor system. This is an unpleasant situation given the fact that these objects are one of the most important tools to determine cosmological parameters. By virtue of empirical calibration methods \\citep[e.g.,][]{phillips1993a} they can be used as standardizable candles for distance measurements. This calls for an understanding of the mechanism of SNe~Ia; and indeed, some progress has been made in recent years in understanding the explosion mechanism in terms of thermonuclear explosions of white dwarf (WD) stars \\citep[e.g.,][]{reinecke2002d,gamezo2003a,roepke2005b, roepke2007b, mazzali2007a, roepke2007c}. In order to judge potential systematic errors in SN~Ia cosmology, a theoretical connection between the explosion characteristics and properties of the progenitor system would be desirable. Yet despite all efforts on both the theoretical and on the observational side, the nature of the progenitor system remains enigmatic. The stabilization of WDs against gravity does not depend on a finite energy source such as the nuclear burning in normal stars. Due to the Fermi pressure of a degenerate electron gas, single WD stars are in principle eternally stable. Thus, some additional dynamics is required in order to reach an explosive state. The most likely possibility is a WD being part of a binary system and accreting matter from its companion. Current progenitor models distinguish between the \\emph{single degenerate} and the \\emph{double degenerate} scenario. The former \\citep[proposed by][]{whelan1973a} assumes a ``normal'', non-degenerate star to be the binary companion -- either a main sequence (MS) star or a red giant (RG). In this case, the WD accretes mass from its MS or RG companion via Roche-lobe overflow or by winds (symbiotic systems) until it approaches the Chandrasekhar mass. The densities reached in the core of the WD are then sufficiently high to trigger nuclear reactions which finally cause a thermonuclear explosion of the star \\citep[but note that an explosion before reaching the Chandrasekhar mass may also be possible, e.g.][]{fink2007a}. The double degenerate scenario \\citep{iben1984a, webbink1984a}, on the other hand, assumes a binary system of two WDs. Due to gravitational wave emission, the system becomes unstable at some point, the WDs merge, and may eventually explode in a SN Ia. A summary of arguments in favor of and against both scenarios can be found in \\cite{livio2000a}. In theoretical modeling, the single-degenerate Chandrasekhar-mass scenario has received most attention recently. By constraining the amount of fuel available in the thermonuclear explosion, it provides a natural explanation for the observed uniformity of SNe~Ia. Observationally, its hard to distinguish between the progenitor scenarios. One fundamental difference is the complete absence of hydrogen in the double degenerate scenario as it assumes a merger of two carbon/oxygen WDs. In contrast, in the typical single degenerate scenario, hydrogen is the main constituent of the companion. An exception are helium-accretors in which a more evolved companion star has lost his hydrogen envelope. Thus, the WD accretes helium instead of hydrogen and hydrogen is missing in the system. \\citet{kato2003a} reported a possible detection of such an object. In the standard scenario, however, the companion star features a hydrogen envelope; and at least some part of it is expected to be carried away by the SN~Ia ejecta impacting the companion. This, in principle, causes a problem for the single-degenerate scenario, because the astronomical classification of SNe~Ia rests on the absence of hydrogen-features in the spectra of these events. The hydrogen stripped off from the companion will have rather low velocities. It may thus be detectable in nebular spectra, if abundant enough. For the single-degenerate scenario it is therefore of critical importance that the mass of stripped material is sufficiently low to be still consistent with the observations. There have been a few attempts to search for hydrogen in nebular spectra of SN Ia. \\citet{mattila2005a} studied nebular spectra of SN 2001el. From modeling them, they derived an upper limit of $0.03\\, \\mathrm{M}_{\\odot}$ of solar abundance material at velocities lower than $1000\\, \\mathrm{km}\\, \\mathrm{s}^{-1}$. Recently, \\citet{leonard2007a} studied nebular spectra of SN~2005am and SN~2005cf. Based on the same model as \\citet{mattila2005a}, he estimated $\\lesssim$$0.01\\, \\mathrm{M}_{\\odot}$ of hydrogen material for both objects. Recently also hydrogen has been detected indirectly by \\citet{patat2007a} in circumstellar material of SN 2006X. An alternative to this approach of observationally constraining the nature of the progenitor system is to directly search for the former companion star of the single-degenerate scenario in the remnants of historical galactic SNe~Ia. Such a search has been carried out in the remnant of Tycho Brahe's supernova of 1572 by \\citet{ruiz-lapuente2004a}, who claimed the identification of the binary companion. The star in question is a slightly evolved solar-type star, that moves with a radial velocity of $-108 \\, \\mathrm{km\\ s^{-1}}$ relative to the sun. It also has an atypical large tangential velocity of about $90 \\, \\mathrm{km\\ s^{-1}}$. A significantly larger velocity of the star compared to neighbours is expected as a result of the disappearing binary orbit. Other stars observed in the same area with similar distances move only with average radial velocities of about $-20$ to $-40 \\, \\mathrm{km\\ s^{-1}}$, with a velocity dispersion of about $20 \\, \\mathrm{km\\ s^{-1}}$. On the theory side, \\citet{marietta2000a} presented two-dimensional hydrodynamical simulations of the impact of SNe~Ia on their companions. They found that $0.15 \\, \\mathrm{M}_{\\odot}$ were stripped from a Roche-lobe filling MS companion. This would rule out a MS-WD system for the SN~Ia analysed by \\citet{leonard2007a}, if it was representative. Recently, \\citet{meng2007a} pointed out that considering the effect of the mass transfer phase on the companion star may change the result significantly. They studied the impact of SNe~Ia on different companion stars analytically. In contrast to \\citet{marietta2000a}, who assumed the structure of single MS stars for the companion, \\cite{meng2007a} evolved it through the binary evolution phase before the explosion of the WD. They found at least $0.035\\, \\mathrm{M}_{\\odot}$ of stripped hydrogen for the companion. However, this result is only a lower limit, since they did not include mass loss by vaporization from the hot surface of the star. Thus, taken at face value, the currently available theoretical studies constitute a strong case against MS+WD progenitor systems for SNe~Ia. The aim of the study presented here is to check and update the \\citet{marietta2000a} calculations with the results of recent detailed binary evolution models. \\citet{ivanova2004a} identified possible SN~Ia progenitors from a parameter study of MS+WD binary evolution. There results are in agreement with other studies in this field \\citep[e.g.][]{langer2000a, han2004a}. Based on these results, we present an exploration of the effect of the impact of a SN Ia on different MS companions by 3D hydrodynamical simulations. Section \\ref{sec:code} summarizes the codes used. Section \\ref{sec:tests} demonstrates that our approach reproduces the results of \\cite{marietta2000a} and presents a resolution study. Section \\ref{sec:results} discusses an exploration of different progenitor systems and Section \\ref{sec:observ} derives observational implications of our results. A summary and an outlook conclude this work in Section \\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} We studied the impact of the ejecta of SNe~Ia on main sequence companion stars in the context of the single-degenerate Chandrasekhar-mass scenario with hydrodynamical simulations. To this end, the (cosmological) GADGET2 smoothed particle hydrodynamics code was adapted to the stellar problem and employed in numerical simulations. It was shown that this SPH-based approach is capable of reproducing previous results obtained with a grid-based 2D scheme by \\cite{marietta2000a}. A resolution study indicated that with a few million particles the simulations yield numerically converged results. We showed that the mass stripped from the companion star by the impact of the ejecta depends on the their kinetic energy and the binary separation. While the latter affects the mass of the stripped material significantly, the former has only a minor effect on it. This is due to the fact that the supernova explosion energy can only vary in a relatively narrow range given the restricted amount of fuel available for the nuclear energy generation. The SPH approach was used to analyze the impact in a number of more realistic progenitor models than those employed in previous studies. For these, the companion stars were constructed with the stellar evolution code GARSTEC mimicking binary mass transfers with the parameters given by \\citet{ivanova2004a}. In the hydrodynamical impact simulations, we found about one order of magnitude less hydrogen material stripped off the companion by the impact of the supernova than predicted by previous studies. The main reason for this difference is a modified, more compact stellar structure of the companion star in combination with a resulting variation in the separation distance of the progenitor system. In particular the more compact state of the companion impedes the mass loss in the impact. The reduced amount of hydrogen mixed into the ejecta of the supernova as predicted by our simulations leads to an agreement with observational studies of SN~Ia nebular spectra \\citep{mattila2005a,leonard2007a}. This removes the former disagreement between the available observations and simulations of the WD+MS progenitor system. Thus, to current knowledge, such a progenitor scenario is admissible in the context discussed here. However, since the hydrogen masses predicted by our simulations are not far below the current observational upper limits, it may be possible in the near future to confirm or reject the studied progenitor scenario by either detecting hydrogen in SNe~Ia or lowering the limits by another order of magnitude. A stringent way of analysis would be to calculate synthetic spectra directly from the presented simulations and to compare the results with observations. This will be tackled in a forthcoming study." }, "0807/0807.1334_arXiv.txt": { "abstract": "Investigating the link between supermassive black hole and galaxy evolution requires careful measurements of the properties of the host galaxies. We perform simulations to test the reliability of a two-dimensional image-fitting technique to decompose the host galaxy and the active galactic nucleus (AGN), especially on images obtained using cameras onboard the {\\it Hubble Space Telescope (HST)}, such as the Wide-Field Planetary Camera 2, the Advanced Camera for Surveys, and the Near-Infrared Camera and Multi-Object Spectrometer. We quantify the relative importance of spatial, temporal, and color variations of the point-spread function (PSF). To estimate uncertainties in AGN-to-host decompositions, we perform extensive simulations that span a wide range in AGN-to-host galaxy luminosity contrast, signal-to-noise ratio, and host galaxy properties (size, luminosity, central concentration). We find that realistic PSF mismatches that typically afflict actual observations systematically lead to an overestimate of the flux of the host galaxy. Part of the problem is caused by the fact that the {\\it HST}\\ PSFs are undersampled. We demonstrate that this problem can be mitigated by broadening both the science and the PSF images to critical sampling without loss of information. Other practical suggestions are given for optimal analysis of \\hst\\ images of AGN host galaxies. ", "introduction": "The mass of supermassive black holes (BHs) is strongly correlated with the luminosity (\\lbul; Kormendy \\& Richstone 1995; Magorrian et al. 1998) and the stellar velocity dispersion (\\vel; Gebhardt et al. 2000; Ferrarese \\& Merritt 2000) of the bulge of the host galaxy. These scaling relations are often interpreted to be evidence that central BHs and their host galaxies are closely connected in their evolution (see reviews in Ho 2004). The empirical correlations between BH mass and host galaxy properties can even be used as tools to track the progress of mass assembly during galaxy evolution (Peng et al. 2006a, 2006b; Woo et al. 2006; Ho 2007b). The central BH mass correlates most strongly with the bulge component of a galaxy rather than with its total mass or luminosity. This has motivated detailed bulge-to-disk decompositions of the host galaxies to better quantify the intrinsic scatter of the \\mbh-\\mbul\\ and \\mbh-\\lbul\\ relations in the local Universe (Marconi \\& Hunt 2003; H\\\"{a}ring \\& Rix 2004). To probe when the BH-host galaxy relations were established and how they evolved, it is of paramount importance to extend similar studies out to higher redshifts. However, direct measurement of BH mass based on spatially resolved stellar or gas kinematics is unfeasible for all but the nearest galaxies with low levels of nuclear activity. Accessing BHs and their host galaxies at cosmological distances requires a different approach---one that relies on active galactic nuclei (AGNs). The masses of BHs in type~1 (unobscured, broad-line) AGNs can be readily estimated with reasonable accuracy using the virial technique with single-epoch optical or ultraviolet spectra (Ho 1999; Wandel et al. 1999; Kaspi et al. 2000; Greene \\& Ho 2005b; Peterson 2007). Quantitative measurements of the host galaxies with active nuclei, on the other hand, are less straightforward to obtain because the presence of the AGN introduces significant practical difficulties, as well as potential biases. This is especially problematic for the bulge component of the host, which is maximally affected by the bright AGN core. A variety of techniques have been employed, both kinematical (e.g., Nelson et al. 2004; Onken et al. 2004; Barth et al. 2005; Greene \\& Ho 2005a, 2006; Ho 2007; Salviander et al. 2007; Ho et al. 2008; Shen et al. 2008) and photometric (e.g., McLure \\& Dunlop 2002; Peng et al. 2006b; Greene et al. 2008). \\begin{figure*} \\psfig{file=table1.ps,width=18.5cm,angle=0} \\vskip -0.4cm \\end{figure*} The main challenge for photometric studies of active galaxies lies in separating the central AGN light from the host galaxy. While structural decomposition can be done by fitting analytic functions to the one-dimensional (1-D) light distribution, two-dimensional (2-D) analysis makes maximal use of all the spatial information available in galaxy images (Griffiths et al. 1994; Byun \\& Freeman 1995; de~Jong 1996; Wadadekar et al. 1999; Peng et al. 2002; Simard et al. 2002; de~Souza et al. 2004) and thus provides the most general and most robust method to decouple image subcomponents. This flexibility proves to be especially important in the case of active galaxies, where the contrast between the central dominant point source and the underlying host can be very high. In this regime, achieving a reliable decomposition requires knowing the point-spread function (PSF) to high accuracy. This paper discusses the complications and systematic effects involved in photometric decomposition of AGN host galaxies, especially as it applies to images taken with the {\\it Hubble Space Telescope (HST)}. Although numerous {\\it HST} studies of AGN hosts have been published, very few have explicitly investigated the systematic uncertainties or practical limitations of host galaxy decomposition. We make use of an updated version of the 2-D image-fitting code GALFIT (Peng et al. 2002) to generate an extensive set of simulated images of active galaxies that realistically mimic actual {\\it HST} observations. We then apply the code to fit the artificial images and quantify how well we can recover the input parameters for the AGN and for the host galaxy. Our strategy resembles those employed in the quasar host galaxies studies by Jahnke et al. (2004) and S\\'{a}nchez et al. (2004), except that we have optimized our simulations to be applicable to nearby bright AGNs, a regime of most interest to us (Kim et al. 2008). Nearby bright AGN hosts observed using coarse pixels and low exposure time can be equally challenging to analyze as high-z AGNs observed with high resolution and signal-to-noise. In other words, fundamentally, the difficulty of the analysis depends only on the following 3 relative conditions: angular resolution of the detector vs. the object being studied (i.e. object scale size in pixels), AGN-to-host contrast, and overall object signal-to-noise. Since we cover a wide range of parameter space the simulations described here should be applicable to most {\\it HST}\\ imaging studies of AGN hosts, including distant quasars. We hope that the results of this study can serve as a guide to other investigators working with similar data. We describe the details of PSF variations in \\S{2}. In \\S{3}, we present the simulation procedure and our results from fitting artificial images using PSF models with varying degrees of realism. A discussion and summary of the main results are given in \\S{4}. ", "conclusions": "In this study, we performed 2-D image-fitting simulations of AGN hosts to illustrate how systematics in the PSF mismatch may affect the deblending of the AGN and the host galaxy components. Based on these simulations, we suggest practical strategies for how to observe and analyze host galaxies with bright active nuclei. As seen in Figures 6--7, careful determination of the PSF is needed to extract very high-contrast subcomponents accurately. This point was also underscored in the study of Hutchings et al. (2002). We find that the three factors that most affect PSF structures are spatial distortion, temporal changes, and pixel undersampling. The SED of the PSF matters to a much lesser degree, but it is worthwhile to match it when possible. The spatial variation can be reduced by observing a stellar PSF at the same position as the science images. However, the temporal variance is trickier to avoid without observing PSFs concurrently with the science data, which is observationally expensive and rarely done in practice. Even then, some temporal variability may happen even within a single orbit. Nevertheless, as there is evidence that PSF mismatch grows over time, and may not be completely periodic in nature, it is important to obtain a stellar PSF as close in time as possible to the science images. From our tests (\\S{2.1}), we recommend that stellar PSF images be taken within a month of the science images. In the absence of a well-matched stellar PSF, synthetic PSFs generated with TinyTim are adequate substitutes (compare Fig. 10 to Fig. 7 and Fig. 9). In all scenarios, whether using stellar or TinyTim PSFs, both the science image and the PSF image should be oversampled to reduce errors caused by pixel undersampling and subpixel shifting. This can be accomplished by observing the science image and the PSF star using a four-point ``dither'' pattern to recover finer pixel sampling. Or, if this option is not available, then simply broadening the images through convolution during the analysis stage is an acceptable alternative solution, and certainly better than no treatment at all (\\S{3.4} and Fig. 10). \\begin{figure*} \\figurenum{15} \\psfig{file=f15.eps,width=18.5cm,angle=0} \\figcaption[fig15.ps] {Distribution of residuals for the magnitudes of the host for ({\\it left}) \\hnr $\\geq 0.2$ and ({\\it right}) \\hnr $< 0.2$. After producing artificial images by varying $n$ between 2 and 6, we fit them with $n$ fixed to 4. This simulation gives an estimate of the measurement error for the bulge luminosity. \\label{fig15}} \\end{figure*} When the host galaxies are faint and difficult to fit, it may be useful to hold the \\ser\\ index fixed to a constant value of either $n=1$ or $n=4$. The decision about what value to use might be based on visual morphology, or by comparing $\\chi^2$ values of the bimodal priors. A similar conclusion was reached by McLure et al. (1999), Jahnke et al. (2004), and S\\'{a}nchez et al. (2004) in their analysis of quasar host galaxies. By varying the \\ser\\ index between $n$ = 2 and $n$ = 6, we find that the scatter is $\\sigma \\approx 0.3 $ mag for \\hnr $\\geq 0.2$ and \\nsn\\ $\\approx 1000$ (Fig. 15). At higher contrast, \\hnr $\\leq 0.2$, the scatter increases to $\\sigma \\approx 0.4 $ mag. Lastly, we briefly conducted a three-component bulge/disk/AGN decomposition to characterize uncertainties in estimating the bulge component. We find that when the $S/N$ is high, the contrast is sufficiently low, and the bulges are sufficiently well-resolved, then a B/D/A decomposition can yield reliable bulge luminosity measurements. These simulations are fully compatible with images of $z\\lesssim 0.3$ quasar host galaxies (e.g., McLure \\& Dunlop 1999). Our tests suggest that, when three-component decomposition is required, the uncertainty in the bulge luminosity increases by an additional $\\sim$0.1 mag compared to fits without a disk component. Finally, we note in passing that our simulations allow the sky value to vary during the fitting. For the current application, this choice makes little difference because the images have a large sky area and the profiles are idealized. However, in real science images, the situation will be different. From prior experience with actual \\hst\\ data, keeping the sky value fixed prevents the \\ser\\ index from going up to extremely high values in situations where the light profile of the galaxy is not well represented by the \\ser\\ function. Thus, we recommend keeping the sky parameter fixed to a well-determined value whenever possible." }, "0807/0807.1116_arXiv.txt": { "abstract": "The outer regions of disk galaxies show a drop-off in optical and H$\\alpha$ emission, suggesting a star formation threshold radius, assumed to owe to a critical surface density below which star formation does not take place. Signs of filamentary star formation beyond this threshold radius have been observed in individual galaxies in the H$\\alpha$ and recent GALEX surveys have discovered that 30\\% of disk galaxies show UV emission out to 2-3 times the optical radius of the galaxy. We run smooth particle hydrodynamics simulations of disk galaxies with constant density extended gas disks to test whether over-densities owing to spiral structure in the outer disk can reproduce the observed star formation. We indeed find that spiral density waves from the inner disk propagate into the outer gas disk and raise local gas regions above the star formation density threshold, yielding features similar to those observed. Because the amount of star formation is low, we expect to see little optical emission in outer disks, as observed. Our results indicate that XUV disks can be simulated simply by adding an extended gas disk with a surface density near the threshold density to an isolated galaxy and evolving it with fiducial star formation parameters. ", "introduction": "The most basic description of galactic-scale star formation is the Schmidt relation, which assumes that the star-formation rate (SFR) is a power law function of the gas density \\citep{Schmidt-1959}. \\citet{Kennicutt-1998} showed that this relation accurately describes star formation in galaxies over a range of $10^{5}$ in gas surface density and $10^{6}$ in SFR per unit area. However, beyond a critical galactocentric radius, SFRs determined from H$\\alpha$ emission are truncated relative to the expectation from the Kennicutt-Schmidt relation despite the availability of large reserves of low density cold gas \\citep{Kennicutt-1989, Martin-Kennicutt-2001, Wong-Blitz-2002}. The lack of star formation at large radii has primarily been attributed to dynamical disk stability, i.e., \\begin{equation} Q(r) = \\frac{\\Sigma_{c}(r)}{\\Sigma_{gas}(r)} = \\frac{\\kappa(r) c_{s}}{\\pi G \\Sigma_{gas}(r)} \\gsim 1 \\, , \\end{equation} \\noindent where $\\kappa$ is the epicyclic frequency at a given radius, $\\Sigma_{c}(r)$ is the critical surface density and $c_{s}$ is sound speed of the gas, proportional the velocity dispersion of the gas ( $\\sigma=c_{s}\\gamma^{-1/2}$, where $\\gamma$ is the ratio of specific heats) \\citep{Toomre-1964, Spitzer-1968, Quirk-1972}. The effective velocity dispersion of gas in spirals is roughly constant across the disk \\citep[e.g.,][]{vanderKruit-Shostak-1984, Anderson-et-al-2006} leaving $\\kappa$ as the only radially dependent quantity in the critical surface density. If velocity curves of spiral galaxies are approximately flat, $\\kappa$, and therefore $\\Sigma_{c}$, fall with $r$, implying that at some radius the disk becomes gravitationally stable, which may inhibit star formation. In support of this notion, several studies have shown correlations between the radius at which star formation ceases and the radius at which $Q\\approx 2$ \\citep[][see \\citet{Schaye-2004} for a comparison of these results]{Kennicutt-1989, Martin-Kennicutt-2001, Wong-Blitz-2002}. More sophisticated models including the formation of cold molecular gas have also been successful at reproducing observed SFR truncations \\citep{ Elmegreen-Parravano-1994, Schaye-2004} . Observations indicate the presence of star formation beyond the threshold. Faint, isolated H\\,II regions distributed in spiral structures out to two optical radii have been found in face on galaxies \\citep{Ferguson-et-al-1998} and in cold disks in edge-on galaxies \\citep{Christlein-Zaritsky-2008}. Similar star formation has been seen in the Milky Way \\citep{Brand-et-al-2001} and dwarf galaxies with gas densities below the threshold \\citep{vanZee-et-al-1997}. A systematic survey by the Galaxy Evolution Explorer (GALEX: \\citet{Martin-et-al-2005}) recently discovered that 30\\% of disk galaxies show extended UV emission beyond the optical radius, and presumed star formation threshold, of the galaxy \\citep{Thilker-et-al-2007, Gildepaz-et-al-2005, Thilker-et-al-2005, Zaritsky-Christlein-2007}. In 2/3 of these cases, this emission is structured, lying in spiral or filamentary patterns (Type I XUV disks). In the other 1/3, the emission takes the form of a large zone in the outer regions of the galaxy with an enhanced UV/optical ratio relative to the inner disk (Type II XUV disk) \\citep[][hereafter T07]{Thilker-et-al-2007}. To reconcile outer disk star formation with disk stability, it has been proposed that local densities above $\\Sigma_{c}(r)$ in the outer disk allow star formation beyond the truncation radii \\citep{Kennicutt-1989, Martin-Kennicutt-2001, Schaye-2004, Elmegreen-Hunter-2006, Gildepaz-et-al-2007}. The low SFR implies a low number of O stars producing H$\\alpha$ emission at a given time, or none at all, explaining why this star formation was undetected in the H$\\alpha$ in some galaxies \\citep{Gildepaz-et-al-2007}. Most of these disks have large amounts of H\\,I in their outer regions and UV and H$\\alpha$ complexes are often coincident with local H\\,I over-densities \\citep[][T07]{Ferguson-et-al-1998} supporting this picture. Analysis of the star formation profiles of these XUV disk galaxies indicates that the azimuthally averaged Kennicutt-Schmidt relation could hold beyond the dynamical threshold radius \\citep[][Hereafter B07]{Boissier-et-al-2007}. However, studies of star forming clumps in the XUV disk of M\\,83 indicate that traditional star formation laws, including the threshold radius, hold locally \\citep{Dong-et-al-2008}. Theoretically, \\citet{Elmegreen-Hunter-2006} have shown that XUV disks may not be inconsistent with a star formation density threshold. In their analytical model, gas clumping triggered by spiral density waves, radial variations in the interstellar medium (ISM) turbulence, and gas phase transitions lead to local regions of star formation in the extended gas disk. In this paper, we investigate XUV disks using smooth particle hydrodynamic (SPH) simulations of an isolated disk galaxy with an extended gas disk and a simple, commonly employed, star formation prescription based on the Kennicutt-Schmidt relation that includes a threshold density for star formation. We demonstrate that, even with these fiducial star formation laws, spiral structure propagates from the inner disk to the extended H\\,I disk and produces local regions of enhanced of gas density which trigger star formation similar to that seen in observed Type I XUV disks. ", "conclusions": "\\label{sec:conc} We show that features similar to those seen in Type I XUV disks can be produced with fiducial star formation prescriptions, including a star formation threshold density, if there is gas present at large radii near the threshold density. This is a clear illustration of the fact that a star formation threshold density does not imply the existence of a star formation threshold radius beyond which star formation ceases. As proposed in earlier studies and shown by \\citet{Elmegreen-Hunter-2006}, local over-densities, caused by spiral density waves in our simulation, allow star formation in gas that, when circularly averaged, is below or near the classical star formation threshold density. It is possible that star formation will differ between inner and outer disks. This could arise owing to different properties of the ISM in the inner and outer regions of disks, from, for example, a different ionized fraction owing to various levels of self-shielding against an ionizing UV background. Currently, our simulations do not have the resolution to probe the detailed physics of star formation. Here, we simply show that any simulation whose star formation prescription follows a Kennicutt-Schmidt relation with a star formation threshold density and has a high enough gas density for local over-densities to exceed the star formation threshold density in the outer disk will exhibit UV emission in the outer disk. Understanding the mechanism for generating XUV disks has promising implications for probing galaxy structure in the outer parts of galaxies. If the distribution of H\\,I can be determined based on the observed UV emission, this can be used to infer the extended H\\,I properties of disks. Outer disk H\\,I and its star formation history could also constrain recent gas accretion in galaxies and the formation theory of disks. For these reasons, the generation of XUV disks warrants further exploration. Simulations having different extended gas profiles (e.g. exponential, $1/R$) at various gas densities are needed to test whether we can create the range of properties seen in XUV disks, including Type II XUV disks. Upcoming extragalactic H\\,I surveys such as THINGs \\citep{Walter-et-al-2005} and ALFALFA \\citep{Giovanelli-et-al-2005} will provide better statistics on the types of H\\,I profiles seen on disk galaxies and how they correlate with XUV disk behavior. The role of interactions, particularly flybys, in exciting spiral structure and affecting the morphology of XUV disks also needs to be explored. Finally, the application of a radiative transfer code to our SPH results to compare colors of these clumps to samples such as T07 and \\citet{Zaritsky-Christlein-2007} would provide interesting constraints on star formation such as the initial mass function in outer disks." }, "0807/0807.1927_arXiv.txt": { "abstract": " ", "introduction": "\\paragraph{} Inflation~\\cite{inflation} has emerged as the standard paradigm describing physics of the very early universe. Besides addressing several fine-tuning issues in big bang cosmology such as the flatness and horizon problems, it provides a framework to explain the origin of structure and the cosmic microwave background (CMB) anisotropy~\\cite{books}. While there is a plethora of effective field theory based models of inflation~\\cite{Lyth:1998xn}, many outstanding questions in inflationary cosmology {\\it require} a fundamental microscopic description. Conversely, recent observations of the CMB and large scale structure~\\cite{observation} lead us to increasingly precise measurements of the inflationary parameters. These measurements provide us with an exciting window to probe physics at ultra-high energies~\\cite{reconstruction}, far higher than what current and upcoming terrestrial accelerators can reach. Thus inflationary cosmology has become the perfect arena for fundamental theory to meet experiment. String theory is currently our leading candidate for a quantum theory of gravity. Thus it is worthwhile to explore explicit realizations of inflation within this framework. In this paper, we will focus on one of the most well developed inflationary scenarios in string theory, i.e. $D$ brane inflation~\\cite{DvaliTye} (see also Refs.~\\cite{Dvali:2001fw,Burgess:2001fx}; for reviews, see Ref.~\\cite{reviews} and references therein), where the inflaton field is identified with the position of a space-filling mobile $D$ brane, usually a $D3$ brane, in a warped six dimensional manifold~\\cite{KKLMMT}. In the original scenario of Refs.~\\cite{DvaliTye,Dvali:2001fw,Burgess:2001fx,reviews,KKLMMT}, an additional $\\overline{D3}$ was introduced to drive inflation. The $\\overline{D3}$ brane is localized by the RR fluxes at the tip of a warped throat, thus inflation proceeds as the mobile $D3$ is attracted by a weak $D3$-$\\overline{D3}$ Coulombic force to move slowly along the warped direction. However, it was also noted in Ref.~\\cite{KKLMMT} that because the volume modulus of the compactification couples non-trivially to the canonical inflaton, its stabilized value gives additional Hubble scale correction to the inflaton mass, causing the well known $\\eta$ problem~\\cite{Copeland:1994vg}. An important step towards addressing the $\\eta$ problem explicitly in this concrete setting was recently made in Refs.~\\cite{Baumann0, Baumann1} (see also Refs.~\\cite{Panda:2007ie, McGill1, Krause:2007jk, Pajer:2008uy}). The key ingredient in the construction was the one loop threshold correction to the non-perturbative superpotential obtained in Ref.~\\cite{Gaugino} (see also Refs.~\\cite{Ganor,BHK}). In Ref.~\\cite{KKLT} and other stabilized compactifications, non-perturbative effects are often introduced to stabilize moduli. In the context of Ref.~\\cite{KKLT}, such effects come from instantons on a stack of $D7$ branes (or Euclidean $D3$ branes). Interestingly, the non-perturbative moduli stabilizing force also turns out to give the dominant contribution to the inflaton potential\\footnote{Although the Coulombic force is subdominant in comparison to the moduli stabilizing force, a $\\overline{D3}$ brane was still introduced to end inflation.}. This contribution arises because the mobile $D3$ brane backreacts on the moduli stabilizing $D7$ branes. The correction depends on the holomorphic four cycles within the conifold on which the $D7$ branes wrap. The embedding of the $D7$ branes breaks the isometry of the deformed conifold, and thus the inflationary trajectory depends sensitively on the choice of the embedding function {defining the loci of the $D7$ branes}. As a result, explicit slow-roll models have been constructed by a ``delicate'' tuning of the microscopic compactification parameters. While the broken angular isometry directions are stabilized by the coordinate dependent non-perturbative superpotential, for a given $D7$ brane embedding, there are typically residual isometries preserved by the resultant scalar potential. The potential for the fields associated with these isometries remain flat during the inflationary epoch and so they can take arbitrary values without affecting the inflationary trajectory. Being almost massless, their quantum fluctuations give rise to a nearly scale invariant isocurvature perturbation spectrum. As argued by Lyth and collaborators in Refs.~\\cite{Lyth,Lyth1}, these isocurvature perturbations can be converted to the curvature perturbations at the end of inflation. In the context of $D$ brane inflation, inflation ends when the open string tachyon condenses between the mobile $D3$ and $\\overline{D3}$. The critical value of the canonical inflaton at which inflation ends $\\phi_\\tip$ depends on the residual symmetries as they enter into the tachyon potential. Since $\\phi_\\tip$ picks up spatial dependence through the quantum fluctuations of the light residual symmetries, inflation can end on a spatial slice of non-uniform energy density. {As we will see, this is the case for instance when the inflaton potential is dominated by the moduli stabilizing force towards the end of inflation. Thus,} one could in principle expect potentially significant contribution to the power spectrum and non-Gaussianities at the end of inflation. In this paper, we study these multi-field effects at the end of brane inflation, and outline the necessary conditions for them to be significant. We then perform a case study for the setup considered in Ref.~\\cite{Baumann1}, by explicitly calculating the canonical inflaton potential near the tip of the deformed conifold, and demonstrate that inflation can persist in this region provided that the $D3$-$\\overline{D3}$ Coulombic attraction becomes subdominant. We also discuss various mechanisms to uplift the vacuum energy which results in a subdominant Coulombic potential all the way to the tip of a warped throat. We also show that the angular stable inflationary trajectory for the specific $D7$ brane embedding \\cite{Kuper} used in Ref.~\\cite{Baumann1} can be extended to the entire deformed conifold. However, along the specific trajectory considered in Ref.~\\cite{Baumann1}, we will see explicitly that the corresponding residual angular isometries have vanishing proper separations at the tip. Thus, for this specific $D7$ embedding, {no significant contribution to the curvature perturbation is generated at the end of inflation}. This implies that while multi-field effects can in principle be significant in brane inflation, they can only happen with other $D7$ embeddings, or with more than one stacks of $D7$ branes present. This paper is organized as follows. In Section~\\ref{sec_D3brane}, we review the basic setup of flux compactification and brane inflation, in order to set up our notation. Readers who are familiar with the above topics can skip this section. In Section~\\ref{Lytheffect}, we recast the mechanism proposed in Ref.~\\cite{Lyth} in the context of brane inflation in a warped throat, and outline the necessary conditions for it to take place. In Section~\\ref{sec_scenario}, we discuss various possible uplifting mechanisms in a warp throat and propose a natural scenario for an uplifted potential to realize the effect of Ref.~\\cite{Lyth}. In Section~\\ref{sec_endofinf}, we explicitly calculate the canonical inflaton potential near the tip of the deformed conifold and the resulting slow-roll parameters. The degeneracy of the residual isometries will also be shown. We end with some discussions in Section~\\ref{Discussion}. We relegate most of the calculational details in a number of Appendices. ", "conclusions": "\\label{Discussion} \\setcounter{equation}{0} In this paper, we studied the systematics of multi-field effects at the end of warped $D$ brane inflation. We discussed the necessary criteria for the isocurvature perturbations generated by the angular motion of a mobile $D3$ brane to be converted into the curvature perturbations usually associated with its radial motion in this scenario. We found that the significance of the end of inflation effects considered in Ref.~\\cite{Lyth} depends on the specific mechanism for uplifting the vacuum energy. If the uplifting is due to some distant $\\overline{D3}$ branes or a $D$-term potential, the Coulombic potential can easily become subdominant even towards the end of inflation, and the effects described in Ref.~\\cite{Lyth} can in principle be significant. However, in the most explicit $D$ brane inflation constructed to date \\cite{Baumann0,Baumann1}, the $D7$ brane embedding chosen \\cite{Kuper} does not yield such effects, { regardless of the uplifting mechanism.} This latter result is specific to the embedding of the moduli stabilizing branes as well as the infrared geometry of the throat. % Along the stable trajectory for the embedding considered in Ref.~\\cite{Kuper}, the proper distance for the residual isometry direction vanishes in the entire throat, the moduli space vanishes at the tip. % It would be interesting to examine other $D7$ brane embeddings and/or other warped throats which leave a moduli space of vacua at the tip. Examples of such embeddings for the deformed conifold appeared in Ref.~\\cite{D3vacua}, where the residual isometry directions reside on the finite size $S^3$. However, finding an angular stable trajectory in these examples may remain challenging. Nevertheless, our results underscore the importance of multi-field effects in string inflation, as noted also in the context of DBI inflation recently in Ref.~\\cite{Langlois} (see also earlier discussions in Refs.~\\cite{Huang:2007hh,Easson:2007dh}). As discussed in Section~\\ref{Lytheffect}, {the strength of the Lyth effect} depends on the ratio $\\varepsilon_k/\\varepsilon_e$. Since the flat region of the inflaton potential considered in Refs.~\\cite{Baumann0,Baumann1} is an inflection point, $\\varepsilon_k$ depends sensitively on where around the inflection point corresponds to the CMB scale. Given a $D$ brane inflation model which can yield {the Lyth effect} considered here, a precise determination of the amplitude of such effects would require the use of the full KS metric~\\cite{KS}. {This is yet another context in which details of the warped geometries in the infrared can have significant effects on the CMB observations~\\cite{inflation_tip}.} {Furthermore, regardless of the {Lyth effect} studied here,} a detailed comparison of the WMAP data with microscopic parameters of $D$ brane inflation requires identifying the relevant part of the inflaton potential which generates the observed CMB {anisotropy}, and the full KS metric is essential. Work along these lines is underway. Finally, one may hope to also realize the curvaton mechanism~\\cite{curvaton} using these light fields. In the setup {we discussed}, however, inflation ends as $D3$ and $\\overline{D3}$ annihilate and thus the would-be curvaton fields themselves disappear. For the same reason, any multi-field effect~\\cite{multi} after inflation will not be present as long as they are associated with $D3$ or $\\overline{D3}$ branes. Nevertheless it would be interesting to implement the curvaton scenario in a different setup satisfying a number of constraints~\\cite{Gong:2006hf}. \\subsection*{Acknowledgement} \\paragraph{} We are indebted to Bret Underwood for numerous valuable insights and discussions, and for collaboration at the initial stage of this project. We are also grateful to Misao Sasaki for discussions and comments on the manuscript. We thank Daniel Baumann, Chong-Sun Chu, Min-Xin Huang, David Lyth, Fernando Marchesano, Liam McAllister, Peter Ouyang, Sudhakar Panda, and Fernando Quevedo for helpful discussions. HYC would like to thank KITP at USCB and the Sixth Simons Workshop at SUNY Stony Brook for their hospitalities where part of the work was being carried out. JG is grateful to the Santa Fe 08 Cosmology Summer Workshop for hospitality where this work was being finished. The work of HYC and GS is supported in part by NSF Career Award No. PHY-0348093, DOE grant DE-FG-02-95ER40896, a Research Innovation Award and a Cottrell Scholar Award from Research Corporation, and a Vilas Associate Award from the University of Wisconsin. JG is partly supported by the Korea Research Foundation Grant KRF-2007-357-C00014 funded by the Korean Government. \\appendix" }, "0807/0807.3439_arXiv.txt": { "abstract": "{We discuss the effect of atmospheric dispersion on the performance of a mid-infrared adaptive optics assisted instrument on an extremely large telescope (ELT). Dispersion and atmospheric chromaticity is generally considered to be negligible in this wavelength regime. It is shown here, however, that with the much-reduced diffraction limit size on an ELT and the need for diffraction-limited performance, refractivity phenomena should be carefully considered in the design and operation of such an instrument. We include an overview of the theory of refractivity, and the influence of infrared resonances caused by the presence of water vapour and other constituents in the atmosphere. `Traditional' atmospheric dispersion is likely to cause a loss of Strehl only at the shortest wavelengths (L-band). A more likely source of error is the difference in wavelengths at which the wavefront is sensed and corrected, leading to pointing offsets between wavefront sensor and science instrument that evolve with time over a long exposure. Infrared radiation is also subject to additional turbulence caused by the presence of water vapour in the atmosphere not seen by visible wavefront sensors, whose effect is poorly understood. We make use of information obtained at radio wavelengths to make a first-order estimate of its effect on the performance of a mid-IR ground-based instrument. The calculations in this paper are performed using parameters from two different sites, one `standard good site' and one `high and dry site' to illustrate the importance of the choice of site for an ELT.} ", "introduction": " ", "conclusions": "Atmospheric dispersion is a well known source of error in astronomical observations at visible and near-infrared wavelengths. With the advent of ELTs and their integrated AO facilities, atmospheric refractivity becomes a relevant quantity even in the mid-infrared. This is particularly the case if the traditional refractivity formulations are recalculated to include the effects of IR water and CO$_2$ resonances, which causes the refractivity to deviate substantially from the NIR values extrapolated to longer wavelengths. The calculations in this paper, though in many places approximate in nature, show the likely magnitude of refractive effects on the AO correction for a mid-IR ELT instrument. While `simple' dispersion will only cause a significant drop in Strehl at the short-wavelength end of the region, differential refraction effects between visible/NIR WFS and mid-IR science instrument may cause a significant image degradation over long exposures, even at moderate zenith distances. This problem in particular will be investigated more closely, taking into account the effect of multiple laser guide stars which can only be sensed in the visisble. Water vapour turbulence has recently been of interest in the interferometry and submillimetre communities, however, a thorough understanding of the phenomenon and experimental data are lacking (and non-existent in the framework of an ELT-sized filled aperture). Our first order estimates show that the Strehl loss can be significant, if not large, depending on the size of the outer scale. However, given the poor understanding of this type of turbulence both in the spatial and temporal domain, and of its correlation with dry air turbulence, these results show be regarded with caution. Further experimental results, specifically designed for ELT-type observations in the mid-IR, are required to help assess the problem. The comparison of results between the a Paranal-like site and a `high and dry' location like Cerro Macon is also very important, and reinforces the notion that a high and dry location is more suitable for mid-IR astronomy. However, we expect some of these advantages to apply to NIR also, and would invite NIR colleagues to examine site-related effects more closely. Correcting atmospheric dispersion on an ELT is a challenging task, even in today's telescopes. Solutions have been suggested for ELTs in recent years, but none have the capability of providing correction over a wide wavelength range (from visible to infrared). The problems will have to be addressed by a combination of technology (e.g. K-band wavefront sensing, NIR laser guide stars) and careful calibration. The first step, however, is a detailed theoretical understanding, and we intend to continue this work to make this possible." }, "0807/0807.1266_arXiv.txt": { "abstract": "Large-scale diffuse radio emission is observed in some clusters of galaxies. There is ample of evidence that the emission has its origin in synchrotron losses of relativistic electrons, accelerated in the course of clusters mergers. In a cosmological simulation we locate the structure formation shocks and estimate their radio emission. We proceed as follows: Introducing a novel approach to identify strong shock fronts in an SPH simulation, we determine the Mach number as well as the downstream density and temperature in the \\MNUniverse\\ simulation which has $2 \\times 1024^3$ particles in a $500\\:h^{-1}\\,{\\rm Mpc}$ box and was carried out with non-radiative physics. Then, we estimate the radio emission using the formalism derived in \\citet{hoeft:07} and produce artificial radio maps of massive clusters. Several of our clusters show radio objects with similar morphology to large-scale radio relics found in the sky, whereas about half of the clusters show only very little radio emission. In agreement with observational findings, the maximum diffuse radio emission of our clusters depends strongly on their X-ray temperature. We find that the so-called accretion shocks cause only very little radio emission. We conclude that a moderate efficiency of shock acceleration, namely $\\xi_{\\rm e} \\lesssim 0.005$, and moderate magnetic fields in the region of the relics, namely 0.07 to $0.8\\,{\\rm \\mu G}$ are sufficient to reproduce the number density and luminosity of radio relics. ", "introduction": "\\label{sec-intro} The large-scale structure of the Universe, composed of clusters, superclusters, filaments, and sheets of galaxies, is still in the process of formation. Overdense regions such as clusters and filaments keep accreting matter. Gas streams out of cosmic voids onto the sheets and filaments. When the newly accreted gas collides with the denser gas within these structures, shock fronts arise, dissipating the kinetic energy. In sheets and filaments, the gas follows the gravitational potential towards the clusters of galaxies. Eventually, the gas collides with the intra-cluster medium (ICM) with a few $1000\\:{\\rm km\\,s^{-1}}$ an gets heated to temperatures of $10^{7}$ to $10^{8}\\,{\\rm K}$. \\\\ The flow of gas is not as steady as the above picture suggests. A significant fraction of the gas accretion onto clusters is in the form of groups and clusters. The mergers of rich clusters are -- to our knowledge -- the most energetic events after the Big Bang. Kinetic energies of the order of $10^{64}$ erg are dissipated in giant shock waves. Only in recent years, X-ray telescopes have reached the necessary spatial and spectral resolution to detect the signatures of such shock waves in a few massive clusters. \\\\ A number of diffuse, steep-spectrum radio sources without optical identification have been observed in galaxy clusters. These sources have complex morphologies and show diffuse and irregular emission \\citep{kempner:01, slee:01, bacchi:03, feretti:05, giovannini:06}. They are usually subdivided into two classes, denoted as `radio halos' and `radio relics'. Cluster radio halos are unpolarised and have diffuse morphologies that are similar to those of the thermal X-ray emission of the cluster gas \\citep{giovannini:06}. Examples for clusters with radio halos are the Coma cluster \\citep{kim:89,deiss:97}, the galaxy cluster 1E\\,0657-56 \\citep{liang:00}, the X-ray luminous cluster A2163, and distant clusters such as CL\\,0016+16. The cluster A520 shows a halo with a low surface brightness with a clumpy structure \\citep{govoni:01}. Other examples can be found in \\citet{giovannini:99}. In general, radio halos are found in clusters with significant substructure and rich clusters with high X-ray luminosities and temperatures. The radio power correlates strongly with the cluster luminosity \\citep{feretti:05}. \\\\ Unlike halos, radio relics are typically located near the periphery of the cluster. They often exhibit sharp emission edges and many of them show strong radio polarisation \\citep{giovannini:04}. The sizes of relics and the distance to the cluster centre vary significantly. Examples for radio relic with sizes of one Mpc or even larger have been observed in Coma and A2256, which contain both a relic and a halo (as do A225, A1300, A2744 and A754). The cluster A3667 \\citep{roettgering:97} contains two very luminous, almost symmetric relics with a separation of more than five Mpc. The cluster A3376 shows an almost ring-like radio emission \\citep{bagchi:06}. The clusters A115 and A1664 show relics only at one side of the elongated X-ray distribution \\citep{govoni:01}. The relic source 0917+75 is particularly puzzling as it is located at 5 to $8\\,{\\rm Mpc}$ from the most nearby clusters. Other clusters show rather small relics as for example A85 \\citep{slee:01}. \\\\ The spectra of the diffuse radio sources indicate that their origin lies in synchrotron losses of relativistic electrons. The cooling time of the electrons which cause observable emission is of the order of one hundred Myr \\citep[see \\eg\\ the models in][]{slee:01}. The origin of the relativistic electron population which causes the emission is still not clear. There are essentially two classes of models that explain the presence of relativistic electrons. Either they are injected into the ICM via AGN activity or stellar feedback or they obtain their energy from particle acceleration at large-scale shock fronts in galaxy clusters. As discussed above, structure formation in the universe causes a variety of shock fronts in the intergalactic medium and the ICM \\citep{miniati:00, ryu:03, ryu:08}. These shock fronts are expected to be collisionless and capable of accelerating protons and electrons to relativistic energies. The correlation between the presence of diffuse radio emission in galaxy clusters and signs for a recent merger supports the scenario in which merger shocks generate the necessary relativistic electrons \\citep{feretti:06}. A3367 may serve as another piece of evidence: The radio relic is located where the merger induced bow shock is expected \\citep{roettiger:99}. Three mechanisms have been proposed for the action of the shock wave: (i) in-situ diffusive shock acceleration of electrons by the Fermi\\,I process \\citep[primary electrons,][]{ensslin:98, roettiger:99, miniati:01}, (ii) re-acceleration of electrons by compression of existing cocoons of radio plasma \\citep{ensslin:01, ensslin:02, hoeft:04}, or (iii) in-situ acceleration of protons and the production of relativistic electrons and positrons by ineleastic p-p collisions (secondary electrons). In the first two cases, the diffuse radio emission is roughly confined to the region of the shock fronts. In contrast, in the latter scenario the relativistic protons have long life times and can travel a large distance from their source before they release their energy. Hence, secondary electrons may be lead to radio halos. \\\\ In principle, observations of non-thermal cluster phenomena could provide an independent and complementary way of studying the growth of structure in our Universe and could shed light on the existence and the properties of the warm-hot intergalactic medium (WHIM), provided the underlying processes are understood. Sheets and filaments are predicted to host this WHIM with temperatures in the range $10^5$ to $10^7\\:{\\rm K}$ whose evolution is primarily driven by shock heating from gravitational perturbations breaking on mildly non-linear, non-equilibrium structures \\citep{cen:99}. Low-frequency aperture arrays such as {\\sc Lofar} are ideally suited to detect many diffuse steep-spectrum sources. In the next two years, the first {\\sc Lofar} survey is expected to chart a million galaxies and may discover hundreds of cluster radio halos \\citep{roettgering:06}. It is thus timely to study the distribution of diffuse radio sources in a cosmological context. \\\\ There have been efforts to simulate the non-thermal emission from galaxy clusters by modelling discretised cosmic ray (CR) energy spectra on top of Eulerian grid-based cosmological simulations \\citep{miniati:01b, miniati:04, miniati:07}. Recently, a series of papers explored the dynamical impact of CR protons on hydrodynamics in a cosmological SPH simulation \\citep{jubelgas:08, pfrommer:06, ensslin:07}. \\citet{skillman:08} presented a new method for identifying shock fronts in adaptive mesh refinement simulations and they computed the production of CRs in a cosmological volume adopting a nonlinear diffusive shock acceleration model. They found that CRs are dynamically important in galaxy clusters. \\citet{pfrommer:08} used {\\sc Gadget} simulations of a sample of galaxy clusters and implemented a formalism for CR physics on top of radiative hydrodynamics. They modelled relativistic electrons that are (i) accelerated at cosmological structure formation shocks and (ii) produced in hadronic interactions of CRs with protons of the ICM. \\citet{pfrommer:08} approximated both the CR spectrum and that of relativistic electrons locally by single power-laws with free parameters for the slope, the normalisation, and the low energy cut-off. Energy and momentum conservation, including source and sink terms, result in evolution equations for the spectra. They found that the radio emission in galaxy clusters is dominated by secondary electrons. Only at the location of strong shocks the contribution of primary electrons may dominate. In the cluster centres they found a radio emission of about $10^2 \\, h^3 \\: {\\rm mJy \\, arcmin^{-2}}$, while in the periphery the azimuthal average the emission is $10^{-3} \\, h^3 \\: {\\rm mJy \\, arcmin^{-2}}$. \\\\ Little is still known about the structure of shock fronts in the ICM. With high resolution imaging one can constrain the width of the transition and a deconvolution of the images gives an estimate for the density and pressure jump. Since the mean free path of protons in the cluster environment is of the order of Mpc, shock fronts in the ICM are collisionless. As studies of the best investigated collisionless shock front, namely the Earth bow shock, have revealed, the dissipation of the upstream kinetic energy is a complex process and depends on several shock properties (see \\eg\\ \\citet{burgess:07} for an introduction). For instance, the angle between the upstream magnetic field and the shock normal determines if an ion can gyrate between the upstream and downstream region, and the Mach number determines if instabilities in the shock region operate efficiently. However, shock fronts in the intra-cluster medium may differ significantly from the bow shock of the Earth as, for instance, the upstream plasma in the ICM is much less magnetised. Unfortunately, it is beyond the scope of current computer resources to study collisionless shock from the first principles, since scales from the electron gyro radius up to the large-scale structure of the shocks are involved. In hybrid simulations an effective small-scale response of electrons and protons is assumed. Using such a method, \\citet{kang:07} found that upstream CRs excite Alfv{\\'e}n waves and thereby amplify the magnetic field. \\\\ In this paper we combine a large cosmological simulation with a simple model for the radio emission of shock accelerated electrons. Our aim is to apply the emission model to the whole range of shock fronts generated during cosmic structure formation. A representative shock front sample is obtained from the \\MNUniverse\\ simulation which has been carried out with TreeSPH code {\\sc Gadget}-2. We have developed a novel approach for locating the shock fronts and to estimate their Mach number. For computing the radio emission we follow Hoeft \\& Br{\\\"uggen} (2007, HB07). They assumed that electrons are accelerated by diffusive shock acceleration and cool subsequently by synchrotron and inverse Compton losses. As a result the radio emission can be expressed as a function of downstream plasma properties, Mach number, and surface area of the shock front. Applying this radio emission model to the \\MNUniverse\\ simulation leads to radio-loud shock fronts with complex morphologies. We visualize these shock fronts to show where the radio emission is generated and compute artificial radio maps. As the \\MNUniverse\\ simulation provides a cosmologically representative cluster sample, we also investigate the relation between radio luminosity and X-ray temperature. \\\\ This paper is organised as follows: In \\Sec{sec-mn-universe} we briefly summarise the main characteristics of the \\MNUniverse\\ simulation. In \\Sec{sec-finding-shocks} we describe our approach for locating strong shock fronts in a SPH simulation and for estimating the Mach number. The radio emission model as worked out in HB07 is outlined and adopted for the SPH simulation in \\Sec{sec-radio-model}. Finally, in \\Sec{sec-results} we show our results for the shock fronts in the \\MNUniverse\\ simulation and for the radio properties of structure formation shocks. ", "conclusions": "\\label{sec-discussion} We have presented a novel method for estimating the radio emission of strong shocks that occur during the process of structure formation in the universe. Our approach is based on an estimate for the shock surface area and we compute the radio emission per surface element using the Mach number of the shock and the downstream plasma properties. The advantage of our method is that we can accurately determine the emission of the downstream regime without using an approximate, discretised electron spectrum. The disadvantage is clearly that we do not include any evolution of the shock front and the downstream medium. However, our main goal here is to compute the radio emission from all shock fronts generated during cosmic structure formation. Our main assumption is that radio emission is caused by primary electrons accelerated at the shock front. To obtain a model suitable for the application in the framework of a cosmological simulation we have to neglect the complexity of collisonless shocks, details of the electron acceleration mechanisms, the amplification of magnetic fields by upstream cosmic rays, and so forth. The model reflects rather an expectation for the average radio emission entailed by structure formation shocks. \\\\ Our results show that in a cosmological simulation one indeed finds textbook examples for large-scale, ring-like radio relics as observed in A3667 and A3376. Our example {\\em A} shows two large half-shell shaped shock fronts and no radio emission in the centre of the merging cluster. However, cluster {\\em A} is not very massive, hence the radio luminosity of the two half-shells is low compared to A3667. Another nice example is cluster {\\em B} which shows a radio relic only one side of the cluster, similar to A115. Beside those spectacular relics in the periphery of galaxy clusters, we find that a lot of clusters show complex, radio-loud shock structures close to the cluster centre. This suggests that part of the central diffuse radio emission in galaxy clusters, \\ie the radio halo phenomena, can be attributed to synchrotron emission of primary electrons. \\\\ We have modelled the radio emission with conservative assumptions for the efficiency of shock acceleration, namely $\\xi_e = 0.005$, and for the strength of magnetic fields in the downstream region, namely of the order of 0.07 to $0.8\\,{\\rm \\mu G}$. Still, we find that the radio luminosity in the cluster region is on average significantly above observed values, see \\Fig{fig-cum-synchro}. Even lower values for the acceleration efficiency or the magnetic field strength would suffice to explain the abundance of diffuse radio objects on the sky. In conclusion, our findings support the scenario in which radio relics are generated by primary electrons. \\\\ In summary, we have analysed one of the largest hydrodynamical simulations of cosmic structure formation with respect to shock fronts in clusters of galaxies. We have developed a method to identify the shock fronts, to estimate their Mach numbers and their orientation. We have applied the method described in \\citet{hoeft:07} to determine the radio emission as a function of the shock surface area and the downstream plasma properties. Our analysis led to the following results: \\begin{itemize} \\item By evaluating the entropy and the velocity field we are able to identify strong shocks in the simulation and to determine their properties. \\item Using conservative values for the efficiency of strong shocks to accelerate electrons and for strengths of magnetic fields, we are able to reproduce the number density and the luminosity of large-scale radio relics. Only a very few of the most massive clusters show such luminous, extended sources. \\item Our model reproduces various spectacular sources of diffuse radio emission at the periphery of galaxy clusters. Other clusters show emission with complex morphology close to the cluster centre. These sources may be classified as part of a central radio halo, following the observational distinction between relics and halos. \\item Our results reproduce the strong correlation between radio luminosity and cluster temperature. The highest radio luminosities occur in dense and hot environments such as massive clusters. In addition we find that a large number of galaxy clusters show only little diffuse radio emission. \\item We find that the abundance of radio relics can be explained with efficiency of diffusive shock acceleration for electrons lower than $\\xi_e = 0.005$ and a strength of magnetic fields in the relic region lower than 0.07 to $0.8 \\, { \\rm \\mu G}$. \\item All of the luminous radio relics belong to internal shock fronts. The accretion shocks are located at larger distances from the cluster centre. Since density and temperature are low at this location, their luminosity is too small to reach a similar flux as internal shock fronts. \\end{itemize} \\noindent {\\sc Acknowledgment } MH acknowledges DLR funding under the grant 50 OX 0201. MB acknowledges the support by the DFG grant BR 2026/3. The \\MNUniverse\\ simulations have been done at the Barcelona Supercomputing Center (Spain) and analysed at NIC J\\\"ulich (Germany). GY thanks MEC (Spain) for financial support under project numbers FPA2006-01105 and AYA2006-15492-C03. Our collaboration has been supported by the European Science Foundation (ESF) for the activity entitled `Computational Astrophysics and Cosmology' ({\\sc AstroSim}). \\newcommand{\\aap }{A\\&A} \\newcommand{\\araa }{ARA\\&A} \\newcommand{\\apj }{ApJ} \\newcommand{\\apjs }{ApJS} \\newcommand{\\apjl }{ApJL} \\newcommand{\\apss }{ApSS} \\newcommand{\\aapr }{A\\&A~Rev.} \\newcommand{\\aj }{AJ} \\newcommand{\\jgr }{JGR} \\newcommand{\\mnras}{MNRAS} \\newcommand{\\nat }{Nature} \\newcommand{\\physrep }{PhysRep}" }, "0807/0807.0110_arXiv.txt": { "abstract": "We present a comprehensive study of the nature of the SDSS galaxies divided into various classes based on their morphology, colour, and spectral features. The SDSS galaxies are classified into early-type and late-type; red and blue; passive, H{\\protect\\scriptsize II}, Seyfert, and LINER, returning a total of 16 fine classes of galaxies. We examine the luminosity dependence of seven physical parameters of galaxies in each class. We find that more than half of red early-type galaxies (REGs) have star formation or AGN activity, and that these active REGs have smaller axis ratio and bluer outside compared to the passive REGs. Blue early-type galaxies (BEGs) show structural features similar to those of REGs, but their centres are bluer than REGs. H{\\protect\\scriptsize II} BEGs are found to have bluer centres than passive BEGs, but H{\\protect\\scriptsize II} REGs have bluer outside than passive REGs. Bulge-dominated late-type galaxies have red colours. Passive red late-types are similar to REGs in several aspects. Most blue late-type galaxies (BLGs) have forming stars, but a small fraction of BLGs do not show evidence for current star formation activity. Differences of other physical parameters among different classes are inspected, and their implication on galaxy evolution is discussed. ", "introduction": "One of the fundamental issues of the observational cosmology is the evolutionary connection between various classes of galaxies. Today, galaxies are classified with various criteria. The most classical classification of galaxies is the Hubble sequence: elliptical galaxies, lenticular galaxies, spiral galaxies, barred spiral galaxies, and irregular galaxies \\citep{hub36,san61,dev74}. The main criterion of this classification is the morphology of galaxies; that is, the existence, size ratio, and appearance of spiral arms, disc, bulge, and bar. The Hubble sequence was established mainly based on the galaxies without nuclear activity, and such galaxies are often called \\emph{normal} galaxies. On the other hand, more and more galaxies showing active nuclei with broad line emission are being found; they are often called \\emph{abnormal} galaxies. According to the features of the line emission, active galactic nuclei (AGN) host galaxies are classified into several sub-classes: Seyfert 1 galaxies, Seyfert 2 galaxies, broad-line radio galaxies, narrow-line radio galaxies, low ionisation nuclear emission regions (LINERs), and so on. Luminosity is another criterion to classify galaxies. \\citet{san84} defined the galaxies with M$_{B} > -18$ as \\emph{dwarf} galaxies. In addition to those general classes of galaxies, there are some unusual galaxy classes with interesting properties: E+A galaxies with the spectral features of both very old stellar populations and very young stellar populations \\citep{dre92,yan06,yan08}; ultra-luminous infrared galaxies (ULIRGs) that are very bright in the mid- and far-infrared bands \\citep{san96,hwa07}; blue compact galaxies that show very compact morphology and high surface brightness with blue colour \\citep{thu97}; extremely red objects (EROs) whose optical $-$ infrared colour is extremely red \\citep{els88}; and so on. Recently, to deal with a large amount of survey data, many astronomers have classified galaxies simply into red sequence galaxies and blue sequence galaxies according to their colour-magnitude relation \\citep[e.g.][]{mar07}. For a long time, the individual properties of galaxies in those various classes have been investigated, and many efforts have been made to answer fundamental questions about galaxy evolution and connections between galaxy classes. Some examples of such fundamental questions are as follows: Why is there a conspicuous bimodality in the colour distribution of galaxies? Why does the colour bimodality not exactly agree with the morphological segregation? How do the environments affect the properties of galaxies? How did active galactic nuclei (AGNs) form? Is there any transition among the different classes of galaxies? Recent studies using large survey data discovered several interesting aspects of galaxy evolution, providing some answers to those fundamental questions. For example, \\citet{bal06} studied the bivariate luminosity functions with galaxy type classification, and found that there is a clear morphological bimodality supporting the idea that merger and accretion are associated with bulges and discs, respectively. \\citet{ber05} showed that both luminosity and colour of early-type galaxies are correlated with stellar velocity dispersion, and that velocity dispersion may be also closely correlated with the age and metal abundance of early-type galaxies. \\citet{cho07} found that late-type galaxies show wider dispersion in several physical quantities than early-type galaxies, and that those physical quantities manifest different behaviours across $M_{\\star}\\pm1$ mag. \\citet{par07} investigated the environmental effects on various physical properties of galaxies, finding that a key constraint on galaxy formation models is the morphology-density-luminosity relation. \\citet{mat06} analysed the colour, 4000{\\AA} break, and age in various spectral classes of galaxies, suggesting that the median light-weighted stellar age of galaxies is directly responsible for the colour bimodality in the galaxy population. Those previous studies, however, are not yet enough to explain the origins of the various galaxy classes and the evolutionary connections between the classes. One of the major limitations is that the galaxy classifications in most previous studies were often limited to only one or two properties. For instance, automatically-classified morphology \\citep[e.g.][]{par05,bal06}, galaxy colour \\citep[e.g.][]{mar07}, or spectral line features \\citep[e.g.][]{mat06} are used frequently in recent studies based on large survey data, but a multilateral classification study using all these criteria at the same time has not yet been seen. Such simplifications may be often very useful to understand the various and complicated phenomena in galaxy evolution. However, a simple classification can not distinguish detailed aspects of galaxy evolution. For example, early-type galaxies are often regarded as red and passive galaxies in most studies, but that is not always true. Some kinds of galaxies with early-type morphology were found to have blue colours possibly originating from young stellar populations or active nuclei \\citep{abr99,im01,fer05,lee06}. It is difficult to understand these kinds of unusual classes in studies with simple classifications. Closely related to galaxy classification, a couple of studies searching for principal components of galaxy properties have been conducted recently. \\citet{ell05} examined the distribution of photometric, spectroscopic and structural parameters for 350 nearby galaxies using the Millennium Galaxy Catalogue \\citep{lis03}, arguing that most properties show a clear distinction between early-type galaxies and late-type galaxies. Later, \\citet{con06} carried out principal-component analyses of the properties of 22000 galaxies at $z<0.05$ using the Third Reference Catalogue of Bright Galaxies \\citep[RC3;][]{dev91}, finding that the three parameters determining a galaxy's physical state may be mass, star formation and interactions/mergers. These studies presented how useful to consider various parameters in galaxy classification, although some important components like AGNs were not considered in their analyses. \\citet{con06} showed that multiple independent components, rather than any single component, may determine the various properties of galaxies. Therefore, to understand the nature of galaxies more comprehensively and to give stronger constraints on galaxy evolution models to \\emph{explain all kinds of galaxies}, it is necessary to investigate the nature of galaxies in various and finely-divided classes, and to find out the evolutionary connections between the classes. We have been doing a comprehensive study on a set of fine galaxy classes in the Sloan Digital Sky Survey \\citep[SDSS;][]{yor00}, based on their morphology, colour and spectral features. In this paper, the first in the series, we present the optical properties of galaxies in various fine classes. The outline of this paper is as follows. Section 2 shows the data set we used, and \\S3 describes the methods to classify the SDSS galaxies and to select volume-limited samples. We present the statistics of selected optical parameters and their luminosity dependence in \\S4. Based on the optical properties, we discuss the nature of galaxies in the fine classes in \\S5. Finally, the conclusions in this paper are given in \\S6. Throughout this paper, we adopt the cosmological parameters $h=0.7$, $\\Omega_{\\Lambda}=0.7$, and $\\Omega_{M}=0.3$. ", "conclusions": "We conducted a comprehensive study of the nature of the SDSS galaxies in various classes based on their morphology, colour and spectral features. Using three criteria, we classified the SDSS galaxies into early-type and late-type; red and blue; passive, H{\\protect\\scriptsize II}, Seyfert and LINER, resulting in 16 fine classes of galaxies in total. We estimated the luminosity dependence of seven physical quantities in each class, and compared the properties among classes, using a sub-sample with the same distribution of the velocity dispersion. From the analysis, we found that each galaxy class has its own distinguishable features. This shows that an analysis based on a simple classification may have a risk of mixing up different kinds of objects with different natures. The red early-type galaxies include well-known typical elliptical galaxies ($p$REGs), but some REGs show evidence for additional star formation in their outer regions ($h$REGs). Some other REGs with AGNs ($s$REGs, $l$REGs) have structural properties showing the existence of larger disc components than $h$REGs, indicating the relationship between AGN activity and gas accretion. The blue early-type galaxies may be in the process of bulge formation. The structural similarity between $p$REGs, $p$BEGs and $h$BEGs, supports an evolutionary sequence of $h$BEGs $\\rightarrow$ $p$BEGs $\\rightarrow$ $p$REGs. Seyfert early-type galaxies have a close relationship with the outer disc components of early-type galaxies, and the disc components in LINER early-type galaxies ($l$REGs, $l$BEGs) are smaller than those in Seyfert early-type galaxies, on average. The blue late-type galaxies have properties in agreement with typical spiral galaxies. Most of them are star-forming ($h$BLGs), but a very small fraction of BLGs do not show any evidence of current star formation ($p$BLGs). Some of BLGs have an AGN ($s$BLGs, $l$BLGs), which are less detected at large inclination, and $s$BLGs show particularly bright centres on average. The median axis ratio in each class shows that some intrinsic BLGs with large inclination may often be classified as red late-type galaxies, due to strong extinction by dust in their disc. In addition to dust extinction, a large bulge-to-disc ratio may make a late-type galaxy red. Many $p$RLGs seem to be bulge-dominated late-type galaxies with small inclination, in which line emission is not detected due to the limited size of the spectroscopic fibre aperture. Like in BLGs, AGN activity is detected in some RLGs ($s$RLGs, $l$RLGs), which have small inclination. This paper is the first in the series of comprehensive studies on the nature of the SDSS galaxies in finely-divided classes. In the following papers, we will inspect various aspects of galaxies in these classes, focusing on their multi-wavelength properties and environmental effects." }, "0807/0807.3219_arXiv.txt": { "abstract": "{} {This study has three principal aims: (i) to increase the number of detected pulsation modes of 44 Tau, especially outside the previously known frequency ranges, (ii) to study the amplitude variability and its systematics, and (iii) to examine the combination frequencies.} {During the 2004/5 and 2005/6 observing seasons, high-precision photometry was obtained with the Vienna Automatic Photoelectric Telescope in Arizona during 52 nights. Together with previous campaigns, a data base from 2000 to 2006 was available for multifrequency analyses.} {Forty-nine pulsation frequencies are detected, of which 15 are independent pulsation modes and 34 combination frequencies or harmonics. The newly found gravity mode at 5.30 cd$^{-1}$ extends the known frequency range of instability. Strong amplitude variability from year to year is found for the $\\ell$ = 1 modes, while the two radial modes have essentially constant amplitudes. Possible origins of the amplitude variability of the $\\ell$ = 1 modes, such as precession of the pulsation axis, beating and resonance effects are considered. The amplitudes of the combination frequencies, $f_i + f_j$, mirror the variations in the parent modes. The combination parameter, which relates the amplitudes of the combination frequencies to those of the parent modes, is found to be different for different parents. } {} ", "introduction": "The asteroseismic comparison between measurements and theoretical models of the radial and nonradial pulsation of stars on and near the main sequence involves the exact matching of the values of the pulsation frequencies, as well as reproducing the frequency range in which instability occurs. However, it is observationally difficult to determine the important borders of the frequency range in which a particular star is unstable. Due to the small amplitude growth rates of modes on the edges of the frequency instability range, only modes with small photometric amplitudes are expected to appear in these frequency regions. If we consider the $\\delta$ Scuti star FG Vir as an example (Breger et al. 2005), all photometric amplitudes in the 30 - 45 cd$^{-1}$ region had amplitudes less than 1.0 millimag. In fact, only the addition of 218 additional nights of photometry to the previous data allowed the detection of the pulsation modes near 40 cd$^{-1}$. In addition to the problem of small amplitudes in specific frequency regions, it is also essential to separate the `real' modes from the combination frequencies, $f_i \\pm f_j$, which occur in these high-frequency and low-frequency regions. This separation requires very high frequency resolution so that the chance of accidental agreements between the expected and measured frequency values is essentially reduced to zero. In practice this requires campaigns covering several years. The requirement of multiyear campaigns is strengthened by the amplitudes of many pulsators showing variability on timescales of months and, especially, years. To determine what types of modes show amplitude variability and to search for the astrophysical origin of this amplitude (and frequency) variability requires extensive observing campaigns. Finally, successful asteroseismology also demands the identification of individual pulsation modes. Both spectroscopic and photometric techniques have been applied: however, the large amounts of observing time required are easier to obtain on smaller telescopes with photometers. The $\\delta$ Scuti star 44 Tau is ideal for a detailed study because of its extremely low rotational velocity of 3 $\\pm$ 2 km s$^{-1}$ (Zima et al. 2007), for which second-order effects of rotation can be neglected, and because its amplitudes place the star between the high-amplitude pulsators (HADS) with dominant radial modes and the average low-amplitude $\\delta$ Scuti variable with mostly nonradial modes. Zima et al. (2007) present high-resolution spectroscopy of 44 Tau, which was used together with radial velocity data to derive the photospheric element abundances and provide $m$ values for the pulsation mode identifications. Six axisymmetric and two prograde modes were identified. In their asteroseismic study of 44~Tau, Lenz et al. (2008) present spherical mode degree identifications for the dominant modes of 44 Tau. Photometric amplitude ratios and phase differences suggest two radial, four $\\ell$ = 1, and three $\\ell = 2$ modes. Due to the measured $\\log g$ value of 3.6 $\\pm$ 0.1, both main sequence and post-main sequence models were examined. The identified modes can be fitted well in both evolutionary stages. The predicted frequency ranges for unstable modes are in good agreement with the observed ranges adopted by Lenz et al. (2008). Due to the different envelope structure of the main sequence and post-main sequence model, the borders of instability differ slightly. Consequently, new independent frequencies in the low-frequency and high-frequency regions possess important information for improving the models. We have previously presented 90 nights of photometric observations obtained during the three observing seasons 2000/2001, 2001/2002, and 2002/3 (Antoci et al. 2007: this paper also lists a detailed observational history of the star). The data from three years led to detecting 29 frequencies, of which 13 were independent pulsation modes. The two comparison stars used showed low-frequency variability at the millimag level, which is, regrettably, quite common. Otherwise, these measurements are very accurate. Consequently, the publication omitted the 0 to 5 cd$^{-1}$ region awaiting future measurements with different comparison stars, as reported in the present paper. ", "conclusions": "Two additional seasons of high-precision photometric data of the $\\delta$ Scuti star 44 Tau were gathered with the Vienna Automatic Photoelectric Telescope in Arizona. A frequency analysis of the total data set including photometry from 2000 to 2006 led to detecting 49 pulsation frequencies, of which 15 are independent pulsation modes and 34 combination frequencies or harmonics. We find that the amplitudes of the combination frequencies, $f_i + f_j$, mirror the variations of the parent modes. The combination parameter, $\\mu$, which relates the amplitudes of the combination frequencies to those of the parent modes, is found to be different for different radial and nonradial parents. The combination frequencies involving the radial fundamental mode have combination parameters, $\\mu$, of 0.003, while those involving the first radial overtone can be as high as $\\approx 0.009$. The time-base of 5 years allows us to study amplitude variability in more detail. Strong amplitude variability from year to year was found for the $\\ell$ = 1 modes, while the two radial modes have essentially constant amplitudes. Moreover, the amplitudes of all of the $\\ell = 1$ modes decrease during most of the observed period. We examined several possible reasons for the amplitude variability of the $\\ell$ = 1 modes: (i) Precession of the pulsation axis: This causes variable amplitudes because the mode visibility depends on the inclination angle. While the amplitudes of the radial modes remain constant, the amplitudes of $\\ell$ = 1 modes will change. However, both axisymmetric and non-axisymmetric modes decrease at the same time; this observation cannot be explained in this scenario because the visibility of $m = 0$ modes behaves opposite to their non-axisymmetric counterparts. (ii) Beating of close frequencies: Due to the slow rotation of 44 Tau, beating between the components of a rotationally split $\\ell=1$ triplet has been examined within the limits of the measured rotation rate. Generally, the predicted beating timescales are too short compared to the observed amplitude modulation. However, beating with hitherto unresolved, additional close frequencies (which are not a component of the same rotational triplet) could produce amplitude modulation at the observed timescales. (iii) Resonance effects: The expected timescales for amplitude modulation in a 44 Tau model were computed. We find that only for a fraction of the variable $\\ell = 1$ modes could the predicted variability timescales match the observed modulation periods. None of these three explanations can explain the observed amplitude variability completely. However, a combination of all these effects cannot be excluded. Even more data would be required to clarify this point. The newly found gravity mode at 5.30 cd$^{-1}$ extends the previously known frequency range to lower frequencies. This detection agrees better with the predicted instability ranges for main sequence models. However, our main sequence models cannot fit the measured fundamental parameters such as effective temperature. Our post-main sequence models provide a good fit of the fundamental parameters but predict stability at 5.30 cd$^{-1}$. The problem with the present models underestimating the observed $\\delta$ Scuti star instability is common to many $\\delta$ Scuti stars and may be related to still incomplete opacity data." }, "0807/0807.3443_arXiv.txt": { "abstract": "We report on an ongoing investigation into a seismic calibration of solar models designed for estimating the main-sequence age and a measure of the chemical abundances of the Sun. Only modes of low degree are employed, so that with appropriate modification the procedure could be applied to other stars. We have found that, as has been anticipated, a separation of the contributions to the seismic frequencies arising from the relatively smooth, glitch-free, background structure of the star and from glitches produced by helium ionization and the abrupt gradient change at the base of the convection zone renders the procedure more robust than earlier calibrations that fitted only raw frequencies to glitch-free asymptotics. As in the past, we use asymptotic analysis to design seismic signatures that are, to the best of our ability, contaminated as little as possible by those uncertain properties of the star that are not directly associated with age and chemical composition. The calibration itself, however, employs only numerically computed eigenfrequencies. It is based on a linear perturbation from a reference model. Two reference models have been used, one somewhat younger, the other somewhat older than the Sun. The two calibrations, which use BiSON data, are more-or-less consistent, and yield a main-sequence age $t_\\odot=4.68\\pm0.02\\,$Gy, coupled with a formal initial heavy-element abundance $Z=0.0169\\pm0.0005$. The error analysis has not yet been completed, so the estimated precision must be taken with a pinch of salt. ", "introduction": "The only way by which the age of the Sun can be estimated to a useful degree of precision is by accepting the basic tenets of solar-evolution theory and measuring those aspects of the structure of the Sun that are predicted by the theory to be indicators of age. The structure measurements must be carried out seismologically, and evidently one expects greatest reliability of the results when all the available helioseismic data are employed. However, the most relevant modes are those of lowest degree, because it is they that penetrate most deeply into the energy-generating core where the relic helium-abundance variation records the integrated history of nuclear transmutation. Moreover, it is also only they that can be measured in other stars. Therefore, there has been some interest in calibrating theoretical stellar models using only low-degree modes. The prospect was first discussed in detail by \\cite[Christensen-Dalsgaard (1984, 88)]{jcd84, jcd88}, \\cite{ulrich86} and \\cite{dog87}, although prior to that it had already been pointed out that the helioseismic frequency data that were available at the time indicated that either the initial helium abundance $Y_0$ or the age $t_\\odot$, or both, are somewhat greater than the generally accepted values \\cite[(Gough 1983)]{dog83}, an inference which is consistent with our present findings. Subsequent, more careful, calibrations were carried out by \\cite{guenther89}, \\cite{dog-nov90}, \\cite{guenther-demarque97}, \\cite{weiss-schlattl98}, \\cite{wd99}, \\cite{dog01} and Bonanno, Schlattl \\& Patern\\`o (2002). Not all of them addressed the influence of uncertainties in $Y_0$ on the determination of $t_\\odot$. As a main-sequence star ages, helium is produced in the core, increasing the mean molecular mass $\\mu$, preferentially at the centre, and thereby reducing the sound speed. The resulting functional form of the sound speed $c(r)$ depends not only on age $t_\\odot$ but also on the relative augmentation of $\\mu(r)$, which itself depends on the initial absolute value of $\\mu$, and hence on $Y_0$. \\cite{dog01} tried to separate these two effects using the degree dependence of the small separation $d_{n,l}=3(2l+3)^{-1}(\\nu_{n,l}-\\nu_{n-1,l+2})$ of cyclic frequencies $\\nu_{n,l}$, where $n$ is order and $l$ is degree. This is possible, in principle, because modes of different degree and similar frequency sample the core differently. However, that difference is subtle, and the sensitivity to the relatively fine distinction between the effects of $t_\\odot$ and $Y_0$ on the functional form of $c(r)$ in the core is low. Consequently the error in the calibration produced by errors in the observed frequency data is uncomfortably high. This lack of sensitivity can be overcome by using, in addition to core-sensitive seismic signatures, the relatively small oscillatory component of the eigenfrequencies induced by the sound-speed glitch associated with helium ionization \\cite[(Gough 2002)]{dog02}, whose amplitude is close to being proportional to helium abundance $Y$ \\cite[(Houdek \\& Gough 2007a)]{hgdog07a}. The neglect of that component in the previously employed asymptotic signature $d_{n,l}$ had not only omitted an important diagnostic of $Y$ but had also imprinted an oscillatory contamination in the calibration as the limits $(k_1, k_2)$, where $k=n+\\frac{1}{2}l$, of the adopted mode range was varied \\cite[(Gough 2001)]{dog01}. It therefore behoves us to decontaminate the core signature from glitch contributions produced in the outer layers of the star (from both helium ionization and the abrupt variation at the base of the convection zone, and also from hydrogen ionization and the superadiabatic convective boundary layer immediately beneath the photosphere). To this end a helioseismic glitch signature has been developed by \\cite{hgdog07a}, from which frequency contributions $\\delta\\nu_{n,l}$ can be computed and subtracted from the raw frequencies $\\nu_{n,l}$ to produce effective glitch-free frequencies $\\nu_{{\\rm s}n,l}$ to which a glitch-free asymptotic formula (\\ref{eq:asymp}) can be fitted. The solar calibration is then accomplished by interpolating the theoretical seismic signatures computed on a grid of solar models to the observations, using a standard grid to compute derivatives with respect to $t_\\odot$ and $Y_0$, and a carefully computed reference solar model designed to be close to the Sun. The result of the first preliminary calibration by this method, using BiSON data, has been reported by \\cite{hgdog07b}. Here we enlarge on our discussion of the analysis, and we augment our results with a calibration based on a second reference solar model. \\vspace{-5mm} \\begin{figure} \\centering \\includegraphics[height=.30\\textheight]{hg_fig1.eps} \\caption{Top left: The symbols (with error bars obtained under the assumption that the raw frequency errors are independent) represent second differences, $\\Delta_2\\nu$, of low-degree solar frequencies from BiSON. Top right: The symbols are second differences $\\Delta_2\\nu$ of adiabatic pulsation eigenfrequencies of solar Model S of \\cite{jcd96}. The solid curves in both panels are the diagnostics\\, (\\ref{eq:delnu}) -- (\\ref{eq:secdiff}), whose eleven parameters have been adjusted to fit the data optimally. Bottom: The symbols denote contributions $\\delta\\nu$ to the frequencies produced by the acoustic glitches of the Sun (left panel) and Model S (right panel).} \\end{figure} ", "conclusions": "" }, "0807/0807.0440_arXiv.txt": { "abstract": "We present high spatial resolution 21~cm HI observations of EA01A and EA01B, a pair of interacting post-starburst, or E+A, galaxies at $z = 0.0746$. Based on optical HST/WFPC2 images, both galaxies are known to display disturbed morphologies. They also appear to be linked by a bridge of stars. Previous HI observations \\citep{chang01} had already uncovered sizable quantities of neutral gas in or near these galaxies but they lacked the spatial resolution to locate the gas with any precision within this galactic binary system. We have analysed deep, high resolution archival VLA observations of the couple. We find evidence for three gaseous tidal tails; one connected to EA01A and two emanating from EA01B. These findings confirm, independently from the optical imaging, that {\\em (i)} EA01A and EA01B are actively interacting, and that, as a consequence, the starbursts that occurred in these galaxies were most likely triggered by this interaction, and that {\\em (ii)} $6.6\\pm 0.9\\times10^9$~{\\Msun} of neutral gas are still present in the immediate vicinity of the optical bodies of both galaxies. The \\HI\\ column density is lowest at the optical positions of the galaxies, suggesting that most of the neutral gas that is visible in our maps is associated with the tidal arms and not with the galaxies themselves. This might provide an explanation for the apparent lack of ongoing star formation in these galaxies. ", "introduction": "Post-starburst galaxies (or PSGs, or E+A galaxies, or k+a/a+k galaxies) are characterized by their optical spectra, independent of any morphological or photometric considerations. PSGs contain a sizable population of very young stars, producing very prominent Balmer lines, but lack ongoing star formation, hence the absence of detectable emission lines. This suggests that PSGs are indeed observed very shortly, within less than 1~Gyr, after the end of a vigorous, abruptly truncated starburst \\citep{dres83,dres99,pog99}. Today, they constitute only a small fraction of the galaxy cluster population ($<1$\\%, \\citet{fab91}); at intermediate-redshifts, however, they constituted a substantial cluster population \\citep{bel95}. Although the trigger and the abrupt end of the starburst is still not fully understood, photometry of PSGs shows evidence for disturbed morphologies, e.g. tidal tails, suggesting that in many cases the trigger is most likely associated to a galaxy-galaxy merger or interaction \\citep{zab96,yang04,blake04,tran03,goto05,yang08}. Using numerical simulations, \\citet{bekki05} has shown that PSGs can be formed via a major merger of two gas-rich spiral galaxies. In these merger simulations, a starburst is triggered that consumes the available gas within a timespan of roughly 1~Gyr, ending the starburst abruptly. Recent \\HI\\ observations have uncovered large quantities of neutral hydrogen gas in a significant fraction of PSGs \\citep{chang01,buyle06,helmboldt07,buyle08}. This opens up the possibility of investigating the hypothesized merger origin of many PSGs. \\begin{figure*} \\includegraphics[width=17cm]{f1.eps} \\caption{Channel maps overplotted on an optical HST/WFPC2 F702W image of EA01A/B. The contour levels are 2$\\sigma$, $3\\sigma$,\\ldots with $\\sigma$=0.2~mJy~beam$^{-1}$. The synthesized beam is depicted in the bottom-right corner of the first map (top left). \\label{channels}} \\end{figure*} Indeed, high-resolution interferometric radio observations of the neutral gas would provide valuable additional information. On the one hand, \\HI\\ morphologies and velocity fields, if they are distorted, could corroborate the merger hypothesis, and, moreover, could reveal whether at least part of the neutral gas is still gravitationally bound and will remain available for future star formation. We selected the PSG binary EA01A/B as an interesting target for our study. Both galaxies have a PSG-type optical spectrum and have the same recession velocity. EA01A is the bluest galaxy of the PSG sample of \\cite{zab96}, suggesting it is also the youngest sample member. HST/WFPC2 images in the F435W and F702W bands \\citep{yang04} provide strong evidence that the couple is interacting. EA01A, positioned about 11$''$ east of EA01B, contains many stellar clusters that are very conspicuous compared with the overall rather low surface brightness of this galaxy. This diffuse appearance suggests it is on the verge of disintegrating due to the injected orbital energy. EA01B, on the other hand, is a bulge dominated early-type spiral galaxy; it sports strongly asymmetric stellar arms. Both galaxies appear to be connected by a stellar bridge. Previous HI observations \\citep{chang01} had already uncovered sizable quantities of neutral gas in or near these galaxies but they lacked the spatial resolution to locate the gas with any precision within this galactic binary system. In section~\\ref{obs}, we describe our data reduction and analysis of deep archival VLA \\HI\\ observations of the EA01A/B galaxy binary. The results are presented and discussed in section~\\ref{discussion} and summarized in section~\\ref{conclusions}. ", "conclusions": "EA01A/B is a close pair of post-starburst (E+A) galaxies, surrounded by some $7 \\times 10^9$~{\\Msun} of neutral gas. Most of this gas resides in what appear to be three tidal arms, two of which are connected with EA01B. Together with optical HST/WFPC2 images, these observations show that EA01A and EA01B are actively interacting. The galaxies themselves show up as minima in the \\HI\\ column density. This lack of galaxy-bound dense neutral gas is most likely connected with the fact that these galaxies are in a post-starburst phase." }, "0807/0807.2029_arXiv.txt": { "abstract": "We present 3.5m Apache Point Observatory\\footnote{The Apache Point Observatory 3.5-meter telescope is owned and operated by the Astrophysical Research Consortium.} second-epoch spectra of four low-metallicity emission-line dwarf galaxies discovered serendipitously in the Data Release 5 of the Sloan Digital Sky Survey (SDSS) to have extraordinary large broad H$\\alpha$ luminosities, ranging from 3 $\\times$ 10$^{41}$ to 2 $\\times$ 10$^{42}$ erg s$^{-1}$. The oxygen abundance in these galaxies is very low, varying in the range 12 + log O/H = 7.36 -- 7.99. Such extraordinarily high broad H$\\alpha$ luminosities cannot be accounted for by massive stars at different stages of their evolution. By comparing with the first-epoch SDSS spectra, we find that the broad H$\\alpha$ luminosities have remained constant over a period of 3 -- 7 years, which probably excludes type IIn supernovae as a possible mechanism of broad emission. The emission most likely comes from accretion disks around intermediate-mass black holes with lower mass limits in the range $\\sim$ 5$\\times$10$^5$ $M_\\odot$ -- 3$\\times$10$^6$ $M_\\odot$. If so, these four objects form a new class of very low-metallicity AGN that have been elusive until now. The absence of the strong high-ionization lines [Ne {\\sc v}] $\\lambda$3426 and He {\\sc ii} $\\lambda$4686 can be understood if the nonthermal radiation contributes less than $\\sim$ 10\\% of the total ionizing radiation. ", "introduction": "Active galactic nuclei (AGN) are thought to be powered by massive black holes at the centers of galaxies, accreting gas from their surroundings. Observations of AGN show that they generally possess a high metallicity, varying from solar to supersolar metallicities \\citep{S98,H02}. While the derived metallicities do depend on the detailed model assumptions, this appears to be a solid conclusion. Gas metallicity is known to be strongly correlated with the stellar mass of the host galaxy \\citep{T04}. Since AGN are usually found in massive, bulge-dominated galaxies that have converted most of their gas into stars by the present epoch, their gas metallicities are generally high. A question then arises: do low-metallicity AGN exist? If so, can we find them in low-mass galaxies? To address these questions, \\citet{Gr06} have searched the Sloan Digital Sky Survey (SDSS) Data Release 4 (DR4) spectroscopic galaxy sample of over 500,000 objects to select out $\\sim$170,000 emission-line galaxies with high S/N spectra. They then use diagnostic line ratios to select out 23,000 Seyfert 2s galaxies. Imposing an upper mass limit of 10$^{10}$ $M_\\odot$ to restrict themselves to low-mass galaxies, they are left with a sample of only $\\sim$ 40 AGN, which they found appear to have metallicities around half that of typical AGN, i.e. having solar or slightly subsolar values. The same high metallicity range is found in the sample of low-mass AGN of \\citet{G07}. Assessing their findings, \\citet{Gr06} are led to another question: ``Why are there no AGN with even lower metallicities?'' In this paper, we suggest that these low-metallicity AGN do exist although they are extremely rare. In the course of a long-range program to search for extremely metal-deficient emission-line dwarf galaxies, \\citet{I07} have used the SDSS DR5 database of 675,000 spectra to assemble a large sample of emission-line galaxies. Two criteria were applied: 1) the [O {\\sc iii}] $\\lambda$4363 line must be detected to allow for a direct determination of element abundances; and 2) obvious high-metallicity AGN spectra are excluded. Thus, contrary to \\citet{Gr06} and \\citet{G07}, \\citet{I07} were not specifically looking for AGN. These criteria resulted in a sample of $\\sim$10,000 emission-line galaxies (ELG). While studying that sample to look for ELGs with broad components in their strong emission lines, \\citet{I07} came across four galaxies with very unusual spectra. The general characteristics of the four galaxies are given in Table \\ref{tab1}. Their absolute magnitudes are typical of dwarf galaxies. Because of their relatively large distance ( $z$ $\\sim$ 0.1-0.3) and relatively small angular sizes ($\\sim$ 1\\arcsec-2\\arcsec, only slightly larger than the seeing disk), their SDSS images (Fig. \\ref{images}) do not show much details. They possess a compact structure. Two galaxies, J1025+1402 and J1047+0739, have an approximately round shape, while the other two more distant galaxies, J0045+1339 and J1222+3602, have a distorted shape suggestive of mergers. Their colors are not blue like the other ELGs, but vary from red to yellow to green. Their spectra shown in Fig. \\ref{spectra}, resemble those of moderately to very low-metallicity high-excitation H {\\sc ii} regions: their oxygen abundances are in the range 12$+$$\\log$ O/H $\\sim$7.4--7.9, i.e. their heavy element mass fractions vary from $Z_\\odot$/19 to $Z_\\odot$/5 if the solar calibration 12$+$$\\log$O/H$=$8.65 of \\citet{A05} is adopted. \\citet{I07} found that there is however a striking difference: the strong permitted emission lines, mainly the H$\\alpha$ $\\lambda$6863 line, show very prominent broad components. These are characterized by somewhat unusual properties: 1) their H$\\alpha$ full widths at zero intensity $FWZI$ vary from 102 to 158 \\AA, corresponding to expansion velocities between 2200 and 3500 km s$^{-1}$; 2) the broad H$\\alpha$ luminosities $L_{br}$ are extraordinarily large, varying from 3$\\times$10$^{41}$ to 2$\\times$10$^{42}$ erg s$^{-1}$. This is to be compared with the range 10$^{37}$--10$^{40}$ erg~s$^{-1}$ found by \\citet{I07} for the other ELGs with broad-line emission. The ratio of H$\\alpha$ flux in the broad component to that in the narrow component varies from 0.4 to 3.4, as compared to 0.01--0.4 for the other galaxies; 3) the Balmer lines show a very steep decrement, suggesting collisional excitation and that the broad emission comes from very dense gas ($N_e$$\\gg$10$^{4}$ cm$^{-3}$). Evidently, these galaxies are exceedingly rare, since they constitute only 4/675,000 or 0.0006\\% of our original sample. To account for the broad line emission in these four objects, \\citet{I07} have considered various physical mechanisms: a) Wolf-Rayet (WR) stars; b) stellar winds from Ofp or luminous blue variable stars; c) single or multiple Supernova (SN) remnants propagating in the interstellar medium; d) SN bubbles; e) shocks propagating in the circumstellar envelopes of \\hbox{type\\,IIn} SNe; and f) AGN. While mechanisms a-d can account for $L_{br}$ $\\sim$ 10$^{36}$ to 10$^{40}$ erg s$^{-1}$, they cannot provide for luminosities that are 30 to 200 times greater. These very large luminosities are more likely associated with SN shocks or AGN. \\citet{I07} have considered \\hbox{type\\,IIn} SNe because their H$\\alpha$ luminosities are larger ($\\sim$10$^{38}$--10$^{41}$ erg~s$^{-1}$) than those of the other SN types, IIp and IIl, and they decrease less rapidly. To decide whether type IIn SNe or AGN are responsible for the broad emission in these galaxies, monitoring of their spectral features on the relatively long time scale of several years is necessary. If broad features are produced by IIn type SNe, then we would expect a decrease in the broad line luminosities. No significant temporal evolution would be expected in the case of an AGN. Additionally, higher signal-to-noise ratio spectra are necessary to put better constraints on the presence of the high-ionization [Ne {\\sc v}] $\\lambda$3426 and He {\\sc ii} $\\lambda$4686 emission lines, good indicators of a source of hard non-thermal radiation. In order to check for temporal evolution, we have obtained second-epoch spectra of the above four galaxies with broad emission, using the 3.5m Apache Point Observatory (APO) telescope\\footnote{In fact, \\citet{I07} have identified 5 objects as AGN candidates. We have not included here the fifth candidate, J2230--0006$\\equiv$PHL 293B, because its broad H$\\alpha$ luminosity is only 8.6$\\times$10$^{37}$ erg s$^{-1}$, some 10$^3$ - 10$^4$ times lower than those of the other four AGN candidates. The broad H$\\alpha$ luminosity of J2230--0006 has been erroneously given in Table 8 of \\citet{I07} as having 10 times its true value. Given this relatively low luminosity and the fact that a new 3.5m APO spectrum of J2230--0006 shows that its broad hydrogen lines have a P Cygni profile with a blue-shifted absorption, the broad emission probably originates from a stellar wind rather than from an accretion disk around a AGN. Most likely, the broad emission in J2230--0006 is caused by a strong outburst in a bright luminous blue variable (LBV) star. Similar broad hydrogen emission has been detected recently by \\citet{P08} in the extremely metal-deficient dwarf galaxy DDO 68.}. We describe the observations in \\S2. We discuss in \\S3 the main properties of the broad emission and show that they can be accounted for by low-metallicity intermediate-mass AGN. Our conclusions are summarized in \\S4. ", "conclusions": " 1. The steep Balmer decrements of the broad hydrogen lines and the very high luminosities of the broad H$\\alpha$ line in all four galaxies (3$\\times$10$^{41}$ to 2$\\times$10$^{42}$ erg s$^{-1}$) suggest that the broad emission arises from very dense and high luminosity regions such as those associated with supernovae (SNe) of type IIn or with accretion disks around black holes. However, the relative constancy of the broad H$\\alpha$ luminosities over a period of 3--7 years likely rules out the SN mechanism. Thus, the emission of broad hydrogen lines is most likely associated with accretion disks around black holes. If so, these four objects would harbor a new class of AGN that are extremely rare (in 0.0006\\% of all galaxies). These AGN would be intermediate-mass black holes residing in low-metallicity dwarf galaxies, with an oxygen abundance that is considerably lower than the solar or super-solar metallicity of a typical AGN. 2. There is no obvious spectroscopic evidence for the presence of a source of a non-thermal hard ionizing radiation in all four galaxies: high-ionization emission lines such as He {\\sc ii} $\\lambda$4686 and [Ne {\\sc v}] $\\lambda$3426 emission lines were not detected at the level $\\leq$ 1 -- 2 percent of the H$\\beta$ flux. We have calculated a series of CLOUDY models with ionizing spectra which include both thermal stellar and nonthermal power-law ionizing radiation in order to account for the absence of the high-ionization lines. We find that the predicted fluxes of the high-ionization lines are below the detectability level if the spectral energy distribution $f_\\nu$ $\\propto$ $\\nu^\\alpha$ of the ionizing nonthermal radiation has $\\alpha$ $\\sim$ --1 and the nonthermal ionizing radiation is significantly diluted by the thermal stellar ionizing radiation contributing $\\la$ 3 percent of the total ionizing radiation, or the ionizing spectrum is steeper ($\\alpha$ $\\sim$ --2), and the nonthermal ionizing radiation contributes $\\la$ 10 percent of the total ionizing radiation. 3. The lower limits of the masses of the central black holes $M_{\\rm BH}$ of $\\sim$ 5$\\times$10$^5$ $M_\\odot$ -- 3$\\times$10$^6$ $M_\\odot$ in our galaxies are among the lowest found thus far for AGN." }, "0807/0807.2503_arXiv.txt": { "abstract": "Several $f(R)$ modified gravity models have been proposed which realize the correct cosmological evolution and satisfy solar system and laboratory tests. Although nonrelativistic stellar configurations can be constructed, we argue that relativistic stars cannot be present in such $f(R)$ theories. This problem appears due to the dynamics of the effective scalar degree of freedom in the strong gravity regime. Our claim thus raises doubts on the viability of $f(R)$ models. ", "introduction": "The current accelerated expansion of the Universe is one of the deepest mysteries in cosmology~\\cite{Acceleration}. This acceleration may be due to some unknown energy-momentum component having the equation of state $p/\\rho \\approx -1$, or may be due to a modification of general relativity. In this paper, we are interested in the latter possibility. The simplest phenomenological way of modifying gravity is to consider a gravitational action described by a function of the Ricci scalar, $f(R)$, instead of the Einstein-Hilbert action. An early attempt is found, e.g., in~\\cite{StInf}, where a modification like $f(R)= R+R^2/\\mu^2$ was used to explain the accelerated expansion in the {\\em early} Universe. More recently, $f(R)$ modified gravity theories are often considered as a possible origin of the {\\em current} acceleration of the Universe~\\cite{Rev}. Any modified theories of gravity must account for the late-time cosmology which is well established by observations, and at the same time must be consistent with solar system and laboratory tests of gravity. However, since $f(R)$ gravity has the equivalent description in terms of the Brans-Dicke theory with the Brans-Dicke parameter $\\omega=0$~\\cite{BD}, naively constructed models would result in violation of the above requirements~\\cite{Chiba, ESK, CSE}. For example, the original proposal of~\\cite{CDTT} employs $f(R) = R-\\mu^4/R$, which admits an acceleratedly expanding solution even in the absence of a dark energy component. In order to accommodate a late time acceleration, however, one must introduce a very low energy scale, $\\mu\\sim H_0$ (the present Hubble scale), leading to a very light scalar field, which predicts the parametrized post-Newtonian (PPN) parameter $\\gamma=(1+\\omega)/(2+\\omega)=1/2$. Obviously, this result contradicts the observational constraint $|\\gamma-1|\\lesssim 10^{-4}$~\\cite{Will} To circumvent this difficulty, it is important to notice that the presence of matter may affect the dynamics of the extra scalar degree of freedom. The key idea is essentially the same as that of the ``chameleon'' model~\\cite{Chameleon, Cham2, Cham3}, in which the effective mass of the scalar field depends on the local matter density. In particular, the scalar field is very light for the cosmological density and is heavy for the solar system density, though the actual mechanism is slightly more complicated. The most successful class of $f(R)$ models~\\cite{Cembranos:2005fi, Hall, NvA, Tegmark, Li, AT, St, Hu, AB} incorporates this chameleon mechanism to evade local gravity tests. The experimental and observational consequences of this kind of $f(R)$ models are found in Refs.~\\cite{Tsujikawa1, Tsujikawa2, Brax}. (See also~\\cite{NO}.) In this paper, we consider the strong gravity regime of the carefully constructed models of~\\cite{St, Hu, AB}. The strong gravity aspects of $f(R)$ theories have not been explored so much before. Recently, Frolov suggested that such $f(R)$ models generically suffer from the problem of curvature singularities which can be easily accessed by the field dynamics {\\em in the presence of matter}~\\cite{Frolov}. In other words, a curvature singularity may be caused not by diverging gravitational potential depth, $|\\Phi|=\\infty$, but rather by a slightly large gravitational field, $|\\Phi|\\lesssim 1/2$. This motivates us to study relativistic stars in the context of $f(R)$ gravity. Spherically symmetric stars in $f(R)$ gravity have been investigated so far in~\\cite{Stars1, Stars2, MV, BB}. (We confine ourselves to a {\\em metric} theory of $f(R)$ gravity. Using the {\\em Palatini} formalism, polytropic stars have been studied in~\\cite{Sot}.) We shall show, both analytically and numerically, that stellar solutions with relatively strong gravitational fields cannot be constructed. Using the specific example of relativistic stars, we clarify how the singularity problem arises in the strong gravity regime of $f(R)$ theories. The singularity problem was also identified in~\\cite{ABCos} in a cosmological setting. This paper is organized as follows. In the next section we derive equations of motion for $f(R)$ modified gravity, and define the specific model we study. In Sec.~\\ref{sec:basic}, we reinterpret the problem of finding the desired stellar configuration as the problem of the particle motion in classical mechanics. We give some analytic arguments in Sec.~\\ref{sec:An} and then we present our numerical results in Sec.~\\ref{sec:Num}. Finally, we draw our conclusions in Sec.~\\ref{sec:Conc}. ", "conclusions": "\\label{sec:Conc} In this paper, we have studied the strong gravity aspect of $f(R)$ modified gravity models that reproduce the conventional cosmological evolution and evade solar system and laboratory tests~\\cite{St, Hu, AB}. It is known that $f(R)$ theories can be recasted simply in the Brans-Dicke theory with $\\omega=0$, but the potential for the effective scalar degree of freedom may play a complicated and nontrivial role. Moreover, the presence of matter may affect dynamics of the scalar field, possibly mimicking the chameleon model~\\cite{Chameleon}. We have explored uniform density, spherically symmetric stars and their exterior geometry in the $f(R)$ model of~\\cite{St}. The main result of the present paper is summarized as follows: given model parameters, there is a maximum value of the gravitational potential produced by a star, above which no asymptotically de Sitter stellar configurations can be constructed. We show this both analytically and numerically. For example, the model with $n=1$ and $\\lambda\\approx 2$ gives $\\Phi_{{\\rm max}}\\approx 0.1$. This raises a warning sign for a class of $f(R)$ theories, because neutron stars cannot be present in such gravity models. The underlying mechanism that hinders strong gravitational fields around matter is explained essentially as follows~\\cite{Frolov}. Consider a static matter distribution. The Newtonian potential obeys the Poisson equation $\\nabla^2\\Phi \\sim G \\rho$, while the equation of motion for the scalar field implies $\\nabla^2\\chi \\sim G\\rho$. From this, one can evaluate the excitation of the scalar degree of freedom around the matter distribution as $\\delta\\chi\\sim{\\cal O}(\\Phi)$. If the de Sitter minimum is located very close to the point $\\chi=1$, which corresponds to $R=\\infty$ in the effective potential, a slightly strong gravitational field will cause the problem of appearance of a curvature singularity. Bearing the above evaluation in mind, let us comment on the other specific models of $f(R)$ gravity. The model of Hu and Sawicki~\\cite{Hu} and Starobinsky's one share the same structure of $f(R)$ in the high-curvature regime, i.e., $f(R)\\approx R-2\\Lambda_{{\\rm eff}}+C/R^{\\alpha}$ with $\\alpha>0$. Therefore, we expect that the same problem arises in the Hu and Sawicki's model. The model by Appleby and Battye is characterized by~\\cite{AB} \\begin{eqnarray} f(R)=\\frac{R}{2}+\\frac{1}{2a}\\ln\\left[ \\cosh(aR)-\\tanh(b)\\sinh(aR) \\right], \\end{eqnarray} where $a$ and $b$ are parameters. Since $\\chi = df/dR = [1+\\tanh(aR-b)]/2$, a positive curvature singularity corresponds to $\\chi=1$. Taking, for example, $b=1.5$, we find that $\\chi_1\\approx 0.93$ at the de Sitter minimum, which is very close to the dangerous curvature singularity (the result is independent of $a$). Since also in this model the effective potential is finite at $R=+\\infty$, we anticipate the same singularity problem. Our choice of the parameters in the present paper gave $\\Phi_{{\\rm max}}\\approx 0.1$, for which neutron stars are unlikely to exist. However, there still remains a possibility that more realistic stellar environments and matter profiles weaken the bound on the potential by a factor of 2 or 3, and at the same time make the chameleon mechanism work~\\footnote{ The neutron star-white dwarf system PSR J1141--6545 can put strong constrains on alternative theories of gravity around relativistic stars~\\cite{Will, E-F}. For this reason, we need to invoke the chameleon mechanism to describe such stars.}. It is technically much more difficult to construct stellar configurations with a realistic equation of state, realistic energy densities, and realistic stellar environments. Such an elaborated modeling of relativistic stars might allow for $\\Phi_{{\\rm max}}$ as large as, say, 0.3, but the parameter space of the theory will be very restricted. To conclude, $f(R)$ theories that reproduce the correct behavior of weak gravity in the solar vicinity do not admit neutron star solutions without special care." }, "0807/0807.2818_arXiv.txt": { "abstract": "We performed, for the first time, the simulation of spiral-in of a star cluster formed close to the Galactic center (GC) using a fully self-consistent $N$-body model. In our model, the central super-massive black hole (SMBH) is surrounded by stars and the star cluster. Not only are the orbits of stars and the cluster stars integrated self-consistently, but the stellar evolution, collisions and merging of the cluster stars are also included. We found that an intermediate-mass black hole (IMBH) is formed in the star cluster and stars escaped from the cluster are captured into a 1:1 mean motion resonance with the IMBH. These ``Trojan'' stars are brought close to the SMBH by the IMBH, which spirals into the GC due to the dynamical friction. Our results show that, once the IMBH is formed, it brings the massive stars to the vicinity of the central SMBH even after the star cluster itself is disrupted. Stars carried by the IMBH form a disk similar to the observed disks and the core of the cluster including the IMBH has properties similar to those of IRS13E, which is a compact assembly of several young stars. ", "introduction": "Young and massive stars have been found within one parsec from the Galactic center (GC) \\citep{Krabbe95,Paumard06}. Some of them are $\\sim$1000 AU from the central SMBH. How these massive stars were brought to the vicinity of the SMBH has been a mystery. One possible scenario is the following. A star cluster formed at a few tens of pc from the GC, and then spiraled in due to the dynamical friction \\citep{Gerhard01}. Previous simulations \\citep{PZ03,KM03,Kim04,GR05} have shown that the timescale of spiral-in of the star cluster can be short enough. However, how close the stars can actually approach the SMBH is not clear. Another possible scenario is the in-situ formation in an accretion disk \\citep{LB03,Nay07,HN08}. Giant molecular clouds fall into the GC and form massive gaseous disks around the central BH. Stars form in the disk if it becomes gravitationally unstable and results in fragmentation. However, accretion disks have difficulty producing stars with eccentric orbits and a compact assembly of stars like IRS 13E, which is located at $\\sim 0.13$ pc in projection from the GC and contains half a dozen young stars and probably an IMBH \\citep{HM03,Maillard04}. \\citet{Mapelli08} argued that if a gas cloud undergoes a very close encounter (pericenter distance of 0.01 pc) with the central SMBH, the tidal compression could trigger the star formation, resulting in stars in close, bound orbits. However, how a cloud can come that close to the GC is not clear. On the other hand, if the star cluster has an eccentric orbit, the orbits of stars escaped from the cluster are also eccentric. The remnant of the core looks like IRS 13E. We performed a fully self-consistent $N$-body simulation in which the internal dynamics of the cluster, that of the parent galaxy, and interactions between cluster stars and galaxy stars are correctly handled. In previous simulations, when the internal dynamics was followed by an accurate $N$-body code, the parent galaxy had to be modeled as a fixed potential with some fitting formulae for the dynamical friction. This means that the orbital evolution is not accurate. \\citet{Fujii08} showed that the actual orbital decay of the cluster is faster than that of previous simulations and the main reason is stars with a mass grater than 90\\% of an initial cluster escaping from this cluster. We describe the method of our $N$-body simulation in section 2. In section 3 we show the results of simulations. Section 4 is for summary. ", "conclusions": "Using $N$-body simulation, we showed that many young and massive stars are carried to the GC by a star cluster due to the 1:1 mean motion resonance with an IMBH which is formed in the cluster. In addition, we found that slingshots in the star cluster throw stars into orbits which pass near the GC. These orbits have very high inclinations and are sometimes retrograde orbits. They are new channels which carry young stars to the central parsec. Our simulation demonstrated the existence of massive stars and we explained why they form a disk-like structure. The possible existence of two counter-rotating disks might suggest that two clusters have spiraled in. The interaction and resonance between stars and IMBHs from multiple clusters might be responsible for the existence of stars which are very close to the central SMBH." }, "0807/0807.0730_arXiv.txt": { "abstract": "The observational characteristics of quasi-periodic oscillations (QPOs) from accreting neutron stars strongly indicate the oscillatory modes in the innermost regions of accretion disks as a likely source of the QPOs. The inner regions of accretion disks around neutron stars can harbor very high frequency modes related to the radial epicyclic frequency $\\kappa $. The degeneracy of $\\kappa $ with the orbital frequency $\\Omega $ is removed in a non-Keplerian boundary or transition zone near the magnetopause between the disk and the compact object. We show, by analyzing the global hydrodynamic modes of long wavelength in the boundary layers of viscous accretion disks, that the fastest growing mode frequencies are associated with frequency bands around $\\kappa $ and $\\kappa \\pm \\Omega $. The maximum growth rates are achieved near the radius where the orbital frequency $\\Omega $ is maximum. The global hydrodynamic parameters such as the surface density profile and the radial drift velocity determine which modes of free oscillations will grow at a given particular radius in the boundary layer. In accordance with the peak separation between kHz QPOs observed in neutron-star sources, the difference frequency between two consecutive bands of the fastest growing modes is always related to the spin frequency of the neutron star. This is a natural outcome of the boundary condition imposed by the rotating magnetosphere on the boundary region of the inner disk. ", "introduction": "} Wave modes in the boundary region in the inner accretion disk are a likely source of the distinct narrow frequency bands of quasi-periodic oscillations (QPOs) (van der Klis 2000) from neutron star sources in low mass X-ray binaries (LMXBs). The dependence of QPO frequencies on accretion rate suggests that the observed QPOs are connected with accretion disk modes. Psaltis, Belloni \\& van der Klis (1999) showed the existence of correlations between the frequencies of different QPO bands extending over a wide span of frequencies. The same correlation encompasses black hole as well as white dwarf and neutron star sources. This strongly implies that the frequency bands are determined by the oscillation modes of the accretion disk. The nature of the compact object, whether it is a black hole, white dwarf or neutron star, probably plays a role in exciting the same disk modes through possibly different mechanisms in different types of compact sources. The QPO bands modifying the X-ray luminosity are likely to belong to the boundary region between the disk and the compact object. For magnetic neutron stars, this boundary region is shaped by the interaction of the disk with the magnetosphere. The region where the rotation deviates from Keplerian flow is not necessarily very narrow; this boundary zone may have a size as large as a few tenths of the inner disk radius. We shall therefore use the terms \\textquotedblleft boundary\\textquotedblright \\thinspace and \\textquotedblleft transition\\textquotedblright \\thinspace region interchangeably. Initial explorations of disk modes underlying QPO frequency bands from LMXBs were provided by Alpar et al. (1992) and Alpar \\& Y\\i lmaz (1997). The characteristics and possible excitation mechanisms of thin-disk oscillations were reviewed and discussed by Kato (2001). In many current models of QPOs, especially those of neutron stars, their frequencies are identified with test-particle frequencies (Stella, Vietri \\& Morsink 1999; Abramowicz et al. 2003). Most parts of an accretion disk, except, significantly, the transition regions at the inner boundary of the disk, are characterized by Keplerian rotation rates to a very good approximation. At any radius beyond the transition region, acoustic, magneto-acoustic and viscous corrections to the test-particle Keplerian orbital frequency are negligible. In the outer disk regime for disks around neutron stars and white dwarfs (and far enough from a central black hole), test-particle frequencies for radial and vertical perturbations of the orbit are degenerate with the Keplerian orbital frequency. The degeneracy is removed if the effects of general relativity are important. A Newtonian field of tidal or higher multi-pole structure would also lead to a split in the degeneracy. However, in Newtonian gravity, distortions of stellar shape even for the most rapidly rotating neutron stars will not introduce a significant level of non-degeneracy between the frequencies of test-particle oscillations that is comparable to the difference between the observed QPO frequency bands. Models employing the Keplerian frequency have been relatively successful in interpreting QPO frequency correlations. The Kepler (test particle) frequency interpretation has also been employed to place constraints on the masses and spins of compact objects. Empirically, predictions based on test-particle frequencies applied to QPO frequency correlations deviate from observations at a level of about 10\\%. The Keplerian frequency represents the basic imprint of rotation on all the dynamical responses of the disk. Albeit simple, the identification of QPOs with test-particle frequencies has a number of shortcomings. First, precisely in a boundary layer where QPOs are expected to be excited, the dynamical frequencies of oscillations are not the test-particle frequencies. In the boundary layer, viscous and magnetic forces lead to deviation of orbital frequencies from Keplerian test-particle frequencies (see, e.g., Erkut \\& Alpar 2004). The resulting band of non-Keplerian rotation frequencies entail viscous and acoustic response and couplings between adjacent rings of the disk fluid. Hydrodynamic effects are therefore essential for an understanding of the frequency bands characterizing the boundary layer. In particular, the degeneracy of test-particle frequencies in the non-relativistic regime is lifted in the hydrodynamic boundary layer, where the radial epicyclic frequency is in fact larger than the orbital frequency. This simple observation (Alpar \\& Psaltis 2005) has important consequences for the interpretation of kHz QPO frequencies, in particular for the constraints on the neutron star mass-radius relation derived from kHz QPOs. Second, test-particle frequencies do not distinguish between azimuthal sidebands. While kinematic models of QPOs like the beat-frequency model involve one specific band of frequencies, say $\\omega $, a large number of sidebands, of frequencies $\\omega _{m}\\cong \\omega -m\\Omega $, where $\\Omega $ is the orbital frequency, are allowed by azimuthal symmetry. Arguments as to why only one or two QPO bands are excited must involve choices imposed by the symmetries of the interaction between the magnetosphere and the accretion disk boundary, leading to resonances with particular frequencies at one or more radial regions in the disk. Without resonances, test-particle frequency spectra present no distinction between the modes. The realistic hydrodynamic modes may, in some parameter ranges, distinguish between both fundamental modes and all their azimuthal sidebands through the different growth or decay rates of these modes, \\emph{even in the case of free oscillations}. A reduction of the number of relevant (easily excitable) modes of free oscillations is certainly an important task for an understanding of accretion disks around neutron stars, white dwarfs and black holes. In the case of black holes, this classification of free hydrodynamic modes in terms of their growth rates is even more important as resonant excitation by the black hole is not available. In this paper, we study the global modes of free oscillations at some position $r$ in the inner disk-boundary or transition region by analyzing the perturbed dynamical equations of a hydrodynamic disk, including pressure gradients, viscous and magnetic stresses. We consider the case of a disk around a neutron star with a magnetosphere. The neutron star is taken to be a \\emph{slow rotator}: the star's rotation rate is less than the value of the Keplerian rotation rate at the inner radius of the disk; $\\Omega _{\\ast }<\\Omega _{\\mathrm{K}}(r_{\\mathrm{in}})$. The actual rotation rate of the disk is set by the boundary condition $\\Omega (r_{\\mathrm{in}})=\\Omega _{\\ast }$. Thus in the steady state solutions for the disk, the rotation rate at the inner edge of the disk and in the transition region beyond is sub-Keplerian. The innermost disk radius $r_{\\mathrm{in}}$ determines the disk magnetosphere interface which may be subject to magnetic Rayleigh-Taylor or interchange instabilities. This instability operates mainly at the disk-magnetosphere interface because of the sharp density contrast across the radius $r=r_{\\mathrm{in}}$. The inner disk region for $r\\gtrsim r_{\\mathrm{% in}}$ consists of a non-Keplerian boundary region that joins the outer Keplerian disk with a continuous density distribution. As we will see in \\S\\ \\ref{ehdfom}, the surface density profile throughout the non-Keplerian boundary region increases with decreasing radii. Thus, our boundary region is not subject to interchange-like instabilities. The free oscillation modes we explore by perturbing a steady state solution embody information about the stability of the disk through their growth or decay rates. Growing modes will provide a clue as to the origin of the QPO frequency bands, which could be excited by resonant forcing of the disk by time dependent interactions with the star's magnetosphere. We find, as expected, that growth or decay rates are determined by the dynamical effects of viscosity, with an additional dependence on the sound speed and the density and rotation-rate profiles in the non-Keplerian boundary region. For some boundary region parameters, not all fundamental modes of free oscillations and not all azimuthal sidebands grow, and among the growing modes the growth rates differ. For other choices of boundary region parameters the growth rates of all sidebands are similar. The frequencies of the modes differ from test-particle frequencies by amounts that can be several times larger than the corresponding QPO width. Most importantly, for a reasonable set of accretion disk parameters, we can show that only a certain few of the hydrodynamic free oscillation modes will grow, and that the frequency bands of these oscillations can be associated with the observed frequency bands. The nature of free oscillation modes in the boundary region depends on whether the neutron star is rotating at a rate slower or faster than the rotation rates prevailing in the inner boundary of the disk. In this paper, we will consider the more common and straightforward case of \\emph{slow rotators}, the case when the neutron star rotation rate $\\Omega _{\\ast }$ is less than $\\Omega _{\\mathrm{K}}(r_{% \\mathrm{in}})$, the Kepler rotation rate at some representative inner disk radius $r_{\\mathrm{in}}$. An investigation of \\emph{forced} resonant excitation of these prevalent modes, and the comparison and association with observed QPO bands will follow in a subsequent paper. \\S\\ 2 lays out the basic assumptions and equations, \\S\\ 3 displays the mode analysis for the free oscillations, including a discussion of hydrodynamic effects, frequency bands, and growth rates. In \\S\\ 4, we discuss the results and present our conclusions. ", "conclusions": "We have studied the global hydrodynamic modes of long wavelength in the boundary or transition region of viscous accretion disks as a possible source of kHz QPOs in neutron-star LMXBs. The stability of the eigenmodes strongly depends on the global disk structure imposed by the boundary conditions as expected from the mode analysis in the long wavelength regime. Our local treatment takes account of the local effects of the global disk parameters on the excitation of hydrodynamic free oscillations. We find that the frequencies and growth rates of the modes are mainly determined by global disk parameters such as $\\kappa /\\Omega $, the ratio of the radial epicyclic frequency to the orbital frequency, the radial surface-density profile $\\beta $, $\\Omega _{\\nu }/\\Omega _{s}$, the ratio of the radial drift velocity to the sound speed, and $\\Omega _{s}/\\Omega $, the ratio of the sound speed to the azimuthal velocity. The local values of $% \\kappa /\\Omega $ and $\\beta $ directly follow from the global solution for the rotational dynamics of a boundary-transition region model. The parameters $\\Omega _{\\nu }/\\Omega _{s}$ and $\\Omega _{s}/\\Omega $ depend additionally on the particular viscosity prescription and the details of the structure of accretion flow in the inner disk. The hydrodynamic effects on the frequencies and growth rates of the modes are reflected by the parameters $\\Omega _{\\nu }/\\Omega _{s}$ and $\\Omega _{s}/\\Omega $. We have found that the growth rates of different modes for a given azimuthal wavenumber $m$ differ significantly when the hydrodynamic corrections are sufficiently large, i.e. $\\Omega _{\\nu}/\\Omega _{s}\\lesssim 1 $ and $\\Omega _{s}/\\Omega \\lesssim 1$. In this regime, the growing mode frequencies significantly exceed test-particle frequencies for a plausible range of $\\Omega _{s}/\\Omega $ at any particular radius within the boundary region. This is because the effect of hydrodynamic corrections on the growth rates of the modes is expected to be strong in a magnetic boundary layer or transition zone as discussed in \\S\\ 3.2. In the limit of small hydrodynamic corrections, i.e. $\\Omega _{\\nu}/\\Omega _{s}\\ll 1$ and $\\Omega _{s}/\\Omega \\ll 1$, the eigenmodes have test-particle frequencies with negligible growth rates. Our analysis shows that taking account of hydrodynamic effects has an important outcome: the frequencies of the growing modes gain higher values above test-particle frequencies as their growth rates increase for larger hydrodynamic corrections. Throughout the boundary region, we have identified the growing mode frequencies by the eigenvalues that approach the test-particle frequencies $% \\omega _{1}^{(0)} =\\kappa $, $\\omega _{1}^{(1)} =\\kappa +\\Omega $, and $% \\omega _{2}^{(1)} =\\kappa -\\Omega $ in the limit of small hydrodynamic corrections (see panel a of Fig.~2). Modes such as $\\omega _{3}^{(0)} =0$ and $\\omega _{3}^{(1)}\\simeq \\Omega $ are not excited for the range of $% \\Omega _{s}/\\Omega $ used in Fig.~2 (panel b) and Fig.~3 (panel b). The difference frequency between successive bands of growing modes is $\\Delta \\omega \\simeq \\Omega _{*}$ as observed for the relatively slowly rotating neutron stars in LMXBs that show kHz QPOs (\\S\\ 3.2). In a boundary layer or transition region, the orbital frequency of the disk matter is close to the rotation frequency of the star ($\\Omega \\sim \\Omega _{*}$). As $\\Delta \\omega \\simeq \\Omega$, the separation between consecutive mode frequencies is always related to the spin frequency of the neutron star. This is the result of the boundary condition imposed by the rotating magnetosphere on the innermost disk. The modes of the test-particle frequencies $\\omega _{1}^{(0)}\\cong \\kappa $ and $\\omega _{2}^{(1)}\\cong \\kappa -\\Omega $ were originally proposed by Alpar \\& Psaltis (2005) to be associated with the upper and lower kHz QPOs observed in neutron-star LMXBs. The frequency separation between the higher and lower-frequency kHz QPO peaks decreases by a few tens of Hz as both QPO frequencies increase by hundreds of Hz in a range of $200-1200$ Hz per source in all observed sources, namely, Sco~X-1, 4U~$1608-52$, 4U~$1728-34$ and 4U~$1735-44$ (van der Klis 2000). This pair of frequencies satisfies the observed behavior that the difference frequency $\\Delta \\omega $ between two kHz QPOs decreases as both frequencies increase. According to our present analysis, the pairs of observed kHz QPO bands could be either $\\omega _{1}^{(1)}\\simeq \\kappa +\\Omega $ and $\\omega _{1}^{(0)}\\simeq \\kappa $ or $\\omega _{1}^{(0)}\\simeq \\kappa $ and $\\omega _{2}^{(1)}\\simeq \\kappa -\\Omega $, respectively, since the identification of the observed QPOs with $\\omega _{1}^{(1)}\\simeq \\kappa +\\Omega $ and $\\omega _{1}^{(0)}\\simeq \\kappa $ also meets the observed condition that $\\Delta \\omega $ decreases when both QPO frequencies increase. The growth rates of free oscillations produce a wide spectrum of sidebands, $\\omega _{1,2}^{(m)}\\cong \\kappa \\pm \\vert m\\vert \\Omega $, some of which fall in the range of observed power spectra. The reduction to only two of kHz QPO bands as observed in neutron star sources must therefore be a product of forced oscillations, resonances, and boundary conditions. The role of forced oscillations and resonances in the excitation of particular modes will be discussed in future work. We find that the fastest growing modes with frequency branches $\\kappa $ and $\\kappa \\pm \\Omega $ are excited near the disk radius $r\\lesssim r_{0}$, where the orbital frequency is maximum. In the boundary layer, the loss of centrifugal support near $r_{0}$ leads to some radial acceleration of the disk matter. The subsequent rise in the radial drift velocity of the disk matter is accompanied by a sudden drop in the surface density as the vertically integrated dynamical viscosity is minimized (see \\S\\ 3.2). This picture is common to all boundary layers or transition zones for which magnetic braking is efficient (see Erkut \\& Alpar 2004 and references therein). Thus, independent of the boundary conditions and the magnetic field configuration in the boundary region, there exists a specific disk radius where the hydrodynamical background quantities such as the surface density $\\Sigma _{0}$ and the radial drift velocity $v_{r0}$ change dramatically. The steepness in the change of the surface density in the radial direction is reflected through the parameter $\\beta $ (see equation % \\ref{bet}). The sign and magnitude of $\\beta $ (the surface density profile in the disk) control whether the free oscillation modes will grow or decay. We see that the modes of frequency bands $\\kappa $ and $\\kappa \\pm \\Omega $ can be excited with the largest growth rates only for the disk radii $% r\\lesssim r_{0}$, where $\\beta $ is minimum and $\\Omega $ is maximum (see \\S\\ 3.2 for the range of $\\beta $). The inverse timescale, $\\Omega _{\\nu} $, associated with radial accretion also obtains its highest value near the same radii yielding the maximum growth rates in the accretion disk. The boundary layers with sub-Keplerian rotation rates are usually expected to be realized in the innermost regions of accretion disks around \\emph{slow rotators}. For a \\emph{slow rotator}, the Kepler rotation rate at the inner disk radius $\\Omega _{\\mathrm{K}}(r_{\\mathrm{in}})$ prevails over the stellar rotation rate $\\Omega _{*}$. As we have illustrated in \\S\\ 3.2, the difference frequency between two consecutive bands of growing modes in the boundary region is close to the spin frequency of the neutron star ($% \\Delta \\omega \\simeq \\Omega _{*}$). This result agrees with the kHz QPO observations of the relatively slow rotators with spin frequencies below $% \\sim 400$~Hz in neutron-star LMXBs. In these sources, for example, in 4U~$1728-34$, the lower kHz QPO frequencies increase from 600 to 900 Hz and the upper kHz QPO frequencies increase from 950 to 1200 Hz, while the difference between the two observed kHz QPO frequency bands decreases slightly, of the order of 10 Hz, while remaining still commensurate with the spin frequency of the neutron star (M\\'{e}ndez \\& van der Klis 1999). A second class of kHz QPO sources, such as the sources KS~$1731-260$, Aql~X-1, 4U~$1636-53$, which have spin frequencies above $\\sim 400$~Hz is usually depicted as \\emph{fast rotators} (see Wijnands et al. 2003). In this second class of sources, for example in 4U~$1636-53$, the lower kHz QPO frequencies increase from 900 to 950 Hz and the upper kHz QPO frequencies increase from 1150 to 1190 Hz, while the difference between the two observed kHz QPO frequency bands again decreases by amounts of the order of 10 Hz (see van der Klis 2000 and references therein). The frequency separation between the kHz QPOs is close to half the spin frequency of the neutron star if it is a \\emph{fast rotator}. The emergence of two seemingly different classes of neutron-star LMXBs could be related to whether the rotation rate of the neutron star is less or greater than that of the inner disk matter. It is probable that $\\Omega _{*}>\\Omega _{\\mathrm{K}}(r_{\\mathrm{in}})$ for the relatively fast rotating neutron star sources for which the difference frequency between twin QPO peaks is $\\Delta \\omega \\simeq \\Omega _{*}/2$. Recent work by M\\'{e}ndez \\& Belloni (2007) actually suggests a more continuous distribution between $\\Delta \\omega \\simeq \\Omega _{*}$ and $% \\Delta \\omega \\simeq \\Omega _{*}/2$. The rotational dynamics of accretion flow in the innermost regions of accretion disks around these \\emph{fast rotators} could be quite different from those of the transition regions we consider here. We plan to address this issue in a subsequent paper." }, "0807/0807.3418_arXiv.txt": { "abstract": "We explore CP violation effects on the neutrino propagation in dense environments, such as in core-collapse supernovae, where the neutrino self-interaction induces non-linear evolution equations. We demonstrate that the electron (anti)neutrino fluxes are not sensitive to the CP violating phase if the muon and tau neutrinos interact similarly with matter. On the other hand, we numerically show that new features arise, because of the non-linearity and the flux dependence of the evolution equations, when the muon and tau neutrinos have different fluxes at the neutrinosphere (due to loop corrections or of physics beyond the Standard Model). In particular, the electron (anti)neutrino probabilities and fluxes depend upon the CP violating phase. We also discuss the CP effects induced by radiative corrections to the neutrino refractive index. ", "introduction": "One of the major open issues in neutrino physics is the possible existence of CP violation. As with the recent crucial experimental discoveries on neutrino oscillations, the answer to this question has fundamental implications in high-energy physics, astrophysics and cosmology e.g. to understand the matter versus anti-matter asymmetry in the Universe. The observation that the weak interaction violates the CP symmetry in the quark sector was first established in 1964 \\cite{Christenson:1964fg}. Future strategies to search for CP violation in the lepton sector depend upon the actual value of one yet unknown neutrino oscillation parameter, i.e. $\\theta_{13}$ \\cite{Ardellier:2006mn}, and require long term accelerator projects producing very intense neutrino beams \\cite{Volpe:2006in}. It is therefore essential to explore alternative avenues to get clues on this fundamental question, such as indirect effects in dense environments like core-collapse supernovae. Core-collapse supernovae emit about 10$^{53}~$erg as neutrinos of all flavours during their rapid gravitational collapse. Such neutrinos might play a role on the two major supernova unsolved problems, namely understanding how the explosion finally occurs and where the nucleosynthesis of the heavy elements, produced during the {\\it r}-process, takes place. While neutrinos from a massive star were first observed during the SN1987A explosion, future observations of (extra)galactic or relic supernova neutrinos will help unravelling supernova physics and/or unknown neutrino properties. For example, in \\cite{Schirato:2002tg,Tomas:2004gr,Fogli:2004ff,Kneller:2007kg} the imprint of the shock wave on the neutrino time signal is investigated; while avenues for extracting information on the third neutrino mixing angle are discussed in \\cite{Dighe:1999bi,Engel:2002hg}. This searches require advances in the modelling of supernova dynamics, of neutrino propagation in dense environments and of our knowledge on neutrinos. Impressive developments are currently ongoing in our understanding of neutrino propagation in dense matter. While solar experiments \\cite{Davis:1964aa,Ahmad:2001an,Eguchi:2002dm,Arpesella:2007xf} have beautifully confirmed the oscillation enhancement induced by the coupling with matter - the Mikheev-Wolfenstein-Smirnov or MSW effect \\cite{Wolfenstein:1977ue,Mikheev:1986wj} -, recent theoretical investigations have shown that the inclusion of the neutrino self-interactions in dense environments introducing a non-diagonal neutrino refractive index \\cite{Pantaleone:1992eq} gives rise to a wealth of new phenomena, as first pointed out in \\cite{Samuel:1993uw}. Various regimes have been identified : the synchronized one \\cite{Pastor:2001iu,Duan:2005cp}, the bipolar oscillations \\cite{Duan:2005cp,Hannestad:2006nj} and the spectral split phenomenon \\cite{Raffelt:2007cb,Raffelt:2007xt}. Since numerical calculations become more involved, analytical treatments for the three flavour case are being proposed (see e.g. \\cite{Dasgupta:2007ws}). The importance of the loop corrections to the neutrino refractive index, the $V_{\\mu \\tau}$ potential \\cite{Botella:1986wy}, is underlined in \\cite{EstebanPretel:2007yq}. Moreover constraints on neutrino mixing from shock re-heating and the {\\it r}-process nucleosynthesis, including the neutrino-neutrino interaction, are investigated in \\cite{Qian:1994wh,Sigl:1994hc}. The impact of the neutrino-neutrino interaction on the electron fraction relevant for the {\\it r}-process is investigated in \\cite{Pastor:2002we,Balantekin:2004ug}. In a previous work \\cite{Balantekin:2007es} we have investigated the CP effects on the neutrino fluxes, on the electron fraction in a supernova (relevant for the {\\it r}-process) as well as the possible impact on the supernova neutrino signal in an observatory on Earth. In particular we have shown analytically that no effects can be found on the electron (anti)neutrino fluxes, when muon and tau neutrino have the same fluxes at the neutrinosphere; while significant effects are obtained numerically on the fluxes when they differ. The calculations in \\cite{Balantekin:2007es} are obtained considering interaction with matter at tree level only. In this work we explore for the first time CP violation effects on the neutrino propagation in dense environments including the standard MSW effect, the neutrino self-interactions and the {\\it V}$_{\\mu \\tau}$ refractive index. We first show analytically that if the muon and tau neutrinos interact similarly, the electron (anti)neutrinos are not sensitive to the CP violating phase, even in presence of neutrino self-interactions. This result is general and valid for any matter density profile and/or initial neutrino luminosity. We present numerical results, obtained within the three flavour formalism, on the neutrino oscillation probabilities and fluxes within the star. In particular we show that both the probabilities and the fluxes become sensitive to the CP violating phase if muon and tau neutrino interact differently with matter (e.g. because of loop corrections or physics beyond the Standard Model.) The paper is structured as follows. Section II presents the theoretical framework for describing the neutrino propagation including the coupling with matter as well as the neutrino-neutrino interaction. Section III gives the analytical and numerical results. Conclusions are drawn in Section IV. ", "conclusions": "We have investigated possible effects of the CP violating phase on the neutrino propagation in dense matter when interaction with matter without/with loop corrections and the neutrino self-interaction are included. Our analytical results demonstrate that, at tree level, even when the neutrino-neutrino interaction is included there are no CP violating effects on the electron (anti)neutrino fluxes in the star unless $\\nu_{\\mu}$ and $\\nu_{\\tau}$ fluxes differ at the neutrinosphere. If such condition is not satisfied, a totally new feature arise, namely that the electron (anti)neutrino oscillation probabilities (and fluxes) become sensitive to the CP violating phase $\\delta$. The latter is also true when the loop corrections to the refractive index are included. We find numerically that, in most cases studied, the modifications introduced by the CP violating phase are larger (smaller) at low (high) energies than in the case where the neutrino-neutrino interaction is not included, and at the level of a few percent. We also find numerically that, even assuming that the muon and tau neutrinos have the same fluxes at the neutrinosphere, the CP effects induced by $V_{\\mu\\tau}$ only are amplified by the neutrino self-interactions up to several percent. \\begin{figure}[t] \\vspace{.6cm} \\centerline{\\includegraphics[scale=0.3,angle=0]{nuefluxratioTeqVmutau.eps}\\hspace{.2cm} \\includegraphics[scale=0.3,angle=0]{nuefluxratioTdifVmutau.eps}} \\caption{Ratios of the $\\nu_e$ fluxes for a CP violating phase $\\delta=180^{\\circ}$ over $\\delta= 0^{\\circ}$ as a function of neutrino energy. They correspond to inverted hierarchy and small $\\theta_{13}$ and different distances from the neutron star surface, i.e. 200 km (dotted), 500 km (dashed), 750 (dot-dashed), 1000 (solid). The results include the $\\nu$-$\\nu$ interaction and the $V_{\\mu \\tau}$ refractive index. They are obtained using equal $\\nu_{\\mu}$ and $\\nu_{\\tau}$ fluxes at the neutrinosphere (left) or taking $T_{\\nu_{\\mu}}=1.05~T_{\\nu_{\\tau}}$ (right). For the $\\bar{\\nu}_e$ fluxes deviations up to 10$\\%$ are found at energies lower than 10 MeV.} \\label{fig:nueIHSA} \\end{figure} \\vspace*{0.5cm} The authors acknowledge the support from \"Non standard neutrino properties and their impact in astrophysics and cosmology\", Project No. ANR-05-JCJC-0023. \\textit{} \\vspace*{-0.25cm}" }, "0807/0807.1301_arXiv.txt": { "abstract": "{We analyze the data of Kamiokande-II, IMB, Baksan using a parameterized description of the antineutrino emission, that includes an initial phase of intense luminosity. The luminosity curve, the average energy of $\\bar\\nu_e$ and the astrophysical parameters of the model, derived by fitting the observed events (energies, times and angles) are in reasonable agreement with the generic expectations of the delayed scenario for the explosion.} \\normalsize\\baselineskip=15pt ", "introduction": "We begin with a rapid historical excursus, with emphasis on the issues that are relevant for data analysis. \\begin{itemize} \\item Colgate \\& White 1966\\cite{cg} propose the paradigm for the explanation of the core collapse supernovae, where neutrinos are the key agents. \\item Nadyozhin 1978\\cite{nad} concludes a detailed calculation that demonstrates an initial phase of intense neutrino luminosity. \\item Bethe \\& Wilson 1985\\cite{bw} suggest that the energy deposition on a scale of half a second can re-energize the stalled shock wave. \\item 1987: Kamiokande-II\\cite{kii}, IMB\\cite{imb}, Baksan\\cite{bak} and LSD\\cite{lsd} observe several events in correlation with SN1987A. \\item Several authors--e.g., Bahcall 1989\\cite{bah}--remark that non-LSD data generically meet the expectations. Then the main interest shifts on the relevance of oscillations. \\item Lamb \\& Loredo 2002\\cite{ll} (LL) discuss whether SN1987A data indicate specific imprints of the delayed scenario\\footnote{With the term `delayed scenario' we refer here and in the following to the scenario for the explosion put forward by Bethe and Wilson, that incorporates the initial phase of intense neutrino luminosity discussed by Nadyozhin (this is also called `standard scenario' or `neutrino assisted explosion').} such as the initial phase of intense luminosity. \\item Imshennik~\\&~Ryazhskaya~2004\\cite{ir} suggest a 2 stage scenario with essential role of rotation and possible explanation of LSD events detected 4.5 hours earlier. \\end{itemize} Next we summarize the present status: although there is mounting evidence that the delayed scenario is correct, a conclusive proof is still missing; while most people is convinced that SN1987A neutrinos confirmed the general picture of the explosion, there are annoying doubts on whether SN1987A was a standard object or not; while we continue sharpening our theoretical tools, we hope intensely in a Galactic supernova event to progress in our understanding. In short, one should be aware that until the theoretical picture will be definitively assessed, the interpretation of the results from SN1987A will continue to contain elements of uncertainty. But if we want to test the expectations, we are forced to answer the question: {\\em What are the generic expectations for the delayed scenario?} We summarize the features that are relevant for our analysis: \\begin{itemize} \\item In conventional detectors (scintillators and water Cherenkov) the main detection reaction is the inverse-beta decay of electron antineutrinos on free protons (IBD): $\\bar\\nu_e p\\to n e^+ $. \\begin{figure}[t] $$\\includegraphics[width=0.5\\textwidth]{vissaniFig1.eps}$$ \\caption{\\it Sketch of $\\nu_e$ and $\\bar\\nu_e$ (higher and lower) luminosity curves. The excess of $\\nu_e$ in the first 50 ms (`neutronization' phase) gives a small contribution to the total number of observable events. \\label{fig1}} \\end{figure} \\item The main part of the emission happens in two stages as shown in Fig.~\\ref{fig1}. 10-20\\% of the energy is radiated in an early phase, here called {\\bf accretion}, that should last about half-a-second; 80-90\\% of the energy is radiated later, during the phase of neutron star {\\bf cooling}, namely in a quiet thermal phase. \\item The main reactions for energy radiation during accretion are $e^- p\\to n \\nu_e$ and $e^+ n\\to p \\bar\\nu_e$, the second being the inverse of IBD. The presence around the nascent neutron star of an abundant amount of $e^+e^-$ plasma ensures a large flux of $\\nu_e$ and $\\bar\\nu_e$, that are key ingredients to revive the stalled shock wave. [See Janka\\cite{j} for a review.] \\item During the cooling phase, neutrinos of all species ($\\nu_e,\\nu_\\mu,\\nu_\\tau,\\bar\\nu_e,\\bar\\nu_\\mu,\\bar\\nu_\\tau$) are radiated with similar luminosities; this feature is occasionally called `equipartition'. \\end{itemize} In this talk, based on a work in collaboration with M.L.~Costantini and A.~Ianni\\cite{paglia}, we present an analysis of SN1987A data along these lines. We will compare two models for neutrino emission, namely the conventional one, when the neutrino emission is described by a `one-component' (cooling) model, and a `two-component' (cooling+accretion) model which resembles more closely the expectations of the delayed scenario. Preliminary results have been documented in\\cite{mosca,prep,ifae,brasile}; see also\\cite{jcap}, in particular Sect.~3.2 there. ", "conclusions": "Our study confirms earlier results with the 1-component model. The refined treatment of background, cross section, description of the scattering angle, inclusion of Baksan data, {\\em etc.}\\ do not lead to important changes. We confirm the results of Lamb \\& {}Loredo in particular the very strong evidence for accretion, when their 2-component model is adopted. We discussed an improved 2-component model, where the average energy and luminosity curves are constrained to be continuous, cooling follows accretion, oscillations (not very important {\\em a posteriori}) are included. The best fit of $\\tau_a$, $M_a$, {\\em etc.}\\ are close to expectations; the binding energy $2.2\\times 10^{53}$ erg is lower than for the 1-component model; the evidence of accretion is not as strong as for the Lamb \\& Loredo model, but still important." }, "0807/0807.3074_arXiv.txt": { "abstract": "{Fourier transform (or lag) correlators in radio interferometers can serve as an efficient means of synthesising spectral channels. However aliasing corrupts the edge channels so they usually have to be excluded from the data set. In systems with around 10 channels, the loss in sensitivity can be significant. In addition, the low level of residual aliasing in the remaining channels may cause systematic errors. Moreover, delay errors have been widely reported in implementations of broadband analogue correlators and simulations have shown that delay errors exasperate the effects of aliasing.} {We describe a software-based approach that suppresses aliasing by oversampling the cross-correlation function. This method can be applied to interferometers with individually-tracking antennas equipped with a discrete path compensator system. It is based on the well-known property of interferometers where the drift scan response is the Fourier transform of the source's band-limited spectrum.} {In this paper, we simulate a single baseline interferometer, both for a real and a complex correlator. \\revd{Fringe-rotation usually compensates for the phase of the fringes to bring the phase centre in line with the tracking centre. Instead, a modified fringe-rotation is applied. This enables an oversampled cross-correlation function to be reconstructed by gathering successive time samples.}} {Simulations show that the oversampling method can synthesise the cross-power spectrum while avoiding aliasing and works robustly in the presence of noise. An important side benefit is that it naturally accounts for delay errors in the correlator and the resulting spectral channels are regularly gridded} {} ", "introduction": "\\label{sec:intro} The observing band of radio interferometers frequently needs to be split into sub-bands, either for spectral line observations or to reduce the effects of chromatic aberration. Fourier transform correlators offer an efficient method for dividing the observation band. In this scheme, the signal from the two arms of the interferometer are correlated at discrete delay steps, making direct measurements of the cross-correlation function (see Fig.~\\ref{fig:ccf-spec}). Taking the Fourier transform of the cross-correlation function gives the complex cross-power spectrum. For a signal of bandwidth $\\Delta\\nu$, Nyquist sampling theorem requires the signal to be sampled at \\revd{time intervals of $1/(2\\Delta\\nu)$} to avoid aliasing. But the cross-correlation function of a band-limited signal extends over an infinite delay range. So the Nyquist sampling theorem holds true only if we sample the cross-correlation function over an infinite delay range. Clearly this is not practical so the cross-correlation function is sampled over a finite range. This results in a recovered signal spectrum with tapered band edges which will overlap with its images in the spectral domain. This overlap causes aliasing and will corrupt the recovered cross-power spectrum. \\begin{figure}[hbtp] \\centering \\mbox{\\subfigure[Real correlator]{ \\includegraphics[width=3.5in,trim=0 0 0 65,clip=true]{pic/8117fig1a.eps} }} \\mbox{\\subfigure[Complex correlator]{ \\includegraphics[width=3.5in]{pic/8117fig1b.eps} }} \\caption[The real and complex CCF]{{\\bf a)} {\\it Top plot:} The cross-correlation function measured by a real correlator (black) and the envelope (grey). This is a graphical representation of the cross-correlation function in Eq.~(\\ref{eqn:corr-response}) with the geometrical delay $\\taug = 0$. The open circles indicate the position of the detectors in instrument delay $\\taui$. {\\it Bottom plots:} The cross-power spectrum recovered by applying the Fourier transform on the cross-correlation function. We have assumed the telescope parameters in Table~\\ref{tab:tel-param}. {\\bf b)} The in-phase (real) and quadrature (imaginary) cross-correlation functions sampled by a complex correlator. The cross-power spectrum recovered in this case is single-sided.}\\label{fig:ccf-spec} \\end{figure} \\revd{As an illustration, the plots in Fig.~\\ref{fig:under-vs-oversamp} show snapshots of the cross-correlation functions and the recovered spectra.} The left-most plots \\revd{(case~1)} show the cross-correlation function when the source transits an east-west baseline (that is when the path difference between the two arms of the interferometer is zero). The circles represent the measurements at discrete delay steps of the correlator. The next plots to the right \\revd{(case~2)} show the cross-correlation function a short while later. The right-most plots show the spectrum calculated by taking the Fourier transform of the discrete measurements. The filled grey circles indicate the spectrum at transit \\revd{(case~1)} and the dark circles are for \\revd{case~2}. In the top row of Fig.~\\ref{fig:under-vs-oversamp}a, the cross-correlation function was \\revd{sampled at 16 delays. This critically samples the {$\\Dnu=6$\\GHz} signal bandwidth but undersamples the {0--12\\GHz} basebandsignal,} so the positive and negative halves of the {6--12\\GHz} bands lie side-by-side. In \\revd{case~1}, the cross-correlation measurements trace out a delta function and give the characteristic flat spectrum. But in \\revd{case~2}, the amplitudes are perturbed. Ideally, the amplitudes should \\revd{not change} so this points to a fundamental problem. This could be worked round by oversampling the \\revd{baseband signal} at 64 or more delays as illustrated by Fig.~\\ref{fig:under-vs-oversamp}b. Now the two sets of amplitudes at different times match perfectly. Oversampling introduces a buffer between the two halves of the passband and also between their spectral images. This suppresses aliasing by reducing the overlaps between the signal and image bands. But clearly, sampling at 64 delays in hardware is not practical. \\revd{This is manifested in the Fourier Transformed data (the spectrum) as temporal modulations in both the amplitude and phase (see Fig.~\\ref{fig:alias-cycles}).} In aliased signals, the noise components between the channels will also be correlated so the individual channels cannot be strictly treated as independent measurements. These effects are particularly pronounced in the edge frequency channels as seen in Fig.~\\ref{fig:alias-cycles} so these channels are usually rejected. This may be acceptable in systems with tens to hundreds of channels. But some correlators, particularly broadband analogue Fourier transform correlators with channels of order 10 (for example, \\citealt[][ hereafter {\\paperone}]{li2004,roberts2007,holler2007}), the loss constitutes a significant portion of the total bandwidth. In principle, there are three ways to suppress aliasing: \\begin{enumerate} \\item Increasing the delay range over which the cross-correlation function is sampled. The spectral channel width will be narrower and the edges of the passband will be sharper. So the overlap between the signal spectrum and its images will be narrower. The edge channels will still have to be rejected but it will be a smaller portion of the whole data. \\item Oversampling the cross-correlation function at finer delay steps as already illustrated with Fig.~\\ref{fig:under-vs-oversamp}. \\item Reducing the bandwidth so that it is not critically-sampled. \\end{enumerate} Implementing either of the first two modifications in hardware would be costly and the third approach would lose sensitivity. In Sect.~\\ref{sec:os-method}, we propose a software-based approach to oversample the cross-correlation function and avoid aliasing. This is based on the well-known notion that the spectrum of a source can be obtained from the cross-correlation function measured by a drift scan. We can reconstruct the cross-correlation function using a combination of source tracking, path compensation and fringe-rotation. In Sect.~\\ref{sec:simulations}, we will illustrate an application of this technique with simulations. The system requirements are: (1) Individually-tracking antennas and (2) discrete-delay path compensation. We have modelled an analogue correlator here but the principles could equally apply to digital correlators. \\revd{A} number of groups recently reported broadband analogue Fourier transform correlators suffering from delay errors (for example, \\citealt{harris2001}; {\\paperone}; \\citealt{roberts2007}). Simulations by {\\paperone} showed that delay errors make the effects of aliasing worse. Although the delay errors can be calibrated a sample at a time (recently by \\citealt{harris2001}; {\\paperone}), the method described here \\revd{will be} a natural way of accounting for delay errors. This \\revd{side benefit} is perhaps as significant as the alias suppression aspect of the method. The resulting spectral channels are also regularly gridded at the desired frequencies. However there are a number of issues that must be considered in a practical system and these are discussed in Sect.~\\ref{sec:practical-issues}. This paper is a follow up to {\\paperone} where we described the development of a broadband 6--12{\\GHz} Fourier transform correlator for the Arcminute Microkelvin Imager \\citep[AMI;][]{kaneko2006,zwart2008}. AMI is a new interferometer designed to survey for clusters of galaxies by exploiting the Sunyaev-Zel'dovich effect \\citep{sunyaev1972}. The effects of aliasing in the correlator used in this instrument are discussed further in {\\paperone}. Before going into a detailed discussion of the oversampling method, we will first give a brief overview of how the data is processed in a conventional interferometric system. \\begin{figure*}[hbtp] \\centering \\mbox{\\subfigure[The \\revd{baseband signal ({0--12\\GHz}) under and critically-sampled. The spectra show signs of aliasing. The cross-correlation function was sampled at 16 and 32 delays.}]{ \\includegraphics[width=17cm]{pic/8117fig2a.eps}}} \\mbox{\\subfigure[\\revd{Oversampling the baseband signal by $2\\times$ and $4\\times$ critical-sampling reduces aliasing.}]{ \\includegraphics[width=17cm]{pic/8117fig2b.eps}}} \\caption[Critical and oversampling]{The effects of different oversampling rates on the recovered spectra. The left-most plots show the cross-correlation function \\revd{from Eq.~(\\ref{eqn:corr-response4}) as a function of the residual delay {\\Dtau}. This is a snapshot when the source transits an east-west baseline so the geometrical delay $\\taug = 0$ (case~1).} The next \\revd{set of} plots to the right \\revd{(case~2) show the cross-correlation function a little later when the centre of the envelope has drifted by a quarter of a delay spacing}. The measurements at transit \\revd{(case~1)} are indicated by the grey circles and the measurements later on \\revd{(case~2)} are indicated by the black circles. The horizontal axis is the delay in {\\Dtau}, which will be defined in Sect.~\\ref{sec:real-comp-corr}. The right-most plots are the amplitudes calculated from the delay measurements. The filled grey marks are at transit \\revd{(case~1)} and the dark circles are a while later \\revd{(case~2)}. The two sets of marks should match but it can be seen in {\\bf a)} that they do not. This mismatch shows up as the temporal modulation seen in Fig.~\\ref{fig:alias-cycles}. In the upper row of {\\bf a)}, \\revd{although the {6\\GHz} signal bandwidth} is critically-sampled at 16 delays, \\revd{the} {0--12\\GHz} \\revd{baseband signal is undersampled.} The two halves of the spectrum are side-by-side. The spectrum is aliased at both the high and low-frequency ends. In the lower row of {\\bf a)}, the highest frequency component ({12\\GHz}) is critically-sampled with 32 delays. Now only the upper frequency range suffers from aliasing. In {\\bf b)}, \\revd{the baseband signal is oversampled by a factor of 2 and 4 (sampled at 64 and 128 delays respectively)}. Now the two sets of spectral amplitudes match. Oversampling ensures a buffer between the two halves of the signal spectrum as well as with the image spectra. This reduces the spectral overlap and suppresses aliasing. In this case, oversampling the \\revd{baseband frequency} ({0--12\\GHz}) by a factor of 2 is sufficient (\\revd{an equivalent} sampling frequency of {48\\GHz} at 64 delays). Oversampling above this does not reduce aliasing any further. As a minor technical detail, the forms of the spectra in {\\bf b)} are slightly different from the one in Fig.~\\ref{fig:ccf-spec}. Here (and also in Fig.~\\ref{fig:alias-cycles}), we shifted the spectra by half a sub-band using the shift theorem of Fourier transforms. This ensures that the sub-bands span the desired frequency range and is discussed further in \\cite{holler2007}. Subsequent simulations will assume channel-shifting.}\\label{fig:under-vs-oversamp} \\end{figure*} ", "conclusions": "We have described a software-based method for overcoming aliasing in critically-sampled Fourier transform correlators. The method reduces aliasing by reconstructing the oversampled cross-correlation function from successive samples. Data from complex correlators can be fringe-rotated directly and efficiently in software. If using this method on real correlators, the signal should be fringe-rotated in hardware to avoid systematic errors. The edge spectral channels are usually discarded because they are the worst affected by aliasing. For an 8-channel complex correlator simulated here, the sensitivity could be improved by up to 15~percent by retaining the edge channels. The additional computational overhead and complexity are offset by an improvement in sensitivity and reduction in systematic errors from aliasing. This could be a significant benefit for Fourier transform correlators with less than 10 spectral channels. Some data with low fringe rates will have to be rejected, but for AMI-like telescopes with fractional bandwidths $\\Dnu / \\nuRF \\gtrsim 0.4$ and maximum baseline to dish ratios of $D/d \\lesssim 10$, the fraction of data lost is relatively small. We have also shown that we can naturally compensate for delay errors to recover regularly-gridded frequency channels at the design frequencies." }, "0807/0807.2244_arXiv.txt": { "abstract": "{ We review what can (and cannot) be learned if dark matter is detected in one or more experiments, emphasizing the importance of combining LHC data with direct, astrophysical and cosmological probes of dark matter. We briefly review the conventional picture of a thermally produced WIMP relic density and its connection with theories of electroweak symmetry breaking. We then discuss both experimental and theoretical reasons why one might generically expect this picture to fail. If this is the case, we argue that a combined effort bringing together all types of data -- combined with explicitly constructed theoretical models -- will be the only way to achieve a complete understanding of the dark matter in our universe and become confident that any candidate actually provides the relic density.} \\preprint{} \\begin{document} ", "introduction": "One of the most compelling hints for physics beyond the standard model is the cosmological observation that nearly a quarter of our universe consists of cold dark matter. In the next few years, the Large Hadron Collider (LHC) shows the promise of producing these elusive particles and possibly measuring their microscopic properties \\cite{Brhlik:2000dm,list}. This will be challenging and using LHC observations to reconstruct a complete theory of cosmological dark matter could prove even more challenging, if possible at all. New information will soon come from PAMELA, GLAST, and other experiments and the interpretation of the reported DAMA annual modulation effect may be clarified. One reason is because before we can claim that we know what the dark matter actually is, it will be essential to calculate the relic density of each candidate and compare the total with the cosmological relic density determined from the dark matter's gravitational influence -- because it is impossible to measure the relic density directly. This calculation will require a detailed knowledge of {\\em all} dark matter particles that contribute to the relic density, their interactions, and a knowledge of the cosmological expansion history. For example, we already know that neutrinos make up roughly a percent of the dark matter, and most complete theories have in addition to massive neutrinos also Weakly Interacting Massive Particles (WIMPs) and axions, so naively there is no convincing reason to expect one form to entirely dominate, as first emphasized in \\cite{Brhlik:2000dm}. In addition, to calculate the relic density of dark matter candidates we must know how they annihilate as the universe cools, so we can calculate the number left today. Which diagrams are important depends on the particle interactions. For example, if the dark matter particles of interest are the lightest superpartner it could behave like the partners of the W boson (wino, $\\tilde{W}$), the Higgs boson (higgsino, $\\tilde{h}$), the photino $\\tilde{\\gamma}$, or the sneutrino $\\tilde{\\nu}$, etc. These interact differently, and annihilate differently, so it is necessary to know the relevant composition and couplings, which means measuring them, perhaps in the context of a theory that allows some to be calculated. Each direct, indirect and collider channel can contribute to this information. It is also necessary to know the cosmological history of the universe from the end of inflation until the relic density is no longer changing. Of course, as soon as there are candidates there will be acclaim for the discovery of dark matter, but the question is too important to ignore the need for a proper calculation of the relic density. In the remainder of the paper we present a review of dark matter, emphasizing the issue of reconstructing a complete picture from both the particle physics and cosmological perspectives -- referring the reader to \\cite{Bertone:2004pz} for a more general review, particularly of candidates and assuming an entirely thermal history. In Section \\ref{section1} we review the cosmological evidence for dark matter and its expected distribution inside our galaxy. In the next section, after a critical overview of the standard calculation of thermal relic density of dark matter -- noting the numerous assumptions that are involved -- we then discuss how this naturally connects the dark matter with new physics near the scale of electroweak symmetry breaking. In Section 4 we discuss what we call the ``dark matter inverse problem''. That is, if we were able to reconstruct signatures of dark matter at LHC (the possibility of which we review) is it possible to also reconstruct the dark matter relic density from this information alone and be convinced that we have obtained a complete understanding of dark matter? The answer is most likely no, and we will give many reasons -- both phenomenological and theoretical-- why this is not only unlikely, but also not expected from complete and self consistent theories. Instead, as we discuss in Section 5, by combining LHC data with other direct and indirect detection experiments it may be possible to attain a complete understanding of dark matter. Indeed, this is a particularly exciting time, since a number of experiments such as GLAST and PAMELA should be reporting data soon, many other experiments will come online in the near future, and when these astrophysical probes are combined with collider data and cosmological probes a more complete picture of dark matter will emerge. In the last section we conclude with a summary of how one should proceed given a detection of a dark matter candidate at LHC, direct, or indirect detection experiment. ", "conclusions": "Both cosmological observations and theoretical consistencies of electroweak symmetry breaking seem to suggest the presence of WIMPs. The expectation of data coming from LHC makes this an exciting time to ask whether these two notions of dark matter are one and the same -- or if the story is perhaps more complicated. We have seen in this review that theory would suggest the latter, but that by combining data from colliders with that coming from cosmological and astrophysical probes of dark matter a complete understanding of dark matter can probably be achieved. In fact, one is tempted to compare this situation with the concordance approach of modern precision cosmology where one experiment alone (such as WMAP) is not enough to overcome uncertainties, but when combined with other probes (such as supernovae, baryon acoustic oscillations, galaxy surveys, etc.) one can obtain a more constrained and precise fit to the cosmological parameters. Thus, by combining the data from different experiments in this way -- along with explicit and complete theoretical models that not only predict specific dark matter candidates, but also the cosmological expansion history -- we can hope to obtain a complete understanding of dark matter. \\\\ \\noindent {\\bf How should one proceed once a dark matter candidate is established?} \\\\ One good path is to calculate the relic density in models that are sufficiently complete to have a cosmological history, and accommodate electroweak symmetry breaking. Presumably several different models would emerge as consistent with dark matter and LHC data, but all of them would have other predictions and tests that would allow one class of them to eventually become convincing. Another approach would be to calculate the relic density as if the history were thermal (using the experimentally inferred self annihilation cross section). Then there are several possibilities (as discussed above): \\bi \\item The relic density could come out about right (i.e., agree with the cosmological abundance determined by e.g. WMAP), in which case it would belong to the previous paragraph. \\item The relic density could come out smaller than the WMAP number, in which case either: \\bi \\item There is another source of dark matter, e.g. axions or some heavy hidden sector matter. \\item And/Or, the dark matter was non-thermally produced yielding an increased relic density. \\item And/Or, at the time the dark matter was produced the cosmic expansion was not entirely radiation dominated, e.g. kination. \\ei \\item The relic density could come out larger than the WMAP number, in which case either: \\bi \\item There was non-thermal production yielding less than the thermal relic density. \\item And\\footnote{Note: As we discussed above, at the time non-thermal production occurs (e.g. from decay of a heavy scalar) this will produce substantial entropy. This will dilute the thermal abundance, which will also contribute to the total relic density. One could imagine more complicated scenarios where all of these factors play a role and the relic density of dark matter gets both thermal and non-thermal contributions (as is expected for example in the case of axions \\cite{arxiv:0807.1726}).}/Or, there were additional sources of entropy after thermal freeze-out (e.g. decays of heavier particles, such as scalars (moduli) or gravitinos). \\item And/Or the cosmic expansion was modified after thermal freeze-out of the dark matter (e.g. thermal inflation). \\ei \\ei In all cases, we will be learning, amazingly, about the cosmological history of the universe, from LHC, direct, and indirect detection of dark matter." }, "0807/0807.1717_arXiv.txt": { "abstract": "Near Earth Objects (NEOs) are fragments of remnant primitive bodies that date from the era of Solar System formation. At present, the physical properties and origins of NEOs are poorly understood. We have measured thermal emission from three NEOs --- (6037) 1988~EG, 1993~GD, and 2005~GL --- with Spitzer's IRAC instrument at 3.6, 4.5, 5.8, and 8.0~$\\mu$m (the last object was detected only at 5.8~and 8.0~$\\mu$m). The diameters of these three objects are 400~m, 180~m, and 160~m, respectively, % with uncertainties of around 20\\% % (including both observational and systematic errors). For all three the geometric albedos are around~0.30, in agreement with previous results that most NEOs are S-class asteroids. For the two objects detected at 3.6~and 4.5~$\\mu$m, diameters and albedos based only on those data agree with the values based on modeling the data in all four bands. This agreement, and the high sensitivity of IRAC, show the promise of the Spitzer Warm Mission for determining the physical parameters for a large number of NEOs. ", "introduction": "Near Earth Objects (NEOs) are bodies whose orbits pass within a few tenths of an AU of the Earth's orbit. As of this writing, there are around 5000 NEOs known. The Pan-STARRS program is likely to increase the number of known NEOs to $\\sim$10,000 or more by 2013. These bodies are of critical interest to both the scientific community and the public. The NEO population is the source of potential Earth-impacting asteroids (hence a Congressional mandate to study these objects), and some may be easily reached by spacecraft, enabling our exploration of the nearby Solar System. Because NEOs have only recently been perturbed out of orbits in the main asteroid belt, and so are relatively primitive objects, they contain information that records the origin of our Solar System and that may offer insight into both the past (via delivery of organic material) and future (via impact-caused extinctions) of life on Earth. However, the physical characterization of these objects is by far outpaced by discoveries. The NEO size and albedo distributions, crucial inputs for Solar System studies as well as the assessment of the NEO Earth impact hazard, are only poorly constrained, especially at the smallest sizes \\citep[e.g.,][]{StuartBinzel2004}. The diameter and albedo of asteroids can be determined from thermal-infrared observations together with appropriate thermal modeling \\citep[e.g.,][]{STM,LebofskySpencer1989,HarrisLagerros2002}, provided the absolute magnitude $H$ \\citep[optical magnitude at a standardized observing geometry; see][]{HG} is known. A suitable thermal model for NEOs is the Near-Earth Asteroid Thermal model \\citep[NEATM, ][]{neatm}, which allows for simultaneous fits of the asteroid diameter, albedo, and effective surface temperature (parametrized through the beaming parameter $\\eta$) \\citep[e.g.,][]{neatm,HarrisLagerros2002}. We have measured thermal emission from three NEOs with the Spitzer Space Telescope \\citep{werner04} and present our data (\\S2) and results (\\S3) here. Using the NEATM, we derive albedos and diameters for all three objects (\\S3). In \\S4 we comment on the apparent thermal inertias for these objects and demonstrate that a study of NEOs could profitably be carried out with the Spitzer Warm Mission. ", "conclusions": "All three objects have diameters less than 500~meters, making them among the smallest NEOs with known albedos and diameters, and among the smallest individual objects studied with the Spitzer Space Telescope. All three objects also have albedos close to~0.3, in agreement with the idea that the NEO population is dominated by S-class asteroids \\citep[e.g.,][]{binzel04}. \\citet{binzel04} also found that the albedos for S-class (and related classes) NEOs rise from their main belt average value around~0.22 to greater than~0.3 for objects $\\lesssim$500~m. Our results appear to confirm this trend (Figure~\\ref{sed}), though with small numbers and not insignificant error bars. It is quite premature to discuss the reality of the potentially interesting downward turn at even smaller sizes. The best-fit (floating) $\\eta$ values found for 6037 and 1993~GD are roughly consistent with empirical expectations \\citep{Delbo2003}, which were recently used by \\citet{Delbo2007} to determine the typical thermal inertia of $D\\sim 1$~km NEAs. Thermal inertia is indicative of the presence or absence of loose material (regolith) on the surface and is a key parameter for model calculations of the Yarkovsky effect, a non-gravitational force that severely influences the orbital dynamics of small asteroids. (Note that \\citet{Vokrouhlicky2005} list 6037 as a potential target for direct observations of the Yarkovsky effect.) Our results suggest that our targets have unremarkable thermal inertias and may be similar to the 320~meter diameter S-type NEO (25143) Itokawa \\citep{ThMueller,MuellerDiss}, the target of the Hayabusa mission. However, more work and a systematic, large survey are needed to determine the typical thermal inertia of sub-km NEAs. For 6037 and 1993~GD the diameters and albedos we derive using only 3.6~and 4.5~$\\mu$m data are in agreement with our other model solutions, particularly for 6037, which is strongly detected (SNR$>$10) in both bands. This agreement has important implications for the Spitzer Warm Mission. After Spitzer's onboard cryogen is exhausted, observations in IRAC bands 1 and 2 (3.6 and 4.5~$\\mu$m) can still be made with essentially no loss of sensitivity. Our results show the promise of capitalizing on the superior sensitivity of IRAC to determine the physical properties of a large number of NEOS during the Spitzer Warm Mission." }, "0807/0807.4304_arXiv.txt": { "abstract": "The critical issue in cosmology today lies in determining if the cosmological constant is the underlying ingredient of dark energy. Our profound lack of understanding of the physics of dark energy places severe constrains on our ability to say anything about its possible dynamical nature. Quoted errors on the equation of state, $w(z)$, are so heavily dependent on necessarily over-simplified parameterisations they are at risk of being rendered meaningless. Moreover, the existence of degeneracies between the reconstructed $w(z)$ and the matter and curvature densities weakens any conclusions still further. We propose consistency tests for the cosmological constant which provide a direct observational signal if $\\Lambda$ is wrong, regardless of the densities of matter and curvature. As an example of its utility, our flat case test can warn of a small transition from $w(z)=-1$ of $20\\%$ from SNAP quality data at 4-$\\sigma$, even when direct reconstruction techniques see virtually no evidence for deviation from~$\\Lambda$. It is shown to successfully rule out a wide range of non-$\\Lambda$ dark energy models with no reliance on knowledge of $\\Omega_m$ using SNAP-quaility data and a large range for using $10^5$ supernovae as forecasted for LSST. ", "introduction": " ", "conclusions": "" }, "0807/0807.4852.txt": { "abstract": "The increased sensitivity of future radio telescopes will result in requirements for higher dynamic range within the image as well as better resolution and immunity to interference. In this paper we propose a new matrix formulation of the imaging equation in the cases of non co-planar arrays and polarimetric measurements. Then we improve our parametric imaging techniques in terms of resolution and estimation accuracy. This is done by enhancing both the MVDR parametric imaging, introducing alternative dirty images and by introducing better power estimates based on least squares, with positive semi-definite constraints. We also discuss the use of robust Capon beamforming and semi-definite programming for solving the self-calibration problem. Additionally we provide statistical analysis of the bias of the MVDR beamformer for the case of moving array, which serves as a first step in analyzing iterative approaches such as CLEAN and the techniques proposed in this paper. Finally we demonstrate a full deconvolution process based on the parametric imaging techniques and show its improved resolution and sensitivity compared to the CLEAN method. \\\\ {\\bf Keywords}: Radio astronomy, synthesis imaging, parametric imaging, minimum variance, robust beamforming, convex optimization, CLEAN. ", "introduction": "The future of radio astronomical discoveries depends on achieving better resolution and sensitivity while maintaining immunity to terrestrial interference which is rapidly growing. To achieve the improved sensitivity and higher resolution new instruments are being designed. The Square Kilometer Array (SKA) \\footnote{For information on the SKA project the reader is referred to http:$//$www.skatelescope.org.} \\cite{hall2005} and the Low Frequency Array (LOFAR) \\cite{bregman2005} are two of these advanced instruments. Achieving higher sensitivity to observe faint objects results in high dynamic range requirements within the image, where strong sources can affect the imaging of the very weak sources. On the other hand, Moore's law \\cite{moore65} together with recent advances in optimization theory \\cite{boyd2004} open the way to the application of more advanced computational techniques. In contrast to hardware implementation, these image formation algorithms, that are implemented in software can benefit from the continuing increase in computational power, even after the antennas and the correlators will be built. In this paper we extend the parametric deconvolution approach of \\cite{leshem2000a} to obtain better power estimation accuracy and higher robustness to interference and modeling uncertainty. The algorithms presented here can also be used in conjunction with real-time interference mitigation techniques as described in \\cite{leshem2000a} and \\cite{leshem2000b}. We briefly describe the current status of radio astronomical imaging techniques. For a more extensive overview the reader is referred to \\cite{thompson86}, \\cite{yen85} or \\cite{taylor99}. A good historic perspective can be found in \\cite{kellerman2001}, whereas \\cite{sault2007} provides a very recent perspective. The principle of radio interferometry has been used in radio astronomy since 1946 when Ryle and Vonberg constructed a radio interferometer using dipole antenna arrays \\cite{ryle52}. During the 1950's several radio interferometers which use the synthetic aperture created by movable antennas have been constructed. In 1962 the principle of aperture synthesis using earth rotation has been proposed \\cite{ryle62}. The basic idea is to exploit the rotation of the earth to obtain denser coverage of the visibility domain (spatial Fourier domain). The first instrument to use this principle was the five kilometer Cambridge radio telescope. During the 1970's new instruments with large aperture have been constructed. Among these we find the Westerbork Synthesis Radio Telescope (WSRT) in the Netherlands and the Very Large Array (VLA) in the USA. Recently, the Giant Microwave Telescopes (GMRT) has been constructed in India and the Allen Telescope Array (ATA) in the US. Even these instruments subsample the Fourier domain, so that unique reconstruction is not possible without some further processing known as deconvolution. The deconvolution process uses some a-priori knowledge about the image to remove the effect of ``dirty beam'' side-lobes. Two principles dominate the astronomical imaging deconvolution. The first method was proposed by Hogbom \\cite{hogbom74} and is known as CLEAN. The CLEAN method is basically a sequential Least-Squares (LS) fitting procedure in which the brightest source location and power are estimated. The response of this source is removed from the image and then the process continues to find the next brightest source, until the residual image is noise-like. During the years it has been partially analyzed \\cite{schwarz78}, \\cite{schwarz79} and \\cite{tan86}. However full analysis of the method is still lacking due to its iterative nature. The CLEAN algorithm has many recent flavors which are capable of faster performance e.g., the Clark version \\cite{clark80} and the Cotton-Schwab \\cite{schwab84}. In these versions the dirty image is recomputed only after several point sources have been estimated. Furthermore the sources are subtracted from the ungridded visibility. This results in better suppression of the sources. Interestingly it is well known that when the noise model is non-white (as in radar clutter modeled by ARMA processes in SAR applications) algorithms such as RELAX \\cite{li96} outperform the CLEAN algorithm. A second approach proposed by Jaynes \\cite{Jaynes57} is maximum entropy deconvolution (MEM). The basic idea behind MEM is the following. Not all images which are consistent with the measured data and the noise distribution satisfy the positivity demand, i.e., the sky brightness is a positive function. Consider only those that satisfy the positivity demand. From these select the one that is most likely to have been created randomly. This idea has also been proposed in \\cite{frieden72} and applied to radio astronomical imaging in \\cite{gull78}. Other approaches based on the differential entropy have also been proposed \\cite{ables74}, \\cite{wernecke77}. An extensive collection of papers discussing the various methods and aspects of maximum entropy can be found in the various papers in \\cite{roberts84}. Briggs \\cite{briggs95} proposed a non-negative least squares approach (NNLS) which eliminates the need for iterative processing. However, the computational complexity is very large. In this paper we use a reformulation of the image formation problem as a parameter estimation problem using a set of covariance matrices, measured at the various observation epochs \\cite{leshem2000a}. This yields a model where the array response is time varying. Previous research on time varying arrays and their application to direction-of arrival (DOA) estimation includes \\cite{zeira95}, \\cite{zeira96} and \\cite{sheinvald98}. In \\cite{leshem2000a} we proposed a simplified ML estimator. Lanterman \\cite{lanterman2000} developed a full EM algorithm for implementing the MLE proposed in \\cite{leshem2000a}. The algorithm performs quite well although it is quite complex compared to the solutions described in this paper. The above mentioned algorithms assume perfect knowledge of the instrumental response (point spread function). Due to various internal and external effects this assumption holds only approximately. One way to overcome this problem is the use of calibrating sources. An unresolved source with known parameters is measured, and by relating the model errors to the array elements a set of calibration equations is solved. A much more appealing solution is to try to improve the fitting between the data and the sky model by adjusting the calibration parameters. Another possibility \\cite{noordam82} is to use the redundant structure of the array to solve for the calibration parameters (this is possible only for some arrays which have redundant baselines, such as the WSRT). A good overview of the various techniques is given in \\cite{pearson84}. In this paper we extend the above methods in several directions. First, we extend the parametric formulation of \\cite{leshem2000a} to the non-coplanar array case and to polarimetric imaging. Then we propose relatively ``low'' complexity approaches for the image deconvolution problem, based on Minimum Variance Distortionless Response (MVDR) and its robust extensions. We call these extensions Least-Squares Minimum Variance Imaging (LS-MVI). We provide a new type of dirty image that has isotropic noise response, something desirable in imaging applications. This is done by generalizing the work of Borgiotti and Kaplan \\cite{borgiotti79} to the moving array case. We discuss acceleration techniques, related to the Clark \\cite{clark80} and the Cotton-Schwab \\cite{schwab84} approaches to CLEAN. However, in contrast to the classical CLEAN case, the accelerated algorithm involves semi-definite programming of low order, ensuring that the covariance matrices remain positive semi-definite after the subtraction. We provide analytic expressions for the asymptotic bias of the MVDR based imaging. We also relate the classical self-calibration technique to a novel extension of the robust Capon beamformer to the moving array case, showing that self-calibration can be cast in terms of semi-definite programming. This has advantage over previous approach to self-calibration since the covariance structure maintains its positive definite structure. We also demonstrate full parametric deconvolution process based on the proposed technique and compare CLEAN and LS-MVI on simulated images. We will show in simulations that LS-MVI, indeed, significantly outperforms the classical CLEAN algorithm over a wide range of parameters, in terms of better resolution, higher sensitivity and dynamic range. The structure of the paper is as follows. In section \\ref{sec:astron} we describe the astronomical measurement equation. The measurement equation is subsequently rephrased in a more convenient matrix formulation both for non co-planar and co-planar arrays. It is then extended with the effect of noise and unknown calibration parameters. We also discuss extension to polarimetric measurements. In section \\ref{sec:LS_MVI} we discuss the new Least Squares Minimum Variance Imaging and extensions to self calibration using robust Capon techniques. In section \\ref{sec:MVDR_analysis} we present bias analysis of the MVDR imaging technique. In \\ref{sec:simulation} we describe several computer simulations demonstrating the gain in parametric deconvolution relative to the CLEAN in terms of resolution, sensitivity and capability to model extended structures. We also provide example of the tightness of the statistical bias analysis of the MVDR DOA estimation with a moving array. We end up with some conclusions. ", "conclusions": "In this paper we extend the matrix formulation of \\cite{leshem2000a} to non co-planar arrays and polarimetric measurements. Then we propose a new parametric imaging technique that improves the resolution and sensitivity over the classical CLEAN algorithm. The method is based on several improvements: A new type of dirty image, LS estimation of the powers and semi-definite constraints. We show how the technique can be combined into self-calibration using semi-definite programming. Our semi-definite self-calibration algorithm also provides a new approach to robust beamforming with a moving array, extending the techniques of \\cite{li03}, \\cite{vorobyov03} and \\cite{lorenz05}. We provide statistical analysis of the location estimator. Simulated examples comparing full deconvolution using LS-MVI and comparing them to the CLEAN method are presented. These simulations demonstrate that the parametric approach has higher resolution, is more robust to source structures, and performs better in noisy situations. The great potential of the methods proposed in this paper is a first step, towards the development of more advanced imaging techniques, capable of providing higher dynamic range and interference immunity as required by the radio telescopes of the future. %" }, "0807/0807.1467_arXiv.txt": { "abstract": "{} {We searched for the presence of extended emission-line regions (EELRs) around low-redshift QSOs.} {We observed a sample of 20 mainly radio-quiet low-redshift quasars ($z<0.3)$ by means of integral field spectroscopy. After decomposing the extended and nuclear emission components, we constructed \\Ox\\ $\\lambda5007$ narrow-band images of the EELR to measure the total flux. From the same data we obtained high S/N ($>$50) nuclear spectra to measure properties such as \\Ox/\\Hb\\ flux ratios, \\Fe\\ equivalent widths and H$\\beta$ line widths.} {A significant fraction of the quasars (8/20) show a luminous EELR, with detected linear sizes of several kpc. Whether or not a QSO has a luminous EELR is strongly related with nuclear properties, in the sense that an EELR was detected in objects with low \\Fe\\ equivalent width and large H$\\beta$ FWHM. The EELRs were detected preferentially in QSOs with larger black hole masses. There is no discernible relation, however, between EELR detection and QSO luminosity and Eddington ratio.} {}{} ", "introduction": "Quasars may show extended emission-line regions (EELRs) stretching over characteristic scales of ten to hundred kpc. Luminous EELRs have been found in particular around steep-spectrum radio-loud quasars \\citep[RLQs,][]{Boroson:1984,Heckman:1991,Crawford:2000,Stockton:2002,Fu:2006} and also radio galaxies \\citep[e.g.][]{Villar-Martin:2006,Villar-Martin:2007}. There are also indications that radio-quiet quasars (RQQs) can have EELRs \\citep[e.g.,][]{Boroson:1985,Stockton:1987,Bennert:2002}. Most existing studies, however, are based either on long-slit spectroscopy or on narrow-band imaging; while the former technique will generally capture only part of any extended emission, the latter does not provide any spectroscopic information. Integral field spectroscopy is a relatively new and powerful tool to study AGN host galaxies. The combined diagnostic power of imaging and spectroscopy allows for a flexible treatmeant of the data, facilitating the construction of broad- or narrow-band images, of 2-dimensional velocity maps as traced by emission or absorption lines, or the spectral analysis of selected regions within the field of view. Here we present preliminary results from a study of 20 low-redshift QSOs, most of which are RQQs, observed with an integral field unit (IFU). After decomposing nuclear and extended emission we identified 8 quasars with clearly detectable EELRs. We focus on exploring the relation between the nuclear spectral properties of a QSO and the existence or non-existence of a luminous EELR. A full account of our IFU study will be given in a separate paper (Husemann et al., in preparation; hereafter Paper~II). ", "conclusions": "By means of integral field spectroscopy, we have revealed the presence of extended emission line regions (EELRs) in a sample of predominantly radio-quiet low-redshift QSOs. The sizes and luminosities are much higher than those of the well-known EELRs in Seyfert galaxies, extending out to several kpc and therefore stretching over the entire host galaxy (and possibly beyond). Some 40~\\% of the objects in our sample show a prominent EELR, while we failed to detect any EELR signature in the remaining 60~\\%. The detection or non-detection of an EELR can be linked to the \\emph{nuclear} spectral properties of the quasar itself, independent of their radio classification, in excellent agreement with what \\citet{Boroson:1984} and \\citet{Boroson:1985} found from their off-nuclear long-slit spectra. Our observations go further in that our integral field spectral data map out the entire nuclear-subtracted EELRs and derive their luminosities. Correlations between various nuclear spectral properties were already studied by many authors \\citep[e.g.][]{Boroson:1992,Sulentic:2000,Netzer:2004}. Here we provide new evidence that properties of the nuclear spectra, depending on the sub-pc scale conditions in AGN, are linked to properties of the host galaxies extending over several kpc. It thus appears possible to predict with high confidence whether or not a QSO will show a very extended emission line region just from its nuclear spectrum: If it has broad \\Hb\\ and weak \\Fe, it will probably have not only strong nuclear, but also strong extended \\Ox\\ emission. It is tempting to interpret this relation between sub-pc and super-kpc scales in the context of the current discussion about AGN feedback affecting the host galaxies. Since we can rule out the QSO luminosity as a main driver of the differences between presence or absence of an EELR, the remaining possibilities are that there could be differences in the structure of the nuclear region (obscuration in connection with possible inclination effects); or differences due to an outflow or jet, or the host galaxies of the two QSO classes are intrinsically different, in particular with respect to their overall distribution of warm ionised gas. We conclude that QSO host galaxies come in at least two flavours: with or without extended ionized gas. Whether this highlights different initial conditions or a sequence of evolutionary stages we cannot say at present. More and better quality data will be needed to elucidate these intriguing trends." }, "0807/0807.1698_arXiv.txt": { "abstract": "We present optical time series spectroscopy of the pulsating white dwarf star G~29-38 taken at the Very Large Telescope (VLT). By measuring the variations in brightness, Doppler shift, and line shape of each spectrum, we explore the physics of pulsation and measure the spherical degree ($\\ell$) of each stellar pulsation mode. We measure the physical motion of the g-modes correlated with the brightness variations for three of the eight pulsation modes in this data set. The varying line shape reveals the spherical degree of the pulsations, an important quantity for properly modeling the interior of the star with asteroseismology. Performing fits to the H$\\beta$, H$\\gamma$, and H$\\delta$ lines, we quantify the changing shape of the line and compare them to models and previous time series spectroscopy of G~29-38. These VLT data confirm several $\\ell$ identifications and add four new values, including an additional $\\ell$=2 and a possible $\\ell$=4. In total from both sets of spectroscopy of G~29-38, eleven modes now have known spherical degrees. ", "introduction": "DAVs, or ZZ Cetis, are variable white dwarf stars with hydrogen atmospheres. They are the coolest known white dwarf pulsators and reside in an instability strip near 12,000~K. Their pulsations are multi-periodic with amplitudes as high as 5~\\percent\\ and periods between 70~s and 1100~s. In the past few years the number of known DAVs has increased from a few dozen to more than a hundred \\cite[see][]{voss07,castanheira06,mullally05,mukadam04}. The pulsators at the hot end of the instability strip show lower amplitudes and more stable pulsations, while the others show more complex pulsations including variability of the excited modes, and an abundance of harmonic and combination modes. The high gravity (log(g)$\\sim$8 cgs) of the DAVs favor gravity-mode type pulsations \\citep{rkn82}, where motion along the surface of the star produces regions with varying temperatures. To describe the spatial distribution of these regions, each of the pulsations on the star are characterized by a spherical harmonic (Y$_{\\ell,m}$) and the radial order, $n$. The spherical degree, $\\ell$, and azimuthal order, $m$, describe the number of nodal lines across the surface of the star, and directly affect observations of the pulsations. The pulsation's radial component allows asteroseismological models to determine the internal structure of the star. The DAVs have evaded unambiguous asteroseismological modeling in part because of the difficulty in measuring the spherical degree of their pulsation modes. Usually a value of $\\ell$=1 is assumed in these models. However, it is possible to fit modes with higher values of $\\ell$ and this can significantly change the parameters of the model. For example, \\citet{bradley06} showed that using the identification of $\\ell$=4 \\citep{thompson04}, instead of $\\ell$=2 \\citep{cast04} for one mode, changes the modeled depth of the Hydrogen layer in G~185-32 by a factor of 100. One method to determine $\\ell$ independent of asteroseismological modeling was introduced by \\citet{robinson95}. They showed that $\\ell$ can be obtained from time series optical and UV spectroscopy. The effects of limb darkening increase in the blue wavelengths and decrease through the spectral lines, changing the fraction of integrated light coming from the limb. As such, the pulsation amplitude varies with both wavelength and the spherical degree of the mode. \\citet*{c00} (hereafter C00) measured chromatic amplitudes (pulsation amplitude at each wavelength bin) across the Hydrogen Balmer lines and demonstrated that spectral variations can indeed be used to identify spherical degree. The same method was applied to the DAVs HS~0507+0434B, G~117-B15A, and G~185-32 \\citep{kotak02a, kotak04, thompson04} and the DBV GD~358 \\citep{kotak03}. Generally, the method worked well, although for one mode -- the 272~s mode in G~117-B15A -- the chromatic amplitudes are extremely puzzling \\citep{kotak04}. Also, they found that the chromatic amplitudes provided unambiguous identifications only for the strongest modes; for weaker ones, noise obscured the expected visible differences. In order to improve the precision, \\citet{kotak03} attempted to use the slope and curvature in selected wavelength ranges of the chromatic amplitudes as measures of spherical degree, with somewhat limited success. A different approach was demonstrated by \\citet{thompson04} where they analysed fits to the spectral lines rather than the observed fluxes in each spectral bin. The chromatic amplitudes created from these fits did allow for mode identifications from lower signal-to-noise data. G~29-38 is one of the brightest DAVs with complex pulsations and some of the largest amplitudes, making it a prime candidate for time series spectroscopy. It has an abundance of modes, but they are not all excited simultaneously and many observation runs are necessary to measure them all. \\citet{kleinman98} compiled 10~years of photometric observations and found a possible series of $\\ell$=1 modes on G~29-38. Time series spectroscopy taken at the Keck Observatory with the LRIS spectrograph provided chromatic amplitudes of G~29-38 that directly detected an $\\ell$=2 mode at 776~s (C00) and gave strong evidence for another at 920~s \\citep{kotak02a}. Using the same data, \\citet*{vk00} (hereafter VK00) measured the periodic line-of-sight velocities associated with the longitudinal motion of the gas. For the 776~s mode they found relatively large motion as compared to the measured flux amplitude, again indicating that the mode is $\\ell$=2 \\citep{dz82}. The measurements of the pulsation velocities were confirmed by \\citet{thompson03} using a series of high resolution spectra of G~29-38 and provided observational evidence regarding the convective driving theory proposed for DAVs \\citep{bhill83, gw1}. The velocities also hint at the nature of the many combination and harmonic modes present on these pulsators. These combinations and harmonics most likely result from nonlinear mixing at the surface of the star and are not real modes that probe the interior \\citep{wu01, montgomery05}. VK00 reinforced this idea by finding no velocity signature associated with the combinations and harmonics on G~29-38. However, \\citet{thompson03} reported one combination mode with a significant velocity detection. In this paper we show the results of time series spectroscopy of G~29-38 taken at the VLT (Very Large Telescope) in 1999. We present the light and velocity curves measured from the series of spectra. We fit 13 modes in the light curve and find variations at the same period in the velocity curve for the largest amplitude pulsations. We then perform fits to the Balmer lines to quantify the changing line shape. Since we are using fits to the observed spectral lines, we can present the line shape variations in terms of the fitted values: the normalized equivalent width variations of the Gaussian and Lorentzian functions. We perform the same analysis on the 1996 Keck spectral series presented by VK00 and C00 as a check of our method and recover their results of an $\\ell$=2 mode among a series of $\\ell$=1 modes. Application of the same method to our VLT data adds four additional mode identifications to the G~29-38 pulsation spectrum. ", "conclusions": "" }, "0807/0807.2522_arXiv.txt": { "abstract": "The Guide Star Catalog II (GSC-II) is an all-sky database of objects derived from the uncompressed Digitized Sky Surveys that the Space Telescope Science Institute has created from the Palomar and UK Schmidt survey plates and made available to the community. Like its predecessor (GSC-I), the GSC-II was primarily created to provide guide star information and observation planning support for {\\it Hubble Space Telescope}. This version, however, is already employed at some of the ground-based new-technology telescopes such as GEMINI, VLT, and TNG, and will also be used to provide support for the James Webb Space Telescope (JWST) and Gaia space missions as well as the Large Sky Area Multi-Object Fiber Spectroscopic Telescope, one of the major ongoing scientific projects in China. Two catalogs have already been extracted from the GSC-II database and released to the astronomical community. A magnitude-limited ($R_F$ = 18.0) version, GSC2.2, was distributed soon after its production in 2001, while the GSC2.3 release has been available for general access since 2007. The GSC2.3 catalog described in this paper contains astrometry, photometry, and classification for 945,592,683 objects down to the magnitude limit of the plates. Positions are tied to the International Celestial Reference System; for stellar sources, the all-sky average absolute error per coordinate ranges from $0.''2$ to $0.''28$ depending on magnitude. When dealing with extended objects, astrometric errors are ~20\\% worse in the case of galaxies and approximately a factor of 2 worse for blended images. Stellar photometry is determined to 0.13-0.22 mag as a function of magnitude and photographic passbands ($R_F$, $B_J$, $I_N$). Outside of the galactic plane, stellar classification is reliable to at least 90\\% confidence for magnitudes brighter than $R_F=19.5$, and the catalog is complete to ~$R_F$=20. ", "introduction": "The Guide Star Catalog II (GSC-II) is an astronomical database constructed from the scanned images of 9541 Palomar and UK Schmidt photographic sky survey plates digitized at Space Telescope Science Institute (STScI). These same plate images are also known as the Digitized Sky Survey (DSS), and are accessible directly from the STScI eb site as well as from a number of major astronomical data centers around the world. A subset of the original images, based on the second-epoch Palomar surveys only (also known as DPOSS), is available separately from Caltech \\citep{2003BASBr..23Q.197D}. Whilst all the images represent over 8 terabytes of data that are archived at STScI, the use of the H-transform compression technique \\citep{1992doss.conf..167W,1994SPIE.2199..703W} has enabled a reduction in data volume to $\\approx 1$ terabyte for distribution to data centers. Each individual uncompressed image was processed to detect the objects, and calibrations were obtained for each plate by polynomial modeling against photometric and astrometric reference catalogs. The average accuracy in position for stellar objects of intermediate magnitude ($V\\lesssim 18.5$) is $\\sim 0.3''$, whilst the corresponding astrometric precision, on a scale of approximately 0.5 degrees, is close to $0.2''$. Photometric calibrations, carried out in the natural plate passband, show typical errors of 0.1-0.2 mag and systematic offsets lower than 0.1 mag. Stars are correctly classified as such with at least 90\\% confidence for magnitudes brighter than $R_F=19.5$. The plate catalogs were loaded into a database system known as Catalog of Objects and Measured Parameters from All Sky Surveys (COMPASS), \\citep[see][]{1998asal.confE...3L}. This database was built using \"Objectivity\", a commercial Object-Oriented Database Management System; it constitutes a repository for $\\approx 5$ billion measurements, and is $\\approx2.5$ terabytes in size. Once the individual observations have been cross-matched to link different measurements of unique astronomical objects, a catalog can be exported from the database as a collection of FITS binary tables. Then, the export catalog is made available to the community via Web services and also provided to a number of data centers. Over the last few years there have been a number of interim releases from the GSC-II, primarily to consortium members for telescope operations. The latest public release, the GSC~2.3.2, contains 945,592,683 unique objects and is $\\approx 170$ GB in size. Details about the release content and access methods are available from the ST ScI and Osservatorio Astronomico di TOrino (OATo) Web sites. In this paper we describe the construction, calibration, and overall quality of GSC~2.3.2 , the current catalog release (for simplicity, GSC~2.3 in the rest of this paper), which was investigated through internal tests and a number of external comparisons to suitable literature data. \\subsection{Background} In the age of new-technology telescopes and space-based missions, it is useful to recall that an important part of observational astronomy has historically been the creation of catalogs containing the reference and (often) target objects required to support observing programs. From a historical viewpoint it is fascinating that astrometry---the oldest branch of astronomy---is so vital to the success of modern high-tech observatories. As the technological complexity (and cost) of building and operating telescopes has increased enormously over the last 20 years, so has the effort to provide the best scientific return for these investments. One of the important issues is the optimization of observing efficiency, which depends on the use of proper pointing and input catalogs as well as on the access to digitized versions of the optical sky surveys. The need for a deeper, all-sky catalog was highlighted early by the creation of the first Guide Star Catalog (GSC-I) to satisfy the pointing requirements of the {\\it Hubble Space Telescope} ({\\it HST}). An in-depth description of the GSC-I catalog may be found in a set of three papers \\citep{1990AJ.....99.2019L,1990AJ.....99.2059R,1990AJ.....99.2082J}. The GSC-I was used for {\\it HST} observation planning as well as target acquisition and tracking and has proved to be very reliable for its intended purpose. In addition, since its publication on CD-ROM, this catalog had become widely used by many ground-based telescopes to speed up the process of finding guide stars. Although the GSC-I has been used with great success operationally and scientifically, it became clear (even during its construction) that it was possible to improve its usefulness by addressing the known systematic calibration errors. Indeed, it was clear that an increase in scope to include multicolor, multi-epoch data would eventually lead to the requirements for a new much improved archive, the GSC-II. In addition to the risk of damaging the UV-sensitive MAMA detectors in the second generation of {\\it HST} instruments by O stars as faint as 19th magnitude, it was generally acknowledged that the proper motions of the guide stars during the proposed lifetime of {\\it HST} would pose a potential problem for the accurate pointing of the telescope during the latter years of its operation. It was also realized that a fainter archive with multicolors such as the GSC-II would address a number of other astronomical needs such as: supporting adaptive optics on the next generation of large telescopes, remote or queue scheduling capabilities, and improved detector instruments requiring color information of the target objects. In order to be prepared for this, STScI negotiated access to the original plates of the Palomar Observatory Sky Survey (POSS-II) and, in partnership with the Anglo-Australian Observatory (AAO), undertook the Second Epoch Southern (SES) red survey with the UK Schmidt Telescope Unit (UKSTU). These new surveys, when combined with the plate material used in the GSC-I and the earlier POSS-I and UKSTU surveys, provide the material to generate a GSC-II based on, at least, two epochs and three passbands. \\begin{figure*}[t] \\centering \\includegraphics[scale=.70,angle=0]{gsc2paper_filter.ps} \\caption{Transmission curves of the photographic passbands $B_J, R_F,$ and $I_N$ for the Palomar (solid lines) and AAO (dashed line) Schmidt surveys, compared to the Johnson--Kron--Cousins BVRI$_{\\rm c}$ filters. \\label{fig:photometry:filters}} \\end{figure*} \\subsection{The GSC-II Project} STScI and OATo first began a collaboration to start developing GSC-II in 1989. STScI was primarily motivated by telescope operations, and OATo was interested in the scientific applications of this catalog for galactic structure. Once the project was underway, additional resources were contributed by ESO and GEMINI (who wished to use GSC-II for VLT and GEMINI telescope operations respectively), ST-ECF (as part of the NASA-ESA {\\it HST} funding agreement), and the Astrophysics Division of ESA (for science projects). A significant fraction ($\\sim 30\\%$) of the plate processing was performed at ST-ECF using the same pipeline software system described later. ", "conclusions": "In Table \\ref{GSCI-IIreqs}, we compare the requirements provided in the GSC-II {\\it implementation plan}---originally foreseen to reach magnitude 18 in $V$---with the actual performance of the current version of the catalog down to the same magnitude limit. In all items we have met or exceeded the specifications except for the proper motions. Table \\ref{GSCI-IIspecs} summarizes the main characteristics of GSC~2.3, which essentially superseeds the requirements in that it reaches 2-3 mag depeer. The astrometric and photometric errors reported in the table are derived from those of Tables \\ref{tab:GSC23_SDSS_astro} and \\ref{tab:GSC23_SDSS_F}, and positional uncertainties are obtained by summing in quadrature the error in each coordinate. A consideration of the global statistics shows that even with the inclusion of fainter objects we are able to meet the original specifications. Whilst no further improvements are required (or funded) for its use in {\\it HST} operations, we intend to continue development to produce the best possible catalog for scientific as well as operational uses. Our future plans include an astrometric recalibration using the UCAC~3 and applying a magnitude-dependent correction; this should not only reduce the absolute positional errors but amend the systematic error that is affecting the computed proper motions. A photometric re-reduction of all plates will also benefit from the improved overall quality of the available calibrating sequences. Finally, we plan to reclassify the objects splitting the nonstar classification into galaxy and blend as well as better determine the magnitudes of galaxies using an algorithm that does not assume a stellar profile \\citep{2007ApJS..170...33P}. As of cycle 15, GSC~2.3 is the guide catalog for {\\it HST}, and it is a reference catalog for the VLT and Gemini adaptive optics programs, where real guide stars are still preferred to laser ``guide stars''; it has also been provided to the {\\it Chandra}, {\\it Galex}, {\\it XMM-Newton} and {\\it Swift} missions. In the preparation for the Gaia mission, the current catalog version is used as a snapshot of what Gaia is expected to observe \\citep{2006MmSAI..77.1172D} and as a base for the Initial Gaia Source List being compiled by OATo for the Gaia data reduction. Finally, GSC-II future release(s) will be at the heart of the Astrometric Support System of the very ambitious Large Sky Area Multi-Object Fiber Spectroscopic Telescope (LAMOST) undertaken by the Chinese Academy of Science, and will be part of the guide system for JWST \\citep{Spagna2001,2003AAS...202.0410S}. In addition to the operational uses of the catalog itself, the GSC-II database has also been mined for many scientific studies, such as planetary nebulae \\citep{2004A&A...420..207K}, young open clusters \\citep{2000A&A...357..460S}, halo white dwarfs \\citep{2006A&A...448..579C}, peculiar objects \\citep{2001Natur.413..139M,2002A&A...393L..45C}, X-ray pulsars \\citep{Panzera2003}, and the structure and kinematics of stellar populations of our Galaxy \\citep{2006A&A...451..125V,2007MNRAS.375.1381K}. Finally, it is worth going back to the density map presented in Figure \\ref{fig:skymap}, as it does illustrate GSC~2.3: it provides accurate positions, magnitudes in three bands, and stellar classification for 4$\\pi$ steradians to levels of completeness commensurate with the underlying stellar density. Future versions extracted from the GSC-II II database will only improve on this situation, including proper motions and increasing the catalog accuracy as a result of improvements in reference catalogs and reduction procedures." }, "0807/0807.0705_arXiv.txt": { "abstract": "Within the next five years, a number of direct-imaging planet search instruments, like the {\\it VLT SPHERE} instrument, will be coming online. To successfully carry out their programs, these instruments will rely heavily on a-priori information on planet composition, atmosphere, and evolution. Transiting planet surveys, while covering a different semi-major axis regime, have the potential to provide critical foundations for these next-generation surveys. For example, improved information on planetary evolutionary tracks may significantly impact the insights that can be drawn from direct-imaging statistical data. Other high-impact results from transiting planet science include information on mass-to-radius relationships as well as atmospheric absorption bands. The marriage of transiting planet and direct-imaging results may eventually give us the first complete picture of planet migration, multiplicity, and general evolution. ", "introduction": "{\\underline{{\\bf Ground-Based}}} {\\it - HiCIAO}: An adaptive optics coronagraphic simultaneous-differential-imager for the Subaru 8.2 meter telescope (Tamura et al. 2006); currently in commissioning phase. {\\it - Project 1640}: An adaptive optics coronagraphic integral field spectrograph for the Palomar 5 meter Telescope (Hinkley et al. 2008); currently in commissioning phase. {\\it - SPHERE}: An adaptive optics coronagraphic simultaneous-differential-imager and integral field spectrograph for one of the VLT 8.2 meter telescopes (Beuzit et al. 2008); commissioning expected 2011. {\\it - GPI}: An adaptive optics coronagraphic integral field spectrograph for the Gemini South 8 meter telescope (Macintosh et al. 2006); commissioning expected 2011. {\\underline{{\\bf Space-Based}}} {\\it Terrestrial Planet Finder Coronagraph\\footnote[1]{http://planetquest.jpl.nasa.gov/TPF-C}?} {\\it Terrestrial Planet Finder Interferometer\\footnote[2]{http://planetquest.jpl.nasa.gov/TPF-I}/Darwin\\footnote[3]{http://www.esa.int/science/darwin}?} \\vspace{4 mm} All of these systems will rely on results from transiting planet science to accurately interpret their observations. ", "conclusions": "" }, "0807/0807.0475_arXiv.txt": { "abstract": "We present a method to simulate color, 3-dimensional images taken with a space-based observatory by building off of the established {\\em shapelets} pipeline. The simulated galaxies exhibit complex morphologies, which are realistically correlated between, and include, known redshifts. The simulations are created using galaxies from the 4 optical and near-infrared bands (\\textit{B}, \\textit{V}, \\textit{i} and \\textit{z}) of the Hubble Ultra Deep Field (UDF) as a basis set to model morphologies and redshift. We include observational effects such as sky noise and pixelization and can add astronomical signals of interest such as weak gravitational lensing. The realism of the simulations is demonstrated by comparing their morphologies to the original UDF galaxies and by comparing their distribution of ellipticities as a function of redshift and magnitude to wider HST COSMOS data. These simulations have already been useful for calibrating multicolor image analysis techniques and for better optimizing the design of proposed space telescopes. ", "introduction": "As astronomical surveys become deeper and wider, analysis techniques correspondingly become more complex and demanding. To calibrate these methods, extensive work has already been invested in the simulation of monochromatic astronomical imaging. Simulation packages have been developed to incorporate a semi-analytic model of galaxy number counts and evolution \\cite{EWBM}, or to mimic the properties of real observations \\cite{MRCB}. However, there are currently no packages able to create correlated images across several bands. Multi-band image simulations are firstly useful to develop and calibrate analysis methods that use multicolor data. Many measurements in astronomy (for example photometry, astrometry and shape measurement) are ``inverse problems,'' where variation in a signal is easy to introduce but difficult to measure, usually due to complications involving observational seeing and noise. Simulated data provide the best way to calibrate such methods because these variables can be controlled. A known astronomical signal can be inserted into simulated data, and the accuracy of a method can be judged by examining any errors in its recovery. One example is the measurement of weak gravitational lensing. In weak lensing, light from background galaxies is lensed by foreground matter distributions, causing a shear (distortion) of the background galaxies' shapes. The distortion is easy to add during the construction of simulated data. Although the lensing signal is achromatic, color simulations can be used to test sophisticated measurement methods that take advantage of \\begin{itemize} \\item the increased number of shear-measurable galaxies if some galaxies are only sufficiently bright in certain bands, \\item reduced {\\em noise} on shear measurement (by $\\sqrt{N}$) if the intrinsic shapes of galaxies are uncorrelated between $N$ bands, and \\item reduced systematic {\\em bias} on shear measurement if the intrinsic shapes of galaxies are correlated between bands \\cite{JainJarvis}. \\end{itemize} One common challenge in weak lensing measurement is the deconvolution of galaxy shapes from the instrumental point-spread function (PSF). Since the PSF is different in each band, PSF-dependent biases will be averaged out by looking at multiple bands. Conversely, biases inherent to a method will not be ameliorated. Developing multicolor analysis techniques to exploit these tricks requires multicolor simulations. Multi-band image simulations are also useful to optimize the design and improve the science case for planned, multi-band imaging surveys such as SNAP \\cite{Aldering} or Euclid \\cite{DUNE}. These surveys require multiple bands in order to observe different types of galaxies, observe objects typically obscured in other bands, observe objects out to different redshifts, and most importantly to obtain photometric redshifts for galaxies. Engineering requirements for the design of these instruments can be derived via image simulations by measuring the (often complex and subtle) effects of engineering parameters on scientific return \\cite{High}. Predictions for the scientific return of a given mission can be similarly estimated \\cite{WLII,WLIII}. A full demonstration of the potential gains in multicolor shear measurement, or a full optimization of a future space-based lensing mission is beyond the scope of this paper. The purpose of this paper is to present a method for simulating deep, multi-color space-based images with correlated morphologies and redshifts. The simulation pipeline we present here will serve as a basis for performing the optimization of both shear measurement techniques and future space missions in future papers. Our simulation pipeline generalizes the single-color method of \\cite{MRCB}, representing complex galaxy morphologies as ``shapelets'' \\cite{Cartesian,Polar}. Shapelet-based simulations are already widely used for weak lensing. The Shear TEsting Program (STEP) used similar simulated data to test and improve shape measurement and PSF correction methods \\cite{Mass2006}. Our generalization to multi-band, 3-dimensional simulations thus increases the realism and utility of a well established technique. This paper is organized as follows. In \\S 2 we give a brief review of shapelets and how they can be used to generate simulated images. In \\S 3 we present the methodology by which we create multi-band, 3-dimensional simulations. In \\S 4 we test the realism of our simulations through comparison to the real HST data. Lastly, in \\S 5, we discuss the conclusions and summarize our findings. ", "conclusions": "We presented a method to create an arbitrary amount of 3-dimensional, color, simulated, unique deep space images. The simulations are created by perturbing a galaxy's polar shapelet coefficients in such a way as to create unique but realistic objects. The previous simulation pipeline has been expanded to correlate morphologies across four wavelength bands and now also include a redshift distribution. Though we currently use four wavelength bands, our simulation pipeline is flexible enough to include an arbitrary number of colors, should such a data set become available and useful. Our simulations were tested by comparing them to the original UDF images. They were found to have similar morphologies to the original galaxies. Additionally, the weak lensing cosmological properties of the simulations were tested against COSMOS HST data. Within reasonable error, our simulations were found to be consistent." }, "0807/0807.2646_arXiv.txt": { "abstract": "{} {For the first time, the lunar occultation technique has been employed on a very large telescope in the near-IR with the aim of achieving systematically milliarcsecond resolution on stellar sources. } { We have demonstrated the burst mode of the ISAAC instrument, using a fast read-out on a small area of the detector to record many tens of seconds of data at a time on fields of few squared arcseconds. We have used the opportunity to record a large number of LO events during a passage of the Moon close to the Galactic Center in March 2006. We have developed and employed for the first time a data pipeline for the treatment of LO data, including the automated estimation of the main data analysis parameters using a wavelet-based method, and the preliminary fitting and plotting of all light curves. } { We recorded 51 LO events over about four hours. Of these, 30 resulted of sufficient quality to enable a detailed fitting. We detected two binaries with subarcsecond projected separation and three stars with a marginally resolved angular diameter of about 2 milliarcseconds. Two more stars, which are cross-identified with SiO maser, were found to be resolved and in one case we could recover the brightness profile of the extended emission, which is well consistent with an optically thin shell. The remaining unresolved stars were used to characterize the performance of the method. } { The LO technique at a very large telescope is a powerful and efficient method to achieve angular resolution, sensitivity, and dynamic range that are among the best possible today with any technique. The selection of targets is naturally limited and LOs are fixed-time events, however each observation requires only a few minutes including overheads. As such, LOs are ideally suited to fill small gaps of idle time between standard observations. } ", "introduction": "The method of observing lunar occultations (LO) of background stars to derive their angular diameter, as well as other characteristics such as binarity, has been employed for decades and provides angular resolution at the milliarcsecond (mas) level independently of the diameter of the telescope used. This surpasses the diffraction limit of even the largest single telescopes and rivals the resolution of long-baseline interferometry (LBI) even with baselines of hundreds of meters. Basically, the LO technique relies on the lunar limb as a diffracting edge, rather than on the diameter of the telescope. The diffraction fringes are generated in space, and due to their relatively fast motion over the telescope they must be sampled at rates of order 0.1-1\\,kHz. Combined together, these factors greatly diminish the adverse effects of atmospheric turbulence, which are the main limit of other high angular resolution techniques. LO also permit not only a model-dependent derivation of source parameters such as the angular diameter and binary parameters, but also a model-independent reconstruction of the brightness profile of the source according to maximum-likelihood, or even a unique reconstruction by light curve inversion in special cases. The time required for observation (dominated by overheads, since an occultation lasts much less than 1\\,s) and for data analysis is significantly shorter than for most other high angular resolution methods. Of course, LO suffer from a number of significant drawbacks, first among them that the sources cannot be chosen at will. The Moon covers only about 10\\% of the celestial sphere in its apparent orbit. LO are fixed time events, and as such subject to weather and instrumental downtimes. Finally, the lunar limb only provides a 1-D scan of the source, although observations from different sites, or at different epochs, can in principle be combined under favorable circumstances. Due mainly to the chromatic properties of the scattered light background around the Moon and of the diffraction fringes, the near-IR is ideally suited to observe LO. More details on the method, its performance, the data analysis and the results can be found in Richichi (\\cite{AR_thesis}, \\cite{richichi96}) and Fors (\\cite{thesis}). The CHARM2 catalogue alone (Richichi et al. \\cite{charm2}) lists 1815 LO results in the field of high angular resolution. While the angular resolution of LO is set mainly by the lunar limb rather than by telescope size, this latter parameter obviously has a crucial role in determining the limiting sensitivity. Richichi (\\cite{1997IAUS..158...71R}) studied the performance of LO under a number of circumstances, including the then-futuristic use of IR array detectors on a 8-m class telescope. This has recently become reality: Richichi et al. (\\cite{Messenger2006}) and Fors et al. (\\cite{SEA2007}, \\cite{fors08}) reported in a preliminary form on the use of the ISAAC instrument in the so-called burst mode at the ESO VLT 8.2\\,m Antu telescope. In the present paper we provide a detailed account of the observations carried out during a few hours in the night of March 21, 2006, when 51 LO events were recorded. A second batch of observations was carried out in August 2006: due to their large number and their different nature, these sources will be discussed in a separate paper. In Sect.~\\ref{data} we describe the observations and we provide details of the data reduction. In Sect.~\\ref{results} we discuss the results, which include new binaries, resolved stars and the near-IR counterparts of two radio maser sources. We also characterize the performance of this specialized observing mode, which combines the highest angular resolution and sensitivity possible today in the near-IR. In the conclusions we describe the integration of the LO technique in the service mode operations scheme in place at the Paranal observatory and our plans for routine LO observations at that site. ", "conclusions": "We have presented the first LO results obtained using the so-called burst mode of the ISAAC instrument at the ESO Antu VLT telescope. This mode permits to record data streams on a small subarray at high temporal frequency (3.2\\,ms, although mostly 6.4\\,ms were used for the present work). We recorded 51 occultation events over about four hours during a close approach of the Moon to the Galactic Center. The events were reappearances from behind the dark limb. This, coupled with initial problems of focussing and image quality due mostly to the commissioning nature of the run, resulted in several sources being missed or without sufficient quality. Nevertheless, 30 of the events resulted in light curves of good, or even excellent quality, including some of the best SNR traces ever recorded with this technique. We have revealed two binaries, three stars with a marginally resolved angular diameter of the order of 2\\,mas, and two resolved masing AGB stars which appear to be in the foreground of the galactic bulge. For one of these, {\\object 2MASS 17453224-2833429}, we were able to recover the brightness profile of the extended emission. This has the typical signature of an optically thin shell, with inner radius of order 10-15\\,mas, corresponding to about 40\\,AU. Follow-up studies are required to obtain further information on these stars, given that they are all heavily reddened and most have no known counterpart or literature entries. Our analysis, including the unresolved sources, has established an unprecedented performance of the LO technique in the near-IR using a very large telescope and a fast read-out mode. We find that the angular resolution varies between 4 and 1\\,mas, and even less in the case of very high SNR. Data quality is generally impressive, compared to smaller telescopes, mostly due to the scintillation reduction effect of the large mirror. The dynamic range is also significantly improved with respect to previous observations, and we have shown that in some cases it would have been possible to detect a companion fainter than the primary by about 8 magnitudes in broad K band on angular scales comparable to the Airy disk of the telescope. This result compares favorably with other AO-assisted high contrast imaging techniques. The random selection of sources to be occulted and the fixed-time nature of the events remain the main downsides of the LO technique. However, thanks to the very short telescope time required for each observation, LO are an ideal filler program for times when the telescope is idle or other programs cannot be conveniently executed. The service mode observations implemented at the VLT provide the right context for this, and we have obtained approval for such a filler program during ESO observing periods 80 and 81 (from October 2007 through September 2008). Several hundreds of OBs have been prepared for each period, and are in the queue awaiting execution whenever five minutes of telescope time become available." }, "0807/0807.1703_arXiv.txt": { "abstract": "{MHD turbulence is known to exist in shearing boxes with either zero or nonzero net magnetic flux. However, the way turbulence survives in the zero-net-flux case is not explained by linear theory and appears as a purely numerical result that is not well understood. This type of turbulence is also related to the possibility of having a dynamo action in accretion discs, which may help to generate the large-scale magnetic field required by ejection processes.} {We look for a nonlinear mechanism able to explain the persistence of MHD turbulence in shearing boxes with zero net magnetic flux, and potentially leading to large-scale dynamo action.} {Spectral nonlinear simulations of the magnetorotational instability are shown to exhibit a large-scale axisymmetric magnetic field, maintained for a few orbits. The generation process of this field is investigated using the results of the simulations and an inhomogeneous linear approach. We show that quasilinear nonaxisymmetric waves may provide a positive back-reaction on the large-scale field when a weak inhomogeneous azimuthal field is present, explaining the behaviour of the simulations. We finally reproduce the dynamo cycles using a simple closure model summarising our linear results.} {The mechanism by which turbulence is sustained in zero-net-flux shearing boxes is shown to be related to the existence of a large-scale azimuthal field, surviving for several orbits. In particular, it is shown that MHD turbulence in shearing boxes can be seen as a dynamo process coupled to a magnetorotational-type instability.} {} ", "introduction": "The problem of angular momentum transport is a central issue of accretion disc theory. Following \\cite{SS73}, angular momentum transport is often modelled assuming the disc is turbulent, using a kind of turbulent viscosity (the so-called $\\alpha$ model). However, the way discs may become turbulent is still a highly debated subject. A possible route to turbulence is the magnetorotational instability \\citep[MRI,][]{BH91a} which appears in discs sufficiently ionised to be coupled with the magnetic field lines \\citep{G96}. This instability has been extensively studied in its nonlinear regime using local \\citep{HGB95,SHGB96} and global \\citep{H00} numerical simulations. More recently, the effect of non-ideal MHD has been investigated numerically in shearing boxes with either zero \\citep{FPLH07} or nonzero \\citep{FSH00,LL07} net magnetic flux, showing a strong dependence on the magnetic Prandtl number $Pm$. These new results bring back the question of the efficiency of MHD turbulence in accretion discs since $Pm$ is believed to vary by several orders of magnitude in real objects \\citep{BH08}. In particular, as turbulence seems to disappear in zero-net-flux simulations for $Pm\\le 1$, the existence of a turbulent flow in low-$Pm$ objects such as protoplanetary discs is questionable. However, as pointed out by \\cite{FPLH07}, today simulations can reach Reynolds numbers that are very small (up to $10^4$) compared to those of real discs ($\\sim 10^{10}$). Therefore, no clear conclusion for accretion discs can be drawn from these numerical results. It has been pointed out by \\cite{PCP07} and \\cite{FP07} that numerical resolution can also play an important role in simulations. In particular, increasing resolution in zero-net-flux simulations \\emph{without} any physical dissipation leads to a weaker turbulence and transport. One might conclude from these results that turbulence should disappear in real astrophysical systems \\citep{PCP07}. We think however that this conclusion is questionable since the ideal MHD model does not hold in turbulent flows, mainly because a dissipation scale necessarily exists, at which non-ideal effects are \\emph{not} negligible. Of course, ideal MHD simulations must include some sort of numerical dissipation in their algorithm, which then implicitly defines a dissipation scale of the order of the numerical grid scale. However, this algorithm-dependent dissipation is quite different from physical dissipation processes \\citep{LB07}, leading potentially to numerical artifacts for large-scale properties, such as turbulent transport. On the other hand, the way MHD turbulence is sustained in these simulations is not well understood. It is known that when a mean vertical field is applied, the magnetorotational instability destabilises the flow and leads to developed three-dimensional turbulence \\citep{BH91a,HGB95}. This picture is not directly applicable to zero-net-flux simulations since the magnetic field can be dissipated either by a finite resistivity, or by the turbulence itself (see \\cite{HGB96} for an extensive discussion of this point). Therefore, even if one assumes that the MRI appears locally because of a given magnetic field configuration, one has to regenerate this field with some sort of dynamo mechanism. The whole process sustaining turbulence in zero-net-flux simulations is therefore a \\emph{nonlinear} mechanism, potentially involving a kind of magnetorotational instability at some stage. As already mentioned by \\cite{BH92}, this means that this disc dynamo is intrinsically nonlinear, as it requires a non-negligible Lorentz force, and cannot be described as a kinematic dynamo. Following these ideas, \\cite{ROP07} obtained a steady nonlinear solution to the MHD equations in the zero-net-flux case, using rigid conducting walls as radial boundary conditions. This solution is clearly a first step toward the understanding of the mechanism working in shearing boxes, although the boundary conditions and Reynolds numbers are significantly different compared to turbulent simulations. Note also that \\cite{BNST95} studied this dynamo process in discs, using boundary conditions allowing for mean flux variations. Although a azimuthal field was generated in their simulations, no physical understanding of the underlying process was provided. In this paper, we describe a possible mechanism able to sustain MHD turbulence in the shearing box with zero net flux. First, we recall the resistive MHD equations in the shearing box, and the numerical method used to solve them. Next, we investigate the temporal evolution of some zero-net-flux simulations, and we exhibit a long-timescale cycle for the large-scale magnetic field. The origin of this cycle is described using a spectral decomposition, and we demonstrate that its origin is related to some properties of the nonaxisymmetric structures of the flow. We then study a linear theory of nonaxisymmetric waves in the presence of a large-scale magnetic field similar to the one observed in simulations. We show that this linear theory predicts the right properties for the nonaxisymmetric structures and partially explains the nonlinear cycle. We then provide a phenomenological closure model for the evolution of the large-scale field, based on our previous findings. We show that this model reproduces the basic behaviour of the cycles, and provides a mechanism able to sustain turbulence in dissipative MHD shearing boxes. The existence of a long-timescale cycle and large-scale structures in these simulations is the most significant finding of our investigation, mainly because it shows that MHD turbulence is able to generate large scales, independent of the dissipative scales. This mechanism can also be seen as a way to generate large-scale magnetic fields, providing a dynamo mechanism specific to accretion disc turbulence. These findings are discussed in detail in the concluding section, along with the implications for real astrophysical systems. \\begin{figure*} \\centering \\includegraphics[width=1.0\\linewidth]{0152fg01.eps} \\caption{Time-evolution of $B_x$, averaged in the $x$ and $y$ directions. Notice the long-lived ($T\\sim 50\\, S^{-1}$) vertical structures, mostly found with long vertical wavelengths.} \\label{bxtm}% \\end{figure*} ", "conclusions": "\\subsection{Summary} In this paper, we have investigated the behaviour of the large-scale magnetic field in zero-net-flux simulations of the magnetorotational instability. We have first shown that the large-scale azimuthal field $B_x(z)$ is subject to a long-timescale oscillation when the flow is turbulent. Studying the induction equation, we have seen that these cycles are primarily due to an oscillation of a large-scale radial magnetic field $B_y$ and an azimuthal EMF $\\mathcal{E}_x$, whereas the radial EMF $\\mathcal{E}_y$ acts like a turbulent resistivity on $B_x$. Studying the nonaxisymmetric modes of the system, we found that the axisymmetric EMF is due mostly to the largest nonaxisymmetric structures, showing that the cycles are essentially a large-scale process. To understand the generation of the EMF, we have investigated the behaviour of nonaxisymmetric linear waves in the presence of an inhomogeneous and constant in time azimuthal magnetic field $B_x(z)$. This linear analysis explains most of the properties of the EMF observed in the nonlinear simulations. In particular, we have shown that the large-scale axisymmetric $\\mathcal{E}_x$ can be reversed for large $B_x$, explaining the cycles in the simulations. To summarise these results, we have considered a simplified closure model, encapsulating the linear properties. This model is able to reproduce the main behaviours of the cycles despite its simplicity, showing that the physics involved in the cycles is well described by our linear analysis. \\begin{figure}[h!] \\centering \\includegraphics[width=1.0\\linewidth]{0152fg38.eps} \\caption{Scheme of the mechanism responsible for the large-scale magnetic field cycles. The dashed lines represent the nonlinear interactions which are not discussed in the present work. These interactions are nonetheless required to obtain a self-sustained mechanism (see text).} \\label{mechanism}% \\end{figure} To summarise our findings, we sketch the mechanism responsible for the cycles in Fig.~\\ref{mechanism}. In this sketch, we have assumed that the nonaxisymmetric excitation (seed) is related to the nonlinear coupling of amplified shearing waves (dashed line). This specific process has not been studied in the present work and is just given here as a way to ``close the loop''. Note however that a more complicated mechanism may be involved in this back-reaction without changing the global picture. Moreover, we have described the contribution of $\\mathcal{E}_x(z)$ as a generic ``dynamo'' effect, but one should remember that this term may be either correlated or anticorrelated with $B_x$, as a function of the strength of $B_x$. Finally, we have omitted viscous and resistive effects for simplicity, as they have a small impact on the mechanism itself. \\subsection{Comparison with previous works} We would like to stress that the model presented here does not constitute a full description of a sustaining mechanism for MHD turbulence in discs. Indeed, we have assumed in the linear analysis and in the closure model that the flow is able to generate continuously small-amplitude shearing waves. As mentioned previously, this process involves a nonlinear coupling of nonaxisymmetric waves, which is beyond the scope of this work. However, a fully consistent self-sustaining process would require a description of this effect. According to \\cite{FPLH07}, MHD turbulence appears in zero-net-flux shearing-box simulations only when $Pm>1$. However, most of our analysis does not depend on the magnetic Prandtl number, and the dynamo process described here may work for arbitrarily low $Pm$. This suggests that the shearing-wave excitation mechanism depends on the Prandtl number, and may be too weak in the $Pm<1$ case. This confirms that the excitation should be related to the small turbulent scales, which are still poorly understood. Interestingly, our results indicate that the large-scale field generation mechanism is not destroyed for large Reynolds numbers, provided that the background flow is turbulent. Therefore, the same kind of mechanism might be at work in real astrophysical discs, generating a large-scale field in a sufficiently ionised and turbulent plasma. One should note however that the configuration used in our simulations is not comparable to the real geometry of an accretion disc. In particular, we do not know how the aspect ratio will modify the mechanism of the cycles, nor how the vertical stratification may enter this picture. However, as discovered by \\cite{BNST95} in the case of a stratified flow with nonperiodic vertical boundary conditions, one observes azimuthal field cycles with a long timescale. These cycles have a significantly longer period ($T\\sim 200\\, S^{-1}$) compared to ours and involve a modification of the net azimuthal magnetic flux. Moreover, it has been shown by \\cite{BD97} that in this case, the cyclic behaviour may be explained by a classical $\\alpha$ effect proportional to the vorticity\\footnote{Note that this $\\alpha$ effect might be related to the presence of vertical stratification and the breaking of the reflectional symmetry}. Because of these fundamental differences, we cannot conclude that the mechanism responsible for all these cycles is the same without further investigation. Nevertheless, it seems that a more systematic investigation of this kind of long-timescale behaviour is required in order to better understand the way MHD turbulence sustains magnetic fields in various configurations. As mentioned in the introduction, \\cite{PCP07} suggested that turbulent transport should vanish for large enough resolution and Reynolds number. Therefore, the turbulent transport associated with this mechanism is another important issue. As one can check easily in the numerical simulation and in the toy model, the large-scale $B_x$ and $B_y$ are roughly phase-shifted by $\\pi/2$, leading to a very small Maxwell stress ($\\langle B_x B_y\\rangle\\sim 10^{-4}$). Therefore, in our picture, most of the transport must come from the shearing waves, which have been shown to have nonzero Maxwell and Reynolds stresses. Unfortunately, we cannot put a precise value on the transport due to the shearing waves since it is controlled by their initial excitation. Therefore, this mechanism cannot give a precise answer on the expected transport in a real accretion disc. From a more phenomenological point of view, we expect the large-scale magnetic field strength produced by our mechanism not to depend on the Reynolds numbers if they are sufficiently large (see section \\ref{numlin}). Since this large-scale field couples with all the other modes available, we expect the large-scale nonaxisymmetric structures to have roughly the same amplitude as the large-scale field, in the limit of a fully developed turbulence. If this picture is correct, we would then expect a minimum transport of the order of a few times $10^{-3}$, independently of the Reynolds numbers. Note however that this conclusion relies on the assumption that the large-scale field does not depend on the dissipation coefficients. We have shown that this assumption is plausible, although a precise analytical description of our linear analysis would be required to ascertain this statement. The present description of an accretion disc dynamo is related to the steady self-sustaining solutions obtained by \\cite{ROP07} by numerical continuation methods. Both mechanisms involve a large-scale azimuthal magnetic field with zero net flux, which is generated through the shearing of a radial field. Both also involve a nonaxisymmetric magnetorotational-type instability that regenerates the radial field through its nonlinear feedback. However, the present mechanism is intrinsically local and unsteady, and may work for arbitrarily large Reynolds numbers whereas the steady, highly symmetrical solutions of \\cite{ROP07} depend on the existence of walls and are apparently restricted to low Reynolds numbers. Nevertheless, one may also think that these steady solutions correspond to some fixed points in phase space, around which the shearing box solutions oscillate, creating the dynamo cycles we have described. If this picture is correct, it would be an elegant way to unify these different approaches, and possibly to find other mechanisms of the same kind. This work might also be compared to the shear dynamo simulations of \\cite{YH08}. However, we emphasise that our analysis involves a nonlinear dynamo, whereas the shear dynamo described by \\cite{YH08} is a kinematic (linear) dynamo, in which turbulence is \\emph{forced}. Moreover, the shear dynamo occurs in a non-rotating system in which the MRI does not appear. Applying our linear analysis to such a system does not produce a $\\gamma$ effect that could explain the shear dynamo in either linear or nonlinear regimes. Finally, we would like to stress that the problem of a non-linear dynamo, involving several linear instabilities, has also been described by \\cite{CBC03} in the context of magnetically buoyant flows. Although the underlying instabilities are somewhat different, they also found a cyclic behaviour which, as in the present work, cannot be fully described by a classical $\\alpha$ effect. This may indicate that all these instabilities share some common properties yet to be discovered, allowing for the development of a nonlinear dynamo. \\subsection{Future work} As discussed previously, the main issue raised by our findings is how the turbulence is able to excite shearing waves. Interestingly, the same kind of problem arises in the case of the shearing box with a mean azimuthal field \\citep{BH92}. Therefore, a simpler way to study this effect is to investigate the way turbulence is sustained in the presence of a mean azimuthal field. In particular, an interesting test would be to check that in this case, the presence of turbulence also depends on $Pm$, in a similar way as in \\cite{FPLH07}. This would be a good indication that the large-scale dynamo process \\emph{itself} does not depend on $Pm$ at large $Re$, even though the underlying turbulent regeneration mechanism process does. Moreover, this idea would provide a new way to understand the $Pm$-dependence observed in MRI simulations with a nonzero mean vertical field \\citep{LL07}, and potentially to describe the asymptotic transport in the limit of high and low $Pm$. On the other hand, the dynamo process itself requires further investigation. The linear analysis provided here has been computed using essentially numerical tools, as the analytical description of this problem is somewhat technical. This analytical description is nevertheless required, as it will give a precise and formal constraint on the $\\gamma$ effect and its dependence on the azimuthal field strength. It may also be a way to have a more physical understanding of the underlying mechanism responsible for the resistive and $\\gamma$ effects, leading to a potential generalisation of this mechanism to a wider class of flow. Finally, we would be able to check, at least linearly, if this dynamo process depends on the non-ideal effects, and potentially to confirm our conjecture." }, "0807/0807.1535_arXiv.txt": { "abstract": "We present new multi-wavelength observations of the dwarf Seyfert 1 galaxy POX~52 in order to investigate the properties of the host galaxy and the active nucleus, and to examine the mass of its black hole, previously estimated to be $\\sim 10^{5}$ \\msun. {\\it Hubble Space Telescope} ACS/HRC images show that the host galaxy has a dwarf elliptical morphology ($M_I= -18.4$ mag, \\sersic\\ index $n=4.3$) with no detected disk component or spiral structure, confirming previous results from ground-based imaging. X-ray observations from both \\chandra\\ and \\xmmn\\ show strong (factor of 2) variability over timescales as short as 500 s, as well as a dramatic decrease in the absorbing column density over a 9 month period. We attribute this change to a partial covering absorber, with a 94\\% covering fraction and $N_{H} = 58^{+8.4}_{-9.2} \\times 10^{21}$ cm$^{-2}$, that moved out of the line of sight in between the \\xmmn\\ and \\chandra\\ observations. Combining these data with observations from the VLA, {\\it Spitzer}, and archival data from 2MASS and {\\it GALEX}, we examine the spectral energy distribution (SED) of the active nucleus. Its shape is broadly similar to typical radio-quiet quasar SEDs, despite the very low bolometric luminosity of $\\lbol\\ = 1.3 \\times 10^{43}$ \\ergs. Finally, we compare black hole mass estimators including methods based on X-ray variability, and optical scaling relations using the broad \\hbeta\\ line width and AGN continuum luminosity, finding a range of black hole mass from all methods to be $\\mbh\\ = (2.2-4.2) \\times\\ 10^{5}$ \\msun, with an Eddington ratio of $\\lbol/\\ledd \\approx 0.2-0.5$. ", "introduction": "\\begin{figure*} \\begin{center} \\scalebox{0.25}{\\includegraphics{f1.eps}} \\end{center} \\caption{\\hst\\ ACS/HRC images of POX 52 and GALFIT models. The top and bottom rows show the F814W and F435W images, respectively. In each row, the left panel is the ACS image, the middle panel is the full GALFIT model, and the right panel shows the fit residuals. For the F435W filter, the double-\\sersic\\ host galaxy model is shown.} \\label{image-panel} \\end{figure*} Stellar and gas dynamical studies have proven to be the most secure method of measuring the masses of supermassive black holes in nearby galaxies, and black holes with masses of $\\sim10^6~$--$~10^9$ \\msun\\ have now been detected in a few dozen galaxies (see Ferrarese \\& Ford 2005 for a recent review). With just a few exceptions, however, stellar-dynamical and gas-dynamical techniques for black hole mass measurement cannot be applied to Seyfert galaxies and quasars, because most luminous active galaxies are too distant for their black hole's gravitational sphere of influence to be resolved. Most black hole mass estimates in broad-lined active galactic nuclei (AGNs) are based on indirect methods that rely on scaling relations resulting from reverberation-mapping data \\citep{Kaspi, bentz06}. Using these methods, it has recently become possible to search for active galaxies having extremely low-mass black holes, in order to extend our understanding of black hole demographics to objects having black hole masses of $\\mbh \\lesssim 10^6$ \\msun. The best nearby example of a black hole in the $<10^6$ \\msun\\ mass range is that in the dwarf Seyfert 1 galaxy NGC~4395 \\citep{FilS89,fh03}. NGC~4395 is an essentially bulgeless, late-type dwarf galaxy with type 1 Seyfert characteristics such as broad emission lines \\citep{FilS89, FHS93} and rapid X-ray variability \\citep{Iwa00, Shih03, Moran05}. Located at a distance of only $\\sim4.3$~Mpc \\citep{thim04}, it is an excellent target for multiwavelength observations despite the tiny luminosity of its central engine ($\\lbol \\approx 6\\times10^{40}$ \\ergs; Moran \\etal\\ 1999). A recent ultraviolet (UV) reverberation-mapping measurement found \\mbh~$=~(3.6~\\pm~1.1)~\\times~10^5$~\\msun\\ \\citep{Pet05}. \\citet{GH04, GH07c} searched the Sloan Digital Sky Survey (SDSS) archives to find AGNs with low-mass black holes and have found numerous examples of Seyfert 1 galaxies with $\\mbh < 10^6$ \\msun, which we will refer to as the GH sample. However, most of the SDSS objects found by Greene \\& Ho are much more distant than NGC 4395, making detailed studies of their nuclei and host galaxies more difficult. Although the GH sample does contain examples of AGNs in low-mass disk galaxies \\citep[see also][]{Dong}, NGC 4395 remains an exceptional case due to its extreme late-type morphology as well as its proximity. \\object{POX 52}, also known as G1200-2038 or PGC 038055 ($D~=~93$ Mpc for $H_0= 70$ \\kms\\ Mpc\\per), is an interesting counterpart to NGC~4395. It was originally discovered in an objective prism study by \\citet{KSB87}, who classified it as a dwarf disk galaxy with a Seyfert 2 nucleus, but noted the presence of a weak broad component of the \\hbeta\\ line. \\citet{bar04} re-observed POX 52, classifying it as a type 1 AGN based on clear detections of broad \\hal\\ and \\hbeta\\ and noting that its optical emission-line spectrum is nearly identical to that of NGC 4395. Both NGC 4395 and POX 52 can be classified as narrow-line Seyfert 1 (NLS1) galaxies, since they satisfy the defining criterion of having FWHM(\\hbeta)$<2000$ \\kms. NLS1 galaxies are often, but not always, characterized by strong \\ion{Fe}{2} emission and relatively weak [\\ion{O}{3}] emission when compared to \\hbeta\\, along with steep soft ($0.5 - 2.0$ keV) spectral slopes and rapid variability in the X-ray \\citep{OP85, Boller96}. NGC~4395 and POX~52 are therefore somewhat atypical NLS1 galaxies in that they have very high-excitation spectra with large [\\ion{O}{3}]/\\hbeta\\ ratios and very weak \\ion{Fe}{2} emission. Using the virial scaling relation calibrated by \\citet{Kaspi}, Barth \\etal\\ estimated a black hole mass of $\\mbh\\approx1.6\\times10^{5}$ \\msun, making this one of the least massive black holes identified in any AGN. This mass is consistent with the mass expected based on the \\msigma\\ relation (extrapolated to small \\sigmastar) for the measured stellar velocity dispersion $\\sigmastar = 36 \\pm 5$ \\kms. They also estimated the bolometric luminosity (\\lbol\\ $\\approx 2 \\times 10^{43}$ \\ergs) and the Eddington ratio (\\lbol/\\ledd\\ $\\approx 0.5 - 1$) for POX~52. Thus, although POX~52 and NGC~4395 have very similar black hole masses and emission-line spectra that are nearly identical in appearance, the AGN in POX~52 is $\\sim 300$ times more luminous than that in NGC~4395, which has an estimated \\lbol/\\ledd\\ $\\approx 10^{-3}$ \\citep{Pet05}. \\citet{bar04} used ground-based images to study the morphology of POX~52 and found that the host galaxy was best fit with a \\sersic\\ model with an index of $n = 3.6 \\pm 0.2$, effective radius of $r_{e} = 1\\farcs{2}$, and absolute magnitude $M_{V} = -17.6$. Combined with \\sigmastar\\ measured from the spectrum, these results show POX~52 to fall near the dwarf elliptical sequence \\citep{Geha03} in the fundamental plane projections of \\citet{Burstein97}, making it the first example of a dwarf elliptical to host a Seyfert nucleus. Although the past few years have seen a large increase in the number of known AGNs with low-mass black holes, POX~52 remains a valuable target for further study, because of its relatively small distance and the fact that it is a less luminous and presumably less massive galaxy than the majority of the host galaxies in the GH SDSS sample. Among nearby dwarf ellipticals, there is little evidence for the presence of central black holes. The closest well-studied example is NGC 205 in the Local Group, and stellar-dynamical observations with the \\emph{Hubble Space Telescope (HST)} have set an upper limit to the black hole mass of $\\mbh < 3.8\\times10^4$ \\msun\\ \\citep{Val05}, which is below the mass expected from a simple extrapolation of the \\msigma\\ relation. For dwarf elliptical galaxies outside the Local Group, stellar-dynamical observations lack the spatial resolution needed to detect black holes, and AGN surveys are the best available alternative for black hole searches. A double nucleus in one Virgo dwarf elliptical galaxy has been interpreted as morphological evidence for a stellar disk in orbit around a central black hole \\citep{deb06}, but aside from NGC~205 there are essentially no direct observational constraints on the black hole occupation fraction in dwarf elliptical galaxies. POX~52 represents a unique opportunity to study, in depth, the unambiguous presence of a low-mass central black hole in a nearby dwarf elliptical galaxy. Here, we present new multiwavelength observations of POX 52, including the first targeted \\hst\\ and X-ray observations of this unusual object. \\hst\\ images are used to examine the host galaxy morphology; our results confirm the previous conclusions that the galaxy is consistent with a dwarf elliptical morphology with a \\sersic\\ index of $\\sim 4$. We describe X-ray observations obtained with both the \\chandra\\ and \\xmmn\\ X-ray Observatories in order to study the temporal and spectral properties of the active nucleus, and to constrain its bolometric luminosity and black hole mass. We also include observations from the Very Large Array (VLA) and {\\it Spitzer} Infrared Spectrometer (IRS), as well as archival data from 2MASS and \\emph{GALEX}, in order to investigate the spectral energy distribution (SED) of POX 52 over a wide range in frequency. We use recently updated versions of the broad-line virial scaling relations to obtain new estimates of the mass of the black hole, and in combination with the multiwavelength SED we determine the bolometric luminosity and Eddington ratio for the AGN. In addition to the \\hst\\ data presented here, we were also awarded time as part of the same project to obtain ultraviolet and optical spectra of POX 52 with the Space Telescope Imaging Spectrograph (STIS), but STIS failed before the observations were scheduled to be taken. ", "conclusions": "New \\hst\\ observations have confirmed POX 52 to be a dwarf elliptical galaxy, with a \\sersic\\ index of $n \\approx 4.3$, $r_e \\approx 1.15$, and $M_I = -17.3$. These parameters, along with the previously measured stellar velocity dispersion from \\citet{bar04}, show POX 52 to reside near the dwarf elliptical family in the fundamental plane. POX 52 was the first AGN found to reside in a dwarf elliptical galaxy, but more recent studies of low-mass AGNs and their host galaxies have found a number of similar objects \\citep{GH08}. We see strong variability over $500 - 10^{4}$ s timescales with both \\chandra\\ and \\xmmn\\ and a substantial change in the absorbing column density due to partial covering over a 9 month time period. Data from the Optical Monitor on \\xmmn\\ showed no significant variability in the UV over short $\\sim 10^3 - 10^5$ s timescales. We see no variability within the $2\\sigma$ errors in the 7 month period between the OM and {\\it GALEX} NUV observations, but this does allow variations of $10-35$\\% on the $1\\sigma$ uncertainty level. The {\\it Spitzer} spectrum shows evidence of possible star-formation due to the PAH features and red continuum, but that analysis is deferred to a later paper. Including data from the VLA, {\\it Spitzer}, \\emph{GALEX}, and 2MASS, we compile an extensive SED that shows POX 52 to be consistent with the radio-quiet template of more massive quasars and estimate a bolometric luminosity of \\lbol\\ $= 1.3 \\times 10^{43}$~\\ergs\\ by incorporating all of the available data. Using a variety of mass estimators, we find a range of black hole mass \\mbh~$\\approx (1.8-4.2) \\times 10^{5}$ \\msun. We find that we are unable to determine the break frequency from the \\chandra\\ PSD, either due to a low signal-to-noise ratio or insufficient temporal sampling. If $\\nu_B \\approx 10^{-4}$ Hz, then our mass estimate from the X-ray excess variance of $\\mbh = (3.9 \\pm 1.7) \\times 10^5$ \\msun\\ is consistent with the optical mass estimates. Otherwise, new X-ray observations of POX 52 in an unabsorbed state with the larger effective area of \\xmmn\\ would be needed in order to determine $\\nu_B$ and estimate the black hole mass with this method. We have shown POX 52 to have an SED consistent with a scaled-down version of more massive type 1 quasars and that it is in a different accretion state than NGC 4395, as seen in the two orders of magnitude disparity between Eddington ratios. The number of low-mass AGNs known is increasing with new results from large AGN surveys, such as \\citet{GH04, GH07c}, but POX 52 remains unique in that it is one of the nearest and best studied of the low-mass AGNs." }, "0807/0807.2911_arXiv.txt": { "abstract": "We extend the Hubble diagram up to $z = 5.6$ using 63 gamma-ray bursts (GRBs) via peak energy-peak luminosity relation (so called Yonetoku relation), and obtain constraints on cosmological parameters including dynamical dark energy parametrized by $P/\\rho\\equiv w(z) = w_0 + w_a \\cdot z/(1+z)$. It is found that the current GRB data are consistent with the concordance model, ($\\Omega_m = 0.28, \\Omega_{\\Lambda} = 0.72, w_0 = -1, w_a = 0$), within two sigma level. Although constraints from GRBs themselves are not so strong, they can improve the conventional constraints from SNeIa because GRBs have much higher redshifts. Further we estimate the constraints on the dark-energy parameters expected by future observations with GLAST (Gamma-ray Large Area Space Telescope) and \\swift by Monte-Carlo simulation. Constraints would improve substantially with another 150 GRBs. ", "introduction": "\\label{sec:intro} In our previous paper \\citep{Kodama2008}, we calibrated the peak energy-peak luminosity relation of GRBs with 33 nearby events ($z < 1.62$) whose luminosity distances were estimated from those of large amount of SNeIa \\citep{Riess2007,Wood-Vasey2007,Davis2007}. This calibrated Yonetoku relation, derived without assuming any cosmological models, can be used as a new cosmic distance ladder toward higher redshifts. Then we determined the luminosity distances of 30~GRBs in $1.8 < z < 5.6$ using the calibrated relation and calculated the likelihood varying $(\\Omega_m, \\Omega_{\\Lambda})$. We obtained $(\\Omega_m, \\Omega_{\\Lambda}) = (0.37^{+0.14}_{-0.11}, 0.63^{+0.11}_{-0.14})$ for a flat universe, which is consistent with the concordance cosmological model within one sigma level. Our logic to obtain a new distance ladder is similar to that for SNeIa, that is, we calibrate a new distance indicator (Yonetoku relation) at low redshifts ($z < 1.62$) using the well established indicators (SNeIa). Then we assume the new relation holds at high redshifts ($z > 1.62$), although more detailed analysis is needed for possible selection bias and evolution effects in the relation \\citep{Oguri2006}. Currently the number of GRBs with $z > 1.62$ is relatively small ($\\sim 30$) and the statistical and systematic errors of the Yonetoku relation are not so small compared with SNeIa. Even then, GRBs are still effective to probe dark energy, especially for dynamical dark energy whose energy density becomes large at high redshifts, because the mean redshift of GRBs is higher than that of SNeIa \\citep{Amati2008,Liang2008,Basilakos2008, Schaefer2007,Ghirlanda2006,Firmani2006a}. There would be many ways to characterize the time variation of dark energy. Here we adopt a simple phenomenological model as \\citep{Chevallier2001,Linder2003} $P/\\rho \\equiv w(z) = w_0 + w_a \\cdot z/(1+z) = w_0 + w_a (1-a)$, where $w_0$ and $w_a$ are constants and $a$ is the scale factor of the universe. For this model, GRBs would give strong constraints on $w_a$ which represents the time dependence of dark energy. In this Letter, we present constraints on cosmological parameters such as $w_0$ and $w_a$ using both SNeIa with $z < 1.8$ and GRBs with $z < 5.6$. In \\S~2, we briefly review the main results of the previous paper \\citep{Kodama2008} and the theoretical and observational basis of the analysis. In \\S~3-1, we assume cosmological constant ($w = -1$) and obtain constraints on ($\\Omega_m, \\Omega_{\\Lambda}$). In \\S~3-2, we assume a flat universe and non-dynamical ($w_a = 0$) dark energy and obtain constraints on $(w_0, \\Omega_m)$. In \\S~3-3, we fix $\\Omega_m = 0.28$ for simplicity, and obtain the plausible values of $w_0$ and $w_a$. In \\S~4, we discuss how these constraints will be improved in future observations of high redshift GRBs by such as GLAST. Throughout the paper, we fix the current Hubble parameter as $H_0 = 66~{\\rm km~s^{-1}Mpc^{-1}}$. ", "conclusions": "" }, "0807/0807.2250_arXiv.txt": { "abstract": "Inelastic dark matter, in which WIMP-nucleus scatterings occur through a transition to an excited WIMP state $\\sim$ 100 keV above the ground state, provides a compelling explanation of the DAMA annual modulation signal. We demonstrate that the relative sensitivities of various dark matter direct detection experiments are modified such that the DAMA annual modulation signal can be reconciled with the absence of a reported signal at CDMS-Soudan, XENON10, ZEPLIN, CRESST, and KIMS for inelastic WIMPs with masses $O(100 \\;\\gev)$. We review the status of these experiments, and make predictions for upcoming ones. In particular, we note that inelastic dark matter leads to highly suppressed signals at low energy, with most events typically occurring between 20 to 45 keV (unquenched) at xenon and iodine experiments, and generally no events at low ($\\sim 10$~keV) energies. Suppressing the background in this high energy region is essential to testing this scenario. The recent CRESST data suggest seven observed tungsten events, which is consistent with expectations from this model. If the tungsten signal persists at future CRESST runs, it would provide compelling evidence for inelastic dark matter, while its absence should exclude it. ", "introduction": "The DAMA collaboration has observed an annual modulation in the nuclear recoil rates of their experiment with a confidence level that now exceeds $8\\sigma$ \\cite{Bernabei:2008yi}. This observation is consistent with weakly interacting massive particles (WIMPs) in the halo striking the target nuclei slightly less and more often as the Earth moves with and against the WIMP wind. The WIMP interpretation has been sharply criticized due to its apparent inconsistency with the null results of several other direct detection experiments, including CDMS and XENON. In this paper, we demonstrate that inelastic dark matter \\cite{TuckerSmith:2001hy} continues to offer a compelling explanation of both the annual modulation signal as well as the lack of a reported signal at CDMS, XENON, ZEPLIN, KIMS, CRESST and other experiments. Inelastic dark matter (iDM) \\cite{TuckerSmith:2001hy} was originally proposed to reconcile the DAMA annual modulation observation and null results from CDMS. The basic model is a simple extension of the standard WIMP model. An inelastic WIMP has two basic properties: \\begin{itemize} \\item{In addition to the dark matter particle $\\chi$, there exists an excited state $\\chi^*$, with a mass $m_{\\chi^*}-m_{\\chi} = \\delta \\approx \\beta^2 m_\\chi \\sim 100$~keV heavier than the dark matter particle.} \\item{Elastic scatterings off of the nucleus, i.e., $\\chi N \\rightarrow \\chi N$ are suppressed, compared with the inelastic scatterings $\\chi N \\rightarrow \\chi^* N$}. \\end{itemize} The splitting $\\delta$ is comparable to the kinetic energy of a WIMP in the halo. This causes the kinematics of the scattering process to be significantly modified compared with a WIMP that scatters elastically. The altered scattering kinematics leads to a fundamental difference in how an inelastic WIMP shows up in direct detection experiments: only those with sufficient kinetic energy to upscatter into the heavier state will scatter off nuclei. The minimum velocity to scatter with a deposited energy $E_R$ is \\begin{eqnarray} \\beta_{\\rm min} &=& \\sqrt{\\frac{1}{2 m_N E_R}} \\left( \\frac{m_N E_R}{\\mu} + \\delta \\right), \\label{eq:betamin} \\end{eqnarray} where $m_N$ is the mass of the target nucleus and $\\mu$ is the reduced mass of the WIMP/target nucleus system. This minimum velocity requirement means that experiments tend to probe only the higher velocity region of the WIMP halo velocity distribution. This simple modification leads to three key features that change the relative sensitivities of dark matter experiments. Specifically \\cite{TuckerSmith:2001hy,TuckerSmith:2004jv}, \\begin{itemize} \\item{ {\\it Heavier targets are favored over lighter targets} - For a given $E_R$ and $\\delta$, a range of velocities in the halo will be accessible to a given dark matter experiment. \\Eref{eq:betamin} shows that the range available for heavier target nuclei will be greater than that for lighter targets. Because the WIMP velocity distribution is expected to fall rapidly above the peak velocity of $\\sim 220\\; {\\rm km/s}$, targets with heavier nuclei tend to have greater sensitivity than those with lighter nuclei.} \\item{{\\it Modulation of the signal is significantly enhanced} - For conventional WIMPs, the modulation is typically of the order of a few percent. For inelastic WIMPs, because one is generally sampling a higher velocity component of the WIMP velocity distribution, where numbers are changing rapidly as a function of velocity, the modulation can be much higher. In the extreme limit, it can be such that there are particles available to scatter at DAMA/LIBRA in the summer, but not in the winter, yielding a $100 \\%$ modulation of the signal.% } \\item{{\\it The spectrum of events is dramatically changed, suppressing or eliminating low-energy events} - The spectrum of conventional WIMPs rises exponentially at low energy. As a consequence, significant sensitivity gains can be achieved by pushing the energy threshold of the experiment to ever lower energy. With inelastic DM, however, the kinematical changes cause the signal to be suppressed or even eliminated at sufficiently low recoil energies. The result is a spectrum which can peak at 20 keVr and above.} \\end{itemize} In this paper we demonstrate that together, these effects can reconcile DAMA/LIBRA with the null results of other experiments. The CRESST experiment, with its tungsten target, has a robust sensitivity. We shall see that the seven tungsten events seen there are consistent with expectations from this scenario. Before proceeding to quantitative studies of the different experiments, we consider qualitatively how the above differences from conventional WIMPs reconcile the various experimental results. As a first example, we consider the limits of CDMS (Ge) on DAMA. As germanium is significantly lighter than iodine, the sensitivity of CDMS is considerably reduced. Indeed, for some regions of parameter space, no WIMPs in the halo are capable of inelastic scattering at CDMS, while some are capable of scattering at DAMA/LIBRA. In practice, there is some residual sensitivity in CDMS to inelastic WIMPs at the high-velocity edge of the halo distribution. This suggests that bounds on inelastic WIMPs are rather sensitive to the velocity cutoff; we treat this carefully below. In contrast to CDMS, the XENON experiment should be sensitive to inelastic nuclear recoils in a similar way to the DAMA experiment due to the close proximity of the respective target nucleus masses (Xe and I). However, here there are two crucial differences between the two experiments' reported results: (1) the XENON experiment's energy threshold is far smaller than DAMA's (unquenched) energy threshold, and (2) XENON took their data between October 6 to February 14, which happens to be roughly during the time when the Earth travels \\emph{with} the galactic WIMP wind, thus reducing average WIMP velocity. Due to a combination of effects from kinematics and nuclear form factors, conventional WIMPs have a rapidly falling spectrum as a function of energy. This has led to a significant push to lower the energy threshold of experimental searches, in order to optimize sensitivity. XENON's bound on the cross section, for instance, is especially strong because of the dearth of events from 4.5 - 14.5 keV. By comparison, interpreted as unchanneled iodine scatters, the unquenched energy range for the DAMA annual modulation signal is roughly 22-66 keV. Hypothesis-testing approaches for elastic WIMPs can achieve very stringent limits, even in the presence of background events at high energy, if the low energy region is under good control. In contrast, an inelastic WIMP that leads to an annual modulation in DAMA often recoils off nuclei with energies in a range that is nearly \\emph{orthogonal} to the range used by XENON to set their bounds on elastic scattering WIMPs. This is particularly interesting as XENON's blind search identified nine candidate events between 14.5 keV to 26.9 keV, and (outside their analysis range) another fourteen candidate events between 26.9 keV to 45 keV. This spectrum is certainly inconsistent with the distribution of nuclear recoils expected from an elastic scattering WIMP. Yet, we demonstrate that inelastic WIMPs with masses $O$(100 GeV) can be made consistent with the DAMA annual modulation, leaving no trace in the low energy bins of XENON, while contributing between a few to tens of nuclear recoil events in the XENON high energy bins. Thus, we find that XENON does not rule out inelastic dark matter precisely because they have twenty-three events in the signal region where inelastic WIMPs would be expected to be found. Finally, the heightened sensitivity to the high-velocity part of the WIMP halo velocity distribution causes the modulation effect to be significantly enhanced. This makes the time of year of an experiment's run especially important. If the modulation is a few percent, the precise dates of data taking are of little consequence, but in the presence of $O$(1) modulation, it can clearly change the limits. In the extreme case, where at some times of the year there are no dark matter particles capable of scattering, it can remove all sensitivity of (in this case) XENON to DAMA signals. These qualitative arguments are borne out by more careful analyses. The layout of the paper is as follows: in section \\ref{sec:indep} we discuss, without reference to halo models, the relative implications of unmodulated signals to the DAMA modulation signal, and show that consistency of the DAMA signal arising from quenched iodine scatters requires a model with low-energy events suppressed, and modulation enhanced well beyond the standard $\\sim 2\\%$ level expected for a standard WIMP. In sections \\ref{sec:models} and \\ref{sec:kine}, we discuss models for iDM and the kinematical implications of iDM at various experiments, which include such low-energy suppressions and enhanced modulation. In section \\ref{sec:limits} we discuss the limits arising from various experiments, and find the allowed parameter space for these models. We find significant regions of parameter space open, with masses $\\sim 50-200\\; \\gev$, which are otherwise closed in the absence of inelasticity. Finally, in section \\ref{sec:discussion}, we consider the implications of the limits for current and future experiments, and conclude. ", "conclusions": "\\label{sec:discussion} In light of the recent results from DAMA/LIBRA, it is important to continue to consider whether simple, viable models of dark matter exist which can reconcile the presence of the annual modulation signal with the lack of a reported signal from the many other experiments. Most recent analyses \\cite{Foot:2008nw,Feng:2008dz,Petriello:2008jj,Bottino:2008mf} have focused on the light ($\\sim$ few GeV) regions of parameter space, and have exploited the possibility of channeling \\cite{Bernabei:2007hw} in opening low mass windows. In contrast, the mass range for inelastic dark matter is more typical of a standard WIMP. The inelastic scattering can arise quite simply from particle physics models with approximate symmetries, and so remains an appealing possibility, quite distinct from the low-mass window. Of particular interest is the spectral data from DAMA/LIBRA. The low ($2-2.5\\; \\kev$) bin shows significantly lower modulation than what usually occurs for most WIMPs, where the lowest bin often (but not exclusively) shows the highest modulation. This spectral feature arises automatically with inelastic dark matter, and drives our numerical fits to large values of $\\delta$, where modulation is enhanced, and low energy rates at other experiments are highly suppressed. We have seen that the three basic features of iDM, namely, preference for heavy targets, enhanced modulation, and suppression of low-energy events, allow the DAMA modulation to be consistent with other experiments. CDMS is suppressed due to its light target, while XENON is evaded by the suppression of events between 4.5 and 14.5 keV. Interestingly, experiments such as ZEPLIN and KIMS, which are significantly weaker than XENON and CDMS for conventional WIMPs, are competitive in studies of inelastic dark matter. This is because they have target masses similar to iodine, and their energy range focuses on the range most relevant for studying the DAMA signal. In addition to studies similar to those presented here, studies of modulation in the presence of background events may prove quite effective, as the modulated fraction can be O(10\\%) with inelastic dark matter. Of particular importance in this vein is CRESST. Aside from a xenon-target signal shifted to a higher energy range, the second robust prediction of iDM \\cite{TuckerSmith:2004jv} was the inevitable signal from tungsten scattering events. The seven events arising in the CRESST experiment are consistent with the rate expected from iDM. Should they persist in future exposure, this would be strong evidence for the inelastic nature of dark matter. Conversely, the absence of such events would exclude the inelastic explanation of the DAMA modulated signal. . Ultimately, this scenario makes two clear predictions: a signal rate in the 20-50 keVr range for iodine and xenon targets, and significant rates on tungsten targets, whose signal would naturally be peaked at 20-30 keV, and extending up to possibly 80 keV. Germanium targets still have tremendous reach in their next rounds, and such a signal would be peaked at high ($\\sim$ 70 keV) energies. The nature of dark matter remains one of the most important questions in physics. We have seen here that the inelastic dark matter scenario continues to provide an explanation of the DAMA modulation, consistent with the results of other experiments. The robust predictions of this scenario make it exceedingly testable, and consequently, the next generation of experiments should determine if the dark matter is inelastic. \\vskip 0.15in \\noindent{{\\em Note added:} As this work was being prepared, \\cite{Petriello:2008jj} appeared, which also considered the inelastic scenario and concluded that the 50-200 GeV window for inelastic dark matter was excluded by CDMS and XENON. We disagree with this conclusion. We believe the different results obtained in \\cite{Petriello:2008jj} may have arisen from the approximations made there for the halo escape velocity and for the form of the modulation.} \\vskip 0.2in \\noindent {\\bf Acknowledgments} The authors would like to thank Aaron Pierce for extensive and useful discussions. The authors would like to thank Kaixuan Ni for very insightful discussions on XENON, and providing us the information on nuclear recoils above 27 keV. The authors would further like to thank Sunkee Kim for consultation and information on the KIMS experiment. The authors would like to thank Pierluigi Belli for important information on the DAMA/LIBRA results, and Rafael Lang for discussions on CRESST. NW and GDK acknowledge the support of the KITP Santa Barbara, where some of this work was performed, and the support in part by the National Science Foundation under Grant No. PHY05-51164. NW thanks the Aspen Center for Physics for their hospitality, where this work was completed. GDK was supported by the Department of Energy under contract DE-FG02-96ER40969. The work of DTS was supported by NSF grant 0555421. SC and NW were supported by NSF CAREER grant PHY-0449818 and DOE grant \\# DE-FG02-06ER41417. \\vskip 0.15in" }, "0807/0807.4476_arXiv.txt": { "abstract": "{ New \\HI observations of the nearby radio-loud galaxies Centaurus~A and B2~0258+35 show broad absorption ($\\Delta v_{\\rm{absorp}}\\sim 400$~\\kms ) against the unresolved nuclei. Both sources belong to the cases where blue- and redshifted absorption is observed at the same time. In previous Cen~A observations only a relative narrow range of redshifted absorption was detected. We show that the data suggest in both cases the existence of a circumnuclear disk. For Cen~A the nuclear absorption might be the atomic counterpart of the molecular circumnuclear disk that is seen in CO and H$_2$. Higher resolution observations are now needed to locate the absorption and to further investigate the structure and kinematics of the central region of the AGN and the way the AGN are fueled. ", "introduction": "The central region of AGN and the way AGN are fueled are important topics in current extra-galactic astronomy. Mergers have always been considered important for the triggering of AGN by providing the mechanism that could bring gas to the central regions and fuel the black hole \\citep[e.g.][]{hibbard96,mihos96,barnes02}. While this seems to be likely in powerful radio galaxies, recent observations show that other AGN - e.g. low-power radio galaxies - exist where there is no evidence for accretion through mergers. The activity in these galaxies appears to be associated with the slow accretion of gas from the ISM/IGM \\citep{best05,balmaverde08}.\\\\ A powerful tool to learn more about the structure of the central regions of AGN and the fueling mechanism is the study of the distribution and kinematics of the gas. In order to perform detailed studies of single objects we have selected two \\Hi -rich radio-loud galaxies, the merger remnant Centaurus~A and B2~0258+35 (NGC~1167). The later is unlikely to be a recent merger but has a young radio source suggesting that ``cold accretion'' from the IGM plays the dominant role by providing the fuel for the AGN activity (see Sect.~3).\\\\ Neutral hydrogen is an ideal tracer of accretion, interaction and merging events. \\HI in absorption has been often detected in radio sources \\citep[see e.g.][]{conway97,vermeulen03,morganti05}. These absorption structures have shown a variety of characteristics. \\HI absorption profiles only redshifted (relative to the systemic velocity), were initially found in a number of radio galaxies and interpreted as gas clouds that are falling into the nucleus, possibly indicating accretion of gas which deliver the fuel for the AGN \\citep[e.g.][]{vangorkom89}. However, more sensitive and broader band observations have shown that the picture is more complicated and that blueshifted absorption occurs even more often than redshifted absorption \\citep[e.g.][]{vermeulen03,morganti05}.\\\\ In other cases, the \\HI absorption is centered on the systemic velocity of the galaxy and these cases have been often interpreted as \\HI associated with a circumnuclear disk (or torus), see e.g. \\citet{conway97}, \\citet{peck01}. Support for the picture of a circumnuclear disk comes from theoretical work that indeed predict the existence of such circumnuclear \\HI disks around (active) black holes \\citep[see e.g.][]{maloney96,loeb08}. Thus, the picture is more complicated and the questions concerning the central structure and fueling remains. \\begin{figure*}[t!] \\includegraphics[height=4.6cm]{cena.eps} \\caption{\\footnotesize Centaurus~A: Left panel: \\HI absorption against the unresolved nucleus as derived from the ATCA observations \\citep{morganti08}. The total absorption width ranges from 400 to 800~\\kms . The arrow indicates the previously known (redshifted) absorption. The vertical line gives the systemic velocity. Right panel: Position-velocity plot of the \\HI (grey-scale and thin contours) and superimposed the CO emission (thick contours; from \\citet{liszt01}, taken along position angle 139\\degr . Note that the CO observations do not extend beyond a radius of about 1 arcmin. Figure taken from \\citet{morganti08}.} \\label{cena} \\end{figure*} ", "conclusions": "Using new, broad band observations of Cen~A and B2~0258+35 we have shown the presence of blue- and redshifted \\HI absorption against the unresolved nuclei. Both galaxies belong to the group of radio-loud galaxies where blue- and redshifted absorption is observed at the same time. Previously, for Cen~A only a narrow range of redshifted absorption was detected.\\\\ We have shown \\citep{morganti08} that the \\HI absorption in Cen~A is not evidence of gas infall into the AGN, but instead is due to a cold, circumnuclear rotating disk. The absorption profile of the WSRT data suggests the same interpretation for B2~0258+35. Both observations are in agreement with theoretical work which predicts the existence of such circumnuclear disks. However, these results leave the fueling question open.\\\\ Sensitive, higher resolution observations (VLBI for Cen~A and VLA A-array for B2~0258+35) are now needed (and proposed) to detect and locate the absorption in order to further explore the characteristics of the nuclear \\HI and to test the hypotheses of circumnuclear disks.\\\\ Both sources are unique objects. Cen~A is the closest AGN and therefore the linear scale is much better than for any other radio-loud source. Hence, the inner structure and fueling mechanism can be studied in great detail. Because the background structure of B2~0258+35 is extended around all directions of the black hole, higher spatial resolution data will allow to investigate the complete 2-D kinematics in the vicinity of the AGN. For both sources it will be interesting to investigate the presence of velocity gradients (which might be evidence for circumnuclear disks) as well as radial motions (infall/outflow?)." }, "0807/0807.3060_arXiv.txt": { "abstract": "If there is a fifth force in the dark sector and dark sector particles interact non-gravitationally with ordinary matter, quantum corrections generically lead to a fifth force in the visible sector. We show how the strong experimental limits on fifth forces in the visible sector constrain the direct detection cross section, and the strength of the fifth force in the dark sector. If the latter is comparable to gravity, the spin-independent direct detection cross section must typically be $\\lesssim 10^{-55}$ cm$^2$. The anomalous acceleration of ordinary matter falling towards dark matter is also constrained: $\\eta_{\\rm OM-DM} \\lesssim 10^{-8}$. ", "introduction": " ", "conclusions": "" }, "0807/0807.1912_arXiv.txt": { "abstract": "The interferometric technique known as peeling addresses many of the challenges faced when observing with low-frequency radio arrays, and is a promising tool for the associated calibration systems. We investigate a real-time peeling implementation for next-generation radio interferometers such as the Murchison Widefield Array (MWA). The MWA is being built in Australia and will observe the radio sky between 80 and 300 MHz. The data rate produced by the correlator is just over 19 GB/s (a few Peta-Bytes/day). It is impractical to store data generated at this rate, and software is currently being developed to calibrate and form images in real time. The software will run on-site on a high-throughput real-time computing cluster at several tera-flops, and a complete cycle of calibration and imaging will be completed every 8 seconds. Various properties of the implementation are investigated using simulated data. The algorithm is seen to work in the presence of strong galactic emission and with various ionospheric conditions. It is also shown to scale well as the number of antennas increases, which is essential for many upcoming instruments. Lessons from MWA pipeline development and processing of simulated data may be applied to future low-frequency fixed dipole arrays. ", "introduction": "\\label{Introduction} The Murchison Widefield Array (MWA) is an 80-300 MHz synthesis array that is being built in Western Australia, with construction to be completed in 2010. The shire of Murchison has a quiet radio environment, making it an excellent site for this and other radio facilities \\cite{Bowman2007a}. Each of the 512 antennas will be a $4\\times4$ tile of dipoles. An analogue beamformer at each antenna combines the signals from the 16 dipoles, producing an electronically steerable primary beam with a width of approximately 25$^\\circ$ at 150 MHz. When the signals from all antennas are combined, the array will have a synthesized beam with a width of approximately 4.5$^\\prime$ at 150 MHz. The main science goals of the MWA are the detection of redshifted 21cm emission from the Epoch of Reionization (EoR) \\cite{Bowman2006}, transient detection (for example \\cite{Bhat2007}), and remote heliospheric sensing \\cite{Salah2005}. A schematic of two MWA antenna tiles is shown in Fig. \\ref{plot:schematic}. To make a map of the sky using radio interferometry, one typically builds up an estimate of the 2D Fourier transform of the sky, then applies a Fourier transform to obtain the image. This is known as synthesis imaging, and a good overview of the subject is given in \\cite{Thompson-etal.2001}. The measured data that are used to build up the Fourier interference pattern are spatial cross-correlations -- or visibilities -- that are obtained by correlating voltage streams collected by many pairs of antennas. The visibilities lie in the uvw coordinate frame, where $w$ is the component of the antenna separation vector (or \\emph{baseline} vector) in the direction of the field center (in units of wavelengths), and $u$ and $v$ are orthogonal coordinates in the plane normal to $w$ (aligned with the corresponding image coordinate axes, $l$ and $m$). This situation is not naturally a 2D Fourier transform. For small images, $w$ is multiplied by a term that is approximately zero and the 2D nature holds. For large fields this is not the case, but the problem can still be reduced to 2D transforms (a good overview is given in \\cite{Cornwell2005}).\\footnote{The MWA will produce \\emph{snapshot} images, which means each antenna pair contributes a single visibility to each image. The MWA visibilities will be approximately coplanar, so a 2D Fourier relationship will hold even for wide-field images.} Post processing typically involves calibrating the visibilities, gridding them onto the uv-plane to form a regularly sampled interference pattern, and then applying a 2D FFT to form an image. Techniques such as self-calibration can then be used in an attempt to improve the calibration by iterating back and forth between the visibilities and the image. For a number of reasons, MWA visibilities cannot be processed in this way. Many of these effects are common to all low frequency arrays, and are described in detail in \\cite{Erickson1999} and \\cite{Thompson-etal.2001}. Each antenna has a different direction-dependent response over the field of view, which cannot simply be divided out. These response patterns may also change significantly over the course of an observation. Furthermore, the ionosphere causes direction-dependent phase shifts that effectively change the position and polarization state of sources during an observation. These effects mean that we cannot make a fully calibrated interference pattern in the standard way. What we can do is use the measured visibilities to iteratively fit ionospheric phase shifts and antenna gains towards many bright catalogue sources, and store these fits to aid deconvolution and resampling processes after the images have been made. These measurements are the focus of this paper, but before they are discussed some of these challenges will be reviewed more closely. \\begin{figure}[ht] \\centering \\includegraphics[width=7.8cm]{fig1.png} \\caption{Schematic diagram of an MWA interferometer baseline and sky. Each MWA receiving element is a phased tile comprising a $4\\times4$ grid of crossed dipoles ({\\it vertical planar bowtie structures}). The main response lobes are steered electronically by beam-formers ({\\it BF}) to establish the instrument field of view. BF output signals are sampled at baseband and filtered digitally in receiver electronics ({\\it RX}) and correlated to provide cross-power spectra. In addition to compact objects (e.g., quasars and pulsars), the MWA sky includes foreground emission from the galaxy, which is chiefly synchrotron emission that is linearly polarized and traces turbulent and ordered magnetic fields in the interstellar medium ($\\vec{B}$). Magnetized plasma from solar coronal mass ejections or other activity move outward through the heliosphere. This medium and the Earth ionosphere induce primarily refractive shifts in source positions (higher order distortions in the observing passband are small due to the limited extent of the array) and time-variable Faraday rotation of polarization along different lines of sight. The calibration scheme described here enables polarization-sensitive imaging of the diffuse galactic and compact source populations and assembly of a sky model. Knowledge of the compact background sources is ``confusion limited'' (i.e., constrained by multiplicity of sources within one instrument resolution element and contamination of each image pixel by sidelobes of distant sources). Confusion increases with image sensitivity and the size of a resolution element..} \\label{plot:schematic} \\end{figure} Unlike the radio sky at higher frequencies, which appears sparsely populated at the sensitivity levels of modern instruments, the sky to be observed by the MWA is full of sources. The high density of sources and large angular resolution of the array will result in images that are confusion limited, with significant flux coming from background galaxies in each synthesized beam, as well as from the sidelobes of other sources in the primary beam of each antenna. The sky is also quite complex. There is significant emission on many angular size scales, from compact extragalactic sources and pulsars to diffuse galactic synchrotron radiation \\cite{Erickson1999}. The latter experiences Faraday rotation and depolarization in the ionized interstellar medium, and as a result exhibits a strong linearly polarized component that is highly dependent on position and frequency \\cite{Wieringa1993}. This polarized all-sky signal is at least a few orders of magnitude brighter than the unpolarized EoR signal, and a fully-polarized calibration formalism, such as the Hamaker-Bregman-Sault measurement equation \\cite{Sault-etal.1996}, is required to reduce contamination from the spatially structured, linearly polarized emission from the galaxy. If the raw interferometer data are stored for offline processing, iterative calibration and deconvolution algorithms can be used to address many of the problems described below (see, for example, \\cite{Bhatnagar2008} and \\cite{Nijboer2007}). However, it is impractical to store the 19 GB s$^{-1}$ data stream coming from the MWA correlator,\\footnote{The $N_a=512$ antennas lead to $N_a(N_a-1)/2=130816$ different cross-correlation measurements. They are made in a correlator for 4 different polarization products and up to 3072 different frequency channels every 0.5 seconds. Each visibility is represented by 3 real bytes and 3 imaginary bytes, which leads to $\\sim19$ GB/s.} and the MWA will store images. This means that much of the calibration must take place in real time, before or during the imaging process. At the heart of the calibration system for the MWA is the calibrator measurement loop (CML), which measures apparent angular offsets induced by the ionosphere and the system gain toward known compact astronomical sources across the sky. These measurements are used to fit models of the ionosphere and instrument response, and support subtraction of strong sources that limits sidelobe contamination during calibration. As in \\cite{Noordam2004} and \\cite{vanderTol2007}, measuring and subtracting the contribution of each source is carried out sequentially, so that the stronger sources are removed before measurements of weaker sources are made. In this paper we do not consider multivariate fits of parameters for all of the calibrator sources simultaneously, as described in \\cite{vanderTol2007}, but we will discuss it briefly in section \\ref{Discussion}. Estimation of calibration parameters is greatly over constrained due to the large number of antenna pairs ($1.31\\times10^5$). On the other hand, the wide-field nature of the instrument and real-time computing requirement pose challenges, the most important of which are listed below. \\begin{enumerate} \\item \\emph{Direction-dependent gain and polarization response}. Each MWA receiving element is a $4\\times4$ array of fixed crossed dipoles (Fig. \\ref{plot:schematic}). The phased beams are steered with an analogue beamformer, which will typically be updated every 5 to 10 minutes to compensate for rotation of the Earth. The common approach of assuming that the polarized receptors are orthogonal over the field of view with a small amount of direction-dependent leakage cannot be used. The direction-dependent instrumental polarization of the antenna beams will be significant, and it will be measured along with the direction-dependent gain using many calibrator sources spread over the entire sky. These measurements will be repeated as the field of interest moves across the antenna beams. \\item \\emph{Confusion}. Since the MWA's primary beams cover such a large section of the sky, each field mapped by the MWA will contain hundreds of relatively bright sources. To calibrate the array, we require accurate flux density measurements of known sources. Such measurements can be corrupted by faint sources within the synthesized beam of the array (``confusion'') as well as the sidelobes of brighter sources outside the synthesized beam (``sidelobe contamination''). Confusion and sidelobe contamination can arise from both compact and large-scale sources such as extragalactic radio galaxies and galactic synchrotron emission respectively. The Galactic synchrotron, in particular, has a polarized component and structure on many spatial scales. Since the interferometer baselines will respond to large-scale structure differently depending on their length and orientation, calibration of data that includes bright resolved sources such as the Sun or Galactic plane must be performed carefully. \\item \\emph{Ionosphere}. The 3-dimensional ionosphere can significantly perturb the waves coming from celestial sources. The maximum antenna separation of the MWA is short enough that, for a given source and during normal ionospheric conditions, all of the antennas have approximately the same line of sight through the ionosphere. This assumption will be used throughout. Under these conditions there will be no defocusing and the ionosphere can be described by a two-dimensional phase screen that makes sources appear to move away from their true positions. (Faraday rotation of incident polarization is a second ramification, but this will be considered in a future paper. We limit consideration here to ionospheric calibration using unpolarized sources, which cannot be used to calibrate polarization position angles \\cite{Sault-etal.1996}, \\cite{Hamaker2000a}.) \\item \\emph{Real-time data reduction}. The real-time nature of the MWA means that compute-intensive processes need to avoided where possible. Unfortunately, this means that many of the promising techniques currently being investigated, such as iterative self-calibration and deconvolution algorithms \\cite{Bhatnagar2008}, and some wide-field imaging algorithms \\cite{Cornwell2005}, cannot currently be implemented in real time. They either cannot be used at all or need to be approximated. \\end{enumerate} Calibration occurs in a back-end known as the real-time system (RTS), which consists of a visibility integrator (time and frequency), the CML, and an imaging pipeline. These tasks run sequentially, and as mentioned later the processing load is split over frequency. The imaging pipeline incorporates gridding, imaging FFTs, correction for ionospheric and wide-field distortion of the sky, Stokes conversion of images,\\footnote{Stokes parameters describe the polarization state of a signal as an unpolarized component, I, two linearly polarized components, Q and U, and a circularly polarized component, V. They are used extensively in radio astronomy, see, for example, \\cite{BornWolf1999} and \\cite{Thompson-etal.2001}.} and astronomical coordinate conversion. MWA primary science drivers require that the time and frequency resolution are sufficient to ensure that stationary signals from sources throughout the antenna field of view are coherent for the highest frequencies and longest antenna separations, where interference fringe phases vary most rapidly with time. Most of the key elements of the real-time calibration system have been coded and are regularly tested with simulated data, as described in section \\ref{Examples}. To date, the tests have focused on unpolarized cosmic signals, but the response of the instrument polarizes the signals during reception, and the processing employs the fully-polarized description discussed in the ensuing chapters. Apart from tolerance testing -- to determine optimal bandwidths, number of calibrators, etc. -- and algorithm development for real-time operation, the main piece of outstanding work is the incorporation of polarized calibrators into the system to fully constrain primary beam models and the Faraday rotation component of the ionosphere. While the discussion and examples given below focus on the MWA, the techniques are applicable to other low-frequency array projects, such as the SKA Molonglo Prototype (SKAMP), the Long Wavelength Array (LWA), and the Low Frequency Array (LOFAR). That said, the instantaneous synthesized beam of the MWA does make it particularly well suited to this type of processing. We will not go into specific details on how the techniques can be optimized, which at any rate will be different for the different arrays, and refer to papers such as \\cite{Lonsdale2004} for an overview of the power of large-$N$ radio arrays. After outlining the assumptions and mathematical model in the next section, the steps in the CML are discussed in more detail in section \\ref{The Calibrator Measurement Loop}, followed by a discussion on algorithm convergence and performance. We then finish with an analysis of peeling simulated MWA data in section \\ref{Examples}. ", "conclusions": "\\label{Discussion} \\subsection*{Performance} The relatively short cadence time of the RTS means we need to make compromises when designing real-time peeling process. One of these compromises will be the number of sources peeled every 8 seconds. Several methods for calibrating antenna gains are compared in \\cite{Boonstra2003}. These are alternatives to the technique discussed in section \\ref{Instrumental Gain Measurements}. The number of complex multiplications required by most of the techniques scales as $N_a^3$, where $N_a$ is the number of antennas (or, more generally, the number of receivers being correlated to form visibilities). The technique that required the fewest computations was the logarithmic least squares (LOGLS) algorithm, which scaled as $N_a^2$. The numbers given in \\cite{Boonstra2003} for a single polarization version are $2N_a^2$ multiplications with an additional $16N_a^2$ for weighting. How does this compare to the algorithm discussed in section \\ref{Instrumental Gain Measurements}? In our application we measure two polarizations, so if we consider each antenna as 2 polarized receptors the number of multiplications for the LOGLS algorithm is $2(2N_a)^2$. There will probably be another factor of 2 since correlations between the receptors on the same antenna will most likely need to be considered.\\footnote{The authors of \\cite{Boonstra2003} note that LOGLS is not easily generalized to a dual-polarized telescope array, but, for the sake of comparison, suppose that such a generalization exists.} The number of complex multiplications used to determine the calibration solutions in section \\ref{Instrumental Gain Measurements} is $O(24N_a^2)$, with an additional $O(12N_a^2)$ for weighting. This efficiency appears to be about as high as one might reasonably expect to achieve. We are not just dealing with a single source, however. The CML needs to pre-peel all of the calibrator sources, and then, for each source, unpeel, rotate all of the visibilities, solve for the ionospheric offset, solve for the antenna gains, and peel. Table \\ref{tbl:FLOPS} shows the approximate number of floating-point operations used in each of these steps. These numbers were obtained by listing the main operations in the inner loops of the routines and multiplying each by our best estimate of the associated floating-point operations. These numbers are not exact; they are provided as a rough indication of where processing time will be spent. Also, the number of sources processed by each of the routines need not be the same. For example, we might peel and make gain measurements for 50 sources, but make ionospheric measurements on a few hundred more (which requires only the third and fourth rows). \\begin{table}[htb] \\centering \\caption{Approximate floating-point operations required for each source in the CML (with 512 antenna tiles and a single frequency channel).} \\begin{tabular}{ c c } {\\bf Routine} & {\\bf Floating-point operations (millions)} \\\\ \\hline peeling (applied 3 times) & 31 \\\\ rotate and accumulate & 26 \\\\ ionospheric sums and rerotation & 21.5 \\\\ measure tile gains & 37 \\\\ \\hline total & $O$(180) \\\\ \\end{tabular} \\label{tbl:FLOPS} \\end{table} While the algorithm scales well, there is still on the order of 180 million floating-point operations required during each 8 second calibration cycle. For the whole array, some of the rows in the table need to be multiplied by the number of frequency channels (768), while others need to be multiplied by the number of frequency sub-bands ($\\sim50$). They also need to be multiplied by the number of sources, as discussed in the previous paragraph. We anticipate a few trillion floating-point operations for the CML over the 8 seconds. This does not include various overheads such as memory access that will increase the number of operations by a factor of a few. \\subsection*{Options for Reducing the Load} Each ionospheric refraction measurement can use data from every baseline, polarization and frequency. That is over 1.5 gigasamples for a single 2D offset. For sources that are only used for ionospheric measurements, all of the processing listed in table \\ref{tbl:FLOPS} will be reduced if we only use a subset of the visibilities. Most of the baselines are short, and many of those will be redundant. Furthermore, long baselines do not see as much of the extended galactic structure as short baselines, and they measure the apparent offset with higher angular resolution. Initial investigations suggest that we may be able to ignore more than 99\\% of the short baselines for the stronger sources. Since the tile gains are changing slowly, we do not necessarily need to make measurements for every source every 8 seconds. This can be exploited by making measurements for the strongest sources every 8 seconds, and cycling through subsets of the other sources. This way we still have gain measurements distributed across the sky every few minutes when we make fits for the tile beams, but we only run the full peel algorithm on 20 or 30 sources at a time. Rather than solving for direction-dependent parameters sequentially, one could fit for all of the calibrator sources simultaneously. In fact, the peeling algorithm is really just a robust and efficient method for reducing the number of unknowns and finding the multivariate solutions. One can also reduce the degrees of freedom by changing the fit parameters to quantities that do not change (or change slowly) with time or frequency, and solve using multiple snapshots, as discussed in detail in \\cite{vanderTol2007}. For the MWA, we hope to be able to use slowly varying direction-dependent dipole gains and phases to describe the primary beam of each tile. Once we have high quality measurements of our tile beams in the field, we will look more closely at fitting these dipole parameters directly. However, source subtraction will then be based on these dipole fits, not on direct tile gain measurements. \\subsection*{Foreground Subtraction} One of the primary challenges in the search for a signature from the EoR is that of foreground subtraction. At best the signal will be several orders of magnitude weaker than the galactic foreground and it is important that we understand the nature of the residuals from the calibration and peeling process. One of the drawbacks of peeling strong sources in real time and then averaging the resulting images together is that any residuals are also averaged into the mix. All of the calibration data will be stored in a database for use in offline processing, and peeling errors can be assessed and reduced at that stage. However, if there is any concern that residuals from the peeling process might mimic the EoR signal then peeling can be used for calibration only, and images formed from unpeeled visibilities. There is significant effort going into techniques for removing foregrounds during offline processing. These include techniques that exploit spectral differences in the foregrounds and the EoR signature (see for example \\cite{Morales2006} and the discussion and references in \\cite{Furlanetto2006}), and the rotation measure synthesis techniques discussed in \\cite{Burn1966} \\& \\cite{Brentjens2005}. Direction-dependent deconvolution techniques such as the one described in \\cite{Bhatnagar2008} will also be essential for imaging. \\subsection*{Radio Frequency Interference} Of concern for any telescope operating at MWA frequencies is radio frequency interference (RFI). Even though the Murchison site is extremely radio quiet \\cite{Bowman2007a}, the array will still have to deal with some RFI. This includes communication and military satellite signals, reflections of FM radio broadcasts, and natural interference such as lightning. Due to the low spectral occupancy of the RFI, the high quality polyphase filter banks used to isolate frequencies, and the campaign-mode operation of the array, we will adopt the traditional strategy of flagging and ignoring contaminated data before imaging. Missing frequency or time samples can be accounted for in the weighting of the various least-squares algorithms. \\label{Summary} We have described a general approach for making measurements of strong point sources that can be used in the calibration process of wide-field, low-frequency radio arrays. This approach has been adopted for the MWA, and there is an ongoing effort to develop the required software, as well as to understand the benefits and limitations of the approach. We have used simulated visibility data to show that the peeling algorithm works well in situations that are of major concern for future radio telescopes: crowded fields, strong galactic emission, and ionospheric refraction. The algorithm exhibits fast convergence, which is important since sources will be moving in and out of antenna sidelobes and the algorithm needs to be able to keep up with the antenna gain and phase changes, as well as changes in the ionosphere. Critical parts of the process are shown to be computationally efficient, and parts of the system lend themselves to significant levels of optimization. The next step is a detailed analysis of the convergence properties of the algorithms, and a series of tolerance tests to investigate how the algorithms will behave in the various conditions we expect to encounter. This includes observations of weak emission that is masked by significant polarized diffuse foregrounds, high dynamic range observations close to the sun, and observations in the presence of severe ionospheric conditions, such as during activation and recombination of the ionosphere. As data from the initial deployment of antennas become available in late 2008 and early 2009, we will get a clearer picture of how harmful phenomena such as source variability (due to ionospheric scintillation, for instance) and dipole mutual coupling (which will affect our tile beam models) can be, and these can be worked into the tolerance tests." }, "0807/0807.4530_arXiv.txt": { "abstract": "We study the impact of inhomogeneous hydrogen reionization on the thermal evolution of the intergalactic medium (IGM) using hydrodynamic + radiative transfer simulations where reionization is completed either early ($z\\sim9$) or late ($z\\sim6$). In general, we find that low-density gas near large-scale overdensities is ionized and heated earlier than gas in the large-scale, underdense voids. Furthermore, at a later time the IGM temperature is inversely related to the reionization redshift because gas that is heated earlier has more time to cool through adiabatic expansion and Compton scattering. Thus, at the end of reionization the median temperature-density relation is an inverted power-law with slope $\\gamma-1\\sim-0.2$, in both models. However, at fixed density, there is up to order unity scatter in the temperature due to the distribution of reionization redshifts. Because of the complex equation-of-state, the evolved IGM temperature-density relations for the redshift range $4\\lesssim z\\lesssim6$ can still have significant curvature and scatter. These features must be taken into account when interpreting the Ly$\\alpha$ absorption in high redshift quasar spectra. ", "introduction": "The cosmic reionization of hydrogen fundamentally changes the thermal and ionization conditions for the high redshift ($z\\gtrsim6$) intergalactic medium (IGM), converting a cold and neutral gas into a warm and highly ionized one \\citep[see][for a review]{2008Loeb}. During this inhomogeneous process, the radiative transfer (RT) of the UV field, primarily generated by stellar sources, will proceed such that large-scale, overdense regions near sources are generally ionized and heated earlier than large-scale, underdense regions far from sources \\citep[e.g.][]{2004BarkanaLoeb, 2004WyitheLoeb, 2004FZH, 2008Lee}. The photo-ionization of hydrogen (HI $\\rightarrow$ HII) and helium (HeI $\\rightarrow$ HeII) is expected to heat the IGM to temperatures of $\\gtrsim10^4$ K, with higher temperatures obtained for harder radiation spectra \\citep[e.g.][]{1990MiraldaEscude, 1994MiraldaEscude}. Thereafter, the thermal evolution will be governed by adiabatic cooling/heating for underdense/overdense gas, Compton scattering on the Cosmic Microwave Background (CMB), atomic line cooling, shock-heating, and additional photo-heating only where recombination is efficient. In principle, the temperature of the IGM can be quantified through the thermal broadening of the Lyman alpha ($\\lya$) forest lines \\citep[e.g.][]{1999Schaye, 2000Schaye, 2000Ricotti, 2001McDonald} and the Jeans smoothing of the $\\lya$ forest flux power spectrum \\citep[e.g.][]{2000Theuns, 2001Zaldarriaga, 2004Viel}. Measurements of the temperature can then be used to put constraints on when the reionization of hydrogen and helium occurred \\citep[e.g.][]{2002Theuns, 2003HuiHaiman}. Note that the full reionization of helium (HeII $\\rightarrow$ HeIII) is expected to occur at lower redshifts ($z\\sim3$), triggered by quasars with harder spectra \\citep[see][for recent modelling]{2008Furlanetto, 2008bBolton, 2008McQuinn}. The current observational constraints have large uncertainties, although better measurements will come in time and improved models will be required to interpret them. In this {\\it Letter} we focus on the impact of inhomogeneous hydrogen reionization on the thermal evolution of the IGM. In particular, we study the equation-of-state and measure the temperature-density relation and its scatter using hydrodynamic + RT simulations. \\citet{1997HuiGnedin} have studied the effects of uniform reionization and found that the temperature-density relation is well approximated by a positive power-law slope for the low-density IGM. However, inhomogeneous reionization can give rise to a much different equation-of-state, especially right after completion. Specifically, the large-scale, underdense voids are initially hotter than higher density gas near sources, that having been ionized and heated earlier have also had more time to cool. This gives rise to an equation-of-state characterized by an inverted (negative slope) temperature-density relation. At later times, the underdense gas cools adiabatically faster and the temperature contrast decreases and eventually reverses back to the positive slope. ", "conclusions": "We have studied the photo-ionization and photo-heating of the IGM from hydrogen reionization using two hydrodynamic + RT simulations. We considered two basic models in which reionization is completed early ($z\\sim9$) and late ($z\\sim6$) and found for both models that the temperature of a low-density region at a later time is inversely related to the redshift of reionization of that region. low-density gas near large-scale overdensities is ionized and heated earlier than gas in the large-scale, underdense voids. As a result, at the end of reionization the median temperature-density relation is an inverted power-law with slope $\\gamma-1\\sim-0.2$. There is up to order unity scatter in the temperature at fixed density due to the distribution of reionization redshifts. We expect that both the slope and scatter will depend on the reionization history, especially its duration. Furthermore, it is known that radiative transfer effects from Lyman limit systems can modify the temperature distribution \\citep[e.g.][]{1999AbelHaehnelt}, although more for helium rather than hydrogen reionization because for the latter, the UV spectrum from stellar sources is relatively soft and only weak spectral filtering can occur. We conclude that at the high redshift range $4\\la z\\la 6$, the equation-of-state of the IGM is more complicated than the commonly assumed forms (isothermal or a tight power-law relation). It is important to keep this result in mind when interpreting the Ly$\\alpha$ absorption in quasar spectra \\citep[e.g.][]{2002Fan, 2006Fan, 2006Lidz, 2008Gallerani}. In fact, a recent analysis by \\citet{2007Becker} has already suggested that adopting an inverted temperature-density relation instead of an isothermal model may have profound implications on the interpretation of the reionization process based on the high redshift Ly$\\alpha$ forest. We will study the observational signatures of inhomogeneous hydrogen reionization in an upcoming paper." }, "0807/0807.4750_arXiv.txt": { "abstract": "With the TRIS experiment we have performed absolute measurements of the sky brightness in a sky circle at $\\delta = +42^{\\circ}$ at the frequencies $\\nu =$ 0.60, 0.82 and 2.5 GHz. In this paper we discuss the techniques used to separate the different contributions to the sky emission and give an evaluation of the absolute temperature of the Cosmic Microwave Background. For the black-body temperature of the CMB we get: $T_{cmb}^{th}=(2.837 \\pm 0.129 \\pm 0.066)K$ at $\\nu=0.60$ GHz; $T_{cmb}^{th}=(2.803 \\pm 0.051 \\ ^{+0.430} _{-0.300})K$ at $\\nu=0.82$ GHz; $T_{cmb}^{th}=(2.516 \\pm 0.139 \\pm 0.284)K$ at $\\nu=2.5$ GHz. The first error bar is statistic (1$\\sigma$) while the second one is systematic. These results represent a significant improvement with respect to the previous measurements. We have also set new limits to the free-free distortions, $ -6.3 \\times 10^{-6} < Y_{ff} < 12.6 \\times 10^{-6}$, and slightly improved the Bose-Einstein upper limit, $|\\mu| < 6 \\times 10^{-5}$, both at 95\\% confidence level. ", "introduction": "At decimetric wavelenghts the sky brightness temperature ($T_{sky}$) can be written as the sum of different contributions: cosmic microwave background ($T_{cmb}$), galactic emission ($T_{gal}$), unresolved extragalactic radio sources ($T_{uers}$). The temperature of the cosmic microwave background (CMB), $T_{cmb}$, can be considered, for our purposes, as isotropic. The largest anisotropy component is the dipole term $\\Delta T_{dipole}(\\alpha,\\delta) = T_d \\ \\cos \\theta$, with $ T_d = 3.381 \\pm 0.007$ mK, and $\\theta$ angle between the direction of observation and the maximum of the dipole at ($\\alpha = 11^h 12^m.2 \\pm 0^m.8$; $\\delta = -7^{\\circ}.06 \\pm 0^{\\circ}.16$) (\\cite[]{Bennet_96}; \\cite[]{Fixen_96} and \\cite[]{Fixen_02}). It gives a level of anisotropy well inside the error budget of our absolute measurements. Therefore we will consider in the present paper only the possible dependence of the CMB brightness temperature on the frequency ($T_{cmb}(\\nu)$). The frequency dependence could be related to the spectral distortion of the CMB. The most accurate measurement of the CMB temperature so far made is the one obtained by the COBE-FIRAS team (\\cite[]{Mather_90}; \\cite[]{Mather_94}; \\cite[]{Fixen_96}; \\cite[]{Mather_99}; \\cite[]{Fixen_02}). Between 60 to 600 GHz they found a spectrum fully compatible with a black body emitting at the thermodynamic temperature $T_{cmb}^{th}=2.725 \\pm 0.001 \\ K$ (see \\cite[]{Fixen_02}). Major spectral distortions in this range are excluded by COBE-FIRAS and the parameters of both the Bose-Einstein and Compton distortions are largely constrained: $|\\mu| < 9 \\times 10^{-5}$ and $|y| < 15 \\times 10^{-6}$ (both at 95\\% CL, see \\cite[]{Fixen_96}). Although very accurate the FIRAS result covers only part of the CMB frequency spectrum. In particular it does not cover the low frequency region where TRIS measurements were made and where distortions can be expected. Table \\ref{tab1} collects the measurements of the CMB temperature made at frequencies below few GHz starting from the 80's. At these low frequencies the error bars are much larger than at FIRAS frequencies, essentially because of the large uncertainties associated to the absolute calibration procedures and to the presence of important foregrounds. To improve the low frequency situation it has been proposed in the past to carry on measurements of the CMB temperature from space: LOBO (\\cite[]{Sironi_95}; \\cite[]{Sironi_97}) for which TRIS is a pathfinder and DIMES \\cite[]{Kogut_96}. DIMES has been then transformed in a balloon program (ARCADE), whose results have been published recently (see \\cite[]{Kogut_04}; \\cite[]{Fixen_04}; \\cite[]{Singal_06}). The black-body spectrum of the CMB is the result of the complete thermalization of the Universe in the pre-recombination era. At that time the strong interactions among the different components of the cosmic plasma were efficient enough to re-establish immediately the thermodynamic equilibrium after an energy injection. Nevertheless small deviations from the black-body distribution can survive if the energy injection occurred at a red-shift $z\\lesssim 10^6$. In the pre-recombination era ($z\\gtrsim 10^4 - 10^3$) a kinetic equilibrium is re-established through the Compton scattering, Double Compton scattering and the Bremsstrahlung mechanism. The resulting spectrum assumes a typical shape described by the Bose-Einstein or by the Compton distribution. For a review of these mechanisms see \\cite[]{Zeldovich_69}, \\cite[]{Sunyaev_70}, \\cite[]{Zeldovich_72}, \\cite[]{Illarionov_75a}, \\\\ \\cite[]{Illarionov_75b}, \\cite[]{Sunyaev_80}, while detailed calculations have been performed by \\cite{Burigana_91a}, \\cite{Burigana_91b}, \\cite{Burigana_95} and \\cite{Daly_91}. These types of distortions are described by the chemical potential $\\mu$ and the Comptonization parameter $y$, whose possible values have been already limited by COBE-FIRAS. Energy injections after the recombination era are responsible of spectral distortions without any possibility of re-thermalization. Among these processes there is the photon injection through free-free mechanism in a re-ionized universe (see \\cite[]{Bartlett_91}). This effect depends on the square of the wavelength and therefore is dominant at very low frequencies and is not constrained by the COBE-FIRAS measurements. Other possible distortions are related to the formation of primordial molecules and to the formation of structures at large scale. Both these mechanisms are difficult to be modeled but, at least in the case of the formation of primordial molecules, a not negligible effect could be expected at decimetric wavelengths (see \\cite[]{Varsha_77} and \\cite[]{Dubrovich_95}). The galactic emission ($T_{gal}(\\nu,\\alpha,\\delta)$) is anisotropic and has a power law frequency spectrum: $T_{gal}(\\nu) = T_{gal}(\\nu_0) (\\nu/\\nu_0)^{\\beta}$. Its importance and properties are analyzed in the companion paper \\cite[]{TRIS-III} (hereafter Paper III) from which we take the spectral index $\\beta$ we will use in the following. The contribution of the unresolved extragalactic radio sources (UERS), $T_{uers}$, has also a power law frequency spectrum, but for a low resolution experiment it can be considered isotropic. In the following we will take it from the model developed by \\cite{Gervasi_08} on the basis of the source number counts measurements available in literature. After a brief summary of TRIS experiment (Section \\ref{TRIS}), in Section \\ref{Calculation} we discuss the techniques used to separate the components of $T_{sky}$ and the results obtained for $T_{cmb}$. The implications of these results in terms of spectral distortions are presented in Section \\ref{Discussion}. ", "conclusions": "Starting from the absolute measurements of the sky brightness temperature, performed by the TRIS experiment, and presented in Paper I \\cite[]{TRIS-I}, we have evaluated the absolute temperature of the Cosmic Microwave Background at $\\nu =$ 0.60, 0.82 and 2.5 GHz. The thermodynamic temperatures of the CMB we get are: $T_{cmb}^{th}=(2.837 \\pm 0.129 \\pm 0.066)K$ at $\\nu=0.60$ GHz; $T_{cmb}^{th}=(2.803 \\pm 0.051 \\ ^{+0.430} _{-0.300})K$ at $\\nu=0.82$ GHz; $T_{cmb}^{th}=(2.516 \\pm 0.139 \\pm 0.284)K$ at $\\nu=2.5$ GHz. The first error bar is 1$\\sigma$ statistics, while the second one is the systematic on the zero level assessment. Thanks to improvements of the absolute calibration system and in the foregrounds separation technique TRIS succeded in reducing previous uncertainties by a factor $\\sim 9$ at $\\nu=0.60$ GHz and by a factor $\\sim 7$ at $\\nu=0.82$ GHz. At 2.5 GHz TRIS results are in agreement with the previous measurements. These results, used to look for CMB spectral distortions, give an upper limit to the chemical potential $|\\mu| < 6 \\times 10^{-5}$ (95\\% CL) used to describe the \\emph{BE} distortions. We have also constrained the free-free distortions to: $ -6.3 \\times 10^{-6} < Y_{ff} < 12.6 \\times 10^{-6}$ (95\\% CL), approaching the values suggested by observations of the Lyman-$\\alpha$ forest \\cite[]{Haiman_97}." }, "0807/0807.2176_arXiv.txt": { "abstract": "The flight calibration of the spectral response of CCD instruments below 1.5 keV is difficult in general because of the lack of strong lines in the on-board calibration sources typically available. We have been using \\name\\, the brightest supernova remnant in the Small Magellanic Cloud, to evaluate the response models of the ACIS CCDs on the Chandra X-ray Observatory (CXO), the EPIC CCDs on the XMM-Newton Observatory, the XIS CCDs on the {\\em Suzaku} Observatory, and the XRT CCD on the {\\em Swift} Observatory. E0102 has strong lines of O, Ne, and Mg below 1.5 keV and little or no Fe emission to complicate the spectrum. The spectrum of E0102 has been well characterized using high-resolution grating instruments, namely the XMM-Newton RGS and the CXO HETG, through which a consistent spectral model has been developed that can then be used to fit the lower-resolution CCD spectra. Fits with this model are sensitive to any problems with the gain calibration and the spectral redistribution model of the CCD instruments. We have also used the measured intensities of the lines to investigate the consistency of the effective area models for the various instruments around the bright O ($\\sim570$~eV and $\\sim654$~eV) and Ne ($\\sim910$~eV and $\\sim1022$~eV) lines. We find that the measured fluxes of the O~{\\small VII}~triplet, the O~{\\small VIII}~Ly~$\\alpha$ line, the Ne~{\\small IX}~triplet, and the Ne~{\\small X}~Ly~$\\alpha$ line generally agree to within $\\pm10\\%$ for all instruments, with 28 of our 32 fitted normalizations within $\\pm10\\%$ of the RGS-determined value. The maximum discrepancies, computed as the percentage difference between the lowest and highest normalization for any instrument pair, are 23\\% for the O~{\\small VII}~triplet, 24\\% for the O~{\\small VIII}~Ly~$\\alpha$ line, 13\\% for the Ne~{\\small IX}~triplet, and 19\\% for the Ne~{\\small X}~Ly~$\\alpha$ line. If only the CXO and XMM are compared, the maximum discrepancies are 22\\% for the O~{\\small VII}~triplet, 16\\% for the O~{\\small VIII}~Ly~$\\alpha$ line, 4\\% for the Ne~{\\small IX}~triplet, and 12\\% for the Ne~{\\small X}~Ly~$\\alpha$ line. ", "introduction": "\\label{sect:intro} % This paper reports the progress of a working group within the {\\em International Astronomical Consortium for High Energy Calibration} (IACHEC) to develop a calibration standard for X-ray astronomy in the bandpass from 0.3 to 2.5 keV. A brief introduction to the IACHEC organization, its objectives and meetings, may be found at the web page {\\tt http://www.iachec.org/}. Our working group was tasked with selecting celestial sources with line-rich spectra in the 0.3-2.5~keV bandpass which would be suitable cross-calibration targets for the current generation of X-ray observatories. The desire for strong lines in this bandpass stems from the fact that the quantum efficiency and spectral resolution of the current CCD-based instruments is changing rapidly from 0.3 to 1.5~keV and also significantly around the Si~K~edge, but the on-board calibration sources currently in use typically have strong lines at only two energies, 1.5~keV (Al~K$\\alpha$) and 5.9~keV (Mn~K$\\alpha$). The only option available to the current generation of flight instruments to calibrate any time variable response is to use celestial sources. The missions which have been represented in this work are the {\\em Chandra X-ray Observatory}\\cite{weiss00,weiss02} (CXO), the {\\em X-ray Multimirror Mission}\\cite{jansen2001} (XMM-Newton), the {\\em ASTRO-E2 Observatory} ({\\em Suzaku}), and the {\\em Swift} Gamma-ray Burst Mission. Data from the following instruments have been included in this analysis: the {\\em High-Energy Transmission Grating} (HETG) and the {\\em Advanced CCD Imaging Spectrometer}\\cite{bautz98,garmire03,garmire92} (ACIS) on the CXO, the {\\em Reflection Gratings Spectrometers}\\cite{denherder2001} (RGS), the {\\em European Photon Imaging Camera} (EPIC) {\\em Metal-Oxide Semiconductor}\\cite{turner2001} (MOS) CCDs and the EPIC p-n junction\\cite{strueder2001} (pn) CCDs on XMM-Newton, the {\\em X-ray Imaging Spectrometer} (XIS) on {\\em Suzaku}, and the {\\em X-ray Telescope}\\cite{burrows2005,godet2007} (XRT) on {\\em Swift}. Suitable calibration targets would need to possess the following qualities. The source would need to be constant in time, to have a simple spectrum defined by a few bright lines with a minimum of line-blending, and to be extended so that ``pileup'' effects in the CCDs are minimized but not so extended that the off-axis response of the telescope dominates the uncertainties in the response. Our working group focused on supernova remnants (SNRs) with thermal spectra and without a central source such as a pulsar, as the class of source which had the greatest likelihood of satisfying these criteria. We narrowed our list to the Galactic SNR Cas-A, the Large Magellanic Cloud remnant N132D and the Small Magellanic Cloud remnant \\name\\/. We discarded Cas-A since it is a relatively young ($\\sim350$~yr) SNR with significant brightness fluctuations in the X-ray, radio, and optical over the past three decades, it contains a faint (but apparently constant) central source, and it is relatively large (radius $\\sim3.5$~arcminutes). We discarded N132D because it has a complicated, irregular morphology in the X-rays and its spectrum shows strong, complex Fe emission. We therefore settled on \\name\\ as the most suitable source given its relatively uniform morphology, small size (radius $\\sim0.4$~arcminutes), and comparatively simple X-ray spectrum. \\begin{figure}[h] \\begin{center} \\begin{tabular}{c} \\includegraphics[width=3.0in,angle=0]{4panel.ps} \\includegraphics[width=3.0in,angle=0]{xis_panel.ps} \\end{tabular} \\end{center} \\vspace{-0.15in} \\caption[image] { \\label{fig:image} Images of E0102 from ACIS S3 (top left), MOS(bottom left), XRT(middle top), pn(middle bottom), and XIS(right). The black circles indicate the extraction regions used for the spectral analysis. The high angular resolution of the CXO's mirrors are evident in the structure resolved in the SNR and the small extraction region.} \\end{figure} The SNR \\name\\ (hereafter E0102) was discovered by the {\\em Einstein Observatory}\\cite{seward1981}. It is the brightest SNR in X-rays in the {\\em Small Magellanic Cloud} (SMC). E0102 has been extensively imaged by CXO\\cite{gaetz2000,hughes2000} and XMM-Newton\\cite{sasaki2001}. Figure~\\ref{fig:image} shows an image of E0102 with the relevant spectral extraction region for each of the instruments included in this analysis. E0102 is classified as an ``O-rich'' SNR and has an estimated age of $\\sim1,000$~yr. The source diameter is small enough such that a high resolution spectrum may be acquired with the HETG on the CXO and the RGS on XMM-Newton. The HETG spectrum\\cite{flanagan2004} and the RGS spectrum\\cite{rasmussen2001} both show strong lines of O, Ne, and Mg with little or no Fe. E0102's spectrum is relatively simple compared to a typical SNR spectrum. Figure~\\ref{fig:rgs_spec_lin} displays the RGS spectrum from E0102. The strong, well-separated lines in the energy range 0.5 to 1.5~keV make this source a useful target for calibration observations. The source is extended enough to reduce the effects of photon pileup, which distorts a spectrum. Although moderate pileup is expected in all the non-grating instruments when observed in modes with relatively long frame times. The source is also bright enough to provide a large number of counts in a relatively short observation. \\begin{figure}[h] \\begin{center} \\begin{tabular}{c} \\includegraphics[width=4.9in,angle=270]{rgspn_mod_tbabs_tbvarabs_2apec_line_ratios_jd_v1.9_rgs_lin_ratio_v2.ps} \\end{tabular} \\end{center} \\vspace{-0.15in} \\caption[image] { \\label{fig:rgs_spec_lin} RGS1/2 spectrum of E0102 from a combination of 23 observations. Note the bright lines of O, Ne, and Mg. } \\end{figure} ", "conclusions": "\\label{sect:conclusions} We have use the line-dominated spectrum of the SNR E0102 to test the response models of the ACIS S3, MOS, pn, XIS, and XRT CCDs below 1.5~keV. We have fitted the spectra with the same model in which the continuum and absorption components and the weak lines are held fixed while allowing only the normalizations of four bright lines/line complexes to vary. We have compared the fitted line normalizations of the O~{\\small VII}~For line, the O~{\\small VIII}~Ly~$\\alpha$ line, the Ne~{\\small IX}~Res line, and Ne~{\\small X}~Ly~$\\alpha$ line to examine the consistency of the effective area models for the various instruments in the energy ranges around 570~eV, 654~eV, 915~eV, and 1022~eV. We find that the instruments are in general agreement with 28 of the 32 scaled normalizations within $\\pm10\\%$ of the RGS determined values. We find that the scaled line normalizations agree to within 23\\%, 24\\%, 13\\%, \\& 19\\% for O~{\\small VII}, O~{\\small VIII}, Ne~{\\small IX}, \\& Ne~{\\small X} when all instruments are considered. When only CXO and XMM-Newton are considered, we find that the fitted line normalizations agree to within 22\\%, 16\\%, 4\\%, \\& 12\\% for O~{\\small VII}, O~{\\small VIII}, Ne~{\\small IX}, \\& Ne~{\\small X}. \\begin{figure}[tbh] \\begin{center} \\begin{tabular}{c} \\includegraphics[width=3.5in,angle=270]{mos1-spie.ps} \\end{tabular} \\end{center} \\vspace{-0.15in} \\caption[MOS1] { \\label{fig:mos1_spec} MOS1 spectrum from OBSIDs 0123110201 and 0135720601. Note the excellent spectral resolution of the MOS data. } \\end{figure} \\begin{figure}[bth] \\begin{center} \\begin{tabular}{c} \\includegraphics[width=3.5in,angle=270]{rgspn_mod_tbabs_tbvarabs_2apec_line_ratios_jd_v1.9_P0412980301PNS001_log_lin_ratio.ps} \\end{tabular} \\end{center} \\vspace{-0.15in} \\caption[pn] { \\label{fig:pn_spec} pn spectrum from OBSID0412980301 . The second (lower) curve shows the same data but with a linear axis which has been shifted downwards for clarity. Note the high count rate and the pattern in the residuals which might indicate an issue with the spectral redistribution function. } \\end{figure} \\begin{figure}[tbh] \\begin{center} \\begin{tabular}{c} \\includegraphics[width=3.5in,angle=270]{iachec08c_xis1_fittedspec.ps} \\end{tabular} \\end{center} \\vspace{-0.15in} \\caption[pn] { \\label{fig:xis_spec} XIS1 spectrum from OBSID 100044010 showing the model for the XRB RX~J0103. Note the residuals below 0.5~keV and above 1.5~keV.} \\end{figure} \\begin{figure}[bth] \\begin{center} \\begin{tabular}{c} \\includegraphics[width=3.5in,angle=270]{e0102_xrt_pc_g0to12.ps} \\end{tabular} \\end{center} \\vspace{-0.15in} \\caption[pn] { \\label{fig:xrt_spec} Combined {\\em Swift} XRT spectrum from four observations. Note the excess from 1.5 to 2.5 keV, this is a signature of pileup. } \\end{figure} \\vspace*{-0.1in}" }, "0807/0807.2340_arXiv.txt": { "abstract": "{Recent observations of the Galactic center revealed a nuclear disk of young OB stars near the massive black hole (MBH), in addition to many similar outlying stars with higher eccentricities and/or high inclinations relative to the disk (some of them possibly belonging to a second disk). In addition, observations show the existence of young B stars (the 'S-cluster') in an isotropic distribution in the close vicinity of the MBH ($<0.04$ pc). We use extended N-body simulations to probe the dynamical evolution of these two populations. We show that the stellar disk could have evolved to its currently observed state from a thin disk of stars formed in a gaseous disk, and that the dominant component in its evolution is the interaction with stars in the cusp around the MBH. We also show that the currently observed distribution of the S-stars could be consistent with a capture origin through 3-body binary-MBH interactions. In this scenario the stars are captured at highly eccentric orbits, but scattering by stellar black holes could change their eccentricity distribution to be consistent with current observations. ", "introduction": "High resolution observations have revealed the existence of many young OB stars in the galactic center (GC). Accurate measurements of the orbital paramters of these stars give strong evidence for the existence of a massive black hole (MBH) which govern the dynamics in the GC \\citep{sch+02b,ghe+03a}. Most of the young stars are observed in the central 0.5 pc around the MBH. The young stars population in the inner 0.04 pc (the 'S-stars' or the 'S-cluster') contain only young B-stars, in apparently isotropic distribution around the MBH, with relatively high eccentricities ($0.3\\le e\\le0.95$) \\citep{ghe+03a,eis+05}. The young stars outside this region contain many O-stars residing in a stellar disk moving clockwise in the gravitational potential of the MBH \\citep{lev+03,gen+03a,lu+06,pau+06,tan+06}. The orbits of the stars in this disk have average eccentricity of $\\sim0.35$ and the opening of the disk is $h/R\\sim0.1$, where $h$ is the disk height and $R$ is its radius. \\citet{pau+06} and \\citet{tan+06} suggested the existence of another stellar disk rotating counter clockwise and almost perpendicular to the CW disk. This disk is currently debated as many of the young stars are have intermediate inclinations, and are possibly just outliers that do not form a coherent disk structure \\citep{lu+06}. Here we briefly report on our study of the dynamical evolution of the young stars in the GC, both in the stellar disk and in the S-cluster. We use extensive N-body simulations with realistic number of stars ($10^{3}-10^{5}$) using the {\\tt gravitySimulator}, a 32-node cluster at the Rochester Institute of Technology that incorporates GRAPE accelerator boards in each of the nodes \\citep{har+07}. Thus we are able to probe the dynamics of the stars near the MBH and their stellar environment. We study two basic issues: (1) the long term evolution of the S-stars up to their lifetime of a few $10^{7}$ yrs, including their dynamical interaction with stars in the vicinity of the MBH; (2) The evolution of a realistic stellar disk, taking into account both the effects of non-equal mass stars, as studied earlier, and more importantly the effect of the interactions of disk stars with the stellar cusp around the MBH. As we show, the latter component proves to be more important than the other components discussed in previous studies. A detailed report of our complete set of out simulations (not shown here), in addition to analytic calculations will be presented in upcoming papers (Perets et al., in preparation) ", "conclusions": "\\label{sec:summary} The dynamical evolution of the young stars in the GC both in the stellar disk(s) and in the inner S-cluster is not yet understood. We used N-body simulations to study the dynamics and origin of these stars. We found that the S-stars close to the MBH in the GC could be stars that were captured following a binary disruption by the MBH, and later on dynamically evolved due to scattering by other stars, or stellar black holes, to obtain their currently observed orbits. We also show the the young stellar disk could have formed as a cold (thin) circular disk and evolve to its currently observed thick (hot) disk, mostly due to scattering by cusp stars, whereas self relaxation of the disk plays a more minor role, especially in regard to the more massive stars seen in observations." }, "0807/0807.0729_arXiv.txt": { "abstract": "% This first Subaru international conference has highlighted the remarkably diverse and significant contributions made using the 8.2m Subaru telescope by both Japanese astronomers and the international community. As such, it serves as a satisfying tribute to the pioneering efforts of Professors Keiichi Kodaira and Sadanori Okamura whose insight and dedication is richly rewarded. Here I try to summarize the recent impact of wide field science in extragalactic astronomy and cosmology and take a look forward to the key questions we will address in the near future. ", "introduction": "% It's a pleasure to summarize the first Subaru international conference. The conference has been striking in many ways. Before I arrived in Hayama, I was impressed by one of the two conference posters: the woodblock print (Ukiyo-e) `The Great Wave off Kanagawa\" by the artist Katsushika Hokusai (1760-1849). This portrays desperate fishermen struggling with a tsunami against the backdrop of Mt Fuji. I thought this seemed an alarming image to use to encourage overseas visitors to come to the meeting, particularly given the proximity of the conference venue to the ocean and the location depicted! Fortunately, I'm glad to report our views of Mt Fuji and the ocean have been peaceful and serene throughout our stay. The second striking thing about the meeting is the cause for celebrating the achievements of the Subaru telescope. Although we've heard panoramic results from many observatories and facilities, it's clear much of the focus has been with Subaru. We're congratulating Professor Keiichi Kodaira for his leadership in bringing this remarkable facility to fruition, and Professor Sadanori Okamura and his colleagues for their insight in developing the remarkable prime focus camera, Suprime-Cam. Subaru remains the only fully-steerable 8 meter class telescope with a prime focus imager; we've seen ample evidence of the wisdom of this decision. With further instruments such as MOIRCS, FMOS and hopefully eventually WFMOS, the Subaru telescope is destined to remain an unique facility in wide-field astronomy. In an interesting presentation, Iye (2007) examined a recent article by Grothkopf et al (2007) which analyzed the Hirsch (2005) $h$-index\\footnote{The $h$-index is the number $n$ of refereed articles each of which has more than $n$ citations} of the four major 8-10m telescopes. Subaru ranks third in this analysis (behind Keck and the VLT) with a $h$-index of 40. However, as citations are cumulative, older facilities naturally garner more citations. The $m$-index takes this into account by dividing the number of citations by the years of routine operation. Grothkopf et al find that when this is done, the newer VLT is having a comparable impact to Keck, both have $m\\simeq$10, but Iye was dismayed to find Subaru remains third with $m\\simeq$6. Not one to be defeated easily, Iye invented the $i$-index which is the $m$-index divided by the number of telescopes in the observatory. As the VLT comprises 4 independent telescopes, and Keck and Gemini comprise two each, Subaru is clearly disadvantaged by this factor. When this correction is made, as a single telescope, Iye was pleased to find Subaru leads the pack! Even if citations are never a fully-accurate guide of the discovery rate of a telescope, I think we'd all agree that Subaru has, through its unique wide field capabilities, contributed enormously to the growth and international standing of Japanese astronomy. ", "conclusions": "" }, "0807/0807.0203_arXiv.txt": { "abstract": "{ We have acquired a deep i-band image of the BL Lacertae object \\object{S5 0716+714} while the target was in an low optical state. Due to the faintness of the nucleus, we were able to detect the underlying host galaxy. The host galaxy is measured to have an I-band magnitude of 17.5 $\\pm$ 0.5 and an effective radius of (2.7 $\\pm$ 0.8) arcsec. Using the host galaxy as a ``standard candle'', we derive z = $0.31 \\pm 0.08$ (1$\\sigma$ error) for the host galaxy of \\object{S5 0716+714}. This redshift is consistent with the redshift z = 0.26 determined by spectroscopy for 3 galaxies close to \\object{S5 0716+714}. The effective radius at z = 0.31 would be 12 $\\pm$ 4 kpc, which is consistent with values obtained for BL Lac host galaxies. An optical spectrum acquired during the same epoch shows no identifiable spectral lines. } ", "introduction": "BL Lacertae objects (BL Lacs) are active galactic nuclei (AGN) characterized by a featureless nonthermal continuum, high optical and radio polarization, and variability across the entire electromagnetic spectrum. These properties arise from a relativistic jet pointed almost towards the observer \\citep{1979ApJ...232...34B}. The strong continuum emission from the jet in the optical sometimes presents a problem when attempting to properly characterize the underlying host galaxy or measure its redshift, especially since by definition the emission lines are very weak in BL Lacs. Hence, it is unsurprising that even after very deep searches \\citep[e.g.][]{2006AJ....132....1S} many BL Lacs remain without a spectroscopically determined redshift. The redshift is a fundamental property of any extragalactic object, without which e.g. the luminosity or intrinsic variability properties of the target cannot be determined. Furthermore, many BL Lacs have been recently detected at TeV gamma-rays using ground-based Cherenkov telescopes MAGIC, HESS, and VERITAS \\citep[e.g.][]{2007arXiv0712.3352H}. Since VHE gamma-rays can be absorbed by the interaction with low energy photons of the EBL via pair production, this opens the possibility of measuring the amount of extragalactic background light \\citep[EBL, ][]{nikishov1962,1992ApJ...390L..49S} in the optical through to the far infrared, which provides important information about the galaxy and star formation history. The absorption depends strongly on the distance of the source and the energy of the gamma-rays. If the redshift of the source is known, the VHE gamma-ray spectrum can be used to derive limits on the EBL \\citep[e.g.][]{2006Natur.440.1018A,2007A&A...471..439M}. On the other hand, if the redshift is unknown, the VHE gamma-ray spectra can be used to set an upper limit to the redshift of the source \\citep[e.g.][]{2007ApJ...655L..13M}. After its discovery as a bright (S$_{\\rm 5 Ghz}$ $>$ 1 Jy) flat-spectrum ($\\alpha \\leq -0.5$, S$_{\\nu} \\propto \\nu^{\\alpha}$) radio source, the BL Lac object \\object{S5 0716+714} has been studied intensively at all frequency bands. The object is highly variable with rapid variations observed from radio to X-ray bands \\citep{1996AJ....111.2187W,2006A&A...451..797O}. The nucleus of \\object{S5 0716+714} is typically bright in the optical \\citep[see e.g.][]{2005AJ....130.1466N}, thus previous attempts to either characterize its host galaxy \\citep{1993A&AS...98..393S, 2000ApJ...532..816U, 2002A&A...381..810P} or to determine its redshift spectroscopically \\citep{1993A&AS...98..393S, 1996MNRAS.281..425M} have not been successful. Thus, in spite of numerous variability studies of \\object{S5 0716+714}, it has never been possible to determine reliably e.g. the linear dimensions and luminosities of the varying components. \\object{S5 0716+714} has also been detected in TeV gamma-rays \\citep{atel1500}. Given that previous estimates \\citep[e.g.][]{1996AJ....111.2187W} place \\object{S5 0716+714} at z $>$ 0.3, this would make the target one of the most distant TeV sources detected so far. Given the obvious importance of \\object{S5 0716+714} to EBL studies, it is be important to secure an accurate spectroscopic redshift for \\object{S5 0716+714} or at least constrain the redshift by some other means. \\object{S5 0716+714} is regularly monitored at Tuorla Observatory as part of the blazar monitoring program\\footnote{http://users.utu.fi/$\\sim$kani/1m/index.html}. On December 17, 2007 we observed \\object{S5 0716+714} to go into a fairly deep optical minimum (R $\\sim$ 14.8) and initiated prompt i-band imaging at the Nordic Optical Telescope (NOT) to detect its host galaxy. The imaging was performed five days after the minimum and the results are reported in this Letter. Throughout this Letter, we use the cosmology $H_0 = 70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{M}$ = 0.3 and $\\Omega_{\\Lambda}$ = 0.7. ", "conclusions": "A deep i-band image of the BL Lacertae object \\object{S5 0716+714} acquired at the Nordic Optical Telescope while the target was in an low optical state has enabled us to detect the underlying host galaxy. The host galaxy has an I-band magnitude of 17.5 $\\pm$ 0.5 and an effective radius of (2.7 $\\pm$ 0.8) arcsec. Using the host galaxy as a ``standard candle'' and a method proposed by \\cite{2005ApJ...635..173S}, we derive a redshift of z = $0.31 \\pm 0.08$ (1$\\sigma$ error) for the host galaxy of \\object{S5 0716+714}. This redshift is consistent with the redshift z = 0.26 determined by spectroscopy for 3 galaxies close to \\object{S5 0716+714}. The corresponding effective radius of the host galaxy at z = 0.31 would be $12 \\pm 4$ kpc, also consistent with the values obtained for BL Lac host galaxies in other studies. An optical spectrum obtained at the same epoch shows a featureless continuum with no identifiable spectral lines." }, "0807/0807.4036_arXiv.txt": { "abstract": "The Dark UNiverse Explorer (DUNE) is a wide-field imaging mission concept whose primary goal is the study of dark energy and dark matter with unprecedented precision. To this end, DUNE is optimised for weak gravitational lensing, and also uses complementary cosmological probes, such as baryonic oscillations, the integrated Sachs-Wolf effect, and cluster counts. Immediate additional goals concern the evolution of galaxies, to be studied with groundbreaking statistics, the detailed structure of the Milky Way and nearby galaxies, and the demographics of Earth-mass planets. DUNE is a medium class mission consisting of a 1.2m telescope designed to carry out an all-sky survey in one visible and three NIR bands (1deg$^2$ field-of-view) which will form a unique legacy for astronomy. DUNE has been selected jointly with SPACE for an ESA Assessment phase which has led to the Euclid merged mission concept. ", "introduction": "\\label{intro} Dark energy and dark matter comprise the bulk of the mass-energy budget of the Universe and pose some of the most fundamental questions in physics. The Dark UNiverse Explorer\\footnote{{\\tt http://www.dune-mission.net}} (DUNE) is a wide field mission concept designed to study these dark cosmic components with unprecedented precision. To do so, DUNE will use weak gravitational lensing along with other cosmological probes. In these proceedings, we give a brief summary of the DUNE mission concept which was recently proposed to ESA's Cosmic Vision programme. A description of the focal plane instrumentation can be found in an adjoining paper\\cite{jeffmark} and a more detailed description of the ESA proposal is provided in a previous publication\\cite{ref08}. An earlier and simpler version of DUNE was described in previous SPIE Proceedings\\cite{gr06,ref06}. Gravitational deflection of light by intervening dark matter concentrations causes the images of background galaxies to acquire an additional ellipticity of order of a percent, which is correlated over scales of tens of arcminutes. Utilisation of this cosmological probe relies on the measurement of image shapes and redshifts for several billion galaxies, both requiring space observations for PSF stablity and photometric measurements over a wide wavelength range in the visible and near-IR (NIR). Furthermore, in order to break as many degeneracies as possible, and to provide independent constraints, complementary approaches should be used. DUNE has thus been designed to provide three additional cosmological probes : Baryon Acoustic Oscillations (BAO), the Integrated Sachs-Wolfe effect (ISW), and galaxy Cluster Counts (CC). It is therefore a unique mission to probe the dark Universe in different independent ways. DUNE will tackle the following questions: What are the dynamics of dark energy? What are the physical characteristics of the dark matter? What are the seeds of structure formation and how did structure grow? Is Einstein's theory of General Relativity the correct theory of gravity? DUNE will combine its unique space-borne observations with existing and planned ground-based surveys, and hence greatly increases the science return of the mission while limiting costs and risks. The panoramic visible and NIR surveys required by DUNE's primary science goals will afford unequalled sensitivity and survey area for further studies. Additional surveys at low galactic latitudes and in deep patches of the sky will open new scientific windows. DUNE will explore the nature of Dark Matter by measuring precisely the sum of the neutrino masses and by testing the Cold Dark Matter paradigm. It will test the validity of Einstein's theory of gravity. In addition, DUNE will investigate how galaxies form, survey all Milky-Way-like galaxies in the 2$\\pi$ extra-galactic sky out to $z \\sim 1$ and detect thousands of galaxies and AGN at $61$ are known, DUNE will find hundreds of Virgo-cluster-mass objects at $z>2$, and several thousand clusters of M=$1-2 \\times 10^{13}$Mo; the latter are the likely environments in which the peak of QSO activity at $z\\sim2$ takes place, and will hold the empirical key to understanding the peak period of QSO activity. Using the Lyman-dropout technique in the near-IR, a deep survey (DUNE-MD) will be able to detect the most luminous objects in the early Universe ($z>6$): $\\sim 10^4$ star-forming galaxies at $z\\sim8$ and up to $10^3$ at $z\\sim10$, for SFRs $>30-100$Mo/yr. It will also be able to detect significant numbers of high-$z$ quasars: up to $10^4$ at $z\\sim7$, and $10^3$ at $z\\sim9$. By applying the Gunn-Peterson test to this large statistically relevant sample of objects we will put stringent constraints on the end of the period of reionisation of the Universe. DUNE will also detect a very large number of strong lensing systems: about $10^5$ galaxy-galaxy lenses, $10^3$ galaxy-quasar lenses and 5000 strong lensing arcs in clusters\\cite{meneg07}. It is also estimated that several tens of galaxy-galaxy lenses will be double Einstein rings\\cite{gav08}, which are powerful probes of the cosmological model as they simultaneously probe several redshifts. In addition, during the course of the DUNE-MD survey (over 6 months), we expect to detect $\\sim 3000$ Type Ia Supernovae with redshifts up to $z\\sim0.6$. This will yield a measurement of the SN rate with unprecedented accuracy, thus providing information on their progenitors. This survey will also discover a comparable number of Core Collapse SNe (Types II and Ib/c), out to $z\\sim0.3$, whose rate provides an independent measurement of the star formation history. \\subsection{Studying the Milky Way} DUNE will also leas to breakthroughs in Galactic astronomy. The extragalactic survey, DASS-EX, complemented by the shallower survey of the Galactic plane (DASS-G with $|b|<30$) will provide all-sky high resolution (0.23'' in the wide red band, and 0.4'' in YJH) deep imaging of the stellar content of the Galaxy, allowing the deepest detailed structural studies of the thin and thick disk components, the bulge/bar, and the Galactic halo (including halo stars in nearby galaxies such as M31 and M33) in bands which are relatively insensitive to dust in the Milky Way. DUNE will be little affected by extinction and will supersede all of the ongoing surveys in terms of angular resolution and sensitivity (photometric depth and low background). DUNE will thus enable the most comprehensive stellar census of late-type dwarfs and giants, brown dwarfs, He-rich white dwarfs, along with detailed structural studies, tidal streams and merger fragments. DUNE's sensitivity will also open up a new discovery space for rare stellar and low-temperature objects via its H-band imaging. Studying the Galactic disk components requires the combination of spatial resolution (crowding) and dust-penetration (H-band) that only DUNE can deliver. It will also yield the most detailed and sensitive survey of structure and substructure in nearby galaxies especially of their outer boundaries, thus constraining merger and accretion histories. \\subsection{Search for Extra-Solar Planets} Using the microlensing effect, DUNE can provide a statistical census of exoplanets in the Galaxy with masses over $0.1 M_\\oplus$ from orbits of 0.5 AU from their parent star to free-floating objects. This includes analogues to all the solar system's planets except for Mercury, as well as most planets predicted by planet formation theory. Microlensing is the temporary magnification of a galactic bulge source star by the gravitational potential of an intervening lens star passing near the line of sight. A planet orbiting the lens star, will have an altered magnification, showing a brief flash or a dip in the observed light curve. Because of atmospheric seeing, and poor duty cycle even using networks, ground-based microlensing surveys are only able to detect a few to 15 $M_\\oplus$ planets in the vicinity of the Einstein ring radius (2-3 AU). A dedicated survey (DUNE-ML), using the high angular resolution of DUNE, and the uninterrupted visibility and NIR sensitivity afforded by space observations will provide detections of microlensing events using as sources G and K bulge dwarfs stars and therefore can detect planets down to $0.1-1 M_\\odot$ from orbits of 0.5 AU. Moreover, there will be a very large number of transiting hot Jupiters detected towards the galactic bulge as a free ancillary science. ", "conclusions": "\\begin{table} \\begin{center} \\caption{DUNE Baseline summary} \\label{baseline} \\begin{tabular}{|l|l|} \\hline Science objectives & Cosmology and Dark Energy\\\\ & Galaxy formation, Extra-solar planets\\\\ \\hline Surveys & 20,000 deg$^2$ extragalactic 20,000 deg$^2$ galactic\\\\ & 100 deg$^2$ medium-deep, 4 deg$^2$ planet hunting\\\\ \\hline Requirements & 1 visible band (R+I+J) for high-precision\\\\ & shape measurements\\\\ & 3 NIR bands (Y, J, H) for photometry\\\\ \\hline Payload & 1.2m telescope, Visible \\& NIR cameras \\\\ & with 0.5 deg$^2$ FOV each\\\\ \\hline Service module & Mars/Venus express, Gaia heritage \\\\ \\hline Spacecraft & 2013kg launch mass\\\\ \\hline Orbit & Geosynchronous\\\\ \\hline Launch & Soyuz S-T Fregat\\\\ \\hline Operations & 4 year mission\\\\ \\hline \\end{tabular} \\end{center} \\end{table} The DUNE mission concept can be seen as the next step in precision cosmology. ESA's Planck mission will bring unprecedented precision to the measurement of the high redshift Universe. This will leave the dark energy dominated low redshift Universe as the next frontier in high precision cosmology. Constraints from the radiation perturbation in the high redshift CMB, probed by Planck, combined with density perturbations at low redshifts, probed by DUNE, will form a complete set for testing all sectors of the cosmological model. In this respect, a DUNE+Planck programme can be seen as the next significant step in testing, and thus challenging, the standard model of cosmology. DUNE will offer high potential for ground-breaking discoveries of new physics, from dark energy to dark matter, initial conditions and the law of gravity. DUNE will (i) measure both effects of dark energy by using weak lensing as the central probe; (ii) place this high precision measurement of dark energy within a broader framework of high precision cosmology by constraining all sectors of the standard cosmology model (dark matter, initial conditions and Einstein gravity); (iii) through a collection of unique legacy surveys be able to push the frontiers of galaxy evolution and the physics of the local group; and finally (iv) be able to obtain information on extrasolar planets, including Earth analogues. The DUNE concept has been recently proposed to ESA's Cosmic Vision programme and has been selected jointly with SPACE\\cite{space} for an ongoing ESA Assesment Phase which has led to the merged {\\it Euclid} mission concept." }, "0807/0807.1049_arXiv.txt": { "abstract": "{When low-mass stars form, the collapsing cloud of gas and dust goes through several stages which are usually characterized by the shape of their spectral energy distributions. Such classification is based on the cloud morphology only and does not address the dynamical state of the object.} {In this paper we investigate the initial cloud collapse and subsequent disk formation through the dynamical behavior as reflected in the sub-millimeter spectral emission line profiles. If a young stellar object is to be characterized by its dynamical structure it is important to know how accurately information about the velocity field can be extracted and which observables provide the best description of the kinematics. Of particular interest is the transition from infalling envelope to rotating disk, because this provides the initial conditions for the protoplanetary disk, such as mass and size.} {We use a hydrodynamical model, describing the collapse of a core and formation of a disk, to produce synthetic observables which we compare to calculated line profiles of a simple parameterized model. Because we know the velocity field from the hydrodynamical simulation we can determine in a quantitative way how well our best-fit parameterized velocity field reproduces the original. We use a molecular line excitation and radiation transfer code to produce spectra of both our hydrodynamical simulation as well as our parameterized model.} {We find that information about the velocity field can reasonably well be derived by fitting a simple model to either single-dish (15$''$ resolution) lines or interferometric data (1$''$ resolution), but preferentially by using a combination of the two. The method does not rely on a specific set of tracers, but we show that some tracers work better than others. Our result shows that it is possible to establish relative ages of a sample of young stellar objects using this method, independently of the details of the hydrodynamical model.} {} ", "introduction": "Low-mass young stellar objects (YSOs) are well-studied phenomena. Observationally, these objects have been measured in many different wavelength regimes in order to determine intrinsic properties such as their mass, luminosity, mass accretion rate, etc. YSOs are traditionally classified by the shape of their spectral energy distribution (SED)~\\citep{lada1984}, which reflects their general morphology. Subsequently, the envelope and disk densities, temperatures and chemistry can be studied using millimeter continuum and molecular line measurements. All of these derived quantities have been used to piece together a consistent chronology of the process of star formation and to establish a reference from which the age of a YSO can be reliably obtained~\\citet{mardones1997,gregersen2000}. A theoretical evolution scenario of YSOs has been established for many years \\citep[e.g.,][]{ulrich1976,shu1977,cassen1981,terebey1984,bodenheimer1990, basu1998,yorke1999}. In these models (with the exception of the Shu model, which is spherically symmetric) a circumstellar disk is formed, after the collapse of the initial cloud sets in, due to conservation of angular momentum. It is therefore natural to use the kinematic properties of the models to characterize the evolutionary stage of the object, because the distribution of angular momentum changes in a monotonous way as the cloud contracts and matter spins up. Unfortunately, measuring the velocity field observationally is not such a simple matter. Kinematic information can only be derived from spectral emission lines and there is no way to directly solve for the velocity field from measurements of such lines. The reason for this is that the spectral profile depends on a range of physical parameters, including the local density, temperature, molecular abundance, and turbulence in the region where the transition in consideration is excited. All of this can be taken into consideration and self-consistent models, including a general parameterized velocity field, can be built and fitted to the observed spectra. The question is, however, how reliable the derived velocity field is given the input model, and to what extent the derived velocity fields are consistent with the prediction of the models. The velocity field of YSOs has been well studied observationally in the last 20 years. At first, mainly infall were observed \\citep[e.g.,][]{calvet1992, vanlangevelde1994}, but soon afterward, objects showing a mix of infall and rotation was discovered \\citep{Saito1996,Ohashi1997,Hogerheijde2001, belloche2002}. More recently, protoplanetary disks have been studied in very high resolution using sub-millimeter interferometry and many of these have velocity fields with rotation only \\citep[e.g.,][]{Lommen2008}. The common interpretation of these observations is that the objects showing infall only are early-type embedded objects, the rotating disks are the end product of the collapse, and the objects showing both infall and rotation are in some kind of transition. Initial modeling work using a parameterization for the velocity field has been done by \\citet{myers1996} for low-mass star formation and by \\citet{keto1990} for the formation of high-mass stars. However, it has never been attempted to quantify the degree of the transition, how far the object has passed from the embedded phase to the protoplanetary disk phase. In this paper we will address this question, by comparing synthetic spectra, calculated from a hydrodynamical simulation of a collapsing cloud, to a model with a simple parameterized velocity field. The comparison is done in spectral space in the same way as one would compare a model to real observations. By considering different transitions of different molecules in two different resolution settings, we investigate how to obtain the most reliable characterization of the velocity field. The importance of being able to reliably model the observations of YSOs using a simple generic model becomes evident when considering the capabilities of the upcoming ALMA (Atacama Large Millimeter Array). While it is possible to build sophisticated custom models, or even directly use the hydrodynamical simulation as a template model to describe single objects, this is a very time-consuming process. ALMA will produce very high resolution observations ($\\sim$ 0.01$''$) in snapshot mode and in order to deal with such vast samples of YSOs, simple yet reliable models are needed. The layout of this paper is as follows: In Sect.~\\ref{model} we present the hydrodynamical simulation of~\\citet{yorke1999} that we use to make synthetic observations and describes our model and the way we fit this to the synthetic spectra. Section~\\ref{lines} presents the synthetic observables and also a few analytical examples of idealized cases for comparison. Section~\\ref{results} shows our results of the spectral comparison and the derived velocity fields, followed by a discussion and conclusion in Sects. \\ref{discussion} and \\ref{conclusion}, respectively. ", "conclusions": "We have explored the possibility of placing Young Stellar Objects in an evolutionary ordering based on their kinematical configurations by fitting a simple parameterized model to sub-millimeter observations. We have tested the feasibility of this by fitting the simple velocity model to a time series of synthetic observations with known velocity fields, generated by a hydrodynamical simulation. We find that the model reproduces the synthetic spectra reasonably well and that the best fit parameters which describe the velocity field are in agreement with the velocity field in the hydrodynamical simulation. We therefore conclude that it is possible to extract reliable information on astronomical objects if we replace the synthetic spectra with real observations. We find that it is difficult, though not impossible, using single-dish data alone, due to the shallowness of the $\\chi^2$-space, but feasible if interferometric data are used, especially when combined with one or more single-dish lines. Using the latter option, we find that the central dynamical mass can be determined to within 20\\%. However, we have not been able to establish a single ``best way'' to obtain the parameters throughout the collapse. We find that the best result is obtained by trying several different methods to constrain the fits, and then pick the result of the one which shows the most peaked $\\chi^2$-space. That said, most of our best constrained points were obtained using a combination of low-resolution and high-resolution and this is the preferred starting point. We further conclude that the molecular species used are not crucial for the result, as long as the line wings are clearly defined and uncontaminated in the case of the single-dish lines, and also that the transition is optically thick. For the high-resolution data, signal-to-noise is the main concern, so strong lines are preferred (CO, HCO$^+$, etc.). Due to the uncertainty of the absolute time scale of the collapse calculated by our relatively simple hydrodynamical simulation, we cannot yet establish an absolute age of a given object, but relative ages among a sample of objects can be obtained, provided that the velocity field of a collapsing cloud evolves similar to that of the simulation. Indeed, if an age calibration could be made, for example using the chemical properties of one or more objects, this could prove a powerful method to place a large number of young stars in an evolutionary sequence, because of the relatively straight-forward observations needed to perform the analysis. The hydrodynamical simulation presented in this paper is not required in order to analyze real objects using our method. However, with an improved and more realistic hydrodynamical scheme, it might be possible to calibrate the evolution of the velocity field so that an absolute time scale of star formation can be established. \\\\ \\noindent \\emph" }, "0807/0807.0423_arXiv.txt": { "abstract": "We demonstrate the feasibility of uncovering supermassive black holes in late-type, ``quiescent'' spiral galaxies by detecting signs of very low-level nuclear activity. We use a combination of x-ray selection and multi-wavelength follow-up. Here, we apply this technique to NGC 3184 and NGC 5457, both of type Scd, and show that strong arguments can be made that both host AGNs. ", "introduction": "As discussed in Mathur et al.\\ (these Proceedings) and in \\cite{gmff08}, the only way to detect low-mass supermassive black holes (SMBHs) may be by their accretion activity. We choose to use x-ray selection to identify candidate AGNs for the following reasons: First, x-ray emission is a hallmark of AGNs. Second, x-rays can penetrate obscuring material which may be hiding the line emitting regions. Third, there are fewer sources of x-rays in a galaxy than there are of optical and UV emission and so dilution of the AGN signature by host galaxy light is less of a problem, even if the AGN is moderately obscured. Fourth, even if, as expected in some theories \\citep{n00,l03} AGNs that have luminosities or accretion rates below a cut-off value do not have broad-line regions, they should still be detectable in x-rays. The disadvantage is that x-ray observations by themselves cannot always distinguish between AGNs and other x-ray sources, such as x-ray binaries (XRBs) and ultraluminous x-ray sources (ULXs). Multi-wavelength data are needed to determine the type of source. As examples, we present here two late-type spiral galaxies, NGC 3184 and NGC 5457 (M101). Their nuclear optical spectra show no signs of AGNs \\cite{hfs97-3}. But using archival \\cxo\\ data and multi-wavelength data from the literature, we show that the sum total of evidence strongly suggests that the galaxies host AGNs. ", "conclusions": "There are compelling, though not conclusive, arguments that NGC 3184 and NGC 5457 actually host AGNs even though neither shows signs of AGNs in their optical spectra. These AGNs are excellent candidates for being low-mass SMBHs, as they reside in low-mass bulges. A similar approach can be used to uncover more candidate low-mass SMBHs. \\begin{theacknowledgments} We are grateful to D.~A.~Dale for kindly providing \\textit{Spitzer} fluxes for the nucleus of NGC 3184 prior to publication. Support for this work was provided by the National Aeronautics and Space Administration through Chandra Award Number GO7-8111X issued by the Chandra X-ray Observatory Center, which is operated by the Smithsonian Astrophysical Observatory for and on behalf of the National Aeronautics Space Administration under contract NAS8-03060. \\end{theacknowledgments}" }, "0807/0807.0615_arXiv.txt": { "abstract": "Systematic survey for multiperiodicity in the LMC Cepheids (Moskalik, Ko{\\l}aczkowski \\& Mizerski \\cite{MKM04}, \\cite{MKM06}) has led to discovery of several new forms of pulsational behaviour. One of them is periodic amplitude and phase modulation observed in many first/second overtone (FO/SO) double mode Cepheids. In the current paper we present detailed discussion of this newly discovered phenomenon, based on a combined OGLE+MACHO sample of double mode pulsators. ", "introduction": "We searched for additional signal in the data using a standard consecutive pre\\-white\\-ning technique. To that effect, we first fitted the data with the double frequency Fourier sum representing pulsations in two radial modes: \\vskip 0pt $$m(t) = \\langle m\\rangle + \\sum_{j,k} {\\rm A}_{jk} \\sin [2\\pi(j{\\rm f}_1 + k{\\rm f}_2)t + \\phi_{jk}].\\eqno(1)$$ \\smallskip \\noindent The frequencies of the modes, ${\\rm f}_1$ and ${\\rm f}_2$, were also optimized. The residuals of the fit were then searched for additional periodicities. This was done with the Fourier transform, calculated over the range of $0-5$\\thinspace c/d. In the next step, a new Fourier fit with all frequencies discovered so far was performed and the fit residuals were searched for additional frequencies again. The process was stopped when no new frequencies were detected. ", "conclusions": "" }, "0807/0807.0092_arXiv.txt": { "abstract": "{The discovery of true solar analogues is fundamental for a better understanding of the Sun and of the solar system. Despite a number of efforts, this search has brought only to limited results among field stars. The open cluster M67 offers a unique opportunity to search for solar analogues because its chemical composition and age are very similar to those of the Sun.} {We analyze FLAMES spectra of a large number of M67 main sequence stars to identify solar analogues in this cluster.} {We first determine cluster members which are likely not binaries, by combining proper motions and radial velocity measurements. We concentrate our analysis on the determination of stellar effective temperature, using analyses of line-depth ratios and H$\\alpha$ wings, making a direct comparison with the solar spectrum obtained with the same instrument. We also compute the lithium abundance for all the stars.} {Ten stars have both the temperature derived by line-depth ratios and H$\\alpha$ wings within 100 K from the Sun. From these stars we derive, assuming a cluster reddening $E(B-V)=0.041$, the solar colour $(B-V)_\\odot=0.649\\pm0.016$ and a cluster distance modulus of 9.63. Five stars are most similar (within 60 K) to the Sun and candidates to be true solar twins. These stars have also a low Li content, comparable to the photospheric abundance of the Sun, likely indicating a similar mixing evolution.} {We find several candidates for the best solar analogues ever. These stars are amenable to further spectroscopic investigations and planet search. The solar colours are determined with rather high accuracy with an independent method, as well as the cluster distance modulus.} ", "introduction": "\\label{sec:Intro} The specificity of the Sun and of our solar system have been the subject of active investigation over the last 5 decades. How typical is the Sun for a star of its age, mass, and chemical composition? How typical is that solar-type stars host planetary systems? Are they similar at all to ours? The quest to find stellar analogues to the Sun has been going on for a long time (for an extensive review see, e.g., \\citealt{Cayrel1996}), and it stems from the poor knowledge we have of the Sun when seen `as a star' and from how typical the Sun is for a G2 type star, for its age, chemical composition, population. It is, however, after the discovery of the first exo-planets (\\citealt{MayQue1995}) that this quest became even more compelling, because to find stars similar to our own would allow us to answer to fundamental questions related to the origin of the solar system, the frequency of planetary systems similar to ours, and eventually the formation of life in other exo-planetary systems (\\citealt{Cayrel1996}). The need to identify in the night sky solar proxies to be used for spectroscopic comparison is also diffuse, in particular for the analysis of small solar system bodies (B\\\"ohnhardt, private communication). Among the most recent results in this research, \\cite{Melendez2006} used high resolution, high signal-to-noise ratio Keck spectra to show that HD 98618 is a very close solar twin, and \\cite{King2005} proposed HD 143436 after analyzing 4 stars pre-selected from literature. These stars seem to compare well with the best known solar twin, HR 6060, first analyzed by \\cite{PorSil1997}, and subsequently confirmed by \\cite{SoubTri2004}, who made a comparative study of several hundreds of ELODIE spectra. Finally, \\cite{Melendez2007} have shown HIP~56948 to be the best solar twin known to date both in stellar parameters and in chemical composition, including a low lithium abundance. The open cluster M67 is a perfect target to search for solar analogues. Recent chemical analyses (\\citealt{Tau2000, Randich2006, Pace2008}), show that this cluster has a chemical composition (not only Fe, but also all the other elements) extremely similar to the solar one, as close as allowed by the high precision of the measurements. The analysis resulted in [Fe/H]=$-$0.03$\\pm$0.03 for \\cite{Tau2000}, [Fe/H]=0.03$\\pm$0.01 for \\cite{Randich2006}, and [Fe/H]=0.03$\\pm$0.03 for \\cite{Pace2008}. There are other two additional characteristics which make M67 strategical. The first one is that all the determinations of age give for this cluster an age encompassing that of the Sun (3.5-4.8 Gyr; \\citealt{Yadav2008}), while the age determination for field stars is always uncertain. The second characteristic is that M67 is among the very few clusters showing Li depleted G stars (\\citealt{Pasquini1997}). This is an important point because, as pointed out by \\cite{Cayrel1996}, even if many stars appear to have most characteristics similar to the Sun, their Li abundance is usually 10 times higher than in our star. Since Li is likely an indicator of the complex interaction taking place in the past between the stellar external layers and the hotter interior, the choice of stars which also share the same Li abundance with the Sun is an additional property to pinpoint the true analogues. In our opinion, the search of analogues to the Sun and to the solar system can be well performed in open clusters (OCs), which show a homogeneous age and chemical composition, common birth and early dynamical environment. As a consequence, they provide an excellent laboratory for investigating the physics of solar stars and of planetary system evolution, besides being excellent probes of the structure and evolution of the Galactic disk. M67 is a rich cluster, therefore it provides us with the opportunity to find many stars candidates sharing similar characteristics, and not only one. This is fundamental to obtain some meaningful statistics, and the cluster hosts many main sequence (MS) stars of mass around the solar mass, which form a continuous distribution (Fig.~\\ref{fig:cmd_M67}). Finding several solar analogues in M67 will also help in providing an independent estimate of the solar colors, a quantity which still suffers of some relevant uncertainty (see, e.g., \\citealt{Holmberg2006}), as well as an independent estimate of the distance modulus of the cluster. The present paper is the culmination of a work, which involved the chemical determination of this cluster (\\citealt{Randich2006,Pace2008}), photometry and astrometry (\\citealt{Yadav2008}) to obtain membership, and FLAMES/GIRAFFE high resolution spectroscopy to clean this sample from binaries, and to look for the best solar analogues using the line-depth ratios method (\\citealt{GrayJoha1991,Biazzo2007}) and the wings of the H$\\alpha$ line (\\citealt{CayBen1989}) to determine accurate temperatures with respect to the Sun. In addition, the Li line is used to separate Li-rich from Li-poor stars. \\begin{figure} \\centering \\includegraphics[width=9cm]{cmd_M67.ps} \\caption{Portion of the colour-magnitude diagram of M67. Our selected targets encompass the solar colour, and are high probability proper motion (\\citealt{Yadav2008}) and radial velocity single members. The red points refer to the stars observed with FLAMES/GIRAFFE in three nights.} \\label{fig:cmd_M67} \\end{figure} ", "conclusions": "By using selected observations with FLAMES/GIRAFFE at the VLT, we have made a convincing case that the open cluster M67 hosts a number of interesting potential solar twins, and we have identified them. We have computed spectroscopic accurate effective temperatures for all the stars with two methods. The color-temperature relationships we derive can be used to determine temperatures for MS solar-metallicity stars. By computing the average solar twin colours, we have obtained a precise estimate of the solar $(B-V)$: $(B-V)_{\\odot}=0.649\\pm0.016$. By averaging the magnitude of the solar twins, we have determined an accurate distance modulus for M67: 9.63$\\pm0.06_{\\rm stat}\\pm0.05_{\\rm sys}$, which is in excellent agreement with the most recent estimates, which were based on different, independent methods and data sets. We have determined for all the stars Li abundances, confirming the presence of a large Li spread among the solar stars of this cluster, but showing for the first time, that the Li extra-depletion appears only in stars cooler than 6000 K. The candidate solar twins have Li abundance similar to that of our star, indicating that they also share with the Sun a similar mixing history." }, "0807/0807.4678_arXiv.txt": { "abstract": "We construct a new brane-world model composed of a bulk --with a dilatonic field--, plus a brane --with brane tension coupled to the dilaton, cold dark matter and an induced gravity term. It is possible to show that depending on the nature of the coupling between the brane tension and the dilaton this model can describe the late-time acceleration of the brane expansion (for the normal branch) as it moves within the bulk. The acceleration is produced together with a mimicry of the crossing of the cosmological constant line ($w=-1$) on the brane, although this crossing of the phantom divide is obtained without invoking any phantom matter neither on the brane nor in the bulk. The role of dark energy is played by the brane tension, which reaches a maximum positive value along the cosmological expansion of the brane. It is precisely at that maximum that the crossing of the phantom divide takes place. We also show that these results remain valid when the induced gravity term on the brane is switched off. ", "introduction": "\\label{sec1} This is a very exciting time for cosmology with the overwhelming amount of new observational data that theorists try to explain. One of the biggest puzzles that the theoretical community faces is to explain the recent speed up in the universe rate of expansion discovered first through observations from distant type Ia supernova a decade ago \\cite{Riess:1998cb,Perlmutter:1998np}. This late-time acceleration of the universe has been latter on confirmed by other independent observational probes based for example on measurement of the cosmic microwave background radiation, the clustering of galaxies on very large scales and the baryon acoustic oscillations \\cite{Tegmark:2003ud,Rapetti:2004aa,Spergel:2006hy,Percival:2007yw,Giannantonio:2008zi,Komatsu:2008hk}. A plethora of different theoretical models have been so far proposed to explain this phenomenon \\cite{Copeland}, although unfortunately none of the models advanced so far is both completely convincing and well motivated. A cosmological constant corresponding to roughly two thirds of the total energy density of the universe is perhaps the simplest way to explain the late-time speed up of the universe --and in addition would match rather well the observational data. However, the expected theoretical value of the cosmological constant is about 120 orders of magnitude larger than the value needed to fit the data \\cite{Durrer:2007re}. An alternative approach to explain the late-time acceleration is to invoke an infrared modification of general relativity on large scales which, by weakening the gravitational interaction on those scales, allows the recent speed up of the universal expansion. This approach is also motivated by the fact that we only have precisions observations of gravity from sub-millimiter scales up to solar system scales while the Hubble radius, which is the scale relevant for the cosmic acceleration, is many orders of magnitude larger. A quite promising scheme in this approach is the Dvali, Gabadadze and Porrati (DGP) model \\cite{Dvali:2000hr} which corresponds to a 5-dimensional (5D) induced gravity brane-world model \\cite{Deffayet,IG, Sahni:2002dx,LDGP2,Bouhmadi-Lopez:2004ys}, where a low-energy modification occurs with respect to general relativity; i.e. an infrared effect takes place, leading to two branches of solutions: (i) the self-accelerating branch and (ii) the normal branch. As it name would suggest the self-accelerating branch solution gives rise to a late-time accelerating brane universe. The acceleration of the brane expansion arises naturally, even without invoking the presence of any dark energy on the brane to produce the speed-up. Not surprisingly therefore this self-accelerated feature of the DGP model has lead to considerable research activity \\cite{reviewDGP}. Furthermore, it has been recently shown that by embedding the DGP model in a higher dimensional space-time the ghost issue present in the original model \\cite{Koyama:2007za} may be curred \\cite{deRham:2007xp} while preserving the existence of a self-accelerating solution \\cite{Minamitsuji:2008fz}. The normal branch also constitutes in itself a very interesting result of the DGP model however, as it can mimic a phantom-like behaviour on the brane by means of the $\\Lambda$DGP scenario \\cite{Sahni:2002dx}. At this respect we remind the reader that observational data do not seem incompatible with a phantom-like behaviour \\cite{Percival:2007yw} and therefore we should keep an open mind about what is producing the recent inflationary era of our universe. On the other hand, this phantom-like behaviour may well be a property acquired only recently by dark energy. This leads to an interest in building models that exhibit the so called crossing of the phantom divide line $w=-1$; for example in the context of the brane-world scenario \\cite{LDGP2,crossing}. The most important aspect of the $\\Lambda$DGP model is that the phantom-like mimicry is obtained without invoking any real phantom-matter \\cite{phantom} which is known to violate the null energy condition and induce quantum instabilities\\footnote{We are referring here to a phantom energy component described through a minimally coupled scalar field with the wrong kinetic term.} \\cite{Cline:2003gs}. In the DGP scenario it is also possible to obtain a mimicry of the crossing of the phantom divide at the cost of invoking a dynamical dark energy on the brane \\cite{LDGP2}, for example modelled by a quiessence fluid or a (generalised) Chaplygin gas. In this paper we will show that there is an alternative form of mimicking the crossing the cosmological constant line $w=-1$ in the brane-world scenario. More precisely, we consider a 5D dilatonic bulk with a brane endowed with (or without) an induced gravity term, a brane matter content corresponding to cold dark matter, and a brane tension $\\lambda$ that depends on the minimally coupled bulk scalar field. We will show that in this set-up the normal branch expands in an accelerated way due to $\\lambda$ playing the role of \\textit{dark energy} --through its dependence on the bulk scalar field. In addition, it turns out that $\\lambda$ grows with the brane scale factor until it reaches a maximum positive value and then starts decreasing. Therefore, in our model the brane tension mimics crossing the phantom divide. Most importantly no matter violating the null energy density is invoked in our model. The paper can be outlined as follows: in section \\ref{sec2} we describe the general bulk plus brane scenario; in Section \\ref{sec3} we then analyse the vacuum (i.e., $\\rho_{m}=0$) solution to prepare the ground for the analysis carried out in section \\ref{sec4}. There we show that, under some assumptions on the nature of the coupling parameters between $\\lambda$ and $\\phi$, $1+w_{\\rm{eff}}$ changes sign as the brane evolves, with $w_{\\rm{eff}}$ the effective equation of state for the brane tension. In section \\ref{sec5} we show that the the presence of the induced gravity term is not a necessary ingredient in order to mimic the crossing of the phantom divide line in this context. Our conclusions are presented in section \\ref{sec6}. Finally, in the appendix \\ref{appendix} we present an analytical proof for the mimicry of the crossing of the phantom divide in the model presented in section~\\ref{sec2}. ", "conclusions": "\\label{sec6} In this paper we have shown the existence of a mechanism that mimics the crossing of the cosmological constant line $w=-1$ in the brane-world scenario, and which is different from the one introduced in Refs.~\\cite{Sahni:2002dx,LDGP2}. More precisely, we have shown that if we have a 5D dilatonic bulk with or without an induced gravity term on the brane (normal branch), a brane tension $\\lambda$ which depends on the minimally coupled bulk scalar field, and a brane matter content corresponding only to cold dark matter, then under certain conditions the brane tension grows with the brane scale factor until it reaches a maximum positive value at which it mimics crossing the phantom divide, and then starts decreasing. Most importantly no matter violating the null energy condition is invoked in our model. Despite of the transitory phantom-like behaviour of the brane tension no big rip singularity is hited along the brane evolution. In the model with an induced gravity term on the brane (normal branch or non-self-accelerating branch), the constraint equation fulfilled by the brane tension is too complicated to be solved analytically (see Eqs.~(\\ref{Friedmann}) and (\\ref{constraint})). However, we have shown that under certain physical and mathematical conditions -cold dark matter dominates at higher redshifts and it dilutes a bite faster than dust during the brane expansion as well as a mathematical condition that guarantees the non-existence of a local minimum of the brane tension- it is possible for the brane tension to cross the cosmological constant line. The analytical proof (see the appendix) has been confirmed by numerical solutions. Furthermore, we have shown that for some values of the parameters the normal branch inflates eternally to the future due to the brane tension $\\lambda$ playing the role of dark energy through its dependence on the bulk scalar field. On the other hand, in the model without an induced gravity term things are much easier to analyse as it is possible to get an analytical solution for the brane tension \\cite{Kunze:2001ji}. For the analytical solution and under the same physical and mathematical conditions assumed in the model with an induced gravity effect, we have shown that the brane tension grows until it reaches its maximum positive value and then it starts decreasing driving the late-time acceleration of the brane. It is precisely at that maximum value that the brane tension mimics the crossing of the phantom divide as its effective equation of state parameter $w_{\\rm{eff}}$ is such that $1+w_{\\rm{eff}}$ changes its sign in a smooth way. We have also imposed observational bounds -from the dark matter equation of state \\cite{Muller:2004yb} and the big bang nucleosynthesis \\cite{Kunze:2001ji} that constraints the modified Friedmann equation and the effective gravitational constant of the brane- to constraint the parameters of the model. In summary, in the model presented here the mimicry of the phantom divide crossing is based on the interaction between the brane and the bulk through the brane tension that depends explicitly on the scalar field that lives in the bulk. We have also shown that the brane undergoes a late-time acceleration epoch." }, "0807/0807.3744_arXiv.txt": { "abstract": "We present a photometric analysis of the star clusters Lindsay\\,1, Kron\\,3, NGC\\,339, NGC\\,416, Lindsay\\,38, and NGC\\,419 in the Small Magellanic Cloud (SMC), observed with the \\textit{Hubble Space Telescope} Advanced Camera for Surveys (ACS) in the F555W and F814W filters. Our color magnitude diagrams (CMDs) extend $\\sim$3.5 mag deeper than the main-sequence turnoff points, deeper than any previous data. Cluster ages were derived using three different isochrone models: Padova, Teramo, and Dartmouth, which are all available in the ACS photometric system. Fitting observed ridgelines for each cluster, we provide a homogeneous and unique set of low-metallicity, single-age fiducial isochrones. The cluster CMDs are best approximated by the Dartmouth isochrones for all clusters, except for NGC\\,419 where the Padova isochrones provided the best fit. Using Dartmouth isochrones we derive ages of $7.5 \\pm 0.5$~Gyr (Lindsay\\,1), $6.5 \\pm 0.5$~Gyr (Kron\\,3), $6 \\pm 0.5$~Gyr (NGC\\,339), $6 \\pm 0.5$~Gyr (NGC\\,416), and $6.5 \\pm 0.5$~Gyr (Lindsay\\,38). The CMD of NGC\\,419 shows several main-sequence turn-offs, which belong to the cluster and to the SMC field. We thus derive an age range of 1.2-1.6~Gyr for NGC\\,419. We confirm that the SMC contains several intermediate-age populous star clusters with ages unlike those of the Large Magellanic Cloud (LMC) and the Milky Way (MW). Interestingly, our intermediate-age star clusters have a metallicity spread of $\\sim$0.6~dex, which demonstrates that the SMC does not have a smooth, monotonic age-metallicity relation. We find an indication for centrally concentrated blue straggler star candidates in NGC\\,416, while for the other clusters these are not present. Using the red clump magnitudes, we find that the closest cluster, NGC\\,419 ($\\sim$50~kpc), and the farthest cluster, Lindsay\\,38 ($\\sim$67~kpc), have a relative distance of $\\sim$17~kpc, which confirms the large depth of the SMC. The three oldest SMC clusters (NGC\\,121, Lindsay\\,1, Kron\\,3) lie in the north-western part of the SMC, while the youngest (NGC\\,419) is located near the SMC main body. ", "introduction": "Star clusters are powerful tools for probing the star-formation history and the associated chemical evolution of a galaxy. As one of the closest star forming galaxies with star clusters covering a wide range of ages, the Small Magellanic Cloud (SMC) is a preferred location for detailed studies of this class of objects. The SMC is the only dwarf galaxy in the Local Group containing populous intermediate-age star clusters of all ages. The SMC appears to be part of a triple system together with the Large Magellanic Cloud (LMC) and the Milky Way (MW). Its star formation activity may be triggered by interactions with its companions \\citep[e.g.][]{yoshi03}. The proximity of the SMC allows us to resolve individual stars in compact and massive star clusters of intermediate and old age, down to the sub-solar stellar mass regime. The globular cluster (GC) system of the MW exhibits a range of ages between $\\sim$10.5 and 14~Gyr \\citep[e.g.,][]{deang05} with the oldest populations belonging to the most ancient surviving stellar systems. In the Galactic halo, a ''young'' group of star clusters is found with Pal\\,1 being the youngest with an age of $8 \\pm 2$~Gyr \\citep{Rosenberg98}. Theories explaining the origin of these so-called young halo clusters, consider them to have been captured by the MW \\citep{buon95}, to have been formed during interactions between the MW and the Magellanic Clouds \\citep{fupe95}, or to have been accreted from destroyed and/or merged dwarf satellites \\citep[e.g.,][]{zinn93,mack04}. The star formation history of the LMC shows pronounced peaks that coincide with the times of possible past close encounters between the LMC, SMC and MW, indicative of interaction-triggered cluster formation \\citep[e.g.,][]{gir95}. In the LMC, two epochs of cluster formation have been observed that are separated by an ''age gap'' of about 4-9~Gyr \\citep[e.g.,][]{holtz99, john99, Harris01}. In the early epoch a well-established population of metal-poor ($\\langle[Fe/H]\\rangle\\sim -2$) star clusters with comparable properties to Galactic halo clusters \\citep{sunt92, olsen98, dutra99} was formed. These clusters are as old as the oldest globular clusters in the MW and in the Galactic dwarf spheroidal companions \\citep{greb04}. In a second epoch, a large population of intermediate-age clusters with ages less than 3-4~Gyr have developed. In contrast, the SMC contains only one old GC, NGC\\,121, which is 2-3~Gyr younger than the oldest GC in the LMC and MW \\citep{glatt08} (Paper~I). The second oldest SMC star cluster, Lindsay\\,1, has an age of $7.5 \\pm 0.5$~Gyr, and since then compact populous star clusters have formed fairly continuously until the present day \\citep[e.g.,][]{daco02}. Furthermore, the intermediate-age clusters in the SMC might survive for a Hubble time, due to their high mass and the structure of the SMC (no bulge or disk to be passed) \\citep{hunter03,lamers05,gieles07}. The existing age determinations to this point have often been associated with large uncertainties. Stellar crowding, field star contamination and faintness of the main-sequence turnoffs made the measurement of precise ages difficult. These problems affect in particular ground-based data. Another difficulty is the large depth extent of the SMC which exacerbates the distance modulus-reddening degeneracy for each cluster. These uncertainties can affect the age determination considerably. The capabilities of the Advanced Camera for Surveys (ACS) aboard the \\textit{Hubble Space Telescope} (HST) provide an improvement both in sensitivity (depth) as well as angular resolution, which is essential for reliable photometric age determinations in dense clusters. We present improved cluster ages and distance determinations for Lindsay\\,1, Kron\\,3, NGC\\,416, NGC\\,339, Lindsay\\,38, and NGC\\,419. This is part of a ground-based and space-based program to uncover the age-metallicity evolution of the SMC. Our space-based imaging data were obtained with HST/ACS and our ground-based spectroscopy was obtained with \\textit{Very Large Telescope} (VLT). We combine our photometric results with spectroscopic metallicity determinations to obtain a well-sampled age-metallicity relation. The age-metallicity relation determined so far indicated that SMC clusters of similar age may differ by several tenths of dex in metallicity \\citep{daco98}. Previous studies provided ages and metallicities of SMC star clusters using a variety of techniques and telescopes (see $\\S$~\\ref{sec:ana}). Combining all published cluster ages for e.g. Kron\\,3 (5-10~Gyr) \\citep{gasc66,alc96,migh98,udal98,rich00}, we find a wide range of ages for some key star clusters depending on the method used for the determination. \\begin{deluxetable*}{cccccc} \\tabletypesize{\\scriptsize} \\tablecolumns{6} \\tablewidth{0pc} \\tablecaption{Journal of Observations} \\tablehead{ \\colhead{Cluster} & \\colhead{Date} & \\colhead{} & \\colhead{Total Exposure Time} & \\colhead{R.A.} & \\colhead{Dec.} \\\\ \\colhead{} & \\colhead{yy/mm/dd} & \\colhead{Filter} & \\colhead{(s)} & \\colhead{} & \\colhead{} } \\startdata Lindsay\\,1 &2005$/$08$/$21& F555W & 40.0 & $0^h03^m53.19^s$ & $-73\\arcdeg28'15.74''$ \\\\ && & 1984.0 & $0^h03^m52.66^s$ & $-73\\arcdeg28'16.47''$\\\\ && F814W & 20.0 & $0^h03^m53.19^s$ & $-73\\arcdeg28'15.74''$\\\\ &&\t& 1896.0 & $0^h03^m52.66^s$ & $-73\\arcdeg28'16.47''$\\\\ Kron\\,3 &2006$/$01$/$17& F555W & 40.0 & $0^h24^m41.64^s$ & $-72\\arcdeg47'47.49''$\\\\ &&\t& 1984.0 & $0^h24^m41.92^s$ & $-72\\arcdeg47'45.49''$\\\\ && F814W & 20.0 & $0^h24^m41.64^s$ & $-72\\arcdeg47'47.49''$\\\\ &&\t& 1896.0 & $0^h24^m41.92^s$ & $-72\\arcdeg47'45.49''$\\\\ NGC\\,339 &2005$/$11$/$28& F555W & 40.0 & $0^h57^m47.40^s$ & $-74\\arcdeg28'26.25''$\\\\ &&\t& 1984.0 & $0^h57^m47.13^s$ & $-74\\arcdeg28'24.16''$\\\\ && F814W & 20.0 & $0^h57^m47.40^s$ & $-74\\arcdeg28'26.25''$\\\\ &&\t& 1896.0 & $0^h57^m47.13^s$ & $-74\\arcdeg28'24.16''$\\\\ NGC\\,416 &2006$/$03$/$08& F555W & 40.0 & $1^h07^m53.59^s$ & $-72\\arcdeg21'02.47''$\\\\ (WFC)\t &&\t& 1984.0 & $1^h07^m54.09^s$ & $-72\\arcdeg21'01.79''$\\\\ && F814W & 20.0 & $1^h07^m53.59^s$ & $-72\\arcdeg21'02.47''$\\\\ &&\t& 1896.0 & $1^h07^m54.09^s$ & $-72\\arcdeg21'01.79''$\\\\ NGC\\,416 &2006$/$08$/$12& F555W & 70.0 & $1^h07^m58.76^s$ & $-72\\arcdeg21'19.70''$\\\\ (HRC)\t &&\t& 1200.0 & $1^h07^m58.96^s$ & $-72\\arcdeg21'19.30''$\\\\ && F814W & 40.0 & $1^h07^m58.76^s$ & $-72\\arcdeg21'19.70''$\\\\ &&\t& 1036.0 & $1^h07^m58.96^s$ & $-72\\arcdeg21'19.30''$\\\\ Lindsay\\,38 &2005$/$08$/$18& F555W & 40.0 & $0^h48^m57.14^s$ & $-69\\arcdeg52'01.77''$\\\\ &&\t& 1940.0 & $0^h48^m56.76^s$ & $-69\\arcdeg52'03.07''$\\\\ && F814W & 20.0 & $0^h48^m57.14^s$ & $-69\\arcdeg52'01.76''$\\\\ &&\t& 1852.0 & $0^h48^m56.76^s$ & $-69\\arcdeg52'03.07''$\\\\ NGC\\,419 &2006$/$01$/$05& F555W & 40.0 & $1^h08^m12.53^s$ & $-72\\arcdeg53'17.72''$\\\\ (WFC)\t &&\t& 1984.0 & $1^h08^m12.71^s$ & $-72\\arcdeg53'15.49''$\\\\ && F814W & 20.0 & $1^h08^m12.53^s$ & $-72\\arcdeg53'17.72''$\\\\ && & 1896.0 & $1^h08^m12.71^s$ & $-72\\arcdeg53'15.49''$\\\\ NGC\\,419 &2006$/$04$/$26& F555W & 70.0 & $1^h08^m17.93^s$ & $-72\\arcdeg53'03.60''$\\\\ (HRC)\t &&\t& 1200.0 & $1^h08^m17.78^s$ & $-72\\arcdeg53'02.80''$\\\\ && F814W & 40.0 & $1^h08^m17.93^s$ & $-72\\arcdeg53'03.60''$\\\\ && & 1036.0 & $1^h08^m17.78^s$ & $-72\\arcdeg53'08.80''$\\\\ \\enddata \\label{tab:journalobs} \\end{deluxetable*} Here we present the deepest available photometry with HST/ACS, which allows us to carry out the most accurate age measurements obtained so far. We determine the ages of these clusters utilizing three different isochrone models, which also yields distances. In the next Section we describe the data reduction procedure. In $\\S$~\\ref{sec:method} we present the color-magnitude diagrams (CMD) of the clusters and discuss their main features. In $\\S$~\\ref{sec:age} we describe our age derivation method and present our results. We give an estimate of the distances of our clusters long the line-of-sight in $\\S$~\\ref{sec:dist} and present a discussion and a summary in Sections $\\S$~\\ref{sec:ana} and $\\S$~\\ref{sec:sum}, respectively. \\begin{figure*} \\epsscale{1.2} \\plotone{f1.eps} \\caption{Spatial Distribution in 2D of our cluster sample (red circles). The location of eight additional SMC clusters, for which reliable ages from the literature are available, is shown (blue crosses). We obtain a complete sample of all intermediate-age and old SMC star clusters (see $\\S$~\\ref{sec:age}), which we will discuss in $\\S$~\\ref{sec:dist} and~\\ref{sec:ana}. One of the clusters, Lindsay\\,116, lies outside the coordinate boundaries of the Figure. The cluster locations are shown superimposed on a star map of the SMC generated using the point source catalog of the Small Magellanic Cloud Photometric Survey \\citep{zar02} for stars with V$<$16.5~mag.} \\label{fig:clusters} \\end{figure*} ", "conclusions": "\\label{sec:ana} \\subsection{Comparison of our age determination with previous studies} \\label{sec:comp} Previous studies done by several different authors provided ages and metallicities of SMC star clusters using a variety of techniques and telescopes. Therefore, if we combine all published cluster ages, we find a wide range of ages and metallicities for a given cluster, depending on the method used for the determination: Lindsay\\,1 has an age range from 7.3-10~Gyr \\citep{gasc66,gasc80,ole87,sara95,migh98,udal98,alc03}, Kron\\,3 from 5-10~Gyr \\citep{gasc66,rich84,alc96,migh98,udal98,rich00}, and NGC\\,339 from 5-7.9~Gyr \\citep{migh98,udal98,rich00}. No other cluster has such a wide range of different age determinations as NGC\\,416, reaching from 2.5-11.2~Gyr \\citep{dur84, els85, bica86, migh98, udal98, rich00}. The cluster is located close to the SMC main body where a large interstellar extinction is expected. The separation of field stars from the real cluster members was a major problem in the age determination process, among uncertain values for metallicity, reddening and distance. Using photometry obtained with the Wide Field Planetary Camera 2 (WFPC2) aboard HST, \\citet{migh98} found an absolute age of $6.6 \\pm 0.5$~Gyr for NGC\\,416, while \\citet{rich00} derived an age of 7.1 to 11.2~Gyr using the same data set. The only available CMD of Lindsay\\,38 is provided by \\citet{pia01}. The observation was carried out with the Cerro Tololo Inter-American Observatory (CTIO) 0.9~m telescope using the Tektronix 2K \\#~3 CCD. They presented the first age determination of Lindsay\\,38 with $6 \\pm 0.6$~Gyr. For NGC\\,419, the latest CMD was published by \\citet{rich00} based on WFPC2 data. \\citet{udal98} published an age of 3.3~Gyr and \\citet{rich00} give an age range of $1.0-1.8$~Gyr. For Lindsay\\,1, Kron\\,3, NGC\\,339, NGC\\,416, and NGC\\,419, the latest and deepest available CMD was provided with WFPC2 \\citep{migh98,rich00}, while for Lindsay\\,38 only ground-based data existed. \\begin{deluxetable*}{ccccccc} \\tablecolumns{7} \\tablewidth{0pc} \\tablecaption{Age Comparison} \\tablenote{Comparison of our ages derived with the Dartmouth isochrones. For NGC\\,419 Padova isochones provided the best fit. } \\tablehead{ \\colhead{Ref.Source} & \\colhead{Lindsay\\,1} & \\colhead{Kron\\,3} & \\colhead{NGC\\,339} & \\colhead{NGC\\,416} & \\colhead{Lindsay\\,38} & \\colhead{NGC\\,419} \\\\ \\colhead{} & \\colhead{Gyr} & \\colhead{Gyr} & \\colhead{Gyr} & \\colhead{Gyr} & \\colhead{Gyr} & \\colhead{Gyr}} \\startdata This paper\t\t & $7.5 \\pm 0.5$ & $6.5 \\pm 0.5$ & $6 \\pm 0.5$ & $6 \\pm 0.5$ & $6.5 \\pm 0.5$ & $1.2-1.6$ \\\\ \\citet{rich00} \t &\t -\t & $5.6-7.9$ & $5.0-7.9$ & $4.0-7.1$ &\t -\t & $1.0-1.8$ \\\\ \\citet{migh98} \t & $7.7 \\pm 0.4$ & $4.7 \\pm 0.7$ & $5.0 \\pm 0.6$ & $5.6 \\pm 0.6$ &\t-\t &\t-\t \\\\ \\citet{udal98} \t & $9.0$\t & $7.5$\t &\t$4.0$\t & $6.6$\t &\t -\t & $3.3$\t \\\\ \\citet{sara95} \t & $7.3 \\pm 0.6$ &\t -\t &\t -\t &\t -\t &\t -\t &\t-\t \\\\ Alcaino et al. (1996, 2003) & $9-10$\t & $8$\t &\t -\t &\t -\t &\t -\t &\t-\t \\\\ \\citet{pia01}\t\t &\t -\t &\t -\t &\t -\t &\t -\t & $6.0 \\pm 0.6$ &\t-\t \\\\ \\enddata \\label{tab:age_com} \\end{deluxetable*} In Table~\\ref{tab:age_com} we compare our ages using the best-fitting Dartmouth isochrones (except NGC\\,419, for which the Padova isochones provided the best fit) with results published in the most recent studies based on HST/WFPC2 photometry. The data reach $\\sim$2~mag below the turnoff points, while our ACS data have a depth of 3.5~mag below turnoffs. We can see that \\citet{migh98} derived a similar age for Lindsay\\,1, while for the remaining clusters in the overlapping sample they found younger ages than the ones derived here. \\citet{rich00}, who used the same WFPC2 ''snapshots'' as \\citet{migh98}, gave age ranges for certain metallicities for the clusters in their sample, which cover the ages determined in this paper. The ages published by \\citet{udal98} using OGLE (Optical Gravitational Lensing Experiment) data, do not exhibit a general trend to older or younger ages as compared to our results, and the age difference varies for each cluster. The OGLE survey is a shallow ground-based survey with a limiting magnitude of $\\sim$21~mag. \\citet{sara95} used the B-V color difference between the HB and the RGB for star clusters with red HB morphologies for their age determination. The CMDs were obtained using data from the 2048 RCA prime-focus CCD on the CTIO 4~m telescope \\citep{ole87} and the photometry reaches V$\\sim$23~mag. The age found for Lindsay\\,1 is in excellent agreement with our result. Alcaino et al. (1996, 2003) used photometry for Lindsay\\,1 obtained with the 1.3~m Warsaw telescope, Las Campanas Observatory and reaches V$\\sim$22~mag. The age was determined by using the so-called vertical method, based on the difference between the luminosity of the MSTO and the HB level. For Kron\\,3, the photometry was taken with the EFOSC-2 CCD camera at the 2.2~m Max-Planck-Institute telescope of ESO , La Silla, and reaches V$\\sim$23~mag \\citep{alc96}. The age was determined using isochrones. In both studies, the resulting ages are higher than our values. \\citet{pia01} were the first to publish an age for Lindsay\\,38, which is in excellent agreement with the age derived in this paper. Most CMDs published in previous studies do not go deep enough to show a clearly outlined MSTO, which is an essential feature for most age determination techniques. \\citet{migh98} determined their cluster ages relative to the age of Lindsay\\,1, measuring the difference between the RC and the RGB and found similar ages as in this paper. Kron\\,3 is an exception for which the authors derived a younger age due to large error associated with the MS photometry. \\citet{rich00} fitted isochrones to the red clump and also calculated the difference between the MSTO and the RC ($\\Delta V^{RC}_{TO}$) in combination with the calibration of \\citet{walker92}. The cluster ages found in this paper are within the age ranges given by \\citet{rich00}. CMDs using ground-based photometry reach $\\sim$V=20~mag, which is not deep enough to show the SGBs or the MSTOs, which can lead to large age differences. \\subsection{Age range and spatial distribution} \\label{sec:agerange} The intermediate-age SMC star clusters Lindsay\\,1, Kron\\,3, NGC\\,339, NGC\\,416, and Lindsay\\,38 form a continuous age sequence from 6 to 7.5~Gyr. The SMC is the only dwarf galaxy of the Local Group known to contain populous star clusters in this age range. The only ''true'' globular cluster in the SMC, NGC\\,121, has an age of 10.5-11.5~Gyr (Paper~I), but is still 2-3~Gyr younger than the oldest LMC and MW globular clusters \\citep[e.g., ][]{Ole96,olsen98,john99,mack04}. Between NGC\\,121, and the second oldest cluster, Lindsay\\,1, there is a small age-gap ($\\sim$3~Gyr), in which no surviving star cluster has been formed. In our sample, we have four clusters with ages between 6-6.5~Gyr, and one that is significantly younger (1.2-1.6~Gyr). Good quality ages are available from ground-based and space-based observations for ten additional intermediate-age SMC star clusters. Combining them with our star clusters, we obtain a complete sample of all intermediate-age and old SMC star clusters: Kron\\,28, Kron\\,44, Lindsay\\,116, Lindsay\\,32, Lindsay\\,11, NGC\\,152, NGC\\,361, NGC\\,411, Lindsay\\,113, and BS90 (Table~\\ref{tab:lit_ages}). \\begin{deluxetable*}{ccccc} \\tablecolumns{5} \\tablewidth{0pc} \\tablecaption{Literature cluster ages} \\tablenote{Ages for nine additional intermediate-age clusters from the literature. } \\tablehead{ \\colhead{Cluster} & \\colhead{Age} & \\colhead{Data} & \\colhead{Method} & \\colhead{Ref.Source} \\\\ \\colhead{} & \\colhead{Gyr} & \\colhead{} & \\colhead{} & \\colhead{}} \\startdata Kron\\,28 & $2.1 \\pm 0.5$ & CTIO 0.9~m telescope / Tektronix 2K \\#~3 CCD& $\\Delta V_{MSTO}^{RC,HB}$ & \\citet{pia01} \\\\ Kron\\,44 & $3.1 \\pm 0.8$ & CTIO 0.9~m telescope / Tektronix 2K \\#~3 CCD& $\\Delta V_{MSTO}^{RC,HB}$ & \\citet{pia01} \\\\ Lindsay\\,116 & $2.8 \\pm 1.0$ & CTIO 0.9~m telescope / Tektronix 2K \\#~3 CCD& $\\Delta V_{MSTO}^{RC,HB}$ & \\citet{pia01} \\\\ Lindsay\\,32 & $4.8 \\pm 0.5$ & CTIO 0.9~m telescope / Tektronix 2K \\#~3 CCD& $\\Delta V_{MSTO}^{RC,HB}$ & \\citet{pia01} \\\\ Lindsay\\,11 & $3.5 \\pm 1.0$ & CTIO 4.0~m telescope / RCA CCD \t & Isochrones \t\t& \\citet{mould92} \\\\ NGC\\,152 & $1.4 \\pm 0.2$ & HST/WFPC2\t\t\t\t & Isochrones \t\t& \\citet{crowl01} \\\\ NGC\\,361 & $8.1 \\pm 1.2$ & HST/WFPC2\t\t\t\t & Isochrones \t\t& \\citet{migh98} \\\\ NGC\\,411 & $1.2 \\pm 0.2$ & HST/WFPC2\t\t\t\t & Isochrones \t\t& \\citet{alsa99} \\\\ Lindsay\\,113 & $4.0 \\pm 0.7$ & HST/WFPC2\t\t\t\t & $d_{B-V}$\\tablenotemark{b} & Mighell et al. (1998a) \\\\ BS90\t & $4.3 \\pm 0.1$ & HST/ACS \t\t\t\t & Isochrones \t\t& \\citet{sabbi07} \\\\ \\enddata \\tablenotetext{b}{The method used by Mighell et al. (1998a) is defined by the (B-V) color difference between the mean color of the red clump and the RGB at the level of the RGB. This value then was compared with Lindsay\\,1, NGC\\,416, and Lindsay\\,113.} \\label{tab:lit_ages} \\end{deluxetable*} For none of these clusters deep HST photometry is available, thus their ages should be considered with some caution. For NGC\\,152, NGC\\,361, NGC\\,411, Lindsay\\,113, and BS90 ''snapshots'' are available taken with WFPC2 (reaching V$\\sim$23~mag), and ACS (BS90, reaching V$\\sim$26~mag). Looking at Figure~\\ref{fig:clusters}, we clearly see that the youngest clusters are located near the SMC main body, while the clusters with ages higher than $\\sim$4~Gyr lie in the outer parts. NGC\\,361 seems to be an exception, but the cluster age is still uncertain, and the literature age of 8.1~Gyr probably is too high. \\citet{crowl01} determined a distance of $51.7 \\pm 1.8$~kpc for N361 whereby the cluster lies $\\sim$ 7.5~kpc ahead of the SMC center. Another exception is BS90 that lies near the SMC main body, even though the cluster has an age of $\\sim$4.3~Gyr. The three oldest SMC clusters (NGC\\,121, Lindsay\\,1, Kron\\,3) are located in the north-western part of the SMC. We note that Lindsay\\,116 cannot be seen in Figure~\\ref{fig:clusters}, because it is located $6\\arcdeg .1$ south-east of the bar and lies therefore outside the displayed area. The closest cluster in our sample, NGC\\,419, and the farthest cluster, Lindsay\\,38, have a relative radial distance of 17~kpc from each other. We can therefore confirm that the SMC has a large extension along the line-of-sight, as was also found by \\citet{crowl01} based on its star clusters. \\begin{figure} \\epsscale{1.2} \\plotone{f43.eps} \\caption{Age vs distance to the sun (projected distance) including different symbols for different metallicity ranges. For five clusters we found reliable distances \\citet{crowl01}, ages \\citep{alsa99,pia01,crowl01}, and metallicities (Kayser et al. 2008, in prep.). All values for BS90 were adopted from \\citet{sabbi07}. The dashed line represents the SMC distance modulus of $(m-M)_0$ = $18.88 \\pm 0.1$~mag \\citep{Storm04}.} \\label{fig:distribution} \\end{figure} In Figure~\\ref{fig:distribution} we show the distribution of age vs the distance to the sun of the clusters in our sample. The locations are shown relative to our adopted SMC distance and indicate that the clusters generally are distributed within $\\pm$6-7~kpc of the SMC centroid. Interesting exceptions are the younger clusters Kron\\,28, NGC\\,411, and NGC\\,419 that in projection appear near the center of the SMC. In fact, they could be located considerably closer to us (see also Fig.~\\ref{fig:proj}). Further measurements of the distance of younger clusters thus would be worthwhile. Moreover, we included five clusters for which we found reliable ages, distances, and metallicities in the literature. We divided the cluster metallicities into four groups and use different symbols for each group in the plot. Even though our plot contains only 11 clusters, we can see trends in the distributions of their properties. Age and distance from the sun appear to be correlated. The closest cluster, NGC\\,419, is also the youngest and most metal rich cluster, while the most distant cluster, Lindsay\\,38, is also the most metal poor, in spite of not being the oldest cluster. \\begin{figure} \\epsscale{1.3} \\plotone{f44.eps} \\caption{Three dimensional distribution is shown for SMC star clusters with ages and distances derived from isochrone fits to CMDs derived from HST observations. Note that the intermediate age clusters are distributed throughout much of the extended body of the SMC. As discussed in the text, the selection of clusters is biased in that our observations generally avoided clusters in locations with high field star densities. However, this incomplete sample suggests that age and radial distance from the center of the SMC are not correlated; e.g. the younger cluster Kron\\,28, NGC\\,411, and NGC\\,419 are at large radial distances and cover a range in metallicity. The yellow star symbolizes the SMC center.} \\label{fig:proj} \\end{figure} One could speculate that in regions at the outskirts of the double LMC-SMC system the star formation activity has been lower/slower than elsewhere, possibly with more unenriched gas, thus allowing for a more moderate enrichment. The oldest object, NGC\\,121, is not the most metal poor cluster, but the second metal poorest and the second farthest one. Its low metallicity could be the result of both a ''natural'' age-metallicity relation and a ''distance from the system'' effect. Figure~\\ref{fig:proj} illustrates the distribution of SMC star clusters with high quality distances derived from isochrone fits to CMDs derived from HST observations. This is a highly biased sample; star clusters seen in the direction of the SMC `bar' are not preferred for these projects because of their large levels of field star contamination. The exception in this case is the cluster BS90 that was accidently included in observations of NGC\\,346 (see Sabbi et al. 2007). The present limited data for clusters in this project show that the SMC is quite extended along the line of sight, consistent with other studies of individual stars and star clusters (see discussion in $\\S$~\\ref{sec:dist}). This three dimensional distribution of the clusters also demonstrates the lack of trends in cluster age or metallicity with radial distance from the center of the SMC. \\begin{deluxetable}{ccc} \\tablecolumns{3} \\tablewidth{0pc} \\tablecaption{Distances} \\tablenote{The projected distances were calculated in this paper and adopted from \\citet{crowl01,sabbi07}. Using these values and the cluster coordinates, we determined the cluster distances to the SMC center ($\\alpha = 0^h52^m44.8^s$, $\\delta = -72\\arcdeg49'43''$).} \\tablehead{ \\colhead{Cluster} & \\colhead{Projected Dist.} & \\colhead{Dist. to SMC Center} \\\\ \\colhead{} & \\colhead{kpc} & \\colhead{kpc} } \\startdata NGC\\,121 & $64.9 \\pm 1.2$ & $8.76 \\pm 1.1$ \\\\ Lindsay\\,1 & $56.9 \\pm 1.0$ & $13.28 \\pm 1.0$ \\\\ Kron\\,3 & $60.6 \\pm 1.1$ & $7.19 \\pm 1.1$ \\\\ NGC\\,339 & $57.6 \\pm 4.1$ & $0.73 \\pm 2.0$ \\\\ NGC\\,416 & $60.4 \\pm 1.9$ & $3.94 \\pm 1.4$ \\\\ Lindsay\\,38 & $66.7 \\pm 1.6$ & $6.27 \\pm 1.3$ \\\\ NGC\\,419 & $50.2 \\pm 2.6$ & $10.83 \\pm 1.6$ \\\\ NGC\\,411 & $50.1 \\pm 1.7$ & $11.1 \\pm 1.3$ \\\\ NGC\\,152 & $59.0 \\pm 1.8$ & $5.58 \\pm 1.3$ \\\\ Kron\\,28 & $45.2 \\pm 1.7$ & $14.78 \\pm 1.3$ \\\\ Kron\\,44 & $57.7 \\pm 1.8$ & $4.37 \\pm 1.3$ \\\\ BS90\t & $60.3$\t & $1.23$ \t\\\\ \\enddata \\label{tab:lit_dist} \\end{deluxetable} \\subsection{Age distribution and cluster formation history} \\label{sec:agedistribution} By combining the ages of our sample with 9 literature ages for intermediate-age SMC star clusters listed in Table~\\ref{tab:lit_ages}, we obtain a well-observed sample of intermediate-age and old star clusters in the SMC. The cluster NGC\\,361 was excluded from the sample, because the cluster is almost certainly younger than the assumed $\\sim$8~Gyr \\citep{migh98}. The age distribution is shown in Figure~\\ref{tab:age_hist}. In each panel we show our resulting age distribution using ages of different isochrone models (black histrograms) and the combined sample (white histograms). Since the cluster ages from the literature were derived using different data and methods, their distribution does not change. In all three plots of Figure~\\ref{tab:age_hist}, the small age gap between $\\sim$8 and 10~Gyr can clearly be seen. In the first panel we used ages derived with the Dartmouth isochrones. \\citet{rich00} based on HST/WFPC2 found two brief cluster formation intervals with the oldest set $8 \\pm 2$~Gyr ago and the second $2 \\pm 0.5$~Gyr ago, and argued that there were gaps in between. During the older burst the clusters NGC\\,339, NGC\\,361, NGC\\,416, and Kron\\,3, and during the younger burst the clusters NGC\\,411, NGC\\,152, and NGC\\,419 have formed according to \\citet{rich00}. Even though they used the same HST/WFPC2 data as \\citet{rich00}, \\citet{migh98} found no evidence for such cluster formation bursts. We also find no evidence for two significant bursts of star cluster formation in our SMC age distribution, but we do see a slightly enhanced cluster formation activity around 6~Gyr. In the second and the third panel we used our derived Teramo and Padova ages, respectively. The cluster formation at 6~Gyr is even more obvious for both isochrone models than in the upper panel. Apparently, between $\\sim$5 and 6~Gyr no star cluster with sufficient mass to survive has formed, but if Lindsay\\,113 is older than the assumed 4~Gyr adopted from the literature, the cluster lies within the gap. We suggest that the SMC has formed its clusters during its entire lifetime with some epochs of more intense cluster formation activity. More detailed information about the age distribution requires additional deep observations of all remaining intermediate-age SMC star clusters. As shown in Figure~\\ref{fig:proj} there appears to be no simple relationship between cluster position and metallicity in any age range. This perhaps is to be expected given that tidal interactions may have perturbed the orbits of star clusters after they formed or provided opportunities for clusters to form at large radii, as in the present-day SMC wing. We have to emphasize that the cluster sample shown in Figure~\\ref{fig:proj} is not complete. Only for 12 clusters reliable distances have been measured this far (see $\\S$~\\ref{sec:dist}), and these are shown in the Figure. The question of the metallicity distribution of the clusters and how this relates to age and position is more complex and beyond the scope of this paper. \\subsection{Evolutionary history of the SMC as a whole} Looking at the metallicities of our star clusters (Tab.~\\ref{fig:results}), we see that the SMC did not experience a smooth age-metallicity relation, even though the SMC is believed to be well-mixed at the present day \\citep[but see ][]{Grebel92,gonzalez99}. The oldest SMC star cluster, NGC\\,121, has a metallicity of [Fe/H] = $-1.46 \\pm 0.10$ and an age of 10.5-11.5~Gyr, while Lindsay\\,38 is more metal-poor with [Fe/H] = $-1.59 \\pm 0.10$ but has an age of $6.5 \\pm 0.5$~Gyr. SMC star clusters of similar age may differ by several tenths of dex in metallicity (see also Da Costa et al. 1998, Kayser et al. 2008, in preparation). The probably most reasonable explanation involves the infall of unenriched, or less enriched gas. The Magellanic Clouds are surrounded by an extensive HI halo \\citep[e.g.,][]{dickey96}, therefore this possibility may be plausible. Another speculative explanation for the existence of those metal-poor clusters is that the SMC acquired these clusters in a past interaction with another dwarf galaxy, similar to the clusters from the Sagittarius dwarf galaxy being acquired by the Milky Way \\citep[e.g., ][]{carraro07}. The SMC, LMC, and MW form an interacting triple system, which affects each other's star formation history (SFH). However, recent studies have suggested that the Magellanic Clouds only entered the vicinity of the MW fairly recently (e.g., Kallivayalil et al. 2006a/b). It is intriguing that the LMC has a significant age gap between $\\sim$4-9~Gyr, while the SMC formed its clusters continuously during the same time period. Moreover, the SMC appears to have a ''delayed'' globular cluster formation history and formed its first and only globular cluster, NGC\\,121, 2-3~Gyr later then the LMC or the MW. \\begin{figure} \\epsscale{1.2} \\plotone{f45.eps} \\caption{The age distribution of 15 intermediate-age and old SMC clusters (excluding NGC\\,361) with ages derived in this paper, in Paper~I (plotted as black histograms), and adopted from \\citep[Mighell et al. 1998a/b,][]{mould92,alsa99,rich00,pia01,crowl01} (plotted as white histrograms). Since the cluster ages from the literature were derived using different data and methods, their distribution does not change. In the first panel we used the ages found using the Dartmouth models, in the second we use the Padova ages, and in the last panel the Teramo ages were used. The literature ages of NGC\\,152, NGC\\,411, and Lindsay\\,113, are based on HST/WFPC2 data \\citep[Mighell et al. 1998a/b,][]{alsa99,rich00}, while the adopted ages of Kron\\,28, Kron\\,44, Lindsay\\,11, Lindsay\\,32, and Lindsay\\,116 are derived from ground-based photometry \\citep{mould92,crowl01,pia01}. NGC\\,361 is not considered due to its uncertain age. The age distribution illustrates the continuous cluster formation with the small age gap between $\\sim$8 and 10~Gyr.} \\label{tab:age_hist} \\end{figure} Possible orbits of the SMC, LMC, and MW have been modelled by several authors \\citep[e.g., Kallivayalil et al. 2006a/b;][]{Bekki05}. Strong tidal perturbations due to interactions could have triggered the cluster formation \\citep[e.g., ][]{Whitmore99} in the SMC. In the LMC, we find that star clusters have formed in evident bursts. The LMC has two main epochs of cluster formation \\citep[e.g., ][]{Bertelli92} and a well-known age-gap of several billion years, in which no star clusters have formed. A few globular clusters are found with coeval ages like the Galactic globular clusters \\citep[e.g., ][]{ols91,olsen98,john99}. We know only of one star cluster, ESO 121-SC03, that lies within the age-gap, which has an age of 8.3-9.8~Gyr \\citep{mackey06}. A correlation between young star clusters in the LMC and putative close encounters with the SMC and MW have been found by e.g. \\citet{gir95}, although the most recent proper motion measurements indicate that the Magellanic Clouds are currently on their first passage around the MW. For young SMC clusters a relation between close encounters and the cluster formation history is not as obvious as for LMC clusters probably due to a smaller number of clusters \\citep{Chiosi06}. The age distribution in Figure~\\ref{tab:age_hist} shows that a slightly enhanced number of star clusters with ages around 1~Gyr is located in the SMC, which might have been produced through a cloud-cloud collision after a pericenter passage $\\sim$0.5~Gyr ago \\citep{Bekki05}. But evidently massive star clusters older than 1~Gyr have formed continuously until $\\sim$7.5~Gyr ago (Lindsay\\,1). It is not yet understood why populous star clusters older than 4~Gyr have not formed and survived continuously in the LMC, while in the SMC they did. \\citet{Bekki05} explained the different cluster formation histories of the Clouds as a difference in birth locations and initial mass of the host galaxies. Kallivayalil et al. (2006a/b, see also Piatek et al. 2008) measured proper motions for the SMC and the LMC and used Monte Carlo simulations to model the orbits of the Clouds and the MW. While they found bound orbits for the Clouds, they also found that it was difficult to keep the Clouds bound to each other for more than 1~Gyr in the past. It is possible that the Clouds are not a bound system \\citep[see also e.g., ][]{Bekki05}, and that they are making their first passage close to the Milky Way." }, "0807/0807.4508_arXiv.txt": { "abstract": "Measurements of the SNe~Ia Hubble diagram which suggest that the universe is accelerating due to the effect of dark energy may be biased because we are located in a 200-300 Mpc underdense `void' which is expanding 20-30\\% faster than the average rate. With the smaller global Hubble parameter, the WMAP-5 data on cosmic microwave background anisotropies can be fitted without requiring dark energy if there is some excess power in the spectrum of primordial perturbations on 100~Mpc scales. The SDSS data on galaxy clustering can also be fitted if there is a small component of hot dark matter in the form of 0.5 eV mass neutrinos. We show however that if the primordial fluctuations are gaussian, the expected variance of the Hubble parameter and the matter density are far too small to allow such a large local void. Nevertheless many such large voids have been identified in the SDSS LRG survey in a search for the late-ISW effect due to dark energy. The observed CMB temperature decrements imply that they are nearly empty, thus these real voids too are in gross conflict with the concordance $\\Lambda$CDM model. The recently observed high peculiar velocity flow presents another challenge for the model. Therefore whether a large local void actually exists must be tested through observations and cannot be dismissed {\\em a priori}. ", "introduction": "The Einstein-de Sitter (E-deS) universe with $\\Omega_\\mathrm{m}=1$ is the simplest model consistent with the spatial flatness expectation of inflationary cosmology. However, Type Ia supernovae (SNe~Ia) at redshift $z \\simeq 0.5$ appear $\\sim25\\%$ fainter than expected in an E-deS universe \\citep{Riess:1998cb,Perlmutter:1998np}. Together with measurements of galaxy clustering in the Two-degree Field survey \\citep{Efstathiou:2001cw} and of cosmic microwave background (CMB) anisotropies by the Wilkinson Microwave Anisotropy Probe (WMAP) \\citep{Spergel:2003cb}, this has established an accelerating universe with a dominant cosmological constant term (or other form of `dark energy') which presumably reflects the present microphysical vacuum state. This `concordance' $\\Lambda$CDM cosmology (with $\\Omega_\\Lambda \\simeq 0.7$, $\\Omega_\\mathrm{m} \\simeq 0.3$, $h \\simeq 0.7$) has passed a number of cosmological tests, including baryonic acoustic oscillations \\citep{Eisenstein:2005su} and measurements of mass fluctuations from clusters and weak lensing \\citep[e.g.][]{Contaldi:2003hi}. Further observations of both SNe~Ia \\citep{Riess:2004nr,Astier:2005qq,WoodVasey:2007jb} and the WMAP 3-year results \\citep{Spergel:2006hy} have continued to firm up the model. However there is no physical basis for this model, in particular there are two fundamental problems with the notion that the universe is dominated by vacuum energy. The first is the notorious fine-tuning problem of vacuum fluctuations in quantum field theory --- the energy scale of the cosmological energy density is $\\sim10^{-12}$ GeV, many orders of magnitude below the energy scale of $\\sim10^2$ GeV of the Standard Model of particle physics, not to mention the Planck scale of $\\sim10^{19}$ GeV \\citep[see][]{Weinberg:1988cp}. The second is the equally acute coincidence problem: since $\\rho_\\Lambda/\\rho_\\mathrm{m}$ evolves as the cube of the cosmic scale factor $a$, there is no reason to expect it to be of ${\\cal O}(1)$ {\\em today}, yet this is apparently the case. In fact what is actually inferred from observations is {\\em not} an energy density, just a value of ${\\cal O}(H_0^2)$ for the otherwise unconstrained $\\Lambda$ term in the Friedmann equation. It has been suggested that this may simply be an artifact of interpreting cosmological data in the (oversimplified) framework of a perfectly homogeneous universe in which $H_0 \\sim 10^{-42} \\mathrm{GeV} \\sim (10^{28} \\mathrm{cm})^{-1}$ is the {\\em only} scale in the problem \\citep{Sarkar:2007cx}. In fact the WMAP results alone do {\\em not} require dark energy if the assumption of a scale-invariant primordial power spectrum is relaxed. This assumption is worth examining given our present ignorance of the physics behind inflation. We have demonstrated \\citep{Hunt:2007dn} that the temperature angular power spectrum of an E-deS universe with $h \\simeq 0.44$ matches the WMAP data well if the primordial power is enhanced by $\\sim 30\\%$ in the region of the second and third acoustic peaks (corresponding to spatial scales of $k \\sim 0.01-0.1~h~\\mathrm{Mpc}^{-1}$). This alternative model with {\\em no} dark energy actually has a slightly better $\\chi^2$ for the fit to WMAP-3 data than the `concordance power-law $\\Lambda$CDM model' and, inspite of having more parameters, has an {\\em equal} value of the Akaike information criterion used in model selection. Other E-deS models with a broken power-law spectrum \\citep{Blanchard:2003du} have also been shown to fit the WMAP data. Moreover, an E-deS universe can fit measurements of the galaxy power spectrum if it includes a $\\sim 10\\%$ component of hot dark matter in the form of massive neutrinos of mass $\\sim 0.5$~eV \\citep{Hunt:2007dn,Blanchard:2003du}. Clearly the main evidence for dark energy comes from the SNe~Ia Hubble diagram. A mechanism that sets $\\Lambda=0$ is arguably more plausible than one which leads to the tiny energy density $\\rho_\\Lambda\\simeq 10^{-47}$ GeV$^4$ associated with the concordance cosmology.\\footnote{`Quintessence' models, which attempt to address the coincidence problem, also {\\em assume} that every other contribution to the vacuum energy cancels apart from that of the quintessence field.} If $\\Lambda$ is indeed zero then perhaps some effect fools us into wrongly deducing the existence of dark energy by \\emph{mimicking} a nonzero cosmological constant. It is natural to connect this effect with inhomogenities since cosmic acceleration and large scale nonlinear structure formation appear to have commenced simultaneously. This approach offers the possibility of solving the cosmological constant problems within the framework of general relativity and keeps the introduction of new physics to a minimum.\\footnote{In models that seek to explain the observations through modifications of gravity, the relevant scale of $H_0^{-1}$ has to be introduced {\\em by hand}, just as in quintessence models the quintessence field has to be given a mass of order $H_0$ --- these are technically {\\em unnatural} choices since this is an infrared scale for any microphysical theory.} Several different ways in which inhomogenities could potentially mimic dark energy have been considered in the literature --- for reviews see \\citet{Celerier:2007jc,Buchert:2007ik,Enqvist:2007vb}. In an inhomogeneous universe averaged quantities satisfy modified Friedmann equations which contain extra terms corresponding to `backreaction' since the operations of spatial averaging and time evolution do not commute \\citep{Buchert:1999er}. The backreaction terms depend upon the variance of the local expansion rate and hence increase as inhomogenities develop. Whether backreaction can indeed account for the apparent cosmological acceleration is hotly debated and remains an open question at present \\citep{Wetterich:2001kr,Ishibashi:2005sj,Vanderveld:2007cq,Wiltshire:2007fg,Khosravi:2007bq,Leith:2007ay,Behrend:2007mf,Rasanen:2008it,Paranjape:2008jc}. Another possibility is that inhomogeneities affect light propagation on large scales and cause the luminosity distance-redshift relation to resemble that expected for an accelerating universe. This has been investigated for a `Swiss-cheese' universe in which voids modelled by patches of Lema\\^{i}tre-Tolman-Bondi (LTB) space-time are distributed throughout a homogenous background. However, the results seem to be model dependent: some authors find the change in light propagation to be negligible because of cancellation effects \\citep{Biswas:2007gi,Brouzakis:2007zi,Brouzakis:2008uw}, whereas \\citet{Marra:2007pm} claim it can partly mimic dark energy if the voids have radius 250 Mpc \\citep{Marra:2007gc}. \\citet{Mattsson:2007tj} has noted that observers may preferentially choose sky regions with underdense foregrounds when studing distant objects such as SNe~Ia, so the expansion rate along the line-of-sight is then greater than average; such a selection effect he argues can allow an inhomogeneous universe to fit the observations without dark energy. In this paper we are mainly interested in a `local void' (sometimes referred to as ``Hubble bubble'') as an explanation for dark energy; to prevent an excessive CMB dipole moment due to our peculiar velocity we must be located near the centre of the void. An underdense void expands faster than its surroundings, thus younger supernovae inside the void would be observed to be receding more rapidly than older supernovae outside the void. Under the assumption of homogeneity this would lead to the mistaken conclusion that the expansion rate of the Universe is accelerating, although both the void and the global universe are actually decelerating. Henceforth we use the `Hubble contrast' $\\delta_H\\equiv \\left(H_\\mathrm{in}-H_\\mathrm{out}\\right)/H_\\mathrm{out}$ to characterise the void expansion rate, where $H_\\mathrm{in}$ and $H_\\mathrm{out}$ are the Hubble parameters inside and outside the void respectively. (Other authors have used the `jump' ${\\mathcal J}\\equiv\\,H_\\mathrm{in}/H_\\mathrm{out} = 1 + \\delta_H$ to characterise the void.) The reduced Hubble parameter $h$ is defined as usual by $H_\\mathrm{out} = 100h$ km $\\mathrm{s}^{-1}$ $\\mathrm{Mpc}^{-1}$ throughout. The local void scenario has been investigated by several authors using a variety of methods \\citep{Celerier:1999hp,Tomita:1999qn,Tomita:2000rf,Tomita:2000jj,Tomita:2001gh,Iguchi:2001sq,Tomita:2002df,Moffat:2005yx,Moffat:2005zx,Moffat:2005ii,Mansouri:2005rf,Vanderveld:2006rb,Garfinkle:2006sb,Chung:2006xh,Alnes:2005rw,Alnes:2006uk,Alexander:2007xx,Biswas:2006ub,Caldwell:2007yu,Clarkson:2007pz,Uzan:2008qp,GarciaBellido:2008nz,Clifton:2008hv,GarciaBellido:2008gd}. In a series of papers, Tomita modelled the void as a open Friedmann-Robertson-Walker (FRW) region joined by a singular mass shell to a FRW background and found that a void with radius 200 Mpc and $\\delta_H=0.25$ fits the supernova Hubble diagram without dark energy \\citep{Tomita:2001gh}. \\citet{Alnes:2005rw} showed that a LTB region which reduces to a E-deS cosmology with $h=0.51$ at a radius of 1.4~Gpc with $\\delta_H=0.27$ can match both the supernova data and the location of the first acoustic peak in the CMB. \\citet{Alexander:2007xx} attempted to find the smallest possible void consistent with the current supernova results --- their LTB-based `minimal void' model has a radius of 350 Mpc and ${\\mathcal J} \\simeq 1.2$ i.e. $\\delta_H \\simeq 0.2$; a void of similiar size but with $\\delta_H=0.3$ had been discussed earlier \\citep{Biswas:2006ub}. Unfortunately, since this model is equivalent to an E-deS universe with $h=0.44$ {\\em outside} the void where the Sloan Digital Sky Survey (SDSS) luminous red galaxies lie, as it stands it is unable to fit the measurements of the baryonic acoustic oscillation (BAO) peak at $z\\sim0.35$ \\citep{Blanchard:2005ev}. LTB models of much larger voids were considered by \\citet{GarciaBellido:2008nz} (with radii of 2.3 Gpc and 2.5 Gpc and Hubble contrasts of 0.18 and 0.30 respectively) and it was demonstrated they can fit the supernova data, BAO data and the location of the first CMB peak. \\citet{Clifton:2008hv} found the best fit to the SNe Ia data for a void of radius $1.3 \\pm 0.2$ Gpc and an underdensity of about $70\\%$ at the centre and \\citet{Bolejko:2008cm} confirmed that such a void provides an excellent fit to the latest `Union dataset' \\citep{Kowalski:2008ez}. Moreover \\citet{Inoue:2006rd} have shown that the unexpected alignment of the low multipoles in the CMB anisotropy can be attributed to the existence of a local void of radius $300$~$h^{-1}\\,\\mathrm{Mpc}$. These authors also suggested that the anomalous `cold spot' in the WMAP southern sky is due to a similar void at $z \\sim 1$ and some evidence for this has emerged subsequently \\citep{Rudnick:2007kw}. Recently, a large number of voids of varying sizes have been identified in the SDSS Luminous Red Galaxy (LRG) catalog in a search for the late integrated Sachs-Wolfe (ISW) effect due to dark energy \\citep{Granett:2008ju}. How likely is the existence of such huge voids according to standard theories of structure formation? Statistical measures of the void distribution such as the void probability function and underdense probability function have been estimated from the 2dfGRS, SDSS and DEEP2 galaxy redshift surveys \\citep{Hoyle:2003hc,Croton:2004ac,Patiri:2005ys,Conroy:2005cx,Tikhonov:2006it,Tinker:2007zi,Tikhonov:2007di,vonBendaBeckmann:2007wt}. Void probability statistics have also been examined theoretically using analytical methods \\citep{Sheth:2003py,Furlanetto:2005cc,Shandarin:2005ea} and N-body simulations \\citep{Little:1993fe,Schmidt:2000md,ArbabiBidgoli:2001jp,Benson:2002tq,Padilla:2005ea}. However such studies have been restricted to voids with radii of 10-30 Mpc. The scales of the large voids we are considering lie in the linear regime where the variance of the Hubble contrast is directly related to the matter power spectrum $\\mathcal{P}_\\mathrm{m}\\left(k\\right)$. It has been noted (using results from \\citet{Turner:1992yi}) that above 100 Mpc linear theory predictions agree well with N-body simulation results, although on smaller scales the Hubble contrast is underestimated due to non-linear effects \\citep{Shi:1995nq}. Applying linear theory and using the measured CMB dipole velocity, \\citet{Wang:1997tp} obtained the model-independent result $\\langle\\delta_H\\rangle^{1/2}_R < 10.5\\,h^{-1}\\,\\mathrm{Mpc}/R$ in a sphere of radius $R$. (This ought to be an acceptable procedure up to scales of order 800 $h^{-1}$~Mpc --- on larger scales, relativistic corrections become increasingly important.) In this paper we update these results by determining the probability distribution of $\\delta_H$ and the density contrast on various scales using constraints on $\\mathcal{P}_\\mathrm{m}\\left(k\\right)$ from WMAP 5-year data \\citep{Komatsu:2008hk} and the SDSS galaxy power spectrum \\citep{Tegmark:2003uf}. We find that even the `minimal local void' is {\\em extremely} unlikely if the primordial density perturbation is indeed gaussian as is usually assumed and the other LTB model voids even less so. However by the same token, the ISW effect due to the voids seen in the SDSS LRG survey \\citep{Granett:2008ju} appears to be too strong. Moreover, observed large-scale peculiar velocities appear to be much higher than expected \\citep{Kashlinsky:2008ut,Watkins:2008hf}. It would appear that the standard model of structure formation itself needs reexamination hence the existence of a large local void cannot be dismissed on these grounds. ", "conclusions": "A void with $\\delta_H\\simeq 0.2-0.3$ and a radius exceeding $100~h^{-1}\\,\\mathrm{Mpc}$ is required to fit the supernova data without dark energy \\citep{Tomita:2001gh,Biswas:2006ub,Alexander:2007xx}. The probability that we are situated in such a void is less than $10^{-12}$ as can be seen from Fig.\\ref{prob1}. The probability is exponentially smaller for the larger voids of Gpc size that have also been considered \\citep{Alnes:2005rw,GarciaBellido:2008nz,Clifton:2008hv}.\\footnote{ There is a further constraint on Gpc scale voids from the observed absence of a `$y$-distortion' in the spectrum of the CMB \\citep{Caldwell:2007yu} and from the `kinetic Sunyaev-Zeldovich' effect observed for X-ray emitting galaxy clusters \\citep{GarciaBellido:2008gd}. However this has no impact on smaller voids.} However before we dismiss the possibility of a local void on these grounds we should also evaluate the probability of voids which have actually been claimed to exist elsewhere in the universe. For example it has been argued that a void with radius $200-300~h^{-1}\\,\\mathrm{Mpc}$ and an density contrast of $\\delta=-0.3$ at $z\\sim1$ can account for the WMAP `cold spot' in a $\\Lambda$CDM universe \\citep{Inoue:2006rd}. Even if we conservatively take the radius to be $150~h^{-1}\\,\\mathrm{Mpc}$ (and the same underdensity), the probability that one or more such voids lie within the volume out to $z=1$ is only $1.05_{-0.93}^{+5.24}\\times10^{-10}$. It has been argued that the WMAP cold spot may not be a localized feature \\citep{Naselsky:2007yd} and there may be no matching void in the NVSS radio source catalogue \\citep{Smith:2008tc}, however an equally striking anomaly arises if we consider the large number of voids which have been identified in the SDSS LRG survey in a search for the late ISW effect \\citep{Granett:2008xb,Granett:2008ju}. These are of angular radius $\\sim 4^0$ corresponding to a (comoving) radius of $\\sim 50~h^{-1}\\,\\mathrm{Mpc}$ and are tabulated as having $1\\sigma$, $2\\sigma$ or $3\\sigma$ underdensities. These numbers relate to the detection significance (the likelihood of detecting the void by chance out of a Poisson distribution) rather than the likelihood of finding such underdensities in a gaussian field which we have computed in this paper (B. Granett, private communication). Moreover the observed LRGs are biased with regard to the dark matter hence the underdensities in dark matter are likely to be smaller than the quoted values. However if \\citet{Granett:2008xb,Granett:2008ju} have indeed detected the late ISW effect as they assert, we can simply circumvent these uncertainties by requiring that the voids be large enough and/or underdense enough to yield the {\\em observed} CMB temperature decrements. To calculate the late ISW effect we consider the propagation of CMB photons to us from the last scattering surface through an intervening void. The photon temperature change caused by the void is \\begin{equation} \\frac{\\Delta T}{T} = -\\frac{2}{c^2}\\int_{a_\\mathrm{far}}^{a_\\mathrm{near}} \\frac{\\mathrm{d}\\Phi}{\\mathrm{d}a}\\,\\mathrm{d}a, \\end{equation} where $a_\\mathrm{far}$ is the scale factor when the photon crossed the far side of the void and $a_\\mathrm{near}$ is the scale factor when the photon crossed the near side of the void. The gravitational potential of a void with proper radius $r$ is \\begin{equation} \\Phi=\\frac{4\\pi G}{3}r^2\\rho_\\mathrm{b}\\delta\\left(a\\right). \\end{equation} Here the background density is given by $\\rho_\\mathrm{b}=3H^2_0\\Omega_\\mathrm{m}/8\\pi G a^3$ and the density perturbation is given by $\\delta\\left(a\\right)=D\\left(a\\right)\\delta\\left(a_0\\right)$ where $D$ is the linear growth factor. Hence \\begin{equation} \\frac{\\Delta T}{T} = \\Omega_\\mathrm{m}\\left(\\frac{R}{c/H_0}\\right)^2 \\left[\\frac{D(a_\\mathrm{far})}{a_\\mathrm{far}} - \\frac{D(a_\\mathrm{near})}{a_\\mathrm{near}}\\right]\\delta. \\label{isw} \\end{equation} Using this we calculate the expected ISW signal for the 50 highest significance voids in Table 4 of \\citet{Granett:2008xb}, employing the concordance $\\Lambda$CDM cosmology to determine $a_\\mathrm{far}$ and $a_\\mathrm{near}$ for each void from the void redshift measurements. The ISW signal is found to be only $-0.42 \\,\\mu \\mathrm{K}$ on average if the dark matter underdensities are smaller than the observed underdensities in the LRG counts by the bias factor of 2.2 (taking $\\sigma_8=0.8$). This is in contrast to the detected mean signal of $-11.3 \\,\\mu \\mathrm{K}$ which is over 20 times bigger! We must therefore conclude that the void radii and/or underdenities have been significantly underestimated. The void radii can at most be increased by a factor of 1.75 within the quoted uncertainties so the observed signal of $-11.3 \\,\\mu \\mathrm{K}$ can be matched only if the underdensities are increased by a factor of 5 (implying a bias factor of 0.2). The CMB temperature decrements of such model voids calculated using eq.(\\ref{isw}) are shown in Fig.\\ref{hist} and are (by construction) similar to the actual measurements shown in Fig.2 of \\citet{Granett:2008ju}. While such an {\\em underbias} for the observed LRGs may seem implausible, we emphasise that this is the only way in which the temperature decrements observed by \\citet{Granett:2008xb,Granett:2008ju} can be accounted for as being due to the late ISW effect. Fig.\\ref{hist} displays a histogram of the probabilities for finding such voids in the SDSS LRG survey volume ($5\\,h^{-3}\\,\\mathrm{Gpc}^{3}$ in the redshift range $0.4 < z < 0.75$). The most improbable void is at $z=0.672$ --- in order to yield the {\\em observed} average CMB temperature decrement it must have a density contrast of -0.72 (quoted galaxy underdensity of -0.316 multiplied by 5/2.2) and a radius of $230~h^{-1}\\,\\mathrm{Mpc}$ (radius derived from the quoted volume of $10^7~h^{-3}\\,\\mathrm{Mpc}^3$ and multiplied by 1.75). The probability of such a void is $1.9 \\times 10^{-247}$ according to our calculations. Although linear theory may not be applicable for such a deep void, it is clear that its existence is in gross conflict with the standard theory of structure formation from {\\em gaussian} primordial density perturbations. \\begin{figure*}% \\begin{minipage}{\\textwidth} \\centering \\begin{tabular}{cc} \\includegraphics*[angle=0,scale=0.5]{fig29.eps} & \\includegraphics*[angle=0,scale=0.5]{fig30.eps} \\end{tabular} \\caption{\\label{hist} The left panel shows the ISW signals of the 50 voids detected by \\citet{Granett:2008xb}, calculated using eq.(\\ref{isw}); in order to match the observed average ISW signal of $-11.3 \\,\\mu \\mathrm{K}$ it has been necessary to increase the void radii by a factor of 1.75 and the underdensities by a factor of 5. The right panel show the probability of such voids occurring in the SDSS LRG survey volume according to the concordance $\\Lambda$CDM model.} \\end{minipage} \\end{figure*} This conclusion is strengthened by the recent detection of very large peculiar velocities on large scales. As seen in Fig.\\ref{var}, the expected variance of the peculiar velocity as calculated by eq.(\\ref{var3}) is about 200 km~s$^{-1}$ on a scale of 100 $h^{-1}$ Mpc, whereas the measured value is at least 5 times higher, and the discrepancy is even bigger on larger scales up to 300 Mpc \\citep{Kashlinsky:2008ut}. It is also seen from Fig.\\ref{prob1} that if a determination of the Hubble constant is required with say 1\\% accuracy, then measurements extending out to at least $150~h^{-1}\\,\\mathrm{Mpc}$ must be made to overcome local fluctuations. A similar estimate was made by \\citet{Li:2008yj} who noted that the observed variance in measurements of $h$ is in accord, thus consistent with the assumption of a gaussian density field. However the voids observed in the SDSS LRG survey \\citep{Granett:2008ju} call this assumption into question. In particular whether there is a large local void is then an issue that must be addressed observationally and not dismissed on the grounds that it is inconsistent with gaussian perturbations. The Hubble flow is presently poorly measured in the redshift range $0.1 \\la z \\la 0.3$ --- just where the effects of such a local void would be most apparent \\citep{Alexander:2007xx}. Given that dark energy may well be an artifact of such a void, this issue needs urgent attention. The question of how such voids can have been generated without conflicting with the CMB observations is beyond the scope of the present work. Some suggestions have been made in the context of multi-field inflationary models \\citep{Occhionero:1997eh,DiMarco:2005zn,Itzhaki:2008ih}." }, "0807/0807.3328_arXiv.txt": { "abstract": "We report HST/NICMOS coronagraphic images of the HD 15115 circumstellar disk at 1.1\\micron. We find a similar morphology to that seen in the visible and at H band--an edge-on disk that is asymmetric in surface brightness. Several aspects of the 1.1\\micron\\ data are different, highlighting the need for multi-wavelength images of each circumstellar disk. We find a flattening to the western surface brightness profile at 1.1\\micron\\ interior to 2\\arcsec\\ (90~AU) and a warp in the western half of the disk. We measure the surface brightness profiles of the two disk lobes and create a measure of the dust scattering efficiency between 0.55-1.65\\micron\\ at 1\\arcsec, 2\\arcsec, and 3\\arcsec. At 2\\arcsec\\ the western lobe has a neutral spectrum up to 1.1\\micron\\ and a strong absorption or blue spectrum $>$1.1\\micron, while a blue trend is seen in the eastern lobe. At 1\\arcsec\\ the disk has a red F110W-H color in both lobes. ", "introduction": "Raw material from the interstellar medium is processed through circumstellar disks into many different types of planetary bodies. Debris disks are a useful environment to study the composition of planetesimals that form around stars different from our own Sun. Multi-wavelength scattered light observations of disks provide one avenue for determining the compositon of dust caused by collisions of planetesimals. HD~15115 possesses one of a growing number of spatially resolved debris disks amenable to multiwavelength observations. HD 15115 is an F2 star at a distance of 45$\\pm$1~~pc \\citep{vanl} and was observed to have an IR excess \\citep{silverstone00}. Unresolved emission from its disk was detected at 60, 100, and 850 \\micron, implying a ring at $\\sim$35~AU with a temperature of 62~K \\citep{zuckerman04,decin03,williams06}. Recently, \\citet[][hereafter KFG07]{kalas07} discovered circumstellar emission from a combination of Hubble Space Telescope/Advanced Camera for Surveys (HST/ACS) and Keck near-IR AO coronagraphic imaging. Surface brightness (SB) profiles were reported for V and H bands. Furthermore, they observed an extreme size and brightness difference between the two lobes of the disk and shape asymmetries around the disk midplane similar to that seen around $\\beta$ Pictoris and indicative of a second disk at a different inclination \\citep{kalas95,golimowski06}. We have imaged HD 15115 at 1.1\\micron\\ with HST; we present our observations in \\S\\ref{s:obs}, and analyze the results in \\S\\ref{s:analysis}. We discuss the implications of our work in \\S\\ref{s:conc} ", "conclusions": "\\label{s:conc} The new features we have observed in the disk at 1.1\\micron\\ add questions to the nature of the disk and the dust. The increase in PA towards the inner part of the disk could be caused by a genuine warp, or perhaps the presence of a second disk inclined at $>$12$^\\circ$ from the main disk as is seen in $\\beta$-Pictoris. This is supported by the detection of an asymmetry in the northern part of the disk seen in the ACS and H data of KFG07. KFG07 noted an overall blue V-H color for the Western lobe of the disk. The presence of a neutral scattering efficiency out to 1.1\\micron\\ as well as a red F110W-H color interior to 2\\arcsec\\ suggests that the spectrum of the HD~15115 disk is more complicated than the V and H data suggest. The abrupt change in scattering efficiency beyond 2\\arcsec\\ from F110W to H may be due to the presence of strong absorption at $\\sim$1.65\\micron. One candidate would be strong absorption due to water ice, something not yet detected in the scattered light of a disk. If that is the case, even deeper absorption should be present at 2 and 3.5\\micron. Two possibilities exist to explain the red F110W-H color of the inner disk--either there is evidence of strong absorption due to olivine (which has an absorption feature at 1\\micron), or the presence of a very red material in the disk, possibly similar to what is seen in HD~100546 \\citep{ardila07} or HR~4796A \\citep{debes08}. The age of HD~15115 is not well constrained, which would be helpful in understanding the origins of the debris disk. Kinematically, it is marginally consistent with membership in the 12~Myr-old $\\beta$-Pictoris Moving Group (BPMG) \\citep{moor06}. However, this is not borne out by backtracking the position of HD~15115 based on its proper motion and radial velocity using the methods of \\citet{ortega02} and \\citet{song03}, nor is kinematics a reliable selector of a moving group \\citep{song02}. Most other indicators point to a $>$ 100~Myr age. \\citet{nordstrom04} determined an age of 900~Myr, but this age is uncertain given their methodology--the lower limit for the age is 0, while the upper limit is 2.2~Gyr. Ca~II H and K line indicators point to an age of 500~Myr \\citep{silverstone00}. Although not an accurate age indicator for early F-type stars, the Lithium 6708\\AA\\ feature shows an equivalent width of 40~m\\AA\\ (Song~2008, in prep.), which is consistent with $\\sim$100~Myr Pleiades stars. As a comparison, previously recognized 30~Myr old (Tucana-HorA) stars or younger early F-type stars all show equivalent widths for Li of $\\sim$100m\\AA. The evidence to date points to an older age for HD~15115. HD~15115 is a prime example of the potential complexity of nearby circumstellar debris disks. It possesses a strange brightness asymmetry in its outer reaches that extends into about 2\\arcsec (90~AU). This asymmetry, however, is a function of wavelength, and at least part of the cause may be compositional or grain population differences between the two sides of the disk. KFG07 investigated whether the close passage of a star might explain the strange morphology of the disk. They found a candidate object, HIP~12545, that might have passed very close to the debris disk of HD~15115, but while its motion is consistent with being a co-moving object, backtracking of the the two star's positions back in time based on their Hipparcos/Tycho-2 astrometry and radial velocities shows that they are more widely separated in the past than they are now and could not have interacted. The fact that we measure a warp only in the western part of the disk around 1\\arcsec\\ suggests that HD~15115 has an even more complex structure. It would be useful to confirm the presence of this structure in the disk at other wavelengths. This is not easily done in the visible, unless STIS is refurbished in the next servicing mission--it is the only instrument able to image to within $\\sim$1\\arcsec\\ in the visible. Follow-up H and K (or F160W or F205W images with HST) will be useful to look for the warp as well as constrain the ultimate nature of the composition of the grains in the disk. If water ice is indeed present, it will show strong absorption at around 2 and 3.5\\micron. Therefore, K and L band images of the scattered light disk are cruicial. What are the other possibilities to explain both the warp, the asymmetry, and the varying nature of the dust grains? One possibility is that the asymmetry is caused by a relatively recent collision that has injected many particles recently into the disk. Such an event is rare given the collisional times of the parent bodies and the timescale for such an event to be evenly spread in azimuth. Similar asymmetries have been seen in the mid-IR with $\\beta$-Pictoris \\citep{weinberger03}, and it is striking that HD~15115 shares many of traits with this well known disk." }, "0807/0807.0810_arXiv.txt": { "abstract": "Measuring the statistics of galaxy peculiar velocities using redshift-space distortions is an excellent way of probing the history of structure formation. Because galaxies are expected to act as test particles within the flow of matter, this method avoids uncertainties due to an unknown galaxy density bias. We show that the parameter combination measured by redshift-space distortions, $f\\sigma_8^{\\rm mass}$ provides a good test of dark energy models, even without the knowledge of bias or $\\sigma_8^{\\rm mass}$ required to extract $f$ from this measurement (here $f$ is the logarithmic derivative of the linear growth rate, and $\\sigma_8^{\\rm mass}$ is the root-mean-square mass fluctuation in spheres with radius $8h^{-1}$Mpc). We argue that redshift-space distortion measurements will help to determine the physics behind the cosmic acceleration, testing whether it is related to dark energy or modified gravity, and will provide an opportunity to test possible dark energy clumping or coupling between dark energy and dark matter. If we can measure galaxy bias in addition, simultaneous measurement of both the overdensity and velocity fields can be used to test the validity of equivalence principle, through the continuity equation. ", "introduction": "Observations of the accelerating nature of the Universe show that there is fundamental physics at work that we do not understand \\cite{riess98,perlmutter99}. Many possibilities have been postulated including a new form of the vacuum which is not present in the contemporary high energy physics, or a modification of gravity which would revolutionize our understanding of space and time. Discovering which mechanism is correct is one of the key challenges for 21st century science. The detection of acceleration was obtained by using supernovae as standard candles, and therefore relies on measuring the cosmological geometry. The physical process causing the acceleration could also affect structure formation, which provides a complementary way of distinguishing between models. In particular, models in which general relativity is unmodified have different Large-Scale Structure (hereafter LSS) formation timescales compared with Modified Gravity (hereafter MG) models~\\cite{dvali00,carroll03}. Because of this, many previous studies have considered testing the signature on LSS formation predicted by MG models~\\cite{Lue:2003ky,ishak05,jain07}. Direct observations of LSS growth as traced by galaxies are of limited value because galaxies are not expected to be simple tracers of the underlying matter density field, although we return to this point later. Maps of galaxies where distances are measured from spectroscopic redshifts show anisotropic deviations from the true galaxy distribution. These differences arise because galaxy recession velocities include components from both the Hubble flow and peculiar velocities from the motions of galaxies in comoving space. Although these ``redshift-space distortions'' are a nuisance when trying to reconstruct the true distribution of galaxies, they provide a mechanism to measure the build-up of structure, which drives these peculiar velocities on large-scales. In linear theory, and in the absence of bias, a distant observer should expect a multiplicative enhancement of the overdensity field $\\delta$ that is proportional to $1+f\\mu^2$, where $f$ is the logarithmic derivative of the linear growth rate, and $\\mu$ is the cosine of the angle to the line-of-sight \\cite{kaiser87}. With a local linear bias, the real-space galaxy density field is affected, while the peculiar velocity term is not, so the multiplicative factor is changed to $1+\\beta\\mu^2$, where $\\beta\\equiv f/b$. Because of the $\\mu$ dependence, this information can be extracted from galaxy redshift surveys, and a number of methods and their application have previously been considered. Analyses have been undertaken using the 2degree-Field Galaxy Redshift Survey (2dFGRS) \\cite{colless03}, measuring redshift-space distortions in both the correlation function \\cite{peacock01,hawkins03} and power spectrum after decomposing into an orthonormal basis of spherical harmonics and spherical Bessel functions \\cite{percival04}. Using the Sloan Digital Sky Survey \\cite{York00}, an Eigenmode decomposition has been performed to separate real and redshift-space effects \\cite{tegmark04,tegmark06}. In a recent paper, these low redshift analyses were extended to $z\\simeq1$ \\cite{guzzo08} using the VIMOS-VLT Deep Survey (VVDS) \\cite{lefevre05,garilli08}. In addition to measuring $\\beta$ at $z=0.8$, this work has pushed explicitly the idea of using large-scale peculiar velocities for constraining models of cosmic acceleration. These data, and the resulting cosmological constraints, are considered further in Section~\\ref{sec:current}. In particular we argue that it would be better to present results in terms of $b(z)\\sigma_8(z)$ and $f(z)\\sigma_8(z)$, rather than $\\beta$ for local linear bias models. We show in Section~\\ref{sec:pec_vel_meas} that bias-independent constraints on $f(z)\\sigma_8(z)$ are able to discriminate between some models of acceleration as well as $f(z)$, which is commonly extracted from $\\beta$ by applying an independent (and difficult) measurement of bias. Galaxy bias measurements tend to have the same fractional error as the redshift-spce distortion measurements: for example comparing redshift-space distortion results from the 2dFGRS\\cite{hawkins03} with bias measurements from measurements of the 3-pt function\\cite{verde02}. In addition to a direct measurement of $f(z)\\sigma_8(z)$, redshift-space distortion measurements can be used to test diverse aspects of LSS, as proposed by Song $\\&$ Koyama~\\cite{song08a}: geometrical perturbations can be reconstructed from the evolution history of peculiar velocities. With the assumption of an additional measurement of galaxy bias, the continuity equation can be tested, and anisotropic stress can be constrained. Those diverse tests strengthen our power to constrain theoretical models, and are considered further in Section~\\ref{sec:further_tests}. Before we do this, we first review the physics that we hope to test using peculiar velocities (Section~\\ref{sec:models}), and then consider the measurements themselves and what they can directly tell us about structure formation (Section~\\ref{sec:pec_vel_meas}). ", "conclusions": "One of the best ways of measuring this structure growth is to use observations of redshift-space distortions. This is an interesting and timely subject given current observations and those planned on a 10--20 year timescale \\cite{Wang08,sapone07,guzzo08}. In this paper we have reviewed the importance of redshift-space distortion measurements given that they provide a measurement of structure growth that is independent of galaxy density bias. We have argued that peculiar velocity measurements are best presented in terms of $\\sigma_{8}^{\\theta}$, or $f\\sigma_8^{\\rm mass}$ for models where the continuity equation holds. The independence from galaxy density bias has not been widely covered in previous literature. Most previous analyses have considered measuring $\\beta$ and $b\\sigma_8^{\\rm mass}$, although there are some exceptions\\cite{percival04}. Although extremely simple, we have focused on the density bias independent constraint resulting from multiplying these together. The physical origin of such a constraint is that linear velocities, which scale with the derivative of the growth factor, depend only on the matter velocity field. The primary conclusion from our work is that constraints on $\\sigma_{8}^{\\theta}$ or $f\\sigma_8^{\\rm mass}$, are extremely good at helping to distinguish between the dark energy models that we reviewed in Section~\\ref{sec:models}. In fact, as we show in Fig.~\\ref{fig:f_v}, these constraints are equivalent to similar percent measurements of $f$ for some models of cosmic acceleration. They also have the simplicity of not having to model galaxy bias. The simple formula of \\cite{guzzo08} has been adapted to determine constraints on $f\\sigma_8^{\\rm mass}$, and shows future constraints in Fig.~\\ref{fig:sigma8_theta}. Although the \\cite{guzzo08} formula is simplistic, and might not be believed at the percent level, it shows that we can expect a huge step forwards in redshift-space distortion measurements with the next generation of surveys. Going beyond simply obtaining a single measurement of $\\sigma_{8}^{\\theta}$ or $f\\sigma_8^{\\rm mass}$, we have considered how the underlying perturbation evolution can be tested using peculiar velocity measurements. Peculiar velocity measurements are important because they can be used to reconstruct Newtonian potential $\\Psi$ which sources the dynamics of a galaxy given by Euler equation. Weak-lensing only measures $\\Psi$ in the combination $\\Phi-\\Psi$, so redshift-space distortions offer a complementary test of perturbations. We have considered how peculiar velocities can be used to test the continuity equation, which is worthwhile since there are many theoretical models which fail to satisfy this relation. If dark energy couples to matter, then current constraints show that it must couple to the dark matter and not to baryonic material. The coupling of dark energy to dark matter modifies the Euler equation for dark matter, and breaks the equivalence principle between dark matter and baryon. This difference in free-fall breaks the continuity equation in which the peculiar velocity of matter is estimated using baryons, while we consider the growth of fluctuations in all matter. In addition, this test can tell if there is dark energy clustering which deepens the curvature potential well because we measure the peculiar velocity of the matter not of the total energy density. Finally we have considered how we can constrain the anisotropic stress by comparing $\\Phi$ and $\\Psi$, reconstructed from the density fields and peculiar velocity respectively." }, "0807/0807.2354_arXiv.txt": { "abstract": "The purpose of this work is the computation of the cosmic Type Ia supernova rates, namely the frequency of Type Ia supernovae per unit time in a unitary volume of the Universe. Our main goal in this work is to predict the Type Ia supernova rates at very high redshifts and to check whether it is possible to select the best delay time distribution model, on the basis of the available observations of Type Ia supernovae. We compute the cosmic Type Ia supernova rates in different scenarios for galaxy formation and predict the expected number of explosions at high redshift ($z\\geq 2$). Moreover, we adopt various progenitor models in order to compute the Type Ia supernova rate in typical elliptical galaxies of initial luminous masses of $10^{10}$M$_{\\odot}$, $10^{11}$M$_{\\odot}$ and $10^{12}$M$_{\\odot}$, and compute the total amount of iron produced by Type Ia supernovae in each case. In this analysis we assume that Type Ia supernovae are caused by thermonuclear explosions of C-O white dwarfs in binary systems and we consider the most popular frameworks: the single degenerate and the double degenerate scenarios. The two competing schemes for the galaxy formation, namely the monolithic collapse and the hierarchical clustering, are also taken into account, by considering the histories of star formation increasing and decreasing with redshift, respectively. We calculate the Type Ia supernova rates through an analytical formulation which rests upon the definition of the SN Ia rate following an instantaneous burst of star formation as a function of the time elapsed from the birth of the progenitor system to its explosion as a Type Ia supernova (i.e. the delay time). What emerges from this work is that: \\emph{i)} we confirm the result of previous papers that it is not easy to select the best delay time distribution scenario from the observational data and this is because the cosmic star formation rate dominates over the distribution function of the delay times; \\emph{ii)} the monolithic collapse scenario for galaxy formation predicts an increasing trend of the SN Ia rate at high redshifts (mainly due to the contribution by massive spheroids), whereas the predicted rate in the framework of a decreasing cosmic star formation rate, more in agreement with the hierarchical scenario, drops dramatically at high redshift; \\emph{iii)} for the elliptical galaxies we note that the predicted maximum of the Type Ia supernova rate depends on the initial galactic mass. The maximum occurs earlier (at about 0.3 Gyr) in the most massive ellipticals, as a consequence of the assumed downsizing in star formation. In addition, we find that the Type Ia supernova rate per unit mass at the present time is higher in bluer ellipticals (i.e. the less massive ones). ", "introduction": "The supernovae (SNe) of Type Ia are fundamental for understanding a number of astrophysical problems of primary importance, such as \\emph{i)} the SN progenitors, \\emph{ii)} the determination of cosmological parameters, \\emph{iii)} the chemical enrichment of galaxies and \\emph{iv)} the thermal history of the interstellar (ISM) and intracluster medium (ICM). The evolution of the rate of Type Ia SNe with cosmic time is a fundamental ingredient for the study of all these issues. The observed features of SNe Ia suggest that the majority of these objects may originate from the thermonuclear explosion of a C-O white dwarf (WD) of mass $\\sim 1.4$ M$_{\\odot}$ (Chandrasekhar mass) in binary systems (Chandra exploders). So each SN Ia should be the result of the explosion of the same mass. However, Phillips (1993) pointed out that there is a significant intrinsic dispersion in the absolute magnitudes at maximum light of local Type Ia SNe. This result was interpreted to arise from a possible range of masses of the progenitors or from variations of the explosion mechanism. \\\\ Here we will consider only Chandra exploders for which two scenarios have been proposed: a) The \\emph{Single Degenerate} (SD) scenario, i.e. the accretion of matter via mass transfer from a non-degenerate companion, a red giant or a main sequence star (e.g. Whelan \\& Iben 1973). In this scenario the mass range for the secondary components of the binary system is $0.8-8$M$_{\\odot}$, while the primary masses should be in the range $2-8M_{\\odot}$. The upper limit is given by the fact that stars with masses $M>8M_{\\odot}$ ignite carbon in a non-degenerate core and do not end their lives as C-O WDs. The lower limit is instead obviously due to the fact that we are only interested in systems which can produce a Type Ia SN in a Hubble time. The clock to the explosion is given by the lifetime of the secondary component. b) The \\emph{Double Degenerate} (DD) scenario, i.e. the merging of two C-O WDs which reach the Chandrasekhar mass and explode by C-deflagration (e.g. Iben \\& Tutukov 1984). The merging is due to the loss of orbital angular momentum due to gravitational wave radiation. In the Iben \\& Tutukov paper, the progenitor masses were defined in the range 5-8M$_{\\odot}$ to ensure two WDs of $\\sim 0.7$M$_{\\odot}$ and then reach the Chandrasekhar mass. The clock for the explosion in this model is given by the lifetime of the secondary star plus the time necessary to merge the system due to gravitational wave radiation. This scenario requires the formation of two degenerate C-O WDs at an initial separation less than $\\sim 3 R_{\\odot}$ and this can occur by means of two different precursor systems: a close binary, and a wide binary. The two scenarios differ for the efficiency of the common envelope phase during the first mass transfer, and therefore for the separation attained at the end of the first common envelope phase. Different arguments can be found in favor or against both SD and DD scenarios and the issue of DD vs. SD is still debated (e.g. Branch et al. 1995, Napiwotzki et al. 2002, Belczynski et al. 2005). The Type Ia SN rate is the convolution of the distribution of the explosion times, usually called the time delay distribution function (DTD), with the star formation history. Several attempts of comparing the Type Ia SN rate and the cosmic star formation rate evolution with redshift have already appeared (e.g. Gal-Yam \\& Maoz, 2004; Dahlen et al. 2004; Cappellaro et al. 2005; Neill et al. 2006; Forster et al. 2006; Barris \\& Tonry, 2006; Poznansky et al. 2007; Botticella et al. 2008, Blanc \\& Greggio, 2008; among others), but no clear conclusions arised yet on this point.\\\\ In this paper, besides the above mentioned scenarios for Type Ia SN progenitors we will test the empirical DTD suggested by Mannucci et al. (2006),together with different star formation histories. In particular, we will study the Type Ia SN rate in ellipticals and different cosmic star formation rates, including a strongly increasing cosmic star formation with redshift, based on the monolithic scenario of galaxy formation and never considered in previous studies.\\\\ The paper is organized as follows: in Section 2 the adopted formulation for the SN Ia rate is presented, together with the description of the adopted star formation histories. In section 3 the models for the distribution function of the delay times are presented and in section 4 the predicted SN Ia rates are discussed and compared with the availale data. Finally, in section 5 some conclusions are drawn. ", "conclusions": "We calculated the cosmic SN Ia rate densities (i.e. the rate per unit comoving volume) and the rate of the explosion of SNe Ia in typical elliptical galaxies adopting the formalism proposed by G05. This formulation rests upon the definition of the SN Ia rate following an instantaneous burst of star formation as a function of the time elapsed from the birth of the progenitor system to its explosion as a Type Ia SN (i.e. the delay time). This function has been termed as DTD and accounts for the SNe Ia progenitor scenarios and for the initial mass function (IMF). Different DTDs and histories of SF have been considered. In all cases, a Salpeter IMF was adopted. For all galaxies it was assumed a galaxy formation epoch at redshift $z_{f}=6$. The reason for chosing $z_f=6$ is given by the fact that the SFR density has been measured up to this redshift (Cole et al. 2001). We have also shown the results obtained assuming $z_{f}=10$ as epoch of galaxy formation. The evolution of the cosmic SN rate with redshift contains, in principle, unique information on the star formation history of the Universe, the IMF of stars, and the Type Ia SN progenitors. These are essential ingredients for understanding galaxy formation, cosmic chemical evolution, and the mechanisms which determined the efficiency of the conversion of gas into stars in galaxies at various epochs (e.g. Madau et al. 1996; Madau, Pozzetti, \\& Dickinson 1998; Renzini 1997). We computed the cosmic Type Ia SN rates for different cosmic SFR histories. Our main result is the prediction of the expected number of explosions of Type Ia SNe at high redshift ($z\\geq 2$). Here we summarize our conclusions in detail: \\begin{itemize} \\item we analyzed the effects of the various parameters entering the computation of the SN Ia rate and concluded that the realization probability $A_{Ia}$ (the actual fraction of systems which are able to give rise to a Type Ia SN) should be in the range $10^{-4}-10^{-3}$, where the lower value is appropriate for the DD model. This quantity, in all models, is a free parameter and it was chosen by reproducing the present time Type Ia SN rate in galaxies. Other parameters also play a key role in the computation of the Type Ia SN rate and these are: \\emph{i)} the mass range for the secondaries; \\emph{ii)} the minimum mass for the primaries; \\emph{iii)} the efficiency of the accretion and \\emph{iv)} the distribution of the separations at birth. Actually, all the different channels could contribute to the SN Ia events, each with its own probability ($A_{Ia}$). \\item We computed the Type Ia SN rate for specific elliptical galaxies of different initial luminous masses ($10^{10}$M$_{\\odot}$, $10^{11}$M$_{\\odot}$, $10^{12}$M$_{\\odot}$) for all the studied DTDs and for the SFRs suggested by chemical evolution models which best reproduce the characteristics of local ellipticals (Pipino \\& Matteucci 2004). All the DTDs predict an early maximum in the Type Ia SN rate, between 0.3 and 1 Gyr, according to the galaxy mass. In particular, the maximum predicted at about 0.3 Gyr corresponds to a typical elliptical of initial mass of $10^{12}$M$_{\\odot}$ while, for an elliptical of initial mass of $10^{10}$M$_{\\odot}$ the peak is at about 1 Gyr. The dependence of the maximum on the initial galactic mass is due to the efficiency of star formation, which is assumed to be higher for the most massive ellipticals . We have also considered the predicted Type Ia SN rates per unit mass (SNuM) at the present time vs. $(B-K)$ color relations for the three different ellipticals. We found that the bluer is the color (and hence the lower is the galactic mass), the higher are the predicted SN Ia rates in SNuM at the present time. Unfortunately, a real comparison with data is not possible since the data for ellipticals of different masses are not available. \\item We computed the cosmic Type Ia SN rate in a unitary volume of Universe by adopting different cosmic SFRds, as predicted by both models and observations (Hopkins \\& Beacom, 2006). We found that the SF history largely dominates over the assumed DTD in the calculation of the Type Ia SN rate. Therefore, unless we know the cosmic star formation history, we cannot safely decide which DTD is better on the basis of the observed cosmic SN Ia rate, in agreement with previous works (see also Forster et al. 2006; Blanc \\& Greggio 2008, Botticella et al. 2008). In particular, it is not possible to predict the delay time for the explosion of SNe Ia on the basis of the cosmic star formation rate, since for example, in the hierarchical framework, it is predicted a decrease of the cosmic Type Ia SN rate for $z>1$ irrespective of the chosen DTD, even for that of Mannucci et al. (2005; 2006) DTD, where the fraction of prompt Type Ia SNe is 50$\\%$! \\item The cosmic Type Ia SN rates for different SFRds (increasing, decreasing, constant with redshift) differ mostly for redshift $z>1$. We compared our results with the available data, although the highest redshift points are still very uncertain. Therefore, it was not possible to decide which scenario should be preferred, but high redshift data will greatly help to draw conclusions on this point. In particular, we predict that the Type Ia SN rate for $z>2$ should be very high if the monolithic scheme for the formation of ellipticals is assumed and even higher if combined with the empirical DTD of Mannucci et al. (2005), as opposed to models with decreasing or constant SFRds, which all predict a significant drop of the Type Ia SN rate for any chosen DTD. Future observations with JWST of high redshift SNe will help in sheding light on this subject. \\end{itemize}" }, "0807/0807.2448_arXiv.txt": { "abstract": "We present the 3-point function $\\xi_3$ and $Q_3=\\xi_3/\\xi_2^2$ for a spectroscopic sample of luminous red galaxies (LRG) from the Sloan Digital Sky Survey DR6 \\& DR7. We find a strong (S/N$>$6) detection of $Q_3$ on scales of 55-125 Mpc/h, with a well defined peak around 105 Mpc/h in all $\\xi_2$, $\\xi_3$ and $Q_3$, in excellent agreement with the predicted shape and location of the imprint of the baryon acoustic oscillations (BAO). We use very large simulations (from a cubic box of L=7680 Mpc/h) to asses and test the significance of our measurement. Models without the BAO peak are ruled out by the $Q_3$ data with 99\\% confidence. This detection demonstrates the non-linear growth of structure by gravitational instability between $z=1000$ and the present. Our measurements show the expected shape for $Q_3$ as a function of the triangular configuration. This provides a first direct measurement of the non-linear mode coupling coefficients of density and velocity fluctuations which, on these large scales, are independent of cosmic time, the amplitude of fluctuations or cosmological parameters. The location of the BAO peak in the data indicates $\\Omega_m =0.28 \\pm 0.05$ and $\\Omega_B=0.079 \\pm 0.025$ (for $h=0.70$) after marginalization over spectral index ($n_s=0.8-1.2$) linear $b_1$ and quadratic $c_2$ bias, which are found to be in the range: $b_1=1.7-2.2$ and $c_2=0.75-3.55$. The data allows a hierarchical contribution from primordial non-Gaussianities in the range $Q_3=0.55-3.35$. These constraints are independent and complementary to the ones that can be obtained using the 2-point function, which are presented in a separate paper. This is the first detection of the shape of $Q_3$ on BAO scales, but our errors are shot-noise dominated and the SDSS volume is still relatively small, so there is ample room for future improvement in this type of measurements. ", "introduction": "The galaxy three-point function $\\xi_3$ provides a crucial test and valuable statistical tool to investigate the origin of structure formation and the relationship between galaxies and dark matter (see \\citet{bernar} for a review). We will concentrate here on the reduced 3-point function $Q_3 \\simeq \\xi_3/\\xi_2^2$ defined in Eq.\\ref{fiftheq} as the scaling expected from non-linear couplings (with $Q_3 \\simeq 1$). Measurements of the three-point function and other higher-order statistics in galaxy catalogs have a rich history (eg \\citet{peeblesgroth}, \\citet{frypeebles}, \\citet{baumgartfry}, \\citet{gazta1992}, \\citet{bouchet1993}, \\citet{frygazta1994}). In the past decade, three-point statistics have confirmed the basic picture of gravitational instability from Gaussian initial conditions (\\citet{friemangazta1994}, \\citet{jingborner1998}, \\citet{friemangazta1999}, \\citet{feldman2001}). The connection between these observables and theoretical predictions is best done on large scales, where the physics (gravity) is best understood, but the surveys previously available were not large enough to have good statistics on sufficiently large scales. With the completion of large redshift surveys such as 2dFGRS (\\citet{colless2001}) and SDSS (\\citet{York00}) we expect measurement of higher-order statistics to provide tighter constraints on cosmology (\\citet{colombietal1998}, \\citet{szapudietal1999}, \\citet{matarreseetal1997}, \\citet{scoccimarro2004}, \\citet{sefusatti2005}). First measurements of the redshift space $\\xi_3$ in the 2dFGRS (\\citet{gazta2005}) and SDSS (eg \\citet{nichol2006}) show good agreement with expectations (see also \\citet{kayo2004}, \\citet{nishimichi2007}, \\citet{kulkarni2007} and references therein). We will assume here that the initial conditions are Gaussian. Current models of structure formation predict a small level of initial non-Gaussianities that we can neglect here. A popular parametrization of this effect is to assume that initial curvature perturbations, given by the gravitational potential, $\\Phi$, are given by $\\Phi = \\Phi_L + f_{NL} ~(\\Phi_L^2 - <\\Phi_L^2>)$, where $\\Phi_L$ is Gaussian and $f_{NL}$ is a non-linear coupling parameter of order unity. This produces non-Gaussianities in the matter density perturbations at wavenumber $k$ which, using the Poisson equation, are suppressed by the square of the horizon scale $k_H \\equiv H_0/c$, so that: $Q_3 \\simeq 3 f_{NL} (k_H/k)^2 T(k)/D(a) $, where $T(k) \\simeq 1$ is the so called CDM transfer function and $D(a)$ is the growth factor. In our analysis $(k_H/k)^2 \\simeq 10^{-3}$ so that these type of primordial non-Gaussianities produce negligible contribution to $Q_3$ for models with $f_{NL} \\simeq 1$. A more detailed analysis of this will be presented elsewhere (see \\citet{sefusattikomatsu} for a detailed forecast for this model). If the non-Gaussianities come from a non-linear coupling in the matter density field rather than in the gravitational potential, the resulting 3-point function will have a non-Gaussian contribution similar to that produced by non-linear bias $c_2$ (ie see Eq.\\ref{eq:Q3G} below and Conclusions). The shape and amplitude of $Q_3$ depends on galaxy bias, ie how galaxy light traces the dark matter (DM) distribution. This is both a problem and an opportunity. A problem because biasing can confuse our interpretation of the observations. An opportunity because one can try to measure the biasing parameters out of $Q_3$. This idea was first proposed by \\citet{frygazta1993} and has been applied to the 3-point function and bispectrum of real data (\\cite{friemangazta1994,fry1994,feldman2001,scoccimarro2001,verdeetal2002,gazta2005,nichol2006}). This paper is the third on a series of papers on clustering of LRG. In the first two papers \\cite{paper1,paper2} we studied redshift space distortions on the 2-point correlation function. The reader is referred to these papers for more details on the LRG samples, simulations and the systematic effects. Similar LRG samples from SDSS have already been used by different groups to study the 2-point function (eg \\cite{detection,hutsia,percival,padmanabhan2007,blake}) and found good agreement with predictions in the BAO scales, where density fluctuations are at a level of few percent. This is encouraging and indicates that this sample is large and accurate enough to investigate clustering on such large scales. In this paper we follow closely the methodology presented in 3 previous analysis. \\citet{barrigagazta2002} presented a comparison of the predictions for the two and three-point correlation functions of density fluctuations, $\\xi_2$ and $\\xi_3$, in gravitational perturbation theory (PT) against large Cold Dark Matter (CDM) simulations. Here we use the same method and codes to estimate the clustering in simulations. \\citet{gaztascocci2005} extend these results into the non-linear regime and focus on the effects of redshift distortions and the extraction of galaxy bias parameters in galaxy surveys. \\citet{gazta2005} apply this methodology to the 2dFGRS. Here we apply the very same techniques to the LRG data, so the reader is referred to these papers for more details. \\citet{kulkarni2007} have also estimated $\\xi_3$ using LRG galaxies from SDSS DR3, but focusing on smaller scales. We use DR6 which has 3 times the area (and volume) of DR3. We also use a volume limited sample and a different estimator for the correlation functions and errors, focusing on the largest scales. ", "conclusions": "We have studied the large scale 3-point correlation function for luminous red galaxies from SDSS, and particularly the reduced $Q_3=\\xi_3/\\xi_2^2$, which measures the scaling expected from non-linear couplings. We find a well-detected peak at 105Mpc/h separation that is in agreement with the predicted position of BAO peak. This detection is significant since it is also imprinted in $\\xi_2$ and $\\xi_3$ separately. We focus our interpretation in $Q_3$ because it is a measure independent of time, $\\sigma_8$ or growth factor. It only depends on the shape of the initial 2-point function and the non-linear coupling of the gravitational interaction. Our result for $Q_3$ is in excellent agreement with predictions from Gaussian initial conditions. When we use the $Q_3$ data alone (with no fit to the 2-point function) we are able to break the strong degeneracy between $\\Omega_m$ and $\\Omega_B$ (see Fig.\\ref{fig:OmOb}). Our detection shows a clear preference for a high value of $\\Omega_B=0.079 \\pm 0.025$. This value is larger, but still consistent at 2-$\\sigma$ with recent results of WMAP ($\\Omega_B=0.045$). At 3-sigma level, the $\\Omega_m-\\Omega_B$ becomes degenerate. Models with no BAO peak are ruled out at 99\\% confidence level. We have used very large realistic mock simulations to study the errors. These simulations show that $Q_3$ is not significantly modified in redshift space so we can use real-space perturbation theory (see Fig.\\ref{fig:q3sim}). This agreement also indicates that loop corrections are small on BAO scales \\cite{Bernardeau08}. This analysis is independent from 2-point statistics, which tests the linear growth of gravity, since 3-point statistics test the non-linear growth. A high value for $\\Omega_B$ is also consistent with the analysis of the peak in the 2-point correlation function shown in \\citet{detection} and in Paper IV (\\citet{paper4}) of this series, which detect a slightly higher peak than expected. We have done all the analysis with just one set of triangle configurations, with fixed sides of $r_{12}=33 \\pm 5.5$ Mpc/h and $r_{13}=88 \\pm 5.5$ Mpc/h, to center our attention to BAO scale. This is about optimal, but we notice that there is much more to learn from $Q_3$, which will be presented in future analysis. Results on smaller scales are consistent with what we find here. Data is in excellent agreement with Gaussian initial conditions, for which $Q_3=0$. But note that our quadratic bias detection $c_2=0.75-3.55$ is degenerate with a primordial non-Gaussian (hierarchical) contribution. Indeed the mean value of $c_2 \\simeq 2$ seems larger in observations than in halo simulations, for which we find $c_2^h \\simeq 0.2$. We believe that this indicates that halos are sometimes occupied by more than one galaxy, which increases the effective value of $c_2$ \\cite{bhalo}. But if we are conservative we can not rule out a primordial non-Gaussian contribution in the range $Q_3(Primordial)= c_2-c_2^h=[0.55,3.35]$. Note that we have pixelized our data in cubical cells of side $dr=11$ Mpc/h. This results in some lost of small scale information but allows for a very fast method to estimate 3-point function \\cite{barrigagazta2002}. This is important for data, but more for simulations. In the MICE7680 simulation there are close to $N=10^{11}$ particles. A brute force method to estimate 3-point correlation would require $N^3=10^{33}$ operations, while our method based on pixels just needed $5 \\times 10^{12}$ operations. \\begin{figure}[t] \\includegraphics[width=7cm]{figures/fitOmObPAU.eps} \\caption{Same as Fig.\\ref{fig:OmObsim} for a future photometric Survey with photo-z error of $\\Delta z<0.003 (1+z)$, volume of $V = 10 Gpc^3/h^3$ and number density of $\\bar{n}=10^{-3} h^3/Mpc^3$ LRG galaxies.} \\label{fig:OmObPAU} \\end{figure} Future surveys will be able to improve much upon our measurement here. A photometric survey with $\\Delta z<0.003 (1+z)$ precision (corresponding to $dr<9$ Mpc/h at z=0), such as in the PAU Survey \\cite{PAU} should have enough spatial resolution to measure $Q_3(\\alpha)$ as presented in this paper (recall that we are binning our radial distances in $dr=11$ Mpc/h). Such survey could sample over 10 times the SDSS DR6 volume (ie to z=0.9) with 20 times better LRG number density (ie for $L>L_*$). Fig.\\ref{fig:OmObPAU} shows the forecast for such a survey, which we have simulated with the MICE7680 mocks in redshift space with a photo-z of $\\Delta z<0.003 (1+z)$ and for the same triangles as shown in our SDSS analysis. This is just illustrative, as we have not marginalized over biasing and other cosmological uncertainties. But note that the improvement is substantial and shows the potentiality of this method to constrain cosmological parameters and models of structure formation. EG wish to thank Bob Nichol for suggesting the test with the EH no-wiggle model. We acknowledge the use of simulations from the MICE consortium (www.ice.cat/mice) developed at the MareNostrum supercomputer (www.bsc.es) and with support form PIC (www.pic.es), the Spanish Ministerio de Ciencia y Tecnologia (MEC), project AYA2006-06341 with EC-FEDER funding, Consolider-Ingenio PAU project CSD2007-00060 and research project 2005SGR00728 from Generalitat de Catalunya. AC acknowledge support from the DURSI department of the Generalitat de Catalunya and the European Social Fund." }, "0807/0807.2481_arXiv.txt": { "abstract": "In traditional seeing-limited observations the spectrograph aperture scales with telescope aperture, driving sizes and costs to enormous proportions. We propose a new solution to the seeing-limited spectrograph problem. A massively fiber-sliced configuration feeds a set of small diffraction-limited spectrographs. We present a prototype, tunable, J-band, diffraction grating, designed specifically for Astronomical applications: The grating sits at the heart of a spectrograph, no bigger than a few inches on a side. Throughput requirements dictate using tens-of-thousands of spectrographs on a single 10 to 30 meter telescope. A full system would cost significantly less than typical instruments on 10m or 30m telescopes. ", "introduction": "\\label{sec:intro} The next generation of large telescopes is designed, among other scientific requirements, to measure fundamental properties of galaxies from high-redshift ($z$) to today. However, the instrumentation cost for the next generation of telescopes will be enormous. Here, we present an approach that combines several modern technologies to significantly reduce the cost of astronomical spectrographs. For any two functionally equivalent spectrographs, their volume increases, roughly, with the telescope diameter cubed\\cite{Schroeder87}. There is a power-law relationship between telescope diameter (volume) and instrument cost, where the exponent sits between two and three. For example, DEIMOS, on a 10m telescope cost US\\$10m dollars\\cite{faber03} (though it was originally budgeted for less.) The preliminary estimate for an instrument with less multiplexing capability, on a 30m telescope, is US\\$60m\\cite{paz06}. This growth of telescope diameter, and instrument cost and complexity, is schematically illustrated in Figure \\ref{fig:cost}. \\begin{figure}[htp] \\includegraphics{Figure1.pdf} \\caption{\\label{fig:cost}A schematic illustration of the evolution in size and cost of optical spectrgraphs. The top two rows show pictures of the Sloan digital Sky Survey (SDSS; 2.5m primary) multi-object spectrograph, and the Keck (10m primary) spectrograph, DEep Imaging Multi-Object Spectrograph. Notice the relationship between telescope size (going down the page) and instrument volume. At 10m, the spectrograph dwarfs the PI and engineers. For the proposed 30m telescope and wide field optical spectrograph (WFOS), a CAD model is rendered next to DEIMOS. The WFOS is the size of a small house. The device in the bottom left is a Fabry-Perot spectrograph built by Axsun Technologies, it illustrates the principle behind this design: build thousands of miniature spectrographs.} \\end{figure} In this paper we describe the design of a spectrograph, built with tens of thousands of mini-spectrographs. By clever use of fiber multiplexing and slicing, enormous optics disappear, simplifying alignment and test procedures, which ultimately drive down the cost of the instrument package. A single, MEMS, spectrograph, including detector, costs manufacturers roughly US\\$20 per part. It is well understood that industrial replication reduces the cost of an individual component. However, a surprising recent result, is that replicated spectrographs reduce the {\\em total} cost of an instrument. For example, the recently designed Visible Integral-Field Replicable Unit Spectrographs (VIRUS), which is replicated by $\\sim100\\times$, saves approximately a factor of two or three compared to a monolithic design\\cite{hill06}. Micro-electrical-mechanical-systems (MEMS) technology, allows the production of spectrographs even smaller than the VIRUS. The drive to reduce spectrograph size is the drive reduce total instrument cost. Though these MEMS spectrographs may sound exotic, we emphasize that this technology is not new\\cite{onat05, Musca05}, commercial, off-the shelf MEMS spectrographs exist today. However, these commercial spectrographs, are typically tuned for specific applications (e.g. laser communication, television displays and molecular identification.) This work leverages the research of MEMS spectroscopy, to design an instrument tuned for a specific astronomical application. ", "conclusions": "" }, "0807/0807.1161_arXiv.txt": { "abstract": "Recent observations have discovered a giant HI bridge that appears to connect between the outer halo regions of M31 and M33. We propose that this HI bridge can be formed as a result of the past interaction between M31 and M33 based on test particle simulations with different orbits of M31 and M33 for the last $\\sim 9$ Gyr. We show that strong tidal interaction between M31 and M33 about $4-8$ Gyr ago can strip HI gas from M33 to form HI streams around M31 which can be observed as a HI bridge if they are projected onto the sky. We show that the number fraction of models reproducing well the observed HI distribution of the bridge is only $\\sim 0.01$\\% (i.e., $\\sim 10$ among $\\sim 10^5$ models) and thus suggest that the observed structure and kinematics of the HI bridge can give some constraints on the past orbits of M31 and M33. We suggest that the observed outer HI warp in M33 could be fossil evidence for the past M31-M33 interaction. We also suggest that some of high velocity clouds (HVCs) recently found in M31 could be HI gas originating from M33. We briefly discuss other possible scenarios for the formation of the HI bridge. ", "introduction": "Recent observational studies on HI gas around M31 based on the Green Bank Telescope (GBT) have reported the presence of at least 20 discrete HI clouds within 50 kpc of the M31 disk (e.g., Thilker et al. 2004). Furthermore possible candidates of HVCs around M31 with heliocentric velocities of $-520$ to $-160$ km s$^{-1}$ have been discovered in a blind HI survey for the disk and halo regions of M31 (e.g., Westmeier et al. 2007). Braun \\& Thilker (2004, hereafter BT04) have discovered a faint bridge-like HI structure that appears to connect between M31 and M33. Although these recent results have provided new clues to formation and evolution of M31 (BT04), no theoretical works have yet clarified the origin of the HI properties surrounding M31, in particular, the intriguing bridge-like structure between M31 and M33. Corbelli et al. (1989) revealed that the HI gas disk of M33 extends out to twice the optical radius and shows complicated kinematical properties indicative of the presence of a warped disk. By using a tilted ring model, Corbelli \\& Schneider (1997) demonstrated that the observed distribution and rotation curve profile of HI in the outer part of M33 are consistent with the presence of a warped gas disk. Although they suggested the formation of the gaseous warp in M33 due to tidal force from M31, the observed unique HI properties of M33 have not been discussed by theoretical and numerical studies in terms of the past M31-M33 interaction. The purpose of this Letter is to first show that the observed HI bridge between M31 and M33 can be formed from the past tidal interaction between M31 and M33 based on a larger number ($>10^6$) of test particle simulations for the past interaction. We here try to choose orbital models of M31 and M33 (among a large number of those) which can reproduce well the observed distribution of the HI bridge (BT04). We therefore do not intend to use hydrodynamical and chemodynamical simulations that are numerically costly and were used in our previous studies for the formation of HI streams and bridges around galaxies (e.g., Bekki et al. 2005a, b; 2008). ", "conclusions": "Although the present study has presented a scenario that the observed apparent HI bridge between M31 and M33 is a tidal stream of HI gas around M31 formed from the past M31-M33 interaction, this scenario would be only one of several possible scenarios. For example, BT04 suggested that the observed bridge can be a ``cosmic web'' which extends between massive galaxies as predicted in previous numerical simulations on structure formation. It would be also possible that the bridge is a tidal stream originating not from M33 but from other gas-rich dwarfs that may have already been destroyed by the strong tidal field of M31. If fully self-consistent hydrodynamical simulations for the bridge formation in our future studies can also reproduce the observed sharp HI edge and outer HI warps in M33 (e.g., Corbelli \\& Salpeter 1993; Corbelli et al. 1997), then the present interaction scenario can be regarded as more realistic and reasonable. Ibata et al. (2007) have recently found an extended metal-poor stellar halo around M33 in the deep photometric survey of M31, though the projected distribution of the halo is not so clear owing to the very limited spatial extent of the survey. If the stellar halo of M33 has a relatively homogeneous spatial distribution with no signs of disturbance and extends more than 35 kpc from the center of M33, then M33 is highly unlikely to have lost HI gas initially within 35 kpc: the present tidal interaction scenario needs to be dramatically modified. It would be possible that the observed outer metal-poor halo of M33 can be the very outer part of the M31 stellar halo. Future kinematical studies of metal-poor stars around M33, which can confirm that the stars belong to M33 rather than to M31, will enable us to discuss how far the M33 stellar halo extends and thereby assess the viability of the present tidal interaction scenario. The present models have shown that HI gas clouds originating from M33 can be located in tidal streams within $\\sim 100$ kpc of M31 and show systematic rotation with respect to M31. One of implications from this result is that at least some of the observed HVCs around M31 (e.g., Westmeier et al. 2007) can originate from M33. Metallicities of the M31 HVCs originating from M33 may well be as small as those of HI gas in the outer part of the gas disk in M33. Also the HVCs from M33 can have systematic rotation with the amplitude significantly smaller than that of the M31 disk ($V_{\\rm c} \\sim 250$ km s$^{-1}$). Thus future observational studies on chemical abundances and kinematical propertied of HVCs will help us to reveal the M31 HVCs originating from M33. Although the present gas-rich late-type disk galaxies are observed to have extended HI disks (e.g., Broeils \\& van Woerden 1994), it remains observationally unclear how and when the extended HI disks were formed. The present model M1 (with the first M31-M33 interaction about 8 Gyr ago) would not be reasonable, if the extended HI disk in M33 was formed relatively recently (4-5 Gyr ago). It should be thus stressed that the viability of the present scenario for the origin of the M31-M33 HI bridge would depend on whether {\\it M33 had already developed its extended HI disk before it experienced the first tidal interaction with M31}. Loeb et al. (2005) proposed a proper motion amplitude of $100\\pm 20$ km s$^{-1}$ form M31 based on orbital models of M31 and M33 and on the observed proper motion of M33 by B05. Recently van der Marel and Guhathakurta (2007) have investigated line-of-sight velocities of the seventeen M31 satellite galaxies in order to derive the M31 transverse velocity. They have found that the Galactocentric tangential velocity of M31 is highly likely to be less than 56 km s$^{-1}$ and suggested that M31 and M33 is in a tightly bounded system. These results appear to be consistent with the present tidal interaction scenario for the HI bridge. As shown in the present study, the observed location of the HI bridge alone can not allow us to provide very precise predictions on the present 3D velocities of M31 and M33. We plan to use the observed kinematical data sets for the HI bridge in order to give much stronger constraints on the 3D orbits of M31 and M33 by comparing the observations with more sophisticated numerical simulations." }, "0807/0807.3955_arXiv.txt": { "abstract": "Traditionally, the distance to \\n4038/39 has been derived from the systemic recession velocity, yielding about 20 Mpc for $H_0 = 72$ \\kmsmpc. Recently, this widely adopted distance has been challenged based on photometry of the presumed tip of the red giant branch (TRGB), which seems to yield a shorter distance of $13.3\\pm 1.0$ Mpc and, with it, nearly 1 mag lower luminosities and smaller radii for objects in this prototypical merger. Here we present a new distance estimate based on observations of the Type Ia supernova (SN) 2007sr in the southern tail, made at Las Campanas Observatory as part of the Carnegie Supernova Project. The resulting distance of $\\dsnia = \\DIa$ Mpc [$(m-M)_0 = 31.74\\pm 0.27$ mag] is in good agreement with a refined distance estimate based on the recession velocity and the large-scale flow model developed by Tonry and collaborators, $\\dflow = 22.5\\pm 2.8$ Mpc. We point out three serious problems that a short distance of 13.3 Mpc would entail, and trace the claimed short distance to a likely misidentification of the TRGB. Reanalyzing {\\em Hubble Space Telescope} (\\hst\\,) data in the Archive with an improved method, we find a TRGB fainter by 0.9 mag and derive from it a preliminary new TRGB distance of $\\dtrgb = 20.0\\pm 1.6$ Mpc. Finally, assessing our three distance estimates we recommend using a conservative, rounded value of $D = 22\\pm 3$ Mpc as the best currently available distance to The Antennae. ", "introduction": "\\label{sec1} The Antennae (\\n4038/39) are the nearest example of a major merger involving two gas-rich disk galaxies of comparable mass. They allow us to study processes of dissipational galaxy assembly from close up, thus providing a valuable glimpse of what must have been more frequent events in the early universe. The Antennae have been observed extensively at all wavelengths (e.g., X-rays: \\citealt{fabb04}; UV: \\citealt{hibb05}; optical: \\citealt{whit99}; IR: \\citealt{gilb07}; and 21-cm line: \\citealt{hibb01}) and have also been repeatedly modeled via N-body and hydrodynamical simulations \\citep[e.g.,][]{tt72,barn88,miho93,hibb03}. An accurate distance to this prototypical merger is of obvious importance in assessing its structure and dynamics, as well as for determining the physical properties of its myriad of stars, star clusters, peculiar objects such as ULXs (ultraluminous X-ray sources), and gas clouds. Traditionally, the distance to \\n4038/39 has most often been derived from the systemic recession velocity and an adopted Hubble constant $H_0$, with or without corrections for deviations of the Hubble flow from linearity due to various attractors. A frequently used modern value for the distance is 19.2 Mpc \\citep{whit99}. This value is based on a systemic recession velocity relative to the Local Group of $\\czlg = 1439$ \\kms \\citep{ws95,rc3}, a linear Hubble flow, and $H_0 = 75$ \\kmsmpc. Updated to the \\hst\\ Key Project's derived $H_0 = 72\\pm 8$ \\kmsmpc\\ \\citep{freed01}, this often-used distance to The Antennae becomes $20.0_{-2.0}^{+2.5}$ Mpc. A significantly shorter distance of $13.3\\pm 1.0$ Mpc has recently been determined by \\cite{savi08} via photometry of the TRGB in a region near the tip of the southern tidal tail. This photometry---performed on new, deep images obtained with \\hst\\,'s Advanced Camera for Surveys (ACS)---seems to support the short distance derived earlier from \\hst/WFPC2 images of the same region by the same method \\citep{savi04}. If the short distance is correct, The Antennae ``diminish'' in physical size, mass, and luminosity, as do their stars, clusters, and gas clouds. However, their recession velocity then deviates from the best large-scale-flow model \\citep[hereafter TBAD00]{tonr00} by about 500 \\kms\\ or 2.7\\,$\\sigma$ \\citep{savi08}. Clearly, new measurements of the distance to The Antennae by independent methods are highly desirable. In the present paper we derive a new distance from observations of the Type Ia SN 2007sr, which appeared in the midsection of the southern tail in 2007 December. Section~\\ref{sec2} presents our observations, data analysis, and new distance. Section~\\ref{sec3} discusses problems with the short distance and points out a likely error in the identification of the true TRGB by \\citet{savi08}. Finally, \\S~\\ref{sec4} presents our conclusions and recommendation. \\begin{figure*} \\centering \\includegraphics[scale=0.545]{n4038newdist_fig01_aph.eps} \\caption{ Supernova 2007sr in the southern tidal tail of \\n4038. {\\em Left:} Image of The Antennae reproduced from the Digital Sky Survey; the box marks the field of view corresponding to the right panel. {\\em Right:} Image of SN 2007sr in the $V$ passband, produced by stacking 29 exposures of 50~s duration each obtained over many nights with the CCD camera of the Swope 1.0~m telescope at Las Campanas Observatory. \\label{fig01}} \\end{figure*} ", "conclusions": "\\label{sec4} We have presented $u'Bg'Vr'i'JH$ light curves of SN 2007sr beginning 6 days after $B$ maximum and obtained during four months by the Carnegie Supernova Project. Analysis of these light curves in conjunction with \\sneia\\ data from the first-year campaign by the Project yields a formal true distance modulus of $\\mu_0=31.74\\pm 0.07$ mag. With all known systematic errors and uncertainties included, this distance modulus becomes $\\dmodsnia=31.74\\pm 0.27$ mag, corresponding to a distance of $\\dsnia=\\DIa$ Mpc to \\n4038/39. (Differential depth effects between SN 2007sr and the center of mass of the galaxies are negligible). This new \\snia\\ distance agrees well with the distance $D_{\\rm flow}= 22.5\\pm 2.8$ Mpc estimated from the systemic recession velocity of The Antennae and the large-scale flow model by TBAD00, where the quoted error reflects the cosmic random radial velocity ($\\sigvelth = 187$ \\kms) of the model. On the other hand, \\dsnia\\ disagrees strongly with the short distance of $D = 13.3\\pm 1.0$ Mpc estimated by \\citet{savi08} from the TRGB. We have discussed three serious problems with such a short distance, and have pointed out the likely misidentification of the TRGB by these authors as the cause of the short distance. Reprocessing their \\hst/ACS frames and using an improved method of TRGB detection, we have located what we believe to be the true TRGB at $I_0 \\approx \\trgbmag = 27.46\\pm 0.12$ mag, fully 0.9~mag below (i.e., fainter than) the TRGB claimed by \\citeauthor{savi08}\\ \\ Analyzing photometry of this new, fainter TRGB we have derived a preliminary distance of $\\dtrgb = 20.0\\pm 1.6$ Mpc. Clearly, additional distance estimates to The Antennae (e.g., from Cepheids, planetary nebulae, etc.) will be very valuable. However, for the moment we see no reason for adopting any distance shorter than 20 Mpc. Given the concordant new $\\dsnia = \\DIa$ Mpc and recession-velocity based $\\dflow = 22.5\\pm 2.8$ Mpc, both supported by our new, preliminary $\\dtrgb = 20.0\\pm 1.6$ Mpc, we recommend using a conservative, rounded value of $D = 22\\pm 3$ Mpc as the best currently available distance to The Antennae." }, "0807/0807.3352_arXiv.txt": { "abstract": "A number of recent numerical investigations concluded that the remnants of rare structures formed at very high redshift, such as the very first stars and bright redshift $z\\approx 6$ QSOs, are preferentially located at the center of the most massive galaxy clusters at redshift $z=0$. In this paper we readdress this question using a combination of cosmological simulations of structure formation and extended Press-Schechter formalism and we show that the typical remnants of Population III stars are instead more likely to be found in a group environment, that is in dark matter halos of mass $\\lesssim 2 \\cdot 10^{13}~h^{-1} \\mathrm{M_{\\sun}}$. Similarly, the descendants of the brightest $z \\approx 6$ QSOs are expected to be in medium-sized clusters (mass of a few $10^{14}~h^{-1} \\mathrm{M_{\\sun}}$), rather than in the most massive superclusters ($M>10^{15}~h^{-1} \\mathrm{M_{\\sun}}$) found within the typical 1~Gpc$^3$ cosmic volume where a bright $z\\approx 6$ QSO lives. The origin of past claims that the most massive clusters preferentially host these remnants is rooted in the numerical method used to initialize their numerical simulations: Only a small region of the cosmological volume of interest was simulated with sufficient resolution to identify low-mass halos at early times, and this region was chosen to host the most massive halo in the cosmological volume at late times. The conclusion that the earliest structures formed in the entire cosmological volume evolve into the most massive halo at late times was thus arrived at by construction. We demonstrate that, to the contrary, the first structures to form in a cosmological region evolve into relatively typical objects at later times. We propose alternative numerical methods for simulating the earliest structures in cosmological volumes. ", "introduction": "Rare dark matter halos at high redshifts host interesting astrophysical objects, especially before or at the end of the reionization epoch. One example is given by the very first Population~III (PopIII) stars formed in the universe at $z \\gtrsim 40$, which started the metal enrichment of the interstellar medium and the reionization process \\citep{abel02,san02,bromm04,naoz06}, and possibly also produced intermediate mass black hole seeds that grow to become super-massive black holes ($M_{BH}> 10^9 M_{\\sun}$) within the first billion year after the Big Bang \\citep{vol03}. Another example are bright $z \\approx 6$ QSOs \\citep{fan04}, considered to be hosted in the most massive dark matter halos at that time \\citep{MR05}. The luminosity of such an object is powered through accretion onto a supermassive black hole \\citep[e.g. see][]{hop06}, which may be the descendant of one of the first generation of PopIII stars formed in the universe, at $z \\gtrsim 40$ (\\citealt{mad01}; see also \\citealt{ts07a}). Many numerical and theoretical investigations have aimed at characterizing the properties of both PopIII stars \\citep[e.g. see][]{abel02,bromm04,ciar,ree05,gao05,osh07} and of high redshift QSOs \\citep[e.g. see][]{hop06,MR05,dimatteo05,li07}. However, the rarity of these structures makes a fully self-consistent treatment even of the formation of the underlying dark matter halos typically outside the current capabilities of a standard cosmological simulation, as the dynamic range that needs to be resolved is too large. The main limitation at high mass resolution is the box size, $l_{box}$: The enforcement of periodic boundary conditions limits the power spectrum of density fluctuations to modes with wavelengths shorter than $l_{box}$. This results in a severe bias on the measured number density of rare, massive halos, especially before the epoch of the reionization of the universe when the abundance of galaxies derived from numerical simulations can be underestimated by up to an order of magnitude \\citep{bar04}. While the number density of rare objects can be estimated in the context of Press-Schechter-type modeling \\citep{lacey93,mo_white96,she99,bar04}, studying the details of the formation histories of the first galaxies and QSOs, possibly including hydrodynamics and radiative feedback processes, requires high-resolution numerical simulation. Thus, given a simulation box that is large enough to contain one or more of the high-redshift halos of interest, the problem is to identify a sub-region that contains one of them for high-resolution simulation. In passing we note that this challenge is different from that of developing a simulation code that is able to adaptively increase the resolution when following the gas collapse that leads to the formation of the first stars. This has been successfully implemented with adaptive mesh refinement codes (for example ENZO - \\citealt{bry98}). \\citet{gao05} proposed a method to identify rare structures formed at very high redshift by recursively resimulating successively smaller, nested, sub-regions of a simulation box at progressively higher resolutions. Specifically, their method resimulates the region of the box centered on the most massive dark matter halo, reidentified at higher redshift and lower mass within the sub-region at each resimulation step, until the desired resolution is achieved in a small fraction of the entire simulation box. This procedure is well-motivated by the fact that number density of massive halos is increased in regions of large-scale over-density \\citep{bar04}, and thus structure formation in the sub-region on small scales at early times is accelerated by sitting at the top of an over-density, an over-density that is known to exist because it collapsed into the massive halo identified in the previous resimulation step. The \\citet{gao05} method extends high-resolution resimulation techniques used previously \\citep[e.g. see ][]{nav94,whi00}, and has been used recently by several authors to study the first stars and QSOs \\citep{gao05,ree05,gao07,li07}. The recursive resimulation method for volume selection is very effective at identifying a region with a very high-redshift halo when compared to random selection of a region of equal volume within the parent simulation box, but still it does not lead to the rarest high-redshift halos in the volume. This is due to the stochastic nature of the formation and growth of dark matter halos \\citep[e.g. see][]{PS,bond,she99}. The most massive dark matter halo in a given volume at some early time does not in general evolve into the most massive halo at a later time. The probability of the most massive halo to evolve into the most massive halo, quantified following \\citet{lacey93}, decreases as time passes and the characteristic mass scale increases. As we demonstrate, the probability falls far below unity for the evolution of the first stars and QSOs to the present time. The stochastic nature of growth histories of dark matter halos is addressed and discussed in \\citet{gao05}, but unfortunately some of the subsequent studies using their method do not take into account this stochasticity and assume a very strong correlation between the location of the richest clusters at $z=0$ and that of rare halos at very high redshift. In this paper we clarify this issue by presenting some basic, but often overlooked, results of Gaussian random fields in extended Press-Schechter theory to quantify the locations at later times of the first dark matter halos to form in a simulation box. We compare these analytical and numerical results to those published using recursive resimulation and highlight the benefits and limitations of that method. In addition, we propose an alternative method, based on analysis of the density field in the initial conditions, to select the sub-region of a simulation box that contains some of the earliest structure in the box without being biased toward the regions hosting the largest halos at $z=0$. The paper is organized as follows. In Sec.~\\ref{sec:QSOs} we investigate where the descendants of $z=6$ QSOs are today, while in Sec.~\\ref{sec:PopIII} we extend the result to the first Population III stars in the Universe. In Sec.~\\ref{sec:Ic} we discuss the implications of our results for the recursive mesh refinement method and propose a viable alternative. Sec.~\\ref{sec:conc} summarizes and concludes. ", "conclusions": "\\label{sec:conc} By means of both analysis of numerical simulations and of extended Press-Schechter modeling we investigated the relation between the most massive dark matter halos at different redshifts. The main conclusion of this work is that---contrary to expectations from many recent works (e.g. see \\citealt{MR05,ree05,gao05,gao07,li07})---the most massive halo at a redshift $z_1>z_2$ does not necessarily evolve into the most massive at $z=z_2$. This is a robust conclusion that can be naturally understood in the context of growth of dark matter density perturbations, for example by constructing merger trees through the \\citet{lacey93} model. Rare high-redshift objects, such as the remnants of the first PopIII stars and QSOs, are not hosted at $z=0$ in the most massive halos, but rather live in a variety of environments. For example, the typical $z>40$ PopIII star remnant lives in a dark matter halo that at $z=0$ has a mass of $\\approx 2 \\cdot 10^{13} h^{-1} \\mathrm{M_{\\sun}}$, typical for a galaxy group, and not within rich clusters as claimed by \\citet{whi00}. Similarly the descendant of a typical $z\\approx 6$ QSO is not located within the most massive clusters at $z=0$ as assumed by \\citet{li07}. These conclusions have important consequences on the application of the recursive simulation method introduced by \\citet{gao05} to identify high-redshift rare halos based on progressive refinement of regions centered around the most massive $z=0$ cluster. In fact, while the recursive method is indeed effective at identifying a sub-region of the simulation with earlier-than-average structure formation, it finds neither the earliest structures in the box, which are dominated by small-scale density fluctuations, nor typical early structure, as it preferentially identifies objects located in the regions with the highest bias. These limitations may have only a minor effect when the goal is to investigate the formation of one rare Population III halo in the simulation box, as it is done for example in \\citet{gao05} and in \\citet{ree05}. However in different physical scenarios it is important to correctly estimate the rarity of the halo simulated and to assess how typical their growth histories are. This is critical if additional physics beyond gravitational interactions is included, such as star formation and radiative feedback. One such example is the formation of rare high-redshift QSOs: \\citet{li07} use the \\citet{gao05} method to identify at $z\\approx 6.5$ ``the most massive halo in a $\\approx$ 3 Gpc$^3$ volume'' and then conclude that the QSO formed in this halo reproduces the properties of observed QSOs with the same number density. From our analysis in Sec.~\\ref{sec:QSOs} it is clear that the halo identified by \\citet{li07} as progenitor of the largest $z=0$ cluster is not likely to be the most massive at $z \\approx 6.5$. Thus other similar or more massive halos are expected to be present at $z \\approx 6.5$ in their $\\approx$ 3 Gpc$^3$ simulation volume: in principle any of these halos could host a bright QSOs, with important consequences for the comparison between observed and simulated QSO number densities. In addition, when the goal is to study the environment in which QSOs live, selecting host halos with the resimulation method introduces systematic effects in the results difficult to quantify and correct for, because these halos would have above-average growth (and merging) histories. To avoid selecting only the halos with atypical accretion histories and in an attempt to improve over the identification of some of the rarest high-redshift halos in a box, we propose instead to select the initial conditions for high-resolution resimulation based on the analysis of the linear density field at uniform resolution. The method, described in Sec.~\\ref{sec:Ic}, identifies subregions of the simulation box with high-redshift halos as those with the highest peaks in the density field, requiring a mass resolution in the field comparable to that of the mass of the halos one wishes to select. The applicability of the method is thus limited only by the size of the largest density grid that can be accommodated in the available memory, otherwise requiring only a modest amount of computing time compared to the \\citet{gao05} method. This is because our method bypasses the need of a series of N-body simulations to be carried out in addition to the final run. Unfortunately, our method does not guarantee identification of \\emph{the first} halo on the desired mass scale, as highlighted by some preliminary testing we presented in Sec.~\\ref{sec:valid}. This seems still an elusive goal. When the density field is defined over a fixed grid, there is not a perfect match between the halo catalog constructed from the density field and that obtained by increasing the resolution of the field and then following the full non-linear dynamics with an N-body simulation. An extensive validation of our linear density field initial conditions generation and its application to the formation of the first bright QSOs will be discussed in a subsequent paper." }, "0807/0807.3813_arXiv.txt": { "abstract": "In this paper we fully calculate the non-Gaussianity of primordial curvature perturbation of island universe by using the second order perturbation equation. We find that for the spectral index $n_s\\simeq 0.96$, which is favored by current observations, the non-Gaussianity level $f_{NL}$ seen in island will generally lie between 30 $\\sim$ 60, which may be tested by the coming observations. In the landscape, the island universe is one of anthropically acceptable cosmological histories. Thus the results obtained in some sense means the coming observations, especially the measurement of non-Gaussianity, will be significant to make clear how our position in the landscape is populated. ", "introduction": " ", "conclusions": "" }, "0807/0807.2139_arXiv.txt": { "abstract": "The ARGO-YBJ experiment has been designed to study the Extensive Air Showers with an energy threshold lower than that of the existing arrays by exploiting the high altitude location (4300 m a.s.l. in Tibet, P.R. China) and the full ground plane coverage. The lower energy limit of the detector (E $\\sim$ 1 GeV) is reached by the scaler mode technique, i.e. recording the counting rate at fixed time intervals. At these energies, transient signals due to local (e.g. Forbush Decreases) and cosmological (e.g. Gamma Ray Bursts) phenomena are expected as a significant variation of the counting rate compared to the background. In this paper the performance of the ARGO-YBJ detector operating in scaler mode is described and discussed. ", "introduction": "The study of gamma ray sources in the 1 - 100 GeV energy range has so far been only partially covered by observations by the satellite-based EGRET, with a maximum detectable energy of 30 GeV. On the other hand, present ground-based detectors such as Cherenkov Telescopes are still trying to reduce their threshold to energies well lower than 100 GeV. \\\\ In this energy range, Gamma Ray Bursts (GRBs) are certainly the most intriguing transient phenomena. GRBs have been deeply studied in the keV-MeV energy range, but only EGRET, in the last decade, provided some information in the GeV range. Three bursts have been detected at energies $>$1 GeV, with one photon of E=18 GeV \\cite{catelli,hurley}. The study of the high energy emission of GRBs could provide extremely useful constraints to the emission models and the parameters of the surrounding medium. \\\\ The interest for transient emission in the GeV region is not only restricted to extragalactic phenomena such as GRBs but also includes events closer to us - inside our Solar System - such as gamma rays from Solar Flares or high energy solar protons producing short time increases of the detector counting rate at the ground (Ground Level Enhancements). \\\\ The study of these transient phenomena can be successfully performed by ground-based experiments such as Extensive Air Shower (EAS) detectors working in ``single particle mode''\\cite{vernetto}. EAS arrays usually detect air showers generated by cosmic rays of energy E $\\geq$ (1 $-$ 100) TeV; the arrival direction of the primary particle is measured by the time delay of the shower front in different counters, and the primary energy is evaluated by the number of secondary particles detected. \\\\ EAS arrays can work in the energy region E $<$1 TeV operating in single particle mode, i.e. counting all the particles hitting the individual detectors, independently of whether they belong to a large shower or they are the lone survivors of small showers. Because of the cosmic ray spectrum steepness, most of the events detected with this technique are in fact due to solitary particles from small showers generated by 1 $-$ 100 GeV cosmic rays. \\\\ Working in single particle mode, an air shower array could in principle detect a transient emission in the interval 1 $-$ 100 GeV (both by gamma rays or charged primaries) if the secondary particles generated by the primaries give a statistically significant excess of events over the background due to cosmic rays. \\\\ The power of this technique is in its extreme simplicity: it is sufficient to count all the particles hitting the detectors during fixed time intervals (i.e. every second or more frequently, depending on the desired time resolution) and to study the counting rate behaviour versus time, searching for significant increases. The observation of an excess in coincidence with a GRB detected by a satellite would be an unambiguous signature of the nature of the signal. \\\\ The single particle technique does not allow one to measure energy and direction of the primary gamma rays, because the number of detected particles (often only one per shower) is too small to reconstruct the shower parameters. However, it can allow to study the temporal behaviour of the high energy emission, and, with some assumptions on the spectral slope (possibly supported by satellite measurements at lower energies), can give an evaluation of the total power emitted. \\\\ The background rate depends on the altitude and, to a lesser extent, on the geomagnetic latitude; it is not strictly constant and several effects both environmental and instrumental are responsible for variations on different time scales with amplitudes up to a few per cent. For this technique, an accurate knowledge of the detector and its sensitivity to both environmental and instrumental parameters is of crucial importance in order to correctly evaluate the statistical significance of the detected signals. \\\\ Obviously the sensitivity of this technique increases with the detection area and the energy threshold decreases at higher observation levels, so the most suitable detectors have a large area and work at very high altitude. \\\\ The ARGO-YBJ experiment, located at the YangBaJing High Altitude Cosmic Ray Laboratory (Tibet, P.R. China, 4300 m a.s.l.) with a detection area of $\\sim$6700 $m^2$, successfully exploits the full coverage approach at very high altitude with the aim of studying the cosmic radiation with a low energy threshold. \\\\ In order to cover many physics items in different energy ranges, two main operation modes have been designed: (1) shower mode, with an energy threshold of a few hundreds of GeV, and (2) scaler mode, with a threshold of a few GeV, this latter differing from the single particle mode for different particle multiplicities (i.e. not only the single particles) are put in coincidence and counted on different scaler channels. \\\\ The large field of view ($\\gtrsim$2 $sr$, limited only by atmospheric absorption) and the high duty cycle (close to 100$\\%$) typical of EAS detectors, make ARGO-YBJ suitable to search for unknown sources and unpredictable events such as GRBs or other transient phenomena. The site location (latitude 30$^{\\circ}$ 06${'}$ 38${''}$ N, longitude 90$^{\\circ}$ 31${'}$ 50${''}$ E) allows sky monitoring in the declination band $-10^{\\circ}<\\delta<70^{\\circ}$. \\\\ In this paper the performance of ARGO-YBJ working in scaler mode will be described, with some examples of application of this technique in solar physics and in the search for GeV emission from GRBs. ", "conclusions": "The ARGO-YBJ detector operated in scaler mode has a high duty cycle and optimal stability, fulfilling the fundamental requirements of the scaler mode technique. \\\\ The main goal is the measurement of the high energy tail of GRB spectra, and the detection of few events during the experiment life time is expected if the spectrum of the most energetic GRBs extends at least up to 100 GeV. In this case, with four measurement channels sensitive to different energies, valuable information on the high energy spectrum slope and possible cutoff may be obtained. \\\\ In any case, even if no signal is detected, the large field of view and the high duty cycle, allowing for continuous monitoring overhead, make the ARGO-YBJ experiment one of the most sensitive detectors for the study of the high energy spectrum of GRBs, with typical fluence upper limits down to $10^{-5} erg/cm^2$ in the 1 - 100 GeV region, well below the energy range explored by the present generation of Cherenkov telescopes. \\\\ On the other hand, due to the high rigidity cut-off, the large area of the detector and its high mountain location allow the study of the solar modulation of galactic cosmic rays and energetic relativistic particles coming from the Sun with high sensitivity and fine time resolution. This has been shown by the observation of the January 2005 FD with a preliminary discussion on the recovery phase. \\\\ Data collected up to May 2008 have been analyzed according to the procedures outlined in this work and the results of these analyses will be presented in later papers." }, "0807/0807.3765_arXiv.txt": { "abstract": "We present results from $N$-body + magnetohydrodynamical simulations of merging clusters of galaxies. We find that cluster mergers cause various characteristic magnetic field structures because of the strong bulk flows in the intracluster medium. The moving substructures result in cool regions surrounded by the magnetic field. These will be recognized as magnetized cold fronts in the observational point of view. A relatively ordered magnetic field structure is generated just behind the moving substructure. Eddy-like field configurations are also formed by Kelvin-Helmholtz instabilities. These features are similarly seen even in off-center mergers though the detailed structures change slightly. The above-mentioned characteristic magnetic field structures are partly recognized in Faraday rotation measure maps. The higher absolute values of the rotation measure are expected when observed along the collision axis, because of the elongated density distribution and relatively ordered field structure along the axis. The rotation measure maps on the cosmic microwave background radiation, which covers clusters entirely, could be useful probes of not only the magnetic field structures but also the internal dynamics of the intracluster medium. ", "introduction": "Clusters of galaxies have plenty of hot plasma as well as galaxies and dark matter (DM), which is called intracluster medium (ICM). Some observational evidence indicates that ICM is magnetized. For example, some clusters have diffuse non-thermal synchrotron radio emission that is called radio halos or relics, which shows that there are magnetic fields as well as relativistic electrons in the intracluster space \\citep{Giov99, Kemp01}. In addition, Faraday rotation measure observations of polarized radio sources such as radio lobes and AGNs behind and/or in clusters indicate that the magnetic field structures are quite random \\citep{Clar01, Vogt03, Govo06}. Comparing the synchrotron radio flux with the hard X-ray one (or its upper limit) due to the inverse Compton scatterings of cosmic microwave background (CMB) photons, we are able to estimate the volume averaged magnetic field strength (or its lower limit). Typically, the strength of $\\sim 0.1 \\mu$G is obtained with this method \\citep{Fusc99, Fusc05} though those detections of non-thermal hard X-ray components are still controversial \\citep{Ross04, Fusc07}. On the other hand, somewhat higher values ($\\sim$ a several $\\mu$G) tend to be obtained with the Faraday rotation measure method \\citep{Clar01, Vogt03, Govo06} though these results depend on the detailed magnetic field structures that are not fully understood \\citep{Enss03, Murg04}. Although the magnetic energy density is typically $\\sim $ a few percents of the thermal one in the intracluster space, it is believed that the magnetic fields play a crucial role in various aspects of ICM. It is likely that non-thermal particles are accelerated via shocks \\citep{Sara99, TakiNait00, ToKi00, Mini01, Ryu03, Taki03, Inou05} and/or turbulence \\citep{Rola81, Schl87, Blas00, Ohno02, Fuji03, Brun04}, where the magnetic fields play a definitive role in most theoretical models of the particle acceleration. Heat conduction is probably depends on the magnetic field configurations because charged particles cannot freely move in the direction perpendicular to the field lines. As we wrote before, magnetic field seems to be a minor component in global ICM dynamics. However, it is possibly important in relatively small scales, where fluid instabilities such as Rayleigh-Taylor and Kelvin-Helmholtz ones might be suppressed by magnetic tension. Cluster mergers have a significant impact on ICM magnetic field evolution. Turbulent motion excited by mergers would amplify the field strength via dynamo mechanism. Moving substructures are expected to sweep the field lines and form the cold subclumps surrounded by the field lines \\citep{Vikh01}. \\citet{Asai04} and \\citet{Asai07} performed two and three dimensional magnetohydrodynamical(MHD) simulations of moving cold subclumps in hot ICM in rather idealized situations, respectively, which confirmed that expectation. This field structure might be responsible for the suppression of heat conduction and fluid instabilities, which is essential for the maintenance of cold fronts. A lot of numerical simulations about merging clusters have been done, most of which are $N$-body + hydrodynamical simulations \\citep{Roet96, Taki99, Taki00, Rick01, Ritc02, Asca06, Taki06, McCa07, Spri07}. Although these simulations give us a great deal of understanding of the structures, evolution, and observational implications for merging clusters of galaxies, they make only a limited contribution to investigate the magnetic field structures. MHD simulations are essential in this regard. However, $N$-body + MHD simulations are rather rare though \\citet{Roet99} did pioneering work. It is true that Lagrangian particle methods based on smoothed particle hydrodynamics \\citep[see][]{Mona92} are extensively used in cosmological MHD simulations \\citep{Dola99, Dola02}. Considering that Eulerian mesh codes are essentially better at following evolution of magnetic fields, simulations based on such codes are highly desirable. In this paper, we present the results from $N$-body + MHD simulations of merging clusters of galaxies and investigate characteristic magnetic field structures during mergers and their implications. The rest of this paper is organized as follows. In \\S \\ref{s:simu} we describe the adopted numerical methods and initial conditions for our simulations. In \\S \\ref{s:resu} we present the results. In \\S \\ref{s:summ} we summarize the results and discuss their implications. ", "conclusions": "\\label{s:summ} We perform $N$-body and MHD simulations of merging clusters of galaxies. We find that cluster mergers cause various characteristic magnetic field structures because of strong bulk flow motion. The magnetic field component perpendicular to the collision axis is amplified especially between a bow shock and contact discontinuity. As a result, a cool region wrapped by the field lines appears. A relatively ordered field structure along the collision axis appears just behind the moving substructure. Eddy-like field structures are also generated by Kelvin-Helmholtz instabilities. Although the detailed structures change slightly, similar features are seen in off-center mergers. RM maps have some information about the magnetic field structure. The above-mentioned characteristic structures are partly recognized in the RM maps when we observe the merging systems in the direction perpendicular to the collision axis. On the other hand, RM maps observed nearly along the collision axis are less informative in this respect. Typical absolute values of the RM become higher when the system is observed along the collision axis because of both the density distribution elongated toward the axis and ordered magnetic field along the same direction. In minor mergers, amplification of the magnetic field component perpendicular to the collision axis becomes less prominent. The results about magnetic field along the collision axis does not change very much. Although numerical resolution effects have the significant impact on the small scale structures such as widths of a bow shock and contact discontinuity and detailed small scale fluctuations of the magnetic field, our results about overall global structures are reasonably reliable. Some observational and theoretical studies suggest that heat conduction in the ICM is suppressed from the Spitzer value. For example, a sharp temperature jump across a cold front in A2142 requires that the heat conductivity is reduced by a factor of between 250 and 2500 at least in the direction across the front \\citep{Etto00}. The spatial temperature variations in the central region of A754 also suggest that the conductivity is at least an order of magnitude lower than the Spitzer value \\citep{Mark03}. It is well-known that fine-tuning of heat conductivity is necessary in order to reproduce observational natures of cool cores assuming that radiative cooling in the cores balances with the conductive heating from the outer part of the clusters \\citep{Zama03}. Although the detailed process involved there is still unclear, magnetic fields likely play an important role. Heat conduction in the direction across the magnetic field lines is likely suppressed because electrons cannot travel freely in that direction. Using two dimensional MHD simulations with anisotropic heat conduction, \\citet{Asai04} shows that a moving subclump naturally form the magnetic field structure along the contact discontinuity and that the temperature jump there is maintained if the heat conduction perpendicular to the field line is sufficiently suppressed. Similar results are obtained in three dimensional cases \\citep{Asai07}. Basically, our simulations confirmed their results about magnetic field configurations in more realistic situations, though the heat conduction is not included. We show that Faraday rotation measure maps are useful to obtain information about characteristic magnetic field structures formed by cluster mergers. This could be another probe of internal dynamics of clusters. However, there is a serious problem in this method at present. We need polarized radio sources such as radio lobes and/or AGNs in and/or behind the clusters of galaxies to do that. In other words, we are able to obtain RM maps in the limited regions where we have suitable polarized sources by chance. Naturally, these do not always correspond to the regions that we are interested in. However, this difficulty could be overcame if we have suitable polarized sources that covers a cluster entirely. One possible solution is to use the CMB as the polarized source \\citep{Ohno03}, though this is still challenging in the present status of CMB observations. There are a lot of observations that are ongoing and planned for measuring the CMB polarization. We hope that future observations of the CMB polarizations will enable us to make RM maps that cover clusters entirely, which would give us an important clues to understand the internal dynamics as well as the magnetic field structures in clusters of galaxies. In actual clusters, the magnetic field have random components in small scales. It is very difficult to treat such scales by numerical simulations that concern the global structures. Clearly, observed RM maps are influenced by the small scale magnetic field fluctuations, which are not considered in RM maps calculated from our results. As a result, coherent length of the magnetic fields tends to be overestimated effectively, which means absolute values of RM from the simulations are also overestimated. However, it is probable that global spatial patterns of RM maps are nearly independent of such small scale fluctuations. In addition, it is also probably robust that higher RM are expected when the merging systems are observed along the collision axis." }, "0807/0807.4373_arXiv.txt": { "abstract": "We discuss the possible dynamical role of extended cosmic defects on galactic scales, specifically focusing on the possibility that they may provide the dark matter suggested by the classical problem of galactic rotation curves. We emphasize that the more standard defects (such as Goto-Nambu strings) are unsuitable for this task, but show that more general models (such as transonic wiggly strings) could in principle have a better chance. In any case, we show that observational data severely restricts any such scenarios. ", "introduction": "Introduction} Symmetry breaking phase transitions in the early universe are expected to have produced networks of topological defects \\cite{Vilenkin}. The possible roles of these defect networks in key cosmological scenarios will depend both on the type of defect considered and on the corresponding dynamics. For example, if defects are to significantly contribute to the dark energy, they should have a negative equation of state. If so the best possible situation is that of a frustrated network, in which case \\begin{equation} w\\equiv\\frac{p}{\\rho}=-\\frac{N}{3} \\end{equation} where $N$ is the defect's spatial dimension ($N=1,2$ respectively for cosmic strings and domain walls, while $N=3$ corresponds to a cosmological constant). It is clear that only the cases with $N=2$ and $N=3$ can lead to the recent acceleration of the Universe. A crucial property shared by these two cases is that the ratio between the dark energy density and the background density grows rapidly with time, with dark energy being dynamically dominant only around today. The defect network should therefore have a present density close to critical ($\\Omega^0_{de}\\sim1$), but compatibility with other cosmological observables requires it to have a characteristic scale several orders of magnitude below the horizon $\\xi \\ll H^{-1}$. Defects can also act as seeds for structure formation \\cite{Silk}. This may be the case if the ratio between the average defect energy density and the background density is approximately a constant and the characteristic scale of the network is roughly proportional to the Hubble radius ($\\xi \\propto H^{-1}$). Moreover, the defect fluctuations on the Hubble scale must be small ($\\delta\\lsim 10^{-5}$), and if the characteristic scale of the defects is of the order of the Hubble scale then it follows that their average energy density must also be very small ($\\Omega_{def}\\lsim 10^{-5}$) for consistency with cosmic microwave background anisotropies. It is well known that a scaling cosmic string network has all these properties \\cite{Vilenkin}. Notice that the required network properties are mutually incompatible: although there are unified dark energy scenarios in which dark matter and dark energy are described by a single entity, it is not possible for a given defect network to simultaneously be the dark energy and act as a seed for structure formation. The possible role of domain walls as dark energy candidates has been investigated in detail in \\cite{IDEAL1,IDEAL3} where it was shown that the dynamics of realistic domain wall networks appears to be incompatible with a dark energy role. Here we take a closer look at the role of defects on smaller scales. Specifically we will be mostly interested in kiloparsec (that is, galactic) scales. In particular, in doing this we will also explore another possibility that has been recently put forward \\cite{Redington,Alex}, \\textit{viz.} that cosmic strings could be the dark matter and explain the observed flat rotation curves of spiral galaxies. ", "conclusions": "" }, "0807/0807.3553_arXiv.txt": { "abstract": "Various radio observations have shown that the hot atmospheres of galaxy clusters are magnetized. However, our understanding of the origin of these magnetic fields, their implications on structure formation and their interplay with the dynamics of the cluster atmosphere, especially in the centers of galaxy clusters, is still very limited. In preparation for the upcoming new generation of radio telescopes (like EVLA, LWA, LOFAR and SKA), a huge effort is being made to learn more about cosmological magnetic fields from the observational perspective. Here we present the implementation of magneto-hydrodynamics in the cosmological SPH code GADGET \\citep{springel2001,springel2005}. We discuss the details of the implementation and various schemes to suppress numerical instabilities as well as regularization schemes, in the context of cosmological simulations. The performance of the SPH-MHD code is demonstrated in various one and two dimensional test problems, which we performed with a fully, three dimensional setup to test the code under realistic circumstances. Comparing solutions obtained using ATHENA \\citep{2008arXiv0804.0402S}, we find excellent agreement with our SPH-MHD implementation. Finally we apply our SPH-MHD implementation to galaxy cluster formation within a large, cosmological box. Performing a resolution study we demonstrate the robustness of the predicted shape of the magnetic field profiles in galaxy clusters, which is in good agreement with previous studies. ", "introduction": "\\label{sec:intro} Magnetic fields have been detected in galaxy clusters by radio observations, via the Faraday Rotation Signal of the magnetized cluster atmosphere towards polarized radio sources in or behind clusters \\citep[see][for recent reviews]{2002ARA&A..40..319C,2004IJMPD..13.1549G} and from diffuse synchrotron emission of the cluster atmosphere \\citep[see][ for recent reviews]{2004IJMPD..13.1549G,2008SSRv..134...93F}. Our understanding of the origin of cosmological magnetic fields is particularly limited. But their evolution and possible implications for structure formation are also not yet fully understood. In addition their interplay with the large-scale structure formation processes, as well as their link to additional dynamics within the cluster atmosphere is unclear, especially their role in the cool core regions and the influence of these regions on the evolution of the magnetic fields. The upcoming, new generation of radio telescopes (like EVLA, LWA, LOFAR and SKA) will dramatically increase the volume of observational data relevant for our understanding of cosmological seed magnetic fields in the near future. To investigate the general characteristics of the magnetic fields in and beyond galaxy clusters at the level required for a meaningful comparison to current and forthcoming observations, numerical simulations are mandatory. Non-radiative simulations of galaxy clusters within a cosmological environment which follow the evolution of a primordial magnetic seed field have been performed using Smooth-Particle-Hydrodynamics (SPH) codes \\citep{1999A&A...348..351D,2002A&A...387..383D,dolag2005b} as well as Adaptive Mesh Refinement (AMR) codes \\citep{2005ApJ...631L..21B,2008A&A...482L..13D}. Although these simulations are based on quite different numerical techniques they show good agreement in the predicted properties of the magnetic fields in galaxy clusters. When radiative cooling is included, strong amplification of the magnetic fields inside the cool-core region of clusters is found \\citep{2008A&A...482L..13D}, in good agreement with previous work \\citep{2000cucg.confE..75D}. Cosmological, magneto-hydrodynamical simulations were also performed using finite-volume and finite-difference methods. Such simulations are used to either follow a primordial magnetic field \\citep{2008ApJS..174....1L} or the creation of magnetic fields in shocks through the so-called Biermann battery effect \\citep{1997ApJ...480..481K,Ryu..1998}, on which a subsequent turbulent dynamo may operate. The latter predict magnetic field strength in filaments with somewhat higher values \\citep[e.g. see ][]{2004PhRvD..70d3007S} than predicted by simulations which start from a primordial magnetic seed field, but are in line with predictions of magnetic field values from turbulence \\citep{2008Sci...320..909R}. Therefore further investigations are needed to clarify the structure, evolution and origin of magnetic fields in the largest structures of the Universe, their observational signatures, as well as their interplay with other processes acting in galaxy clusters and the large scale structure. The complexity of galaxy clusters comes principly from their hierarchical build up within the large-scale structure of the Universe. In order to study their formation it is necessary, to follow a large volume of the Universe. However, one must also describe cosmic structures down to relatively small scales, thus spanning 5 to 6 orders of magnitudes in size. The complexity of the cluster atmosphere reflects the infall of thousands of smaller objects and their subsequent destruction or survival within the cluster potential. Being the source of shocks and turbulence, these processes directly act on the magnetic field causing re-distribution and amplification. Therefore realistic modelling of these processes critically depends on the ability of the simulation to resolve and follow correctly this dynamics in galaxy clusters. Starting from a well-established cosmological n-body smoothed particle hydrodynamic (SPH) code GADGET \\citep{springel2001,springel2005} we present here the implementation of magneto-hydrodynamics, which allows us to explore the full size and dynamical range of state of the art cosmological simulations. GADGET also allows us to turn on the treatment of many additional physical processes which are of interest for structure formation and make interesting links with the treatment of magnetic fields for future studies. This includes thermal conduction \\citep{2004MNRAS.351..423J,2004ApJ...606L..97D}, physical viscosity \\citep{2006MNRAS.371.1025S}, cooling and star-formation \\citep{springel2003}, detailed modelling of the stellar population and chemical enrichment \\citep{2004MNRAS.349L..19T,2007MNRAS.382.1050T} and a self consistent treatment of cosmic rays \\citep{2007A&A...473...41E,2007MNRAS.378..385P}. The MHD implementation presented here is fully compatible with all these extensions, but here we want to focus on non-radiative simulations. All such processes are expected to increased the complexity and lead to interplay with the evolution of the magnetic field. This would make it impossible to critically check the numerical effects caused by the different SPH-MHD implementations and therefore we will ignore such additional processes in this work. The paper is structured as follows: In section 2 we present the details of the numerical implementation, whereas in section 3 we present various code validation tests, all performed in fully three dimensional setups. In section 4 we present the formation of a galaxy cluster as an example for a cosmological application before we present our conclusions in section 5. In addition we present a convergence test for the code in the appendix. ", "conclusions": "\\label{sec:conc} We presented the implementation of MHD in the cosmological, SPH code GADGET. We performed various test problems and discussed several instability correction and regularization schemes. We also demonstrated the application to cosmological simulations, the role of resolution and the role the regularization schemes play in cosmological simulations. Our main findings are: \\begin{itemize} \\item The combination of many improvements in the SPH implementation, like the correction terms for the variable smoothing length \\citep{springel02} as well as the usage of the signal velocity in the artificial viscosity \\citep{1997JCoPh..136....298S} together with its generalization to the MHD case \\citep{2004MNRAS.348..123P} improve the handling of magnetic fields in SPH significantly. \\item Correcting the instability by explicitly subtracting the contribution of a numerical non-zero divergence of the magnetic field to the Lorenz force from the Maxwell tensor as suggested by \\citet{2001ApJ...561...82B} seems to perform well. Specifically in three dimensional setups where it seems to work much better than other suggestions in the literature. \\item The SPH-MHD implementation performs very well on simple shock tube tests as well as on planar test problems. We performed all tests in a fully three-dimensional setup and find excellent agreement of the results obtained with the SPH-MHD implementation compared to the results obtained with ATHENA in one or two dimensions. \\item With a convergence study we demonstrate that the SPH-MHD results when increasing the resolution are converging to the true solution, especially in the sharp features. However, in some regions it seems that small but systematic differences converge only very slowly to the correct solution. \\item Regularization schemes help to further suppresses noise and $\\mathrm{div}(\\vec{B})$ errors in the test simulations, however one has to carefully select the numerical parameters to avoid too strong smoothing of sharp features. Performing a full set of individual shock tube tests allows one to tune the numerical schemes and to determine optimal values. However they reflect an optimal choice for problems where the local timescales are mostly similar to the global timescale of the problem. For cosmological simulations it turns out that regularization by artificial dissipation leads to questionable results, whereas the regularization by smoothing the magnetic field (which is applied on global timescales) produces reasonable results. \\item The SPH-MHD implementation allows us to perform challenging cosmological simulations, covering a large dynamical range in length-scales. For galaxy clusters, only the shape of the predicted magnetic profiles is, (with the exception of the central part of clusters) converged in resolution and in good agreement with previous studies. Also the structures obtained in synthetic Faraday Rotation maps are in good agreement with previous findings and compare well with observations. \\end{itemize} The results obtained with artificial dissipation in cosmological simulations indicate that physical dissipation could play a crucial role in determining the exact shape of the predicted, magnetic field profiles in galaxy clusters. Future work, especially when including more physical processes at work in galaxy cluster -- as can be done easily with our SPH-MHD implementation -- will reveal an interesting interplay between dynamics of the cluster atmosphere and amplification of magnetic fields. Thus having the potential to shed light on many, currently unknown aspects of cluster magnetic fields, their structure and their evolution." }, "0807/0807.0146_arXiv.txt": { "abstract": "We report the discovery of a circumbinary disk around the Herbig Ae/Be system v892~Tau. Our detailed mid-infrared images were made using segment-tilting interferometry on the Keck-1 Telescope and reveal an asymmetric disk inclined at $\\sim$60$\\arcdeg$ with an inner hole diameter of 250~mas (35~AU), approximately 5$\\times$ larger than the apparent separation of the binary components. In addition, we report a new measurement along the binary orbit using near-infrared Keck aperture masking, allowing a crude estimate of orbital parameters and the system mass for the first time. The size of the inner hole appears to be consistent with the minimum size prediction from tidal truncation theory, bearing a resemblance to the recently unmasked binary CoKu Tau/4. Our results have motivated a re-analysis of the system spectral energy distribution, concluding the luminosity of this system has been severely underestimated. With further study and monitoring, v892~Tau should prove a powerful testing ground for both predictions of dynamical models for disk-star interactions in young systems with gas-rich disks and for calibrations of pre-main-sequence tracks for intermediate-mass stars. ", "introduction": "v892~Tau (Elias-1, Elias 3-1) is a young stellar object in the Taurus-Aurigae star forming region \\citep[$d=140~pc$,][]{kdh1994}. Its visible spectrum is classified as spectral type B8V \\citep{hernandez2004}, making v892~Tau one of the few Herbig Ae/Be stars in Taurus (in addition to AB~Aur, MWC 480). The spectral energy distribution (SED) of v892~Tau is dominated by bright thermal emission in the mid-infrared \\citep{hillenbrand1992}, with relatively large line-of-sight extinction in the visible \\citep[one estimate is A$_{\\rm V}\\sim5.9$,][]{kh1995}. Despite being one of the closest Herbig stars and after decades of multi-wavelength observations, the nature of v892~Tau is still uncertain. The SED suggests v892~Tau is either a young embedded Class I source \\citep{lada1987} still surrounded by its nascent envelope or a more evolved, Class II object (i.e, Herbig Ae/Be star) seen through its edge-on disk. Spatially-resolved imaging can easily distinguish between these two scenarios, and early speckle interferometry \\citep{kataza1991,haas1997,leinert2001} suggested the presence of an extended and elongated nebula in the near-infrared that could be due to scattering in bipolar lobes. The high-resolution speckle imaging of \\citet{smith2005} clearly resolved the K band ($\\lambda=2.2\\mu$m) emission to be coming from two unresolved stars in v892~Tau with little or no sign of a nebula or extended emission. The two stars of roughly equal brightness had an apparent separation of 55~milliarcseconds (7.7~AU) and some evidence of orbital motion was seen between epochs separated by 7 years. Although this binary should have a dramatic effect on the infrared emission, carving out a large hole in the circumbinary disk, recent analysis of the SED of v892~Tau including ISO data \\citep{acke2004} uncovered no distinct signature of the underlying binary\\footnote{\\citet{acke2004} did note strong 11.2$\\mu$m PAH emission and anomalous infrared colors.}. Other workers \\citep{liu2005,liu2007} made use of the technique of nulling interferometry in the mid-IR ($\\lambda=10.3\\mu$m) to marginally resolve v892~Tau (FWHM $\\sim$20~AU) along PA~164$\\arcdeg$ consistent with normal (single-star) disk emission. Here, we report new mid-iIR imaging of the v892~Tau system which resolves these mysteries, discovering very extended and resolved emission that we interpret to be coming from a circumbinary disk. We also confirm the presence of the binary at K band and our new data allow first crude estimates of the orbital elements. Lastly, we report a new SED analysis which provides an improved determination of the system luminosity and our viewing geometry. ", "conclusions": "We have discovered an extensive circumbinary disk around v892~Tau in the mid-infrared. We also independently confirm the binary nature of the underlying stellar system and our new measurement allows us to fit an astrometric orbit, finding a period $\\sim$14~years for system mass of $\\sim$6$\\msun$. Our limited orbital phase coverage and some ambiguity in position angles allow only crude estimates and we strongly urge continued monitoring of this system using near-IR speckle, aperture masking, or adaptive optics. We have proposed a new SED decomposition with a line-of-sight extinction ($A_V\\sim11$) higher than previously thought, implying a system luminosity $\\sim$400$\\lsun$. This result highlights the fact that circumbinary disks (and transitional disks) have much larger opening angles since the puffed-up inner wall causes enhanced extinction of the central stars even at inclinations of 60$\\arcdeg$. v892~Tau is another case where high-resolution imaging has motivated a fundamental shift in our understanding of an individual object. As spectral energy distributions are used to discover ``transitional'' disks implicating planet formation, we must be cognizant of the important role of binarity and the associated circumbinary disks that can mimic signs of planet formation \\citep[e.g.,][]{ireland2008}. Far from merely being spoilers to planet finders, new circumbinary disks offer fresh laboratories for studying dynamical interactions between gas-rich disks and the massive orbiting bodies embedded within them as well as critical opportunities to calibrate pre-main sequence tracks." }, "0807/0807.2469_arXiv.txt": { "abstract": "The Arcminute Microkelvin Imager is a pair of interferometer arrays operating with six frequency channels spanning $13.9$--$18.2$~GHz, with very high sensitivity to angular scales $30\\arcsec$--$10\\arcmin$. The telescope is aimed principally at Sunyaev-Zel'dovich imaging of clusters of galaxies. We discuss the design of the telescope and describe and explain its electronic and mechanical systems. ", "introduction": "\\label{sec:intro} Clusters of galaxies are the best samplers of matter on the largest scales and are sensitive probes of structure formation, both in the linear regime of growth and when merging, shocking and gradual virialization occur. Determining the structures and physics of clusters, their mass function and their evolution is therefore of basic importance. Observationally, however, relatively little is known about the properties of the cluster population at redshifts $z\\gtrsim 1$, due to the fundamental problem of the dimming of surface brightness with redshift. Optical observations suffer from confusion from foreground galaxies, and both optical and X-ray observations suffer from biases towards strong concentrations of mass. Ionized matter at any redshift out to recombination will inverse-Compton scatter the cosmic microwave background (CMB) radiation and so imprint spatial fluctuations upon it. Hot ionized gas in the deep potential wells of galaxy clusters (statistically) up-scatters CMB photons passing through it, giving rise to the Sunyaev-Zel'dovich (SZ) effect (\\citealt{sz70}, \\citealt{sz72}; see \\eg \\citealt{birkinshaw99} and \\citealt*{carlstrom02} for reviews); at frequencies below 217~GHz, one sees a decrease in the CMB temperature towards the cluster. This dip in temperature $\\Delta T$ is given in the Rayleigh-Jeans region by $\\Delta T = -2yT_{\\rm{CMB}}$, where the Comptonization parameter $y=\\int n_{\\rm{e}} \\sigma_{\\rm{T}} \\frac{k_{\\rm{B}}T_{\\rm{e}}}{m_{\\rm{e}} c^2} {\\rm d}l$, $\\sigma_{\\rm{T}}$ is the Thomson cross-section, $l$ is the line-of-sight path, and $m_{\\rm{e}}$, $n_{\\rm{e}}$ and $T_{\\rm{e}}$ are respectively the electron mass, density and temperature. The dip $\\Delta T$ is thus proportional to the line integral of pressure but does \\textit{not} depend on redshift $z$. One way of seeing this is that in such a inverse-Compton scattering process, the power lost by the electrons and given to the CMB photons is proportional to the energy density of the CMB radiation which is proportional to $(1+z)^{4}$; this exactly cancels out the cosmological drop in bolometric surface brightness proportional to $(1+z)^{-4}$. This is an extremely important result --- one can get \\textit{directly} to high redshift using SZ without any intermediate steps. The total SZ decrement in flux density, $ \\Delta S_{\\rm SZ}$, is proportional to the integral of the brightness temperature $ \\int \\Delta T_{\\rm SZ}\\,{\\rm d}\\Omega$ over the solid angle $\\Omega$ subtended by the cluster. Thus $\\Delta S_{\\rm{SZ}} \\propto D_{\\rm{A}}^{-2} \\int n_{\\rm{e}}T_{\\rm{e}} \\, \\rm{d}V, $ where $D_{\\rm{A}}$ is the angular-size distance and $\\rm{d}V$ is an element of volume. For concordance cosmologies, for redshifts between, say, 0.3 and 3, $D_{\\rm{A}}$ is only weakly dependent on redshift. Since $D_{\\rm{A}}$ is always a weaker function of redshift than luminosity distance $D_{\\rm{L}}$, SZ clusters will have a selection function that is much less dependent on redshift then most self-luminous objects. $\\Delta S_{\\rm{SZ}}$ measures the total thermal energy of the cluster, and since temperature is strongly correlated to mass (\\eg $T \\propto M^{2/3}$ for a population of clusters, modelled as virialized gravitationally-collapsed systems), the integrated SZ flux density measures gas mass directly, and without bias to concentrated structure. Detailed studies of clusters in SZ and other wavebands will help to establish the cosmological relationship between $M$ and $T$ and the thermal history of the cluster, and will help to calibrate the cluster scaling relations. These points immediately emphasize the importance of SZ surveys for clusters. These have long been advocated (see \\eg \\citealt*{ksy86}, \\citealt{bm92}, \\citealt{bs94}, et seq.) and it is now widely recognized that such surveys are key measurements for cosmology. For example, it is critical to measure and understand the comoving number density of clusters, $N(M,z)$, as a function of mass $M$ and redshift. $N(M,z)$ is a very strong function of the key cosmological parameter $\\sigma_8$ --- the density contrast on scales of $8\\,h^{-1}$Mpc now --- and of the physics of cluster assembly. The X-ray flux from a cluster is proportional to $n_{\\rm{e}}^{2}f(T_{\\rm{e}})$, where $f(T_{\\rm{e}})$ is not a strong function of $T_{\\rm{e}}$. The SZ flux is proportional to $n_{\\rm{e}} T_{\\rm{e}}$, so the combination of good X-ray and SZ data on a cluster gives robust determinations of density and temperature distributions and the related physics. Since the SZ flux density is proportional to the integrated gas pressure rather than to the density squared, it is relatively more sensitive to the outer parts of clusters than are X-rays, and not biased towards dense regions. All these points also emphasize the importance of pointed, high-resolution SZ observations of clusters known from X-ray, optical/IR or SZ surveying, as well as X-ray/optical/IR observations of SZ-selected targets. For example, cluster observations from XMM-Newton and Chandra show clear entropy structures in a cluster's gas, with contact discontinuities and sloshing motions (see \\eg \\citealt{rossetti07}), and bubbling and reheating of cooling gas by FRI radio jets of low luminosity but considerable bulk kinetic power \\citep[\\eg ][]{sanders07}. Observations in different wavebands tease out different selection effects and allow for cross-calibration. The first unequivocal SZ detections were made using the OVRO 40-m telescope and the NRAO 140-foot telescope \\citep*{birkinshaw81,birkinshaw-plus84,uson86}, and the first SZ image was obtained with the Ryle Telescope (RT; \\citealt{jones93}). The RT and OVRO/BIMA interferometers have since then made many SZ observations of known clusters (see \\eg \\citealt{grainge96}, \\citeauthor{grainge02a} \\citeyear{grainge02b}a,b, \\citeauthor{cotter02a} \\citeyear{cotter02b}a,b, \\citealt{grainger02}, \\citealt{saunders03}, \\citealt{jones05}; \\citealt*{carlstrom96}, \\citealt{grego00}, \\citealt{patel00}, \\citealt{reese00}, \\citealt{joy01}, \\citealt{reese02}, \\citealt{bonamente06}, \\citealt{laroque06}). Each of these observations has however been long (1--60 days), because even the shortest baselines resolved out most of the SZ signal: some 90 per cent of the SZ flux from a massive cluster at $z=0.2$ is missed by a $600$-$\\lambda$ baseline, the smallest possible given the existing dish sizes of the RT and OVRO/BIMA. Several other instruments have successfully observed the SZ effect, including the OVRO 5.5-m Telescope \\citep{herbig95}, SuZIE \\citep{holzapfel97}, ACBAR \\citep{cantalupo02}, CBI \\citep{udomprasert04}, VSA \\citep{lancaster05} and OCRA \\citep{lancaster07}. It has, however, long been clear that, for detailed imaging and blind surveying, a new generation of fast SZ imagers is required, with angular-scale and brightness-temperature sensitivities far better matched to the features of interest in clusters. Of course, such instruments should also be able to measure the CMB power spectrum at high angular scales ($\\ell > 2000$) and thus investigate the origins of, for example, the excess CMB power detected by CBI \\citep{cbi-excess}. They also are pertinent to the search for other non-Gaussian features such as topological defects. Accordingly, we have designed, built and now operate the Arcminute Microkelvin Imager (AMI, see \\citealt{rk01} and AMI Collaboration \\citeyear{ami-a1914}), a pair of interferometer arrays operating around 15~GHz near Cambridge, UK. Before describing AMI in detail, we next review other new SZ instruments, as well as the advantages and disadvantages of interferometers. \\subsection{Fast SZ Instruments} \\label{sec:intro:instruments} Several powerful SZ instruments are now coming online. The majority are direct-imaging, focal-plane arrays of bolometers, operating at frequencies above 90~GHz: ACT (\\eg \\citealt{ACT06}) in Chile uses a 6-m antenna with a resolution of $1\\farcm44$ at \\eg 150~GHz; APEX (\\eg \\citealt{APEX06}) in Chile has a 12-m antenna with $1\\farcm0$ resolution at \\eg 150~GHz; SPT (\\eg \\citealt{SPT04}) at the South Pole has a 10-m antenna with resolution of $1\\farcm06$ at \\eg 150~GHz; the Planck satellite, due for launch in 2009 January, will detect clusters in SZ over the whole sky, but detecting distant clusters will be challenging due to its limited resolution (see \\eg \\citealt*{geisbuesch05}). There are also three interferometers, operating at lower frequencies: AMI, see below; AMiBA (see \\eg \\citealt{amiba06}) in Hawaii, with the array comounted on a hexapod and currently at 90~GHz; and the SZA (see \\eg \\citealt{sza07}) in California with 3.5-m diameter antennas --- the SZA has been operating at 30~GHz with a resolution of $1\\arcmin$ and will be operated at 90~GHz as part of CARMA (\\texttt{www.mmarray.org}). \\subsection{Advantages and Disadvantages of Interferometers} \\label{sec:intro:ints} Interferometry has some significant advantages for observations where high sensitivity and low systematics are required (see \\eg \\citealt{church95}, \\citealt{lay00}, \\citealt{katy-thesis} and \\citealt{zwart2007}): \\begin{enumerate} \\item Stability of receivers. Short-term fluctuations in the gains of the front-end amplifiers are uncorrelated and therefore contribute only to the random noise. \\item Emission from the atmosphere is largely uncorrelated (provided the antenna beams do not overlap within the troposphere), and associated receiver power fluctuations are completely uncorrelated and thus contribute only to the system noise. \\item Fringe-rate filtering. Astronomical signals are modulated at the celestial fringe rate by the Earth's rotation. Unwanted interference from sources at terrestrial fringe rates, such as ground spill and geostationary satellite broadcasts, can be attenuated. Residual correlated atmospheric emission also has very little power in components that are synchronous with the astronomical fringe rate. \\item Interferometers have zero response to power which is uniformly spatially distributed (\\eg most of the atmospheric emission and the CMB average temperature), and hence do not respond to temporal fluctuations in that power. \\item Each interferometer baseline responds to only a narrow range of spatial frequencies; as well as filtering out the large-scale atmospheric signals this can be used to \\eg minimize the response to the primary CMB anisotropies while retaining sensitivity to clusters. \\item Foreground removal. Radio point sources can be detected by baselines chosen to be sensitive to smaller angular scales. This allows them to be measured simultaneously and subtracted. \\end{enumerate} \\noindent The main disadvantages of interferometers are complication (with the number of correlators going as the square of the number of detector elements) and a restriction on observing frequency. The SZ decrement has a maximum amplitude in intensity at approximately 130~GHz; contaminating synchroton emission from radio sources \\textit{generally} decreases with increasing frequency (falling in temperature as $\\nu^{-(\\alpha + 2)}$ where $\\alpha$ is the flux-density spectral index), although the effect of dust in galaxies on SZ observation at, say, 150~GHz is presently uncertain. Radio interferometers employing heterodyne receivers have costs and amplifier noise that rise with frequency, which in the past has limited interferometers to observing frequencies of around 30~GHz. \\subsection{The Rest of this Paper} \\label{sec:intro:rest} We next describe the design of the instrument (section \\ref{sec:design}), including the choice of frequency and the need for two interferometric arrays, the Large Array (LA) and the Small Array (SA). We describe the antennas, optics and array configurations of the SA and LA in section \\ref{sec:ant}. The amplifiers and cryostats are outlined in section \\ref{sec:cryo}, with the signal chain expounded in section \\ref{sec:if}. Section \\ref{sec:corr} covers correlator and readout systems, and section \\ref{sec:control}, the telescope control systems. Some lessons learned are outlined in section \\ref{sec:comm}. \\begin{table} \\centering \\caption{AMI technical summary.}\\label{table:summary} \\begin{tabular}{{l}{c}{c}} \\hline% & SA & LA \\\\ \\hline% Antenna diameter & 3.7~m & 12.8~m \\\\ Antenna efficiency & 0.75 & 0.67 \\\\ Number of antennas & 10 & 8 \\\\ Number of baselines & 45 & 28 \\\\ Baseline lengths (current) & 5--20~m & 18--110~m \\\\ Primary beam (15.7~GHz) & $20\\farcm1$ & $5\\farcm5$ \\\\ Synthesized beam & $\\approx3\\arcmin $ & $\\approx30\\arcsec$ \\\\ Flux sensitivity & 30~mJy~s$^{-1/2}$ & 3~mJy~s$^{-1/2}$ \\\\ Observing frequency & \\multicolumn{2}{c}{13.9--18.2{~GHz}} \\\\ Bandwidth & \\multicolumn{2}{c}{4.3{~GHz}} \\\\ Number of channels & \\multicolumn{2}{c}{6} \\\\ Channel bandwidth & \\multicolumn{2}{c}{0.72{~GHz}} \\\\ System temperature & \\multicolumn{2}{c}{25~K} \\\\ Declination range & $>-15\\degr$ & $>-20\\degr$ \\\\ Elevation limit & $+20\\degr$ & $+5\\degr$ \\\\ Polarization measured & \\multicolumn{2}{c}{I+Q} \\\\ \\hline% \\end{tabular} \\end{table} ", "conclusions": "\\begin{tabular}{{l}{c}{c}} \\hline% & SA & LA \\\\ \\hline% Antenna diameter & 3.7~m & 12.8~m \\\\ Antenna efficiency & 0.75 & 0.67 \\\\ Number of antennas & 10 & 8 \\\\ Number of baselines & 45 & 28 \\\\ Baseline lengths (current) & 5--20~m & 18--110~m \\\\ Primary beam (15.7~GHz) & $20\\farcm1$ & $5\\farcm5$ \\\\ Synthesized beam & $\\approx3\\arcmin $ & $\\approx30\\arcsec$ \\\\ Flux sensitivity & 30~mJy~s$^{-1/2}$ & 3~mJy~s$^{-1/2}$ \\\\ Observing frequency & \\multicolumn{2}{c}{13.9--18.2{~GHz}} \\\\ Bandwidth & \\multicolumn{2}{c}{4.3{~GHz}} \\\\ Number of channels & \\multicolumn{2}{c}{6} \\\\ Channel bandwidth & \\multicolumn{2}{c}{0.72{~GHz}} \\\\ System temperature & \\multicolumn{2}{c}{25~K} \\\\ Declination range & $>-15\\degr$ & $>-20\\degr$ \\\\ Elevation limit & $+20\\degr$ & $+5\\degr$ \\\\ Polarization measured & \\multicolumn{2}{c}{I+Q} \\\\ \\hline% \\end{tabular} \\end{table}" }, "0807/0807.3145_arXiv.txt": { "abstract": "A special type of Hamamatsu MPPC, with a sensitive area of 1.3$\\times$1.3~mm$^2$ containing 667 pixels with 50$\\times$50~$\\mu$m$^2$ each, has been developed for the near neutrino detector in the T2K long baseline neutrino experiment. About 60 000 MPPCs will be used in total to read out the plastic scintillator detectors with wavelength shifting fibers. We report on the basic performance of MPPCs produced for T2K. ", "introduction": "T2K (Tokai-to-Kamioka)~\\cite{T2K} is a long baseline neutrino oscillation experiment in Japan, using an intense beam from J-PARC accelerator at Tokai and the massive Super-Kamiokande detector 295~km away. The main goals of T2K are a sensitive search for the $\\nu_e$ appearance from $\\nu_\\mu$, which is related to the mixing angle $\\theta_{13}$, and precise measurements of `atmospheric' oscillation parameters. In order to achieve the aimed precision, good understanding of the beam properties and $\\nu$-nucleus interaction are indispensable. The `near detector' (ND) complex will be placed in Tokai to provide this information. The T2K-ND~\\cite{T2K-ND280} consists of several sub-detectors with specific and complimentary functions. As the basic elements for particle detection, most of detectors will use the plastic scintillator read out by wavelength shifting (WLS) fibers. This is a widely used technique, especially in recent accelerator neutrino experiments~\\cite{sci-wls}. In those experiments, multi-anode PMTs (MAPMTs) have been used as the photosensor. For T2K, MAPMT is not a good candidate because some of detectors have to operate inside a magnetic field of 0.2~T and cope with a limited space available. The following are major requirements for photosensors in T2K: \\begin{itemize} \\item More or equal photon detection efficiency than that of a MAPMT. \\item Compact to fit the limited space inside the magnet. \\item Operational in a magnetic field. \\item Good stability and low cost for a large number of readout channels (60 000). \\end{itemize}% We decided to use the Multi-Pixel Photon Counter (MPPC)~\\cite{MPPC, MPPC2} in August 2005. Since then, we continued the development collaborating with Hamamatsu and KEK Detector Technology Project. In this paper, we report on the performance of the MPPC developed for T2K. ", "conclusions": "New type of MPPCs have been developed for the T2K experiment. The T2K-MPPC is designed to have 1.3$\\times$1.3~mm$^2$ sensitive area in order to minimize the light loss when coupled to a WLS fiber of 1.0~mm diameter. The size of pixel is 50$\\times$50~$\\mu$m$^2$ and the number of pixels is 667. In T2K, about 60 000 MPPCs will be used in total. The mass production has started in February 2008. The gain, breakdown voltage, noise rate, photon detection efficiency, and cross-talk and afterpulse rate of T2K-MPPCs are measured for each device. The device uniformity is found to be excellent based on the measurement of 5820 MPPCs. All of MPPCs satisfy our requirements. We have established techniques necessary for a large scale application of MPPC to the WLS readout. The T2K experiment is planed to start in 2009. It will be the first experiment to use MPPCs in a large scale." }, "0807/0807.3832_arXiv.txt": { "abstract": "We study the role of the unstable equilibrium points in the transfer of matter in a galaxy using the potential of a rotating triaxial system. In particular, we study the neighbourhood of these points for energy levels and for main model parameters where the zero velocity curves just open and form a bottleneck in the region. For these energies, the transfer of matter from the inner to the outer parts and vice versa starts being possible. We study how the dynamics around the unstable equilibrium points is driven, by performing a partial normal form scheme and by computing the invariant manifolds of periodic orbits and quasi-periodic orbits using the reduced Hamiltonian. In particular, we compute some homoclinic and heteroclinic orbits playing a crucial role. Our results also show that in slow rotating and/or axisymmetric systems the hyperbolic character of the equilibrium points is cancelled, so that no transfer of matter is possible through the bottleneck. ", "introduction": "\\label{sec:intro} This paper focuses on the study of the dynamics around the hyperbolic equilibrium points, $L_1$ and $L_2$, of a non-axisymmetric galactic potential, i.e. a potential whose principal axes on the $(x,y)$ plane are different. In particular, we analyse the role of the invariant manifolds associated with the unstable periodic orbits and with invariant tori around $L_1$ and $L_2$ in the large scale transfer of matter within the system. For this purpose, we use both semi-analytical and numerical techniques to compute the invariant manifolds and to study the role they play in the global morphology. This is a modern approach using dynamical systems which has successfully been applied in celestial mechanics and astrodynamics (e.g. \\cite{gom91,jor99,gom01,koo00,gom04}). A similar technique has already been used to explain the spiral arm and ring morphology in barred galaxies \\cite{rom06,rom07,ath08}. Previous theories believe that spiral arms are density waves in a disc galaxy \\cite{lin63}. The density waves propagate from the centre towards the principal resonances of the galaxy, where they damp down \\cite{too69}. Other replenishment theories have been proposed, therefore, to obtain long-lived spirals (see \\cite{ath84} for a review). Using this innovative approach, we obtain outer rings and long-lived spirals. This paper is intended to explain the geometrical behaviour of such structures in a general manner. Such studies have many applications in galactic dynamics. Elliptical galaxies can be triaxial, (i.e. their principal axes can be all different from each other) while disc galaxies can contain several triaxial components, rotating or non-rotating, and with widely varying degree of non-axisymmetry, such as haloes, bulges, bars, or oval discs. In this paper, we consider a galactic model that describes a triaxial system. By studying its dynamics within a wide range of parameters, we will explain the dynamics of all the components and make links with their morphology. The logarithmic potential is a very suitable model for our purposes, because it has a simple expression, it is analytic, its series expansion can be calculated up to a high order, and it admits both semi-analytical and numerical treatments. We choose a reference frame such that the origin of coordinates coincides with the centre of the galaxy and the triaxial system is fixed, i.e. a reference frame rotating with the system. The idea of the present analysis is based on the fact that if we consider a range of energies for which the particles are confined in the inner region defined by the zero velocity curve, in our case, for energies lower or equal to that of the unstable equilibrium points ($E_J(L_1)$), then there is no possible transfer of matter from the inner region of the galaxy to the outer region, or vice versa (see Fig.~\\ref{fig:bottleneck}(a) and (b)). In this paper, we focus on energies slightly larger than that of the unstable equilibrium point, for which an opening in form of a bottleneck appears. We will hereafter refer to this aperture as bottleneck. Transfer of matter can be possible through this bottleneck (see Fig.~\\ref{fig:bottleneck}(c)), and we study the objects that drive the motion in this region. \\begin{figure} \\begin{center} \\includegraphics[scale=0.43,angle=-90.0]{plotcv0.ps} \\caption{Location of the equilibrium points and the zero velocity curves. {\\bf (a)} Zero velocity curves at an energy level smaller than the energy of $L_1$ and $L_2$. Three regions are thus defined, namely an inner (solid dark grey), an outer (solid white) and the forbidden region (hatched light grey), and no transfer of matter between the inner and outer regions is possible. {\\bf (b)} Zero velocity curves at the energy level of the unstable equilibrium points $L_1$ and $L_2$. {\\bf (c)} Zero velocity curves at an energy level larger than the energy of $L_1$ and $L_2$. A bottleneck appears and matter can transit from the inner to the outer regions (solid white) and vice versa.} \\label{fig:bottleneck} \\end{center} \\end{figure} Therefore, our first goal is to study the neighbourhood of the unstable equilibrium points in the energy range where the bottleneck appears. This energy range is suitable to apply the normal form technique, which is known as the reduction to the centre manifold. As previously mentioned, it has been successfully applied to celestial mechanics problems, in particular to the \\rtbp, e.g. \\cite{gom91,gom01,jor99}. The reduced Hamiltonian obtained from the reduction to the centre manifold process gives a good qualitative description of the phase space near the equilibrium points and uncouples the centre manifold from its hyperbolic behaviour. The procedure is similar to the Birkhoff normal form, usually used to study the stability properties around the central equilibrium point of the galaxy and to compute the families of periodic orbits around it, e.g. \\cite{gus66,mir89,bel07}. Our second goal is to study the behaviour of the hyperbolic invariant manifolds associated with the orbits contained in the centre manifold; these are essentially invariant manifolds of periodic orbits and invariant tori. As is well-known, stable and unstable invariant manifolds are dynamical features, that are responsible for the global dynamics in a dynamical system. In order to obtain the global picture of the transfer of mass, it is necessary to compute the invariant manifolds and to look for intersections between different parts, that is, obtaining the possible heteroclinic and homoclinic connections. These type of connections have been recently studied in the \\rtbpe and applied to obtaining transit and non-transit orbits in celestial mechanics \\cite{koo00,gom04}. In Section~\\ref{sec:mod}, we describe the characteristics of the galactic model. In Section~\\ref{sec:mot}, we present the equations of motion and we study the effect of the main parameters on the linear behaviour of the equilibrium points. In Section~\\ref{sec:red}, we explain in detail the reduction to the centre manifold in the particular case of the logarithmic potential and perform a study of the practical convergence of the reduced Hamiltonian. In Section~\\ref{sec:res}, we compute the invariant manifolds associated with periodic orbits and quasi-periodic orbits and we study the role they play in the transfer of matter. We also perform an analysis on the variation of the free parameters of the potential. Finally, in Section~\\ref{sec:conc} we conclude. ", "conclusions": "\\label{sec:conc} In this paper we use a suitable galactic potential (the logarithmic potential) to perform a semi-analytic study of the neighbourhood of the unstable equilibrium points $L_1$ and $L_2$. For a detailed study of their neighbourhood, we use a partial normal form scheme and we find that the main objects are the planar and vertical families of Lyapunov orbits and invariant tori. We compute the invariant manifolds associated with both periodic orbits and quasi-periodic orbits using the reduced Hamiltonian and we find that the motion around $L_1$ and $L_2$ is mainly driven by the invariant manifolds of the planar Lyapunov orbits. However, we are also interested in the global structure of the galaxy. Thus, we study the possible homoclinic and heteroclinic connections between the planar periodic orbits. For such a purpose, we use suitable Poincar\\'e surfaces of section. We note that this approach has successfully been used in celestial mechanics and in this paper we apply it to a galactic dynamics problem, namely the formation of spiral arms and rings in barred galaxies, in comparison to other theories given so far. Here we are interested in the hyperbolic behaviour of $L_1$ and $L_2$ and, particularly, in determining the role the invariant manifolds play in the transfer of matter for energy levels where the zero velocity curves are open (i.e. a range of energies somewhat larger than the energy of $L_1$ and $L_2$) and a bottleneck appears around $L_1$ and $L_2$. For such purpose, we also study the influence of the main model parameters, namely the pattern speed, $\\Omega$, and the shape parameters, $\\p$ and $\\q$. The logarithmic models are suitable for describing triaxial systems such as haloes, bars, bulges in disc galaxies or elliptical galaxies. The structures we have constructed using the invariant manifolds, however, are not globally dependent on the model characteristics \\cite{rom06,rom07}. This implies, for example, that if a system had some degree of rotation, these kind of structures should be present. Elliptical galaxies are triaxial systems that do not present any external feature, i.e. they are an ellipsoidal distribution of matter with different degrees of ellipticity and they present neither spiral arms nor rings. Observations show that elliptical galaxies barely rotate or do not rotate at all as a figure \\cite{bin87,sch82}. Our results are in agreement with this statement. We have shown that models that rotate slowly or do not rotate cancel the hyperbolic behaviour of the equilibrium points and, thus, no transfer or escape of matter is possible. On the other hand, bars in disc galaxies are non-axisymmetric components usually characterised in the literature by elliptical distributions of density \\cite{pfe84} although both observations and simulations show they might have more rectangular endings \\cite{ath90,ath02}. It is well-known that bars rotate and observations show that spiral arms or rings emanate from the ends of the bar. This characteristic is also consistent with our results which show that if the system rotates at a given angular velocity, the hyperbolic equilibrium points are present, so that the invariant manifolds drive the motion, and therefore, set the global morphology to the galaxy. We have seen that, depending on the rotation velocity and the shape of the bar, the morphology will be that of a barred spiral galaxy or that of a barred ringed galaxy." }, "0807/0807.1538_arXiv.txt": { "abstract": "We analyze the photometric redshift catalog of the Sloan Digital Sky Survey Data Release 5 (SDSS DR5) to estimate the Fisher information in the galaxy angular power spectrum with the help of the Rimes-Hamilton technique. It is found that the amount of Fisher information contained in the galaxy angular power spectrum is saturated at lensing multipole scale $300\\le l\\le 2000$ in the redshift range $0.1\\le$photo-z$<0.5$. At $l=2000$, the observed information is two orders of magnitude lower than the case of Gaussian fluctuations. This supports observationally that the translinear regime of the density power spectrum contains little independent information about the initial cosmological conditions, which is consistent with the numerical trend shown by Rimes-Hamilton. Our results also suggest that the Gaussian-noise description may not be valid in weak lensing measurements. ", "introduction": "Cosmic shear, which refers to the weak gravitational lensing by large-scale structure in the universe, is an invaluable tool in cosmology since it directly probes the gravitational clustering of dark matter in the universe without concerns about light-to-matter bias. Very recently, the three dimensional map of dark matter distribution has been indeed reconstructed on cosmological scales from the measurements of cosmic shear \\citep{Massey07}, which marks an advent of {\\it lensing cosmology}. Among many lensing observables, the angular power spectrum is regarded particularly important, since it is in principle capable of constraining the key cosmological parameters with high precision, including the enigmatic nature of dark energy \\citep[see][for a recent review]{HJ08}. The success of the angular power spectrum as a cosmological probe, however, is subject to one critical issue: how much information does it preserve about the initial conditions of the universe? A natural expectation is that some but not all information might have been destroyed in the subsequent nonlinear evolution. The amount of information contained in the power spectrum about the initial cosmological conditions is directly related to the precision of the cosmological parameters constrained by using the matter power spectrum. In other words, the non-Gaussian errors caused by the loss of information in the power spectrum would propagate into large uncertainties in the determination of the cosmological parameters. The error propagation and precision on cosmological parameters can be quantified in terms of the Fisher information contained in the matter power spectrum \\citep{TTH97}. \\citet[][hereafter, RH05]{RH05} have for the first time estimated the Fisher information in the matter power spectrum. Basically, RH05 have calculated the information content about the amplitude of the linear power spectrum averaged over an ensemble of many N-body realizations and found very little independent information at the translinear scale. Somewhat surprisingly, however, RH05 also found that there is a sharp rise in the amount of information in the nonlinear scale \\citep[see also][]{RH06,RHS06}. This phenomenon has been also noted by the analytic work of \\citet{NSR06}. Yet, according to the recent results derived by \\citet{NS07} based on the halo model, the Fisher information in the density power spectrum about all key cosmological parameters including the initial amplitude is highly degenerate both in the translinear and the nonlinear regime. Their work has indicated that it might not be possible to extract the initial cosmological conditions from the nonlinear dark matter power spectrum to a high statistical accuracy. Thus, the previous numerical and analytic results forecast that the accuracy in the determination of the cosmological parameters in lensing cosmology may be lower due to the non-Gaussian errors. As the next generation of large surveys optimized for weak lensing will soon be on the pipeline, it is imperative to test observationally the information content in the lensing power spectrum. In this Letter we attempt to do this test by applying the RH05 technique to the photometric galaxy catalogs from the Sloan Digital Sky Survey Data Release 5 \\citep[SDSS DR5,][]{AM07} at typical lens redshifts $z\\sim 0.3$. ", "conclusions": "It is worth mentioning here that what you have calculated using the SDSS data is not rigorously the information about the amplitude of the power spectrum. But, it only approximates the information in the linear power spectrum to the extent that $d\\ln P/d\\ln A$ is near unity. Our result provides an observational evidence for the previous theoretical clues that at translinear scale there is very little independent information in the matter power spectrum, consistent with light tracing matter till this scale \\citep{RH05,RH06,RHS06, NSR06,NS07}. Since the multipole scales where the information saturation is detected correspond to the weak lensing regime, this result has a direct impact on the weak lensing analyses. The loss of information in the angular power spectrum would lead to increasing non-Gaussianity in sampling errors. Unlike the usual assumption adopted in most weak lensing analyses that the non-Gaussianity contribution to the sampling errors for the angular power spectrum is marginal \\citep[][and references therein]{WH00,CH01}, our result suggests that it should be quite substantial and thus the Gaussian-noise description for the lensing power spectrum should not be valid." }, "0807/0807.3904.txt": { "abstract": "In this series of four lectures, I discuss four important aspects of AGN host galaxies. In Lecture \\#1, I address the starburst-AGN connection. First, I briefly review the primary diagnostic tools that are used to quantify and distinguish star formation and nuclear activity. Next I describe the best evidence for a connection between these two processes, first at low luminosity and then at high luminosity. In the last section, I summarize the main results and offer possible explanations. In Lecture \\#2, I discuss our current understanding of ultraluminous infrared galaxies (log[L$_{IR}$/L$_\\odot$] $\\ge$ 12; ULIRGs). First, I describe the general properties of ULIRGs, comparing the local sample with their distant counterparts. Then I discuss the role of ULIRGs in the formation and evolution of spheroids and their massive black holes. The discussion of their possible role in the metal enrichment of the IGM through superwinds is postponed until Lecture \\#3. In this third lecture, I discuss the importance of feedback processes in the local and distant universe. The emphasis is on mechanical feedback. I describe the basic physics of winds, a few classic examples of winds in the local universe, the statistical properties of winds, near and far, and their impact on galaxy formation and evolution. A list of potential thesis projects is given at the end. The fourth and final lecture is on elemental abundances as tracers of star formation. First, I explain the basic principles behind chemical evolution, and describe three simple models whose predictions are compared with observations in the Milky Way. Next I discuss and give an interpretation of the results of abundance determinations in local quiescent and starburst galaxies before discussing elemental abundances in the more distant universe. ", "introduction": "\\subsection{Introduction} \\label{} The apparent connection between black hole driven nuclear activity and starburst activity on large scale has been the topic of debates for many years (e.g., see references in review by Veilleux 2001). More than ever, this topic is relevant to help us understand galaxy formation and evolution, the global star formation and metal enrichment history of the universe, and the origin of nuclear activity and associated black hole growth. The existence of an apparently tight relation (Fig. 1) between black hole masses and spheroid masses (or velocity dispersions; Gebhardt et al. 2000; Ferrarese \\& Merritt 2000) points to a causal connection between spheroid formation (via a starburst) and black hole growth (via nuclear activity). A flurry of theoretical papers have tried to make sense of these results. In many scenarios, gas or radiation pressure from a starburst- and/or AGN-driven wind helps shut off the fuel supply to the black hole and terminate star formation in the surrounding galaxy (e.g., Murray, Quataert, \\& Thompson 2005). Regardless of the exact process involved in regulating the black hole and spheroid growths (this topic of negative feedback is covered in Lecture \\#3 of this series, \\S 3), the correlation indicates that the starburst-AGN connection is alive and well and has had a cosmologically important impact on galaxy formation and evolution. \\begin{figure} %\\centerline{\\epsfig{figure=fig1_1.eps,width=1.0\\textwidth,angle=0}} %\\plotone{fig1.eps} \\caption{Black hole mass versus bulge luminosity (left) and the luminosity-weighted aperture dispersion within the effective radius (right). Green squares denote galaxies with maser detections, red triangles are from gas kinematics, and blue circles are from stellar kinematics. Solid and dotted lines are the best-fit correlations and their 68\\% confidence bands. (From Gebhardt et al. 2000)} \\end{figure} To better understand this connection, one first needs to discuss the diagnostic tools that are used to detect and distinguish star formation and nuclear activity. This is done in \\S 1.2. In \\S 1.3, I describe key results from recent studies of low- and high-luminosity AGNs. In \\S 1.4, I summarize the results and offer a few possible explanations. \\subsection{Star Formation Diagnostics} \\label{} A very useful paper here is Kennicutt (1998). The material in \\S\\S 1.2.1 -- 1.2.4 is taken directly from that review and is therefore not described in detail. \\subsubsection{Ultraviolet} Hot, young stars emit copious amounts of UV radiation (e.g., Leitherer et al. 1999). The strength of the UV (1500 -- 2800 \\AA) continuum scales linearly with the luminosity of young stars and therefore with the star formation rate. For solar abundances and a Salpeter Initial Mass Function ($\\equiv$ IMF, 0.1 -- 100 M$_\\odot$): \\begin{eqnarray*} SFR (M_\\odot~{\\rm yr}^{-1}) = 1.4 \\times 10^{-28} L_\\nu ({\\rm ergs}~{\\rm s}^{-1} {\\rm Hz}^{-1}) \\end{eqnarray*} \\subsubsection{Recombination Lines} Hot, young stars emit radiation that ionizes the surrounding ISM. The Str\\\"omgren sphere is the spherical volume of this HII region where the rate of ionizations balances the rate of recombinations. The nebular lines produced in the HII region effectively re-emit the integrated stellar luminosity shortward of the Lyman limit ({\\em i.e.} $\\ge$ 13.6 eV). The intensity of these lines scales linearly with the number of hot, young stars and therefore the star formation rate. For solar abundances and a Salpeter IMF (0.1 -- 100 M$_\\odot$): \\begin{eqnarray*} SFR (M_\\odot~{\\rm yr}^{-1}) = 7.9 \\times 10^{-42} L_{H\\alpha} ({\\rm ergs}~{\\rm s}^{-1}) \\end{eqnarray*} \\subsubsection{Forbidden Lines} H$\\alpha$ is redshifted out of the visible window beyond $z \\sim 0.5$. In principle, H$\\beta$ and the higher order Balmer emission lines could be used to estimate star formation rates, but these lines are weak and stellar absorption more strongly influences their emission-line fluxes than that of H$\\alpha$. Neutral oxygen has the same ionization potential as hydrogen (13.6 eV). This means that ionized oxygen coexists with ionized hydrogen and therefore lines produced by ionized oxygen scales with the number of hot, young stars in HII regions. The strongest emission feature in the blue is the [O II] $\\lambda\\lambda$3726, 3729 forbidden-line doublet. The strength of these collisionally excited lines is sensitive to abundance and ionization state (electron temperature) of the gas, more so than the recombination lines. A rough calibration is: \\begin{eqnarray*} SFR (M_\\odot~{\\rm yr}^{-1}) = (1.4 \\pm 0.4) \\times 10^{-41} L_{[O II]} ({\\rm ergs}~{\\rm s}^{-1}) \\end{eqnarray*} Given its blue wavelength, this doublet is also more sensitive to dust extinction than H$\\alpha$. \\subsubsection{Far-Infrared Continuum} A significant fraction of the bolometric luminosity of a galaxy may be absorbed by interstellar dust and re-emitted in the thermal infrared (10 -- 300 $\\mu$m). The absorption cross-section of the dust is strongly peaked in the UV, so to first order the far-infrared emission scales with the star formation rate. In the limiting case of a dust cocoon surrounding a star-forming galaxy: \\begin{eqnarray*} SFR (M_\\odot~{\\rm yr}^{-1}) = 4.5 \\times 10^{-44} L_{FIR} ({\\rm ergs}~{\\rm s}^{-1}) \\end{eqnarray*} \\subsubsection{AGN Contamination} In many cases (particularly at higher redshifts), it is difficult to spatially resolve the emission produced by star formation from that produced by the AGN. In this situation, one must use spectroscopic methods to disentangle the two processes. In some instance, one may use the strengths of stellar atmospheric features to quantify the starburst. These include the Balmer series, Ca I triplet $\\lambda\\lambda\\lambda$8498, 8542, and 8662 in the visible/deep-red and Si IV $\\lambda$1400, C IV $\\lambda$1550, and He II $\\lambda$1640 in the UV (Robert et al. 1993). The emission lines produced in the ionized gas also bear the signature of the source of energy. The ionizing spectra of all but the hottest O stars cut off near the He II edge (54.4 eV), while AGNs are generally strong X-ray emitters. These high-energy photons have two effects on the emission line spectrum: (1) the material near the AGN is more highly ionized and emit strong high-ionization lines; (2) due to the strong energy dependence of the absorption cross-section ($\\sigma_\\nu \\propto \\nu^{-3}$), these high-energy photons are absorbed deeper into the gas clouds and produce extended partially ionized zones. These zones are strong emitters of collisionally excited low-ionization lines. One therefore expects an enhancement of {\\em both} high- and low-ionization lines in AGN relative to those in HII regions. Several diagnostic diagrams at optical and near-infrared wavelengths have been designed to specifically take advantage of these differences (e.g., Veilleux \\& Osterbrock 1987; Osterbrock, Tran, \\& Veilleux 1992; Kewley et al. 2001). In deeply obscured galaxies, the UV, optical, and near-infrared diagnostics cannot be used to distinguish between starbursts and AGN. One must rely on the relative intensities of the mid-infrared fine structure lines and/or the strengths (equivalent widths) of the polycyclic aromatic hydrocarbon (PAH) features. The principles behind the use of the fine structure lines are roughly the same as for the optical/UV lines {\\em i.e.} use line ratio diagrams that take advantage of the fact that AGN are copious emitters of low- and high-ionization lines. The use of the PAH features nicely complements that of the fine structure lines since they are generally easier to detect in the fainter, more distant galaxies. The PAH features are less visible in AGN due to the much stronger continuum in these objects and possible PAH destruction. \\subsection{Evidence for a Starburst-AGN Connection} \\label{} Since the triggering mechanism for AGN activity probably depends on the luminosity of the AGN, I make a distinction in the following discussion between the nearby, low-luminosity Seyferts and Fanaroff-Riley type I (FR I) radio galaxies and the more distant and powerful quasars, Fanaroff-Riley type II (FR II) radio galaxies, and ultraluminous infrared galaxies (ULIRGs; log [L$_{IR}$/L$_\\odot$] $\\ge$ 12 by definition - this is the subject of Lecture \\#2; \\S 2). \\subsubsection{Low-Luminosity AGN} Direct evidence for recent nuclear star formation exists in a number of Seyfert 2 galaxies ({\\em i.e.} Seyferts without broad recombination lines). Optical and ultraviolet spectroscopy of the nuclear regions of these galaxies often reveals the signatures of young and intermediate-age stars. The stellar Ca II triplet feature at $\\lambda\\lambda\\lambda$8498, 8542, 8662 in Seyfert 2s has an equivalent width similar to that in normal galaxies while the stellar Mg~Ib $\\lambda$5175 is often weaker (Terlevich, Diaz, \\& Terlevich 1990; Cid Fernandes et al. 2004). This result is difficult to explain with a combination of an old stellar population and a featureless power-law continuum from an AGN. The most natural explanation is that young red supergiants contribute significantly to the continuum from the central regions. Evidence for intermediate-age (a few 100 Myrs) stars in Seyfert galaxies is also apparent in the blue part of the spectrum, where the high-order Balmer series and He I absorption lines appear to be present in more than half of the brightest Seyfert 2 galaxies (e.g., Cid Fernandes \\& Terlevich 1995; Joguet et al. 2001; Gonz\\'alez Delgado, Heckman, \\& Leitherer 2001; Fig. 2). A few of these objects may even harbor a broad emission feature near 4680 \\AA, possibly the signature of a population of young (a few Myrs) Wolf-Rayet stars (Gonz\\'alez Delgado et al. 2001). The ultraviolet continuum from some of the brightest UV Seyfert 2s also appears to be dominated by young stars based on the strength of absorption features typically formed in the photospheres and in the stellar winds of massive stars (e.g., Heckman et al. 1997; Gonz\\'alez Delgado et al. 1998; Fig. 2). The bolometric luminosities of these nuclear starbursts ($\\sim$ 10$^{10}$ L$_\\odot$) are similar to the estimated bolometric luminosities of their obscured Seyfert 1 nuclei. This explains why UV-bright stellar clusters are more frequently detected in Seyfert 2s than in Seyfert 1s (Mu\\~noz Mar\\'in et al. 2007). The recent detection of near-infrared CN bands in Seyferts brings support to the idea that star formation is indeed connected to the AGN in these objects (Riffel et al. 2007). \\begin{figure} %\\plottwo{fig2a}{fig2b} \\caption{($left$) UV spectrum of NGC 7130. It is displayed in log ($F_\\lambda$) to show the emission and absorption lines. The most important stellar wind and photospheric absorption lines are labeled. ($right$) Normalized optical spectrum of NGC 7130 (dashed line) plotted with the normalized spectrum of a B0 V star combined with a G0 V star (thick line). 60\\% of the light is from a B0 V and 40\\% from a G0 V star. The comparison shows that most of the stellar features in NGC 7130 are well reproduced by a combination of young (B0 V) and old (G0 V) stars. (From Gonzalez Delgado et al. 1998)} \\end{figure} In recent years, SDSS has contributed significantly to our knowledge of local AGNs. For instance, Kauffmann et al. (2003b) have studied a sample of 22,623 narrow-line AGN with 0.02 $<$ z $<$ 0.3. They find that the hosts of low-luminosity AGN have a stellar population similar to that of normal early-type galaxies, while the hosts of high-luminosity AGN have much younger mean stellar ages. Indeed, young ($<$ 1 Gyr) stellar population appears to be a general property of AGN with high [O~III] luminosity (Figs. 3 and 4). This is true regardless of the presence of broad recombination lines ({\\em i.e.} types 1 and 2). The young stars are spread out over scales of at least a few kpc. \\begin{figure} %\\plotone{fig3.eps} \\caption{ The strengths of the 4000 \\AA\\ break and H$\\delta$ absorption feature are plotted as a function of log $L[O~III]$. The solid line shows the median, while the dashed lines indicate the 16-84 percentiles of the 1/V$_{\\rm max}$ weighted distribution. (From Kauffmann et al. 2003b)} \\end{figure} \\begin{figure} \\epsscale{0.5} %\\plotone{fig4.eps} \\caption{The fraction $F$ of AGN with H$\\delta_A$ values that are displaced by more than 3 $\\sigma$ above the local of star-forming galaxies is plotted as a function of log $L[O~III]$. The dashed line indicates the fraction of such systems in the subsample of normal massive galaxies. The dotted line indicates the fraction of such systems in the subsample of normal massive galaxies with $D_n(4000) < 1.6$. (From Kauffmann et al. 2003b)} \\end{figure} \\subsubsection{High-Luminosity AGN} Abundant molecular gas has been detected in radio galaxies and quasars (e.g., Evans et al. 2001), but is this gas forming stars? The extensive multiwavelength data set on these objects seems to indicate that starbursts are indeed present in local and distant quasars. Approximately 20 -- 30\\% of all PG~QSOs show an infrared excess L$_{IR}$/L$_{blue}$ $>$ 0.4. IR-excess QSOs tend to have large dust and H$_2$ masses, suggesting that the infrared -- submm ``bump'' in the spectral energy distribution of PG~QSOs is due to star formation. More direct evidence for a starburst-AGN connection has recently been found in local quasars from Spitzer mid-infrared spectroscopy (Schweitzer et al. 2006; Shi et al. 2007). PAH emission is detected in 11 of 26 PG QSOs and in the average spectrum of the other 15 PG QSOs (Schweitzer et al. 2006). The strength of the PAHs in these quasars is consistent with the far-infrared luminosity being produced primarily by U/LIRG-like starbursts with star formation rates of order 2 -- 300 M$_\\odot$ yr$^{-1}$ (Fig. 5). The strength of the starburst (measured by the FIR or PAH luminosity) correlates with that of the QSO (based on the 5100 \\AA\\ luminosity, a direct indicator of the mass accretion rate onto the black hole; Netzer et al. 2007; Fig. 6). This suggests a strong starburst -- AGN connection in these objects. \\begin{figure} \\epsscale{1.0} %\\plotone{fig5.eps} \\caption{PAH fluxes $F$(PAH 7.7 $\\mu$) vs. $F$(60 $\\mu$m) for local QSOs and starburst-dominated ULIRGs. (From Schweitzer et al. 2006)} \\end{figure} \\begin{figure} %\\plotone{fig6.eps} \\caption{$Top:$ Correlation of the optical (5100 \\AA) and FIR (60 $\\mu$m) continuum luminosities. $Bottom:$ $L(5100)$ vs. $L$(PAH 7.7 $\\mu$m) showing detections (filled squares) and upper limits (open squares). (From Netzer et al. (2007)} \\end{figure} A recent Spitzer study of high-z quasars by Maiolino et al. (2007a) fail to detect PAH emission in these objects. This would indicate that the correlation between star formation rate and AGN power ``saturates'' at high luminosities (Fig. 7). The ``flattening'' of the relation also seems to be present for CO emission (Maiolino et al. 2007b). It may therefore be that not enough fuel is available at high $z$ to fuel the starburst and match the strength of the AGN in these quasars. \\begin{figure} %\\plotone{fig7.eps} \\caption{ ($a$) PAH(7.7~$\\mu$m) luminosity as a function of the QSO optical luminosity. Blue diamonds are data from Schweitzer et al. (2006). The red square is the upper limit obtained by the average spectrum of luminous, high-z QSOs. ($b$) Distribution of the PAH(7.7~$\\mu$m) to optical luminosity ratio in the local QSOs sample of Schweitzer et al. (2006). The hatched region indicates upper limits. The red vertical line indicate the upper limit inferred from the average spectrum of luminous QSOs at high-z. (From Maiolino et al. 2007a) } \\end{figure} \\subsection{Discussion} \\label{} As explained in \\S 1.3, starbursts often coexist with actively accreting supermassive black holes, but the starburst-AGN relation appears to be tighter at high luminosity than at low luminosity. This suggests that black hole fueling in low-luminosity AGN is more stochastic and does not necessarily scale with the surrounding starburst. The fueling of AGN requires mass accretion rates $\\dot{M}$ $\\approx$ 1.7 (0.1 / $\\epsilon$)(L/10$^{46}$ ergs s$^{-1}$) M$_\\odot$ yr$^{-1}$, where $\\epsilon$ is the mass-to-energy conversion efficiency. A modest accretion rate of order $\\sim$ 0.01 M$_\\odot$ yr$^{-1}$ is therefore sufficient to power a Seyfert galaxy. Only a small fraction of the total gas content of a typical host galaxy is therefore necessary for the fueling of these low-luminosity AGNs. A broad range of mechanisms including intrinsic processes (e.g., stellar winds and collisions, dynamical friction of giant molecular clouds against stars; nuclear bars or spirals produced by gravitational instabilities in the disk) and external processes (e.g., minor galaxy interaction or mergers) may be at work in these objects. So it may not be surprising after all that the power of these AGN does not necessarily scale with the surrounding starburst. The stringent requirements on the mass accretion rates for luminous AGNs almost certainly require external processes such as major galaxy interactions or mergers to be involved in triggering and sustaining this high level of activity over $\\sim$ 10$^8$ years. Starbursts are a natural consequence of major mergers. This topic will be discussed in more detail in Lecture \\#2 (\\S 2). A lack of ``fuel'' (= molecular gas) may explain the apparent break in the starburst-AGN correlation at very high AGN luminosities. ", "conclusions": "" }, "0807/0807.3017_arXiv.txt": { "abstract": "{One of the most dramatic events in the life of a low-mass star is the He flash, which takes place at the tip of the red giant branch and is followed by a series of secondary flashes before the star settles on the zero-age horizontal branch (ZAHB). Yet, no stars have ever been positively identified in this key phase in the life of a low-mass star (hereafter the ``pre-ZAHB'' phase).} {In this paper, we investigate the possibility that at least some pre-ZAHB stars may cross the instability strip, thus becoming variable stars whose properties might eventually lead to positive identifications. In particular, it has been suggested that some of the RR Lyrae stars with high period change rates ($\\dot{P}$) may in fact be pre-ZAHB stars. Here we present the first theoretical effort devoted to interpreting at least some of the high-$\\dot{P}$ stars as pre-ZAHB pulsators.} {We constructed an extensive grid of evolutionary tracks using the Garching Stellar Evolution Code (GARSTEC) for a chemical composition appropriate to the case of the globular cluster M3 (NGC~5272), where a number of stars with high $\\dot{P}$ values are found. We follow each star's pre-ZAHB evolution in detail, and compute the periods and period change rates for the stars lying inside the instability strip, also producing pre-ZAHB Monte Carlo simulations that are appropriate for the case of M3.} {Our results indicate that one should expect of order 1 pre-ZAHB star for every 60 or so bona-fide HB stars in M3. Among the pre-ZAHB stars, approximately 22\\% are expected to fall within the boundaries of the instability strip, presenting RR Lyrae-like pulsations. On average, these pre-ZAHB pulsators are expected to have longer periods than the bona-fide HB pulsators, and 76\\% of them are predicted to show negative $\\dot{P}$ values. While the most likely $\\dot{P}$ value for the pre-ZAHB variables is $\\approx -0.3$~d/Myr, more extreme $\\dot{P}$ values are also possible: 38\\% of the variables are predicted to have $\\dot{P} < -0.8$~d/Myr.} {It appears likely, therefore, that some~-- but certainly not all~-- of the RR Lyrae stars in M3 with high (absolute) $\\dot{P}$ values are in fact pre-ZAHB pulsators in disguise.} ", "introduction": "Low-mass stars, after exhausting their central hydrogen supply, develop increasingly more massive and more highly degenerate helium cores as they continue their evolution accross the Hertzsprung-Russell diagram (HRD). As the mass of the core increases due to the continued supply of He-rich material from the H-burning shell that surrounds the core, it also contracts and becomes increasingly denser and hotter. Eventually, the conditions in the core become ripe for the ignition of the He-burning triple-$\\alpha$ reactions. Since the core at this point is electron-degenerate, the process does not take place quiescently, but rather in the form of a thermonuclear runaway~-- the so-called {\\rm He flash}. The primary He flash, which does not take place at the center proper (due to the very efficient neutrino cooling there), is followed by a series of secondary flashes, each occurring increasingly closer to the star's center, until degeneracy is lifted throughout the core and quiescent He burning finally commences~-- and a so-called {\\em horizontal branch} (HB) star is born \\citep[see][for a recent review]{mc07}. In the process just described, a low-mass star will typically change its luminosity by several orders of magnitude; its temperature may also increase very significantly. The inner structure of the star undergoes dramatic changes in the process as well. And yet, in spite of the importance of this evolutionary phase~-- hereafter the pre-zero-age HB (or pre-ZAHB) phase, to the best of our knowledge not a single such star has ever been positively identified. This dramatically limits the possibility of directly testing current models of pre-ZAHB evolution. The reasons why it has proven so challenging to pinpoint pre-ZAHB stars are twofold: first, the pre-ZAHB phase is very short, compared to other major evolutionary phases in the life of a low-mass star. Second, pre-ZAHB stars are largely expected to overlap the loci occupied by asymptotic giant branch (AGB), HB, and red giant branch (RGB) stars in observed color-magnitude diagrams (CMD). Bearing these difficulties in mind, \\citet{mc05,mc07} suggested an alternative approach for the detection of pre-ZAHB stars, namely, through {\\em stellar pulsations}. Indeed, some pre-ZAHB stars are expected to rapidly cross the Cepheid/RR Lyrae instability strip on their route from the RGB tip to the ZAHB, thus becoming pulsating stars along the way. At least some of these variables may present anomalously large period change rates ($\\dot{P}$), as a consequence of their very high evolutionary speeds, and so at least some pre-ZAHB stars could be singled out in view of their high $\\dot{P}$ values. As well known, high-$\\dot{P}$ values (e.g., $|\\dot{P}| \\gtrsim 0.1-0.15$~d/Myr) have indeed often been observed among field and cluster RR Lyrae stars \\citep[see, e.g.,][for a review]{hs95}, whereas canonical stellar evolution theory does not predict such high $\\dot{P}$ values for these stars, except towards the end of the HB phase when the star is already approaching the AGB \\citep[e.g.,][]{ywl91}. On the other hand, small, random mixing events associated with the composition redistribution in the cores of HB stars have also been proposed as possible causes for such high period changes in RR Lyrae stars \\citep{sr79}. The main goal of the present paper is, accordingly, to perform the first systematic study of the expected properties of pre-ZAHB stars, with a view towards their detection using a combination of pulsation and CMD properties. In this paper we shall, in particular, attempt to describe the expected pre-ZAHB stellar evolution in the case of the Galactic globular cluster M3 (NGC~5272), which is well known for being a particularly RR Lyrae-rich cluster, and also for containing a number of high-$\\dot{P}$ RR Lyrae stars \\citep[see, e.g., the recent work by][and references therein]{cc01}. In \\S\\ref{metodos} we describe the method employed to compute the evolutionary tracks, along with our Monte Carlo technique to produce synthetic pre-ZAHB distributions. In \\S\\ref{empirico} we present the empirical data that we use to compare our theoretical predictions with the observations. In \\S\\ref{resultados} we show the results from this comparison. Conclusions and final remarks are provided in \\S\\ref{conclusiones}. ", "conclusions": "In this paper, we have performed a detailed theoretical study of the expected evolutionary and pulsation properties of pre-ZAHB stars for a metallicity appropriate to the case of the Galactic globular cluster M3. Our results indicate that there should be around one pre-ZAHB star for every 60 or so bona-fide HB stars in the cluster. Based on extensive Monte Carlo simulations we find that, among the pre-ZAHB stars, of order 20\\% should exhibit RR Lyrae-like pulsations~-- to be compared with 21\\% and 57\\% blue HB- and red HB-like pre-ZAHB stars, respectively, and but a very small (of order $1-2\\%$) population of AGB/RGB-like pre-ZAHB stars. The pre-ZAHB, RR Lyrae-like pulsators are characterized by relatively long periods (typically longer than the average for the cluster), and (especially) by period change rates falling mostly in the $[-1,0]$~d/Myr range, with $-0.3$~d/Myr being the most likely value. However, more extreme $\\dot{P}$ values are also possible; in fact, 38\\% of the variables in our simulation have $\\dot{P} < -0.8$~d/Myr. Relatively few (i.e., around 24\\%) of the pre-ZAHB pulsators are found with positive period change rates. Our pre-ZAHB Monte Carlo simulations show, in addition, that for an M3-like HB mass distribution the pre-ZAHB CMD positions should be basically indistinguishable from those of bona-fide HB stars in the cluster. We predict that, in the specific case of M3, $\\sim 9$ pre-ZAHB stars should be present (assuming a total population of 530 HB stars), $\\sim 2$ of which most likely disguised as RR Lyrae stars with relatively long (and negative) period change rates. Since the number of stars in M3 with high $\\dot{P}$ is significantly larger \\citep[e.g.,][]{cc01}, we conclude that pre-ZAHB evolution may explain the high-$\\dot{P}$ phenomenon for some (but not all) of the RR Lyrae stars with high period change rates in this cluster. In the future, we will extend our calculations to other clusters (and dwarf galaxies) with different metallicities and HB morphologies. These calculations should also prove important for the identification of candidate pre-ZAHB pulsators among field stars in the Galaxy and neighboring galaxies. With the era of the survey telescopes quickly approaching, the number of such stars available for study should dramatically increase, thus making empirical studies of pre-ZAHB evolution increasingly more feasible~-- at least in the long term." }, "0807/0807.3684_arXiv.txt": { "abstract": "We have selected a small sample of post-AGB stars in transition towards the planetary nebula and present new Very Large Array multi-frequency high-angular resolution radio observations of them. The multi-frequency data are used to create and model the targets' radio continuum spectra, proving that these stars started their evolution as very young planetary nebulae. In the optically thin range, the slopes are compatible with the expected spectral index (-0.1). Two targets (IRAS 18062+2410 and 17423-1755) seem to be optically thick even at high frequency, as observed in a handful of other post-AGB stars in the literature, while a third one (IRAS 20462+3416) shows a possible contribution from cold dust. In IRAS 18062+2410, where we have three observations spanning a period of four years, we detect an increase in its flux density, similar to that observed in CRL 618. High-angular resolution imaging shows bipolar structures that may be due to circumstellar tori, although a different hypothesis (i.e., jets) could also explain the observations. Further observations and monitoring of these sources will enable us to test the current evolutionary models of planetary nebulae. ", "introduction": "The evolution of planetary nebulae remains a challenging topic in astrophysics. Several studies have focused on the link between the post-Asymptotic Giant Branch (AGB) and planetary nebula phases to investigate the changes that occur in the evolutionary stage between these object classes. Planetary nebulae (PN) have been observed over a wide wavelength range, from X-ray to radio frequencies. Their complex morphologies and the shaping mechanisms that produced them are still a matter of debate. Companion stars, jets from central stars, magnetic fields, dust tori, and interacting winds are some of the possible shaping agents suggested as being responsible for various PN morphologies, and an overlap of their actions cannot be ruled out. Hubble Space Telescope (HST) imaging has demonstrated that pre-PN show non--spherical symmetry even at late spectral types, which implies that the shaping process in these sources starts very early after the AGB \\citep{balick&frank}. Non-spherically symmetric structures have also been observed in AGB stars that are part of a binary system \\citep{karovska}, which could indicate that some shaping agents are already active in this evolutionary phase. In general, the current theory of PN evolution is based on the Interacting Stellar Wind (ISW) model \\citep{kwok1978} and its generalised version \\citep{kahn}. However, this model does not account for shaping because it assumes that an asymmetric distribution of matter is already present when the wind interaction occurs. Other models take into account the possible role of jets, and have been successfully applied to some nebulae \\citep{sahai98}. It has been suggested that the interaction with a companion object, even a massive planet, may provide the necessary asymmetry \\citep{soker06}. Also, large scale magnetic fields, influencing or determining the shapes, might be sustained by a dynamo process \\citep{blackman}. The birth of a PN is defined by the observation of an ionisation front, which is itself a shaping agent and could heavily influence the morphology established in earlier evolutionary phases, disrupting the molecular and dust circumstellar shells. Near- and mid-IR images of PN, which trace the molecular and warm dust emission, have been compared to optical line images, which trace the ionised elements, and have shown the presence of similar structures \\citep{latter}. Therefore it seems that molecular and ionised gas and dust grains can spatially coexist in these sources. This makes the investigation of the spatial distribution and physical properties of the ionised component in these envelopes even more compelling. In this context, important information can be provided by observations of very young PN, or pre-PN, where the physical processes associated with PN formation are still occurring. To investigate the properties of these rare objects in transition from the post-AGB to PN, we have selected a sample of pre-PN and searched for radio emission from ionised shells. The targets were selected from stars classified in the literature as hot post-AGB candidates, showing strong far-IR excess and B spectral type features. We detected radio emission in 10 sources in the selected sample \\citep{umana}. The detection of radio continuum emission is proof of the presence of free electrons and therefore of an ionised shell. For a better understanding of these sources, we have performed a multi-frequency follow-up to build up radio continuum spectra and collect information about the physical conditions in the nebulae. Several nebular parameters can be determined by radio flux density measurements (for example, electron density and ionised mass) but most calculations rely on the assumption of an optically thin emission. Although this is presumably the case for observations at frequencies higher than 5 GHz, only the construction of the whole centimetre continuum spectrum can enable us to distinguish between optically thick and thin regimes. VLA A array observations have also given us insight in the morphologies of the envelopes. ", "conclusions": "We have presented radio continuum observations of a small sample of pre-PN. The observations were carried out to inspect both the morphology and the physical conditions of the ionised component in these nebulae. Compact planetary nebulae are, in principle, only one possible explanation for radio-frequency continuum emission from evolved stars, the others being stellar photosphere, chromosphere, and circumstellar dust \\citep{knapp}. Our observations rule out the possibility of a stellar wind or dust origin for the emission, since the obtained spectra are not compatible with such hypotheses. The dust would exhibit an approximate black body spectrum ($S_\\nu \\propto \\nu^2$ in the radio range), and a stellar wind would have a 0.6 spectral index over a wide range of frequencies \\citep{panagia}. % We have modelled the observed spectra assuming that the emission arises in spherical shells around the central stars, with the density decreasing as $r^{-2}$. The values of electron densities and ionised masses found for our targets confirm their nature as young PN, since they fall within the ranges expected for such objects. The data seem to point to a somewhat steeper than expected slope beyond 5~GHz but, by spectral fitting, we calculated spectral indexes that match, to within errors, the theoretical value of -0.1. IRAS~18062+2410 appears to be optically thick up to 8.4~GHz and also shows flux density variations on a time scale of a few years. Such variations are rarely observed in the post-AGB/young-PN transition phase and indicate that this star belongs to a group of objects having such properties as those observed in CRL 618. IRAS 17423-1755 also seems to be optically thick at high frequency, but the complete spectrum is needed to confirm this result. Quite interestingly, IRAS 20462+3416, whose overall centimetre spectrum is very flat, has a larger flux density at 22.4 GHz than at lower frequencies. We speculate that this may be due to a contribution from cold dust. The high-angular resolution observations that we have presented show that the ionisation in these targets has already involved the walls of the cavity around the central star and is not limited to the tenuous gas within the cavity. All of the targets have been at least partly resolved, except IRAS 18062+2410. Almost all of the envelopes show two peaks of emission in a somewhat extended nebulosity that can be explained as an opacity effect, possibly due to a circumstellar torus, or the onset of jets. Unlike the other targets, the envelope in IRAS 22023+5249 shows more of an elongation along an axis than two bright peaks. Such a morphology would be better explained by the action of jets rather than the presence of a torus." }, "0807/0807.0152_arXiv.txt": { "abstract": "\\noindent \\citet{Venemans2005} found evidence for an overdensity of Ly$\\alpha$ emission line galaxies associated with the radio galaxy MRC 0316--257 at $z=3.13$ indicating the presence of a massive protocluster. Here, we present the results of a search for additional star-forming galaxies and AGN within the protocluster. Narrow-band infrared imaging was used to select candidate [O{\\sc iii}] emitters in a 1.1$\\times$1.1 Mpc$^2$ region around the radio galaxy. Thirteen candidates have been detected. Four of these are among the previously confirmed sample of Ly$\\alpha$ galaxies, and an additional three have been confirmed through follow-up infrared spectroscopy. The three newly confirmed objects lie within a few hundred km s$^{-1}$ of each other, but are blueshifted with respect to the radio galaxy and Ly$\\alpha$ emitters by $\\sim2100$ km s$^{-1}$. Although the sample is currently small, our results indicate that the radio--selected protocluster is forming at the centre of a larger, $\\sim60$ co-moving Mpc super-structure. On the basis of an HST/ACS imaging study we calculate dust-corrected star-formation rates and investigate morphologies and sizes of the [O{\\sc iii}] candidate emitters. From a comparison of the star formation rate derived from UV-continuum and [O{\\sc iii}] emission, we conclude that at least two of the [O{\\sc iii}] galaxies harbour an AGN which ionized the O$^{+}$ gas. ", "introduction": "To understand the formation and evolution of galaxy clusters, it is desirable to find and study their high redshift progenitors. Although galaxy clusters have been found out to a redshift of $z=1.5$ \\citep{Mullis2005,Stanford2005}, their higher redshift progenitors are sparse and difficult to find. A successful technique for finding more distant structures that by--passes the need for surveying very large areas of the sky is to search for emission-line galaxies in the neighbourhood of luminous high--redshift radio galaxies (HzRGs) using narrow-band imaging. Multiwavelength studies of HzRGs have resulted in strong evidence that they are massive forming galaxies \\citep[e.g.][]{Seymour2007,Villar-Martin2006} and are frequently associated with overdensities of emission-line galaxies \\citep{Venemans2007}. These overdense regions in the early universe are the likely progenitors of local galaxy clusters or groups and are termed ''protoclusters\". PROCESS (PROtoCluster Evolution Systematic Study) is a project designed to use a few key radio-selected protoclusters with $2 \\le z \\le 5$ to investigate the formation of and evolution of various populations of galaxies in dense environments \\citep{Overzier2006,Overzier2007,Overzier2008,Venemans2005,Venemans2007}. This article presents observations of the protocluster surrounding the PROCESS radio galaxy MRC\\,0316-257. The associated $1.5 \\ {\\rm Jy}$ radio source was listed in the 408 MHz Molonglo Reference Catalogue \\citep{Large1981} and was optically identified with a galaxy at $z=3.13$ by \\citet{McCarthy1990}. \\citet{LeFevre1996} spectroscopically confirmed two Ly$\\alpha$ emitting companions to the HzRG, indicating that the radio galaxy is located in a dense environment. Recently, \\citet{Venemans2005,Venemans2007} confirmed 31 Ly$\\alpha$ emitters at a similar redshift of MRC\\,0316-257. The corresponding overdensity is approximately 3.3 times the galaxy field density at this redshift. The protocluster redshift of $z \\sim 3.13$ corresponds to an epoch when both the cosmic star formation rate and the quasar luminosity function were at their peak, indicating that this is a key epoch for studying the evolution of different populations of galaxies. We have identified and studied additional galaxies in the MRC\\,0316-257 protocluster on the basis of their redshifted [O{\\sc iii}] emission. The observations that we shall discuss here consist of infrared imaging and spectroscopy with ESO's Very Large Telescope (VLT) and deep optical imaging with the Advanced Camera for Surveys (ACS) on the Hubble Space Telescope (HST). Section 2 of this article is an outline of the observations and the data reduction. In Section 3 we present results from the VLT search programme and the deep ACS images. Corrected star formation rates derived from the UV fluxes are used to discriminate between star-forming galaxies and obscured AGNs and the morphologies and sizes of the candidate emitters are discussed. The implications of our results for the space density of [O{\\sc iii}] emitting galaxies and the origin of the [O{\\sc iii}] emission are discussed in Section 4 and the conclusions of the article are presented in Section 5. We assume a flat cosmology with H$_0$=71 [km sec$^{-1}$ Mpc$^{-1}$] and $\\Omega_{\\rm m}$=0.27 \\citep{Spergel2003}. At the distance of MRC\\,0316-257 an angular scale of 1\\,arcsec corresponds to a projected linear scale of 7.73\\,kpc. All magnitudes are given in the AB system \\citep{Oke1974}. ", "conclusions": "Searching for [O{\\sc iii}]--emitting galaxies is a new feasible method for detecting protocluster members. We have detected a new population of [O{\\sc iii}] emitting galaxies in the neighbourhood of the radio galaxy MRC\\,0316-257 at $z$=3.13. About half of the [O{\\sc iii}] candidates emitters are LBGs and a third were also previously detected by the Ly$\\alpha$ selection technique. The [O{\\sc iii}] technique complements narrow-band searches using Ly$\\alpha$ and H$\\alpha$ emission, and observations of the Lyman and Balmer breaks, for finding members of protoclusters. All of these different galaxy selection techniques are needed to study the different galaxy populations and to obtain a complete understanding of protocluster evolution. 13 candidate [O{\\sc iii}] emitters were detected, including the radio galaxy, and 8 of these were spectroscopically confirmed. Three [O{\\sc iii}] emitting galaxies lie in a small redshift interval at $3.095$2) galaxies has been undertaken, including many in the rest-frame UV/optical (e.g., Steidel et al.\\ 1996, 1999, 2003; Sawicki et al.\\ 1997; Franx et al.\\ 2003; Daddi et al.\\ 2004; Sawicki \\& Thompson 2005; Iwata et al.\\ 2007) and far-IR (e.g., Barger et al.\\ 1998; Hughes et al.\\ 1998; Blain et al.\\ 1999; Smail et al.\\ 2002; Chapman, et al.\\ 2003; Webb et al.\\ 2003; Sawicki \\& Webb, 2005; Coppin et al.\\ 2006). The study of galaxies selected by their strong \\lya\\ emission allows an investigation of a population that's potentially very different from these other high-$z$ galaxy populations. \\lya\\ selection not only gives us the prospect of probing galaxies with very faint spectral continuum levels, but also may select systems in very early stages of a starburst (e.g., Malhotra \\& Rhoads 2002). It is thus not surprising that over the last several years much observational effort has gone into \\lya\\ surveys (e.g., Ajiki et al.\\ 2003; Hu et al.\\ 1998, 2004; Kudritzki et al.\\ 2000; Ouchi et al.\\ 2003; Rhoads et al.\\ 2000, 2003; Rhoads \\& Malhotra 2001; Santos et al.\\ 2004). Indeed, the short rest-frame wavelength of the \\lya1216 line allows galaxies to be detected and spectroscopically confirmed to very high redshifts, $z$$\\lesssim$7, without the need for infrared observations. The ability to push to such high redshifts raises the intriguing possibility of probing the tail end of the epoch of reionization, regarded by many as marking the onset of galaxy formation in the Universe. The expected evolution of the IGM neutral fraction across the reionization epoch has led to the prediction that the observed \\lya\\ luminosity function (LF) will undergo rapid evolution as IGM opacity to \\lya\\ photons drops with decreasing redshift (e.g., Haiman \\& Spaans 1999; Malhotra \\& Rhoads 2004; Kashihawa et al.\\ 2006) and attempts have been made to detect the signature of this effect albeit with contradictory results (Malhotra \\& Rhoads 2004; Kashikawa 2006). At slightly lower redshifts, the LF of high-$z$ galaxies may be evolving due to evolution in the intrinsic properties of galaxies (e.g., Sawicki \\& Thompson 2006; Iwata et al.\\ 2007; see also Bouwens et al.\\ 2007). Consequently even if the observed evolution of the \\lya\\ LF proves to be real, it will remain an open question as to whether it reflects an evolution in the opacity of the IGM, or evolution in the \\emph{intrinsic} properties of galaxies at a time when they were young. To believe the former we must understand the latter, and thus it is important to study the properties of Lyman Alpha Emitters (LAEs) in detail not only at very high redshift, $z$$>$6, but also as a function of time. To date, narrowband imaging searches for high-$z$ LAEs have been the most fruitful approach to finding LAEs and have yielded significant samples of galaxies at \\zs3--6.5, with several dozen objects now confirmed spectroscopically by various teams (e.g., Rhoads et al.\\ 2003; Hu et al.\\ 2004; Dawson et al.\\ 2004, 2007; Shimasaku et al.\\ 2006; Kashikawa et al.\\ 2006). An alternative to narrowband imaging surveys lies in direct, blank-sky spectroscopic searches whose strategy is to bypass the photometry stage and find galaxies directly in blank-sky spectra. One advantage of this approach is the potential saving of the time associated with deep narrowband imaging; another is the gain in sensitivity that results from working at the natural linewidth of the line ($\\sim$10--20\\,\\AA) rather than fighting against the sky background admitted by the full width of the narrowband filter ($\\sim$100--200\\,\\AA) (see Martin \\& Sawicki 2004). Until recently, the chief disadvantage of spectroscopic searches has been the small sky area accessible to most spectrographs. Serendipitous spectroscopic discoveries of single emission-line objects (e.g., Franx et al.\\ 1997; Dawson et al.\\ 2002), while useful, do not yield {\\it samples} of galaxies suitable for statistical analysis. Spectroscopic searches of regions enhanced by gravitational lensing have met with some success (Santos et al.\\ 2004; Stark et al.\\ 2007), probe very deep into the luminosity function but over very small areas. Until recently, multislit blank-sky searches (e.g., Crampton \\& Lilly 1999; Martin \\& Sawicki 2004; Tran et al.\\ 2004) have found many line emitters but no confirmed LAEs; only now are the sky areas accessible to this technique becoming large enough to report successful detections (Martin, Sawicki, Dressler, \\& McCarthy 2008). In the present work we take an approach that is hybrid to the dedicated blank-sky spectroscopic surveys and the serendipitous discoveries. Specifically, we carry out a dedicated search for serendipitous LAEs in the extremely large, existing spectroscopic database of the DEEP2 survey. The present paper reports our first discovery of a sample of \\lya\\ emitters using these data and illustrates some of the constraints that this approach can place on the properties of these objects. The paper is organized as follows. In \\S\\ref{data.sec} we describe our DEEP2 data and their suitability to the task at hand. In \\S\\ref{deep2serendips.sec} we describe our search for LAEs and report on the objects that we have found. In \\S\\ref{ConfidenceTests.sec} we present a number of tests with which we build confidence in the high-redshift nature of our objects. In \\S\\ref{numberdensity.sec} we constrain the number density and luminosity function of high-$z$ LAEs and in \\S\\ref{SpectralProperties.sec} we study the spectral properties of the \\lya\\ emission line at high redshift. In \\S\\ref{discussion.sec} we discuss future prospects for finding LAEs using deep blank-field spectroscopy and summarize our main findings. Throughout this paper we adopt the $\\Omega_M$=0.3, $\\Omega_\\Lambda$=0.7, $H_0$=70~km~s$^{-1}$~Mpc$^{-1}$ cosmology. ", "conclusions": "\\label{discussion.sec} Deep LAE surveys are an important tool for the study of formation and evolution of galaxies. Despite their many very important successes, narrowband imaging surveys for LAEs have their limitations: the numbers of spectroscopically-confirmed objects are still modest and the lack of detailed spectroscopy limits the ability to study in detail such properties as galactic outflows. Moreover, the lack of spectroscopic redshifts for all the LAEs introduces biases in luminosity function and correlation function studies. Thus, while we must continue to press with narrowband imaging surveys, we must also develop other, complementary techniques. Martin et al.\\ (2008) compare in detail the merits of spectroscopic and narrowband imaging imaging LAE surveys and discuss the future prospects for these techniques in the coming era of extremely large telescopes; here, we focus more narrowly on the near-term prospects of exploiting the remaining $\\sim$80\\% of the DEEP2 spectroscopic database. In this paper we have demonstrated that a large blank-sky spectroscopic survey can find significant numbers of LAEs. Blank-sky spectroscopic searches have been attempted in the past (e.g., Crampton \\& Lilly 1999; Martin \\& Sawicki 2004; Tran et al.\\ 2004; Martin et al.\\ 2008), although the number of LAEs fund by them to date is very small. Here we have piggy-backed on a large intermediate-$z$ survey that is DEEP2 and in just $\\sim$20\\% of these data we have found 9 objects that are bona fide LAEs (and have a further 10 of which some may also be such). So far, ours is a modest sample, but we have been able to use it to show that spectroscopic LAE searches can do well at constraining properties of the population. Specifically, we found that the number density of objects in our survey is consistent with those found by narrowband imaging searches (\\S~\\ref{numberdensity.sec}), and --- using only our stacked discovery spectra with no additional follow-up data, we confirmed the interpretation that LAEs are associated with large-scale outflows of material (\\S~\\ref{SpectralProperties.sec}). Much more could be done with a bigger sample and with deep follow-up spectroscopy. A simple extension of the by-eye search described in this paper to the remaining $\\sim$80\\% of the DEEP2 data can be expected to yield five times more LAEs in total, i.e. 42 class 3 and 2 objects (or 88 if class 1 objects are also included). However, further improvements in efficiency over that achieved here are possible. Specifically, in the present by-eye search we are sure to have missed many objects since the human eye is not an optimal tool when confronted with vast amounts of data. An automated search of the available dataset should be much more efficient. Our first experiments with implementing such an automated search are very encouraging and show that we can recover the objects found by eye as well as find additional, missed line emitters. We speculate that with an automated search we should be able to double the number of LAEs we find per unit area --- as is also suggested by a comparison of our number counts with those of narrowband imaging surveys (Fig.~\\ref{LF.fig}). An automated search of DEEP2 would thus give us a sample of $\\sim$90 class 3 and 2 objects, all with spectroscopic redshifts and with high-resolution spectra suitable for the study of the details of gas kinematics. Finally, the spectral coverage of DEEP2 gives us the potential to probe LAEs all the way up to $z$$\\sim$6.6 (Fig.~\\ref{volume_vs_z.fig}), thus allowing us to study the evolution of the population from near the epoch of reionization to $z$$\\sim$4.5 using a single, uniform dataset. Another clear advantage of an automated search is that it would allow us to accurately calibrate the detection efficiency through simulations and thus correct for it in our LF analysis. Moreover, analyzing the full four-field DEEP2 dataset would also allow us to constrain the importance of field-to-field variance. Clustering effects appear important even in large fields such as those from Subaru/Suprimecam and constraining them using the four DEEP2 fields should be a worthwhile endeavor. In summary, we have shown that the systematic examination of a large spectroscopic database can yield significant numbers of faint high-$z$ line emitters which are immediately suitable for further studies such as the determination of their number density (\\S~\\ref{numberdensity.sec}) or kinematics (\\S~\\ref{SpectralProperties.sec}). Automating such a search and applying it to an even larger spectroscopic database should be possible and should yield a large sample of high-quality high-$z$ LAEs suitable for a variety of studies." }, "0807/0807.4151_arXiv.txt": { "abstract": "name{\\color{red}\\bf Abstract} \\begin{abstract} {\\footnotesize We propose a conceptually and computationally simple method to evaluate the neutrinos emitted by supernova remnants using the observed $\\gamma$-ray spectrum. The proposed method does not require any preliminary parametrization of the gamma ray flux; the gamma ray data can be used as an input. In this way, we are able to propagate easily the observational errors and to understand how well the neutrino flux and the signal in neutrino telescopes can be constrained by $\\gamma$-ray data. We discuss the various possible sources of theoretical and systematical uncertainties ({\\em e.g.}, neutrino oscillation parameters, hadronic modeling, {\\em etc.}), obtaining an estimate of the accuracy of our calculation. Furthermore, we apply our approach to the supernova remnant RX J1713.7-3946, showing that neutrino emission is very-well constrained by the H.E.S.S. $\\gamma$-ray data: indeed, the accuracy of our prediction is limited by theoretical uncertainties. Neutrinos from RX J1713.7-3946 can be detected with an exposure of the order ${\\rm km}^2\\times\\,{\\rm year}$, provided that the detection threshold in future neutrino telescopes will be equal to about 1 TeV.} ", "introduction": "Introduction} Under-water and under-ice neutrino telescopes are instruments aiming to discoveries. They could reveal an effective acceleration of cosmic rays (CR) in galactic sources (such a supernova remnants \\cita{ginz} and micro-quasars \\cita{carla}) and/or in extragalactic sources (such as AGN and gamma ray bursts \\cita{halzen}). However, at present it is possible to obtain reliable expectations only for few of these sources, such as the supernova remnants (SNR) discussed in this paper. The idea that CR could originate in supernovae has been put forward already in 1934 but the first quantitative formulation of the conjecture that the young SNRs refurnish the Milky Way of cosmic rays, compensating the energy losses, is due to Ginzburg \\& Syrovatskii \\cita{ginz}. In fact, the turbulent gas of SNRs is a large reservoir of kinetic energy and this environment can support diffusive shock waves acceleration~\\cita{fermi}. The theory of CR acceleration in SNRs is still in evolution, but the generic expectations are stable. The CR flux in SNRs is expected to have a power law spectrum with spectral index $\\Gamma=2.0-2.4$ at low energies with a cutoff at an energy $E_{\\rm c}$ which depends on the details of the acceleration mechanism and on the age of the system. In specific implementations the cutoff energy $E_{\\rm c}$ can be as large as several PeV and, thus, is consistent with the ``knee\" in the CR spectrum at $E\\sim3\\times 10^{15}\\, {\\rm eV}$, which is believed to mark the transition from galactic to extra-galactic origin of CR~\\cita{berez}. In recent times, great progress has been made in the observation of SNRs. In particular, the High Energy Stereoscopic System (H.E.S.S.) \\cita{hessWEB} has determined quite precisely the gamma ray spectra of few SNRs showing that they extend above 10 TeV. In the context of Ginzburg \\& Syrovatskii hypothesis, it is natural to postulate that the observed gammas are produced by the decay of $\\pi^0$ (and $\\eta$) resulting from the collision of accelerated hadrons with the ambient medium. New and crucial observations are being collected and the hadronic origin seems to be favored for certain SNRs, such as Vela Jr \\cita{vela} and RX J1713.7-3946 \\cita{rxj,rxjhadr}.\\footnote{ See also \\cita{waxman} for a recent analysis leading to a different conclusion.} It is not yet possible, however, to exclude that (part of) the observed radiation is produced by electromagnetic processes. The definitive proof that SNRs effectively accelerate CR could be obtained by the observation of high energy neutrinos in neutrino telescopes presently in operation, in construction or in project \\cita{nutel}. As well known, there is a strict connection between photon and neutrino fluxes produced by hadronic processes in transparent sources (see, {\\em e.g.},~\\cita{gaisserbook}), which results from the fact that the same amount of energy is roughly given to $\\pi^0$, $\\pi^+$ and $\\pi^-$ in hadronic collisions. On this basis, one can estimate the neutrino fluxes expected from a sources with known $\\gamma$-ray spectra, trying to identify detectable sources and/or to optimize the detection strategies. The gamma-neutrino connection has been described in various recent papers \\cita{x1,Costantini:2004ap,x2,Vissani:2006tf,kappes,x4,x5,Vissani:2008zz} at a different level of accuracy, relying on different assumptions on the primary cosmic ray spectrum and/or hadronic interaction model. In this particular moment, when high energy gamma ray astronomy is flourishing and the neutrino telescopes are finally becoming a reality, it is clearly important to have solid and transparent predictions. We continue, thus, the work started in \\cita{Vissani:2006tf,Vissani:2008zz,Villante:2007mh}, proposing a method to calculate neutrinos fluxes which is, at the same time, simple, accurate and model-independent. Our results are in essence a straightforward applications of standard techniques \\cita{lipari}, but we believe that that they will be useful since they improve the existing calculations in various respects. In particular, we provide simple analytic expressions for the neutrino fluxes which have a general validity and can be applied directly to gamma ray data since they do not require any parametrization of the photon spectrum. This allows us to propagate easily the observational errors in the gamma ray flux and, thus, to understand how well the neutrino flux and the signal in neutrino telescopes can be constrained by $\\gamma$-ray data. We also discuss the various possible sources of theoretical and systematical uncertainties, obtaining an estimate of the accuracy of our method. This kind of analysis is relevant in the present situation, since the number of events expected in neutrino telescopes from SNRs is quite low (see, {\\em e.g.}, \\cita{Vissani:2006tf}) and even small (downward) revisions of the expected signals may be important and/or require different detection strategies. It is thus important to understand the relevance of different assumptions in the calculations and the origin of (apparently) contrasting results appeared in the literature. The plan of the paper is as follow. In Sect.~\\ref{method} we review the method presented in our previous paper~\\cita{Vissani:2006tf}. The relations obtained in this work -- Eqs.~(\\ref{allkernels}) -- are physically equivalent to those presented in \\cita{Vissani:2006tf}. However, they are more compact and more convenient for numerical computations since we have been able to recast them in such a way that they require only one numerical integration. In Sect.~\\ref{oscillations}, we discuss the effect of neutrinos oscillations (see \\cita{Strumia:2006db} for a review) and we discuss the relevance of uncertainty in neutrino mixing parameters for the predicted neutrino flux. In Sect.~\\ref{finalrelations} we present our main results, {\\em i.e.}, Eqs.~(\\ref{numuflux}) and (\\ref{antinumuflux}) which relate the (oscillated) muon neutrino and antineutrino fluxes to the $\\gamma$-ray flux. As it is explained, the only necessary information to predict neutrino fluxes are the relative production rates of the various mesons in hadronic processes, which are robust predictions of hadronic interaction models. In this paper, we adopt the results from Pythia \\cita{pythia}, estimated by using the parametrization of hadronic cross sections presented in \\cita{Koers:2006dd}. A comparison with SYBILL~\\cita{sybill} and DPMJET-III \\cita{dpmjet} is performed (when possible) by using the parametrization and tabulations of hadronic cross sections presented by \\cita{Kelner:2006tc} and \\cita{Huang:2006bp}. In Sect.~\\ref{fluxes} and \\ref{rates} we propose a procedure to predict neutrino fluxes and event rate in $\\nu$-telescopes directly from $\\gamma$-ray observational data, and we apply it to RX J1713.7-3946 which is presently the best studied SNR. In Sect.~\\ref{conclusions} we summarize our main results. ", "conclusions": "" }, "0807/0807.3547_arXiv.txt": { "abstract": "Hyper-accreting discs occur in compact-object mergers and in collapsed cores of massive stars. They power the central engine of $\\gamma$-ray bursts in most scenarios. We calculate the microphysical dissipation (the viscosity and resistivity) of plasma in these discs, and discuss the implications for their global structure and evolution. At the temperatures ($k_{\\rm b}T > \\mec2$) and densities ($\\rho\\sim 10^{9}-10^{12}~ {\\rm gr~cm^{-3}}$) characteristic of the neutrino-cooled innermost regions, the viscosity is provided mainly by mildly degenerate electrons, while the resistivity is modified from the Spitzer value due to the effects of both relativity and degeneracy. Under these conditions the magnetic Reynolds number is very large ($R_{\\rm eM} \\sim 10^{19}$) and the plasma behaves as an almost ideal magneto-hydrodynamic (MHD) fluid. Among the possible non-ideal MHD effects the Hall term is relatively the most important, while the magnetic Prandtl number, $\\Pr$ (the ratio of viscosity to resistivity), is typically larger than unity: $10 \\lsim \\Pr \\lsim 6 \\times 10^3$. Inspection of the outer radiatively inefficient regions indicates similar properties, with magnetic Prandtl numbers as high as $\\sim 10^{4}$. Numerical simulations of the magneto-rotational instability (MRI) indicate that the saturation level and angular momentum transport efficiency may be greatly enhanced at high Prandtl numbers. If this behaviour persists in the presence of a strong Hall effect we would expect that hyper-accreting discs should be strongly magnetised and highly variable. The expulsion of magnetic field that cannot be dissipated at small scales may also favour a magnetic outflow. We note that there are limited similaries between hyper-accreting discs and X-ray binary discs -- which also have a high magnetic Prandtl number close to the black hole -- which suggests that a comparison between late-time activity in $\\gamma$-ray bursts and X-ray binary accretion states may be fruitful. More generally, our results imply that the possibly different character of high Prandtl number MHD flows needs to be considered in studies and numerical simulations of hyper-accreting discs. ", "introduction": "The fluid dynamics of dense accretion discs is described by the Navier-Stokes and magneto-hydrodynamic equations. Within such a description, the microscopic properties of the fluid are parameterized (in part) via the kinematic viscosity $\\nu$, which describes the rate at which inter-particle collisions damp fluid motions, and the resistivity $\\eta$, which quantifies the dissipation of currents via Ohmic losses. Timescales can be associated with each of these quantities. The viscous timescale of a disc $t_{\\rm acc} = r^2 / \\nu$, for example, is the timescale on which the microscopic viscosity would lead to angular momentum transport over a scale $r$. As is well known \\citep{pringle81}, for astrophysical discs the timescales for evolution driven by the microscopic viscosity are many orders of magnitude larger than the observed evolution timescales. It is therefore generally assumed that evolution is driven by an ``anomalous\" or ``turbulent\" viscosity $\\nu_{\\rm t} \\gg \\nu$ \\citep{shakura73}, possibly augmented in some cases by angular momentum loss in a wind \\citep{blandford82}. In most astrophysical discs $\\nu_t$ is likely generated from turbulence driven by the magneto-rotational instability (MRI; see \\citet{balbus98} and references therein). Since $\\nu_{\\rm t} \\gg \\nu$ most studies of discs make the implicit assumption that the actual values of $\\nu$ and $\\eta$ are unimportant, and consider instead evolution under the action of the effective transport coefficients $\\nu_t$ and $\\eta_t$. Provided that the Reynolds ($R_{\\rm e}$) and magnetic Reynolds ($R_{\\rm eM}$) numbers are large enough, the absolute value of either one may indeed be of little import for accretion discs. Their ratio, however, which we define as the magnetic Prandtl number, \\begin{equation} {\\rm Pm} \\equiv \\frac{\\nu}{\\eta} \\end{equation} is almost certainly {\\em not} ignorable \\citep{balbus98}. An accretion disc serves to convert the large-scale energy present in the shear flow into other energy forms -- turbulent kinetic energy, magnetic field, and heat -- and the dissipation required to generate heat may vary strongly depending upon the microphysical transport coefficients. In particular, if ${\\rm Pm} \\gg 1$ the plasma is highly viscous on the scale at which resistivity operates. In this regime one may expect that dissipation of the magnetic field will be suppressed, with a resulting modification in the saturation level of dynamo generated fields and an inverse cascade of magnetic energy to large scales \\citep{brandenburg01}. The importance of the magnetic Prandtl number for the resulting magnetic field structure has been illustarted analytically \\citep{umurhan07,umurhan07b} and numerically, both in idealised simulations of dynamo action within forced turbulence \\citep{schekochihin04} and in local shearing-box simulations of the MRI within accretion discs \\citep{lesur07,fromang07}\\footnote{Since dissipation in discs typically occurs on small scales for which the fluid is unaware of the large scale shear, one would anticipate that the basic physics of large ${\\rm Pm}$ dynamos is similar. Disc simulations are required, however, to assess the impact of the small-scale physics on MRI saturation and the large scale flow properties. It is unclear whether the shearing-box approach is adequate for study of these effects.}. For discs the current limited set of simulations indicates that the efficiency of angular momentum transport, parameterized via the Shakura-Sunyaev $\\alpha$ parameter, increases with ${\\rm Pm}$, and can approach $\\alpha \\sim 1$ given the combination of modest ${\\rm Pm} = 8$ and a weak net vertical magnetic field. Motivated by these results, \\citet{balbus08} calculated the expected radial variation of ${\\rm Pm}$ within standard Shakura-Sunyaev models of geometrically thin, radiatively efficient accretion discs. They found that ${\\rm Pm}$ exceeds unity within $\\sim 50$ Schwarzschild radii of the central object, and suggested that the qualitatively different behaviour of the MRI in the high ${\\rm Pm}$ limit might be responsible for the time-dependent outbursts and state changes that are observed in X-ray binaries. In this paper, we compute the magnitude and radial dependence of the magnetic Prandtl number in hyper-accreting discs. Such discs, which can support accretion rates of the order of $1 \\ M_\\odot \\ {\\rm s}^{-1}$, form from compact object mergers \\citep[e.g.][]{ruffert99} and when the cores of rapidly rotating massive stars collapse \\citep{woosley93}. The disc structure is shaped by neutrino emission in the regions close to the accreting object, while in the outer parts neither neutrinos (the temperature is too low) nor photons (which are trapped within the flow) can provide cooling. Hyper-accreting discs are invoked to provide the initial energy injection in many models for $\\gamma$-ray bursts (GRBs). Our ultimate motivation for studying the magnetic Prandtl number in these discs is to understand the complex phenomenology (jets, highly variable prompt emission, X-ray flares, plateau phases etc) that is observed in GRBs \\citep[e.g.][]{burrows05} and which may well derive from the evolution of the disc. We anticipate different results from those found by \\citet{balbus08} for photon-cooled discs. In the neutrino-cooled region the differences arise from two reasons. First, since neutrino opacities are vastly smaller than photon opacities, hyper-accreting discs are much cooler and denser than an extrapolation of Shakura-Sunyaev discs would imply, and this on its own would alter the magnetic Prandtl number. Second, the temperatures and densities in hyper-accreting discs are such that the electrons are mildly relativistic and mildly degenerate, while the elastic collision cross-section for the nuclei exceeds the Coulomb cross-section. As a consequence, the classical \\citet{spitzer62} expressions for $\\nu$ and $\\eta$, which are adequate for ordinary accretion flows, no longer apply. They do apply, instead, in the outer regions of hyper-accreting discs, where a different $\\Pr$ behaviour occurs since the flow is radiatively inefficient and photon-pressure dominated. The organization of this paper is as follows. In \\S2 through \\S5 we describe the formalism for calculating the structure of hyper-accreting discs. This is required here in order to determine consistently the plasma conditions within these discs -- including the temperature, density, nuclear composition and degree of degeneracy. Readers familiar with vertically averaged hyper-accreting disc models will find that our approach generally follows standard practice. The structure of the resulting disc models is presented in \\S6. Given these conditions, we then calculate in \\S7 and \\S8 the resisitivity and viscosity in the neutrino cooled inner region of the disc. These Sections contain the principle new results of this paper. Combining the viscosity and resistivity, we show in \\S10 that ${\\rm Pm} > 1 ~{\\rm or} \\gg 1$ across the entire radial range relevant for GRB models. In \\S11 we estimate the strength of other non-ideal MHD effects, and in \\S12 we study the structure of the outer non-radiative region of the disc. \\S13 and \\S14 discuss and summarize our results, and what they may imply for disc evolution and GRB observables. ", "conclusions": "In this paper, we calculated the magnetic Prandtl number for plasma conditions in hyper-accreting discs. This dimensionless number $\\Pr = \\nu/\\eta = \\lambda_{\\nu}^{2}/\\lambda_{\\eta}^{2}$ expresses the relative size of two critical turbulent flow scales: the scale on which velocity fluctuations are viscously damped $\\lambda_{\\nu}$, and the scale $\\lambda_{\\eta}$ on which ohmic losses result in magnetic field dissipation. We find that in the inner neutrino-cooled regime: \\begin{itemize} \\item[(1)] Electric resistivity involves relativistic, mildly degenerate electrons. As a consequence, the resistivity is very low and weakly dependent on density and temperature. \\item[(2)] The main source of viscosity is Coulomb collisions between mildly {\\it degenerate electrons} and protons. It decreases as the flow gets denser or cooler, since the electron fraction decreases. \\item[(3)] The magnetic Prandtl number is always {\\em greater} than unity, unless the angular momentum transport efficiency, parameterized via the Shakura-Sunyaev $\\alpha$, is very small ($\\alpha < 0.01$). $\\Pr$ ranges between a few tens to a few $10^3$, and it typically increases with $\\alpha$. In the optically {\\em thin} regions for neutrinos, the main dependence of $\\Pr$ is on density (inversely, via viscosity) since the disc temperature varies very slowly as a function of $\\dot{M}$ and radius. This causes $\\Pr$ to {\\em decrease} with increasing accretion rates. In the optically {\\em thick} regions, the main dependence is on temperature, that varies more rapidly than density. As a consequence, $\\Pr$ {\\em increases} with $\\dot{M}$ \\end{itemize} \\no In the outer, radiatively inefficient region: \\begin{itemize} \\item[(4)] Resistivity and viscosity have the classical ``Spitzer'' dependencies on temperature and density, since electrons are non-relativistic and non-degenerate and viscosity is mainly given by Coulomb interactions between ionised helium particles. The magnetic Prandtl number is thus a much stronger function of temperature $\\Pr \\propto T^{4}/\\rho$. In contrast, it does not depend on the overall magnitude of accretion (via $\\dot{M}$ and $\\alpha$). The Prandtl number is always much larger than unity ($\\Pr \\simeq 10^4 (100~\\rs)/r$) and it is unlikely to fall below unity for the entire extent of the disc. \\end{itemize} \\no We conclude by noting a number of implications for the global structure of hyper-accreting discs: \\begin{itemize} \\item[(5)] For all plausible values of the magnetic field strength in the disc, the Hall effect provides the largest non-ideal term in the MHD equations. These discs are perhaps a unique example of a high magnetic Reynolds number flow in which the Hall effect is the dominant non-ideal term. Since there are no simulations where these conditions are accounted for, their effect on the magnetic field evolution and MRI is unclear. \\item[(6)] The large values of the magnetic Prandtl number mean that the evolution of magnetic field within hyper-accreting discs is likely to be qualitatively different from discs with lower magnetic Prandtl numbers. The results of forced dynamo and MRI simulations suggest that, in the high $\\Pr$ regime, small scale field dissipation is suppressed and the saturation level of the magnetic field is enhanced, perhaps dramatically \\citep{lesur07,fromang07}. Numerical simulations of hyper-accreting discs that do not account for microphysical dissipation may therefore have underestimated the magnetic field strength. \\item[(7)] The {\\em immediate} consequence of a high value of $\\Pr$ for models of the central engines of GRBs is that disc formation is likely to be accompanied by vigorous expulsion of magnetic flux. This favours models in which energy is liberated in the form of a magnetic jet or outflow. \\end{itemize}" }, "0807/0807.3292_arXiv.txt": { "abstract": "Generic inspirals and mergers of binary black holes produce beamed emission of gravitational radiation that can lead to a gravitational recoil or {\\em kick} of the final black hole. The kick velocity depends on the mass ratio and spins of the binary as well as on the dynamics of the binary configuration. Studies have focused so far on the most astrophysically relevant configuration of quasi-circular inspirals, for which kicks as large as $\\sim 3,300\\, \\KMS$ have been found. We present the first study of gravitational recoil in {\\em hyperbolic} encounters. Contrary to quasi-circular configurations, in which the beamed radiation tends to average during the inspiral, radiation from hyperbolic encounters is plunge dominated, resulting in an enhancement of preferential beaming. As a consequence, it is possible to achieve kick velocities as large as \\kickmax. ", "introduction": " ", "conclusions": "" }, "0807/0807.3771_arXiv.txt": { "abstract": "We show that the proper motion of the Becklin-Neugebauer (BN) object is consistent with its dynamical ejection from the $\\theta^1$~Ori~C binary, contrary to recent claims by G\\'omez et al. Continued radio observations of BN and future precise astrometric observations of $\\theta^1$~Ori~C with SIM and the Orion Nebula Cluster with GAIA can constrain the properties of this ejection event, with implications for theories of how the nearest example of massive star formation is proceeding. ", "introduction": "\\label{S:intro} Understanding massive star formation remains one of the most challenging and important problems of contemporary astrophysics (Beuther et al. 2007; Zinnecker \\& Yorke 2007). The complexity of the process means that massive star formation theories, such as the turbulent core model (McKee \\& Tan 2003), the competitive accretion model (Bonnell \\& Bate 2006) and stellar coalescence model (Bonnell et al. 1998; Clarke \\& Bonnell 2008) require close testing against observed systems. The closest forming (i.e. accreting) massive star is thought to be radio source I (Menten \\& Reid 1995) within the Orion Nebula Cluster (ONC), at a distance of $414\\pm7$~pc (Menten et al. 2007, adopted throughout), in the Kleinmann-Low (KL) region. As reviewed by Tan (2008), this source has been used as observational evidence in support of all three of the above theories. Part of this confusion is due to the Becklin-Neugebauer (BN) object, 9.9\\arcsec to the NW (Fig.~1), which is a fast moving (radio-ONC-frame proper motion of $\\mu_{\\rm BN}=13.2\\pm 1.1\\:{\\rm mas\\:yr^{-1}}$, i.e. $v_{\\rm 2D,BN}=25.9\\pm2.2\\:{\\rm km\\:s^{-1}}$ towards P.A.$_{\\rm BN}=-27^\\circ.5\\pm4^\\circ$, Plambeck et al. 1995; G\\'omez et al. 2008) embedded B star ($L_{\\rm BN}=(2.1 - 8.5)\\times 10^3L_\\odot$, Gezari, Backman \\& Werner 1998, equivalent to a zero age main sequence mass $m_{\\rm BN,zams} = 9.3\\pm2.0\\sm$). This proper motion implies that BN has been moving through the KL region and made a close, possibly coincident, passage with source {\\it I} about 500 years ago. Thus to understand the nearest example of massive star formation, we need to understand the origin of BN's motion. Including the $(+21) - (+8) = +13\\:{\\rm km\\:s^{-1}}$ radial velocity of BN with respect to the ONC mean (Scoville et al. 1983; Walker 1983), the 3D ONC-frame velocity of BN is $v_{\\rm 3D,BN}=29\\pm3\\:{\\rm km\\:s^{-1}}$, and its kinetic energy is $E_{\\rm BN} = (8.3\\pm2.3)\\times 10^{46} (m_{\\rm BN}/10\\sm)\\:{\\rm ergs}$. BN is very likely to have formed somewhere in the ONC and then attained its high speed by a close interaction with a massive multiple stellar system followed by dynamical ejection (Poveda, Ruiz \\& Allen 1967). Tan (2004) proposed BN was launched from the $\\theta^1$~Ori~C binary (also shown in Fig.~\\ref{fig:bn}), since this is the only stellar system in the ONC known to have all the physical properties required by this scenario: (1) a location along BN's past trajectory (\\S\\ref{S:ast}); (2) an (optical)-ONC-frame proper motion ($\\mu_{\\theta^1C}=2.3\\pm0.2\\:{\\rm mas\\:yr^{-1}}$, van Altena et al. 1988, i.e. $v_{\\rm 2D,\\theta^1C} = 4.5\\pm0.4\\:{\\rm km\\:s^{-1}}$, towards $\\rm P.A._{\\theta^1C}=142^\\circ.4\\pm4^\\circ$) that is in the opposite direction to BN (the direction to BN from $\\theta^1$~Ori~C is a P.A.$=-30^\\circ.949$) and is of the appropriate magnitude (the dynamical mass of BN implied by this motion agrees with the estimate of $m_{\\rm BN,zams}$ and is $m_{\\rm BN,dyn}=8.6\\pm1.0\\sm$ assuming negligible error in $m_{\\theta^1C}=49.5\\sm$ and negligible motion of the pre-ejection triple system in this direction; a pre-ejection motion of 0.35~mas/yr along this axis (\\S\\ref{S:high}) would contribute an additional $1.5\\sm$ uncertainty); (3) primary ($m_{\\theta^1C-1}=34\\sm$) and secondary ($m_{\\theta^1C-2}=15.5\\sm$) masses greater than $m_{\\rm BN}$ (Kraus et al. 2007); (4) a semi-major axis of $a=17.0\\pm5.8$~AU (Patience et al. 2008) and thus a total orbital energy ($E_{\\rm tot}=Gm_{\\theta^1C-1}m_{\\theta^1C-2}/(2a)= (2.7\\pm0.9)\\times 10^{47}\\:{\\rm ergs}$) greater than the sum of BN's kinetic energy and $\\theta^1$~Ori~C's kinetic energy ($1.00\\times 10^{46}\\:{\\rm ergs}$) (see Tan 2008 for a review). Note, $\\theta^1$~Ori~C's recoil in this scenario is large enough to remove it from the Trapezium region (see Pflamm-Altenburg \\& Kroupa 2006 for theoretical studies of the dynamical decay of Trapezium-like systems) and may be enough to eject it from the ONC completely, with implications for the effectiveness of its ionizing feedback on disrupting the star cluster formation process. Rodr\\'iguez et al. (2005) and Bally \\& Zinnecker (2005) proposed BN was launched from an interaction with radio source {\\it I}, which would require this system to be a massive binary, recoiling away from any large scale ($\\gtrsim 100$~AU) gas that it was originally accreting. G\\'omez et al. (2008) used the relative motion to BN with respect to source {\\it I} to claim that BN could not have made a close passage with $\\theta^1$~Ori~C, excluding this possibility at the 5-10~$\\sigma$ level. We show in \\S\\ref{S:ast} that if BN's motion is considered in the reference frame of the ONC, then a close (coincident) passage with $\\theta^1$~Ori~C is allowed by the data, which permits the scenario of dynamical ejection of BN from $\\theta^1$~Ori~C. In \\S\\ref{S:high} we discuss the potential of future high precision astrometric measurements to constrain the properties of BN's dynamical ejection, which then constrain BN's interaction distance with source {\\it I}, the mass of source {\\it I}, and thus the strength of tidal perturbations on the massive protostar during this encounter. \\vspace{0.2in} ", "conclusions": "We have reviewed the latest evidence that BN was dynamically ejected from the $\\theta^1$~Ori~C binary, finding that $\\theta^1$~Ori~C has all the physical properties expected in this scenario. We showed that the trajectory of BN is also consistent with this scenario, in contrast to recent claims by G\\'omez et al. (2008). We discussed how high precision astrometry of $\\theta^1$~Ori~C with SIM can yield information on the pre-ejection velocity of the system and the size of any subsequent deflections, in particular that of BN caused by close passage with source {\\it I}, the nearest massive protostar." }, "0807/0807.0291_arXiv.txt": { "abstract": "The amdlib AMBER data reduction software is meant to produce AMBER data products from the raw data files that are sent to the PIs of different proposals or that can be found in the ESO data archive. The way defined by ESO to calibrate the data is to calibrate one science data file with a calibration one, observed as close in time as possible. Therefore, this scheme does not take into account instrumental drifts, atmospheric variations or visibility-loss corrections, in the current AMBER data processing software, amdlib. In this article, we present our approach to complement this default calibration scheme, to perform the final steps of data reduction, and to produce fully calibrated AMBER data products. These additional steps include: an overnight view of the data structure and data quality, the production of night transfer functions from the calibration stars observed during the night, the correction of additional effects not taken into account in the standard AMBER data reduction software such as the so-called \"jitter\" effect and the visibility spectral coherence loss, and finally, the production of fully calibrated data products. All these new features are beeing implemented in the modular pipeline script \\texttt{amdlibPipeline}, written to complement the amdlib software. ", "introduction": "\\label{sect:intro}% The AMBER data reduction software: \\texttt{amdlib}\\cite{2007A&A...464...29T}, features a new type of algorithm, the so-called P2VM algorithm\\cite{2004SPIE.5491.1222M}, which performs a direct fit of the observed fringes using a template fringe pattern recorded during the instrument calibration and produces a series of observables; namely, the squared visibility, the closure phase, and the differential phase\\cite{2006-ITHD_FMillour}. The standard calibration plan handles the AMBER signal from the raw data to raw averaged visibilities OI DATA files. Though a simple calibration script (\\texttt{amdlibDivideOiData}) is included in the \\texttt{amdlib} software, a complete calibration procedure is not yet provided. In this document, we present our approach to produce calibrated OI data files. It consists of several recipes, just like for the standard \\texttt{amdlib} software, to organize, evaluate, plot, and finally, calibrate the data. This philosophy allows one to switch easily from one recipe to another if necessary. This leads to the modular structure of the example pipeline script \\texttt{amdlibPipeline}\\footnote{currently available at \\url{http://www.mpifr-bonn.mpg.de/staff/fmillour/}}, written in the scientific language \\texttt{yorick}, to perform the final steps of data reduction after an \\texttt{amdlib} data reduction session. This paper is organized into three different sections: \\begin{itemize} \\item A section about file management and coping with different types of calibrations (different DITs, different modes, use of BCD), \\item The rise of advanced calibration methods: ``coherence length'' and ``jitter'' effects correction, \\item How to cope with calibration stars and transfer function, as well as a template for data calibration. \\end{itemize} ", "conclusions": "The standard default calibration scheme proposed for AMBER (calibrating one file with another) is obviously insufficient to get realistic error bars, and an overnight calibration quality estimate. We presented a proposition for a user-friendly pipeline framework for the amdlib software. We are able, with the proposed scripts and an Internet connection, to produce science-grade data with realistic error bars in a semi-automated way, as it is already done for the first data reduction steps in \\texttt{amdlib}. This article is thought to be a starting point for further reflexions about AMBER data calibration. Further work could allow one to perform the AMBER data reduction in a fully automated way, allowing in the future the production of archived data products, in the same way as it is done with the ESO imaging and spectroscopic instruments (WFI, FEROS, etc.)." }, "0807/0807.0634_arXiv.txt": { "abstract": "The recent advent of laser guide star adaptive optics (LGS AO) on the largest ground-based telescopes has enabled a wide range of high angular resolution science, previously infeasible from ground-based and/or space-based observatories. As a result, scientific productivity with LGS has seen enormous growth in the last few years, with a factor of $\\approx$10 leap in publication rate compared to the first decade of operation. Of the 54~refereed science papers to date from LGS AO, half have been published in the last $\\approx$2 years, and these LGS results have already made a significant impact in a number of areas. At the same time, science with LGS AO can be considered in its infancy, as astronomers and instrumentalists are only beginning to understand its efficacy for measurements such as photometry, astrometry, companion detection, and quantitative morphology. We examine the science impact of LGS AO in the last few years of operations, largely due to the new system on the Keck~II 10-meter telescope. We review currently achieved data quality, including results from our own ongoing brown dwarf survey with Keck LGS. We assess current and near-future performance with a critical eye to LGS AO's capabilities and deficiencies. From both qualitative and quantitative considerations, it is clear that the era of regular and important science from LGS AO has arrived. ", "introduction": "\\label{sec:intro} % Astronomers have envisioned using laser guide star adaptive optics (LGS AO) to achieve diffraction-limited observations from ground-based telescopes for over two decades\\cite{1985A&A...152L..29F, 1987Natur.328..229T, 1994OSAJ...11..263H}. (See Refs.~\\citenum{2004aoa..book.....R} and~\\citenum{1998aoat.conf.....H} for a historical review.) The realization of this vision has required the dedication, talents, and resources of myriad individuals and organizations. LGS AO is now being used regularly to conduct a broad range of science. The promise of near diffraction-limited imaging and spectroscopy from the ground over most of the sky is coming to fruition. This heralds an important new capability for observational astronomy. The purpose of this paper is to examine the science that has been done using LGS AO, with a critical eye to what has been achieved and what promises and challenges remain. Much has been accomplished since the last review of science from LGS (Ref.~\\citenum{2006SPIE.6272E..14L}), which focused on publication output. Here, we go beyond mere productivity and examine the science {\\em impact} of LGS as demonstrated in the record of published research. ", "conclusions": "\\subsection{A High Angular Resolution Survey of Field Brown Dwarfs} To illustrate what is currently possibly with LGS, we examine the on-sky performance of the Keck LGS system. Since the inception of the Keck system for open science use in 2005, my collaborators and I have been conducting a high angular resolution near-IR study of nearby brown dwarfs. Our goals are (1) to assess the binary frequency of ultracool dwarfs; (2) to test atmospheric models with these coeval systems, \\eg, as associated with the abrupt spectral transition from the L~dwarfs to the T~dwarfs;\\cite{2006astro.ph..5037L} (3) to search for exceptionally low-temperature companions;\\cite{2008A&A...482..961D} and (4) to find and monitor substellar binaries suitable for dynamical mass determinations.\\cite{liu08-2m1534orbit} LGS AO represents a major advance for this science area. Most known brown dwarf binaries have separations of $\\lesssim$0.3\\arcsec,\\cite{2006astro.ph..2122B} hence the need for high angular resolution imaging to find and characterize them via resolved photometry and spectroscopy, but they are too optically faint for natural guide star AO. Our observations have achieved 3--4$\\times$ the angular resolution at $K$-band (2.2~\\micron) compared to \\HST\\ and thus are more sensitive to close companions. In addition, the ability of Keck LGS AO to find tighter binaries means that systems with much shorter orbital periods than the current sample can be found and expeditiously monitored. Finally, many of the key spectral diagnostics of brown dwarfs are in the infrared and thus probing the atmospheric properties of these objects with resolved colors and spectra of binaries is well-suited to current LGS capabilities. \\begin{figure}[h] \\vskip -0.5in \\begin{center} \\begin{tabular}{c} \\hskip -0.7in \\includegraphics[width=3.5in,angle=90]{plot-obs-summary-fwhm-k} \\hskip -1.6in \\includegraphics[width=3.5in,angle=90]{plot-obs-summary-fwhm-ferror-k.ps}\\\\ \\hskip -0.6in \\includegraphics[width=3.5in,angle=90]{plot-obs-summary-strehl-k} \\hskip -1.6in \\includegraphics[width=3.5in,angle=90]{plot-obs-summary-strehl-ferror-k.ps}\\\\ \\end{tabular} \\end{center} \\vskip -4ex \\caption[fig:kecklgs] { \\label{fig:kecklgs} Summary of Keck LGS AO $K$-band (2.2~\\micron) performance, based on our near-IR imaging survey of brown dwarfs (Liu \\etal, in prep). No bad data have been censored, so the compilation comprises a mix of seeing conditions, target airmasses, and technical performance. {\\bf Top panels:} Histogram of $K$-band FWHM and its fractional RMS variation within a given dataset. Typically each dataset is composed of 6--12 individual images taken over an elapsed time of 10--20~minutes. The median FWHM is 0.076\\arcsec\\ with a median RMS variation of 5\\%. {\\bf Bottom panels:} Same histograms for $K$-band Strehl. The median Strehl is 0.17 with an RMS variation of a factor of 1.19 (\\eg, Strehl = $0.17\\pm0.03$).} \\end{figure} Our brown dwarf imaging survey provides an excellent dataset for assessing typical Keck LGS performance in the case of off-axis observations, namely the situation where the LGS is pointed to the science target but tiptilt sensing and correction are derived from an adjacent field star. Brown dwarfs are far too optically faint to serve as their own tiptilt references and hence the need for a nearby tiptilt star -- this is the same observing situation as for many extragalactic LGS applications and thus provides a good reference point. For Keck, the tiptilt star must be within 60\\arcsec\\ of the science target -- in practice, we find that this results in a sky coverage fraction of about 2/3 for an estimated $K$-band Strehl ratio of $\\gtrsim$0.2. Since targets for our brown dwarf program span most of the sky (except for avoidance of the galactic plane), this 2/3 sky coverage estimate is a fair representation of the fraction of any set of generic targets that can be imaged with LGS. Figure~\\ref{fig:kecklgs} summarizes the image quality of a subset of our Keck LGS observations of nearby brown dwarfs, spanning multiple observing runs from 2005--2007. No bad data have been censored, so a mix of seeing conditions, off-axis tiptilt star properties, and technical performance (e.g., LGS projected power and sodium light return flux) are represented. (See also Ref.\\ \\citenum{2006SPIE.6272E..14L} for more performance descriptions.) The median $K$-band image FWHM for our survey is 0.076\\arcsec\\ with a best value of 0.051\\arcsec. The median $K$-band Strehl is 0.17 with a best value of 0.34. LGS images are naturally time-variable, and the shape and detailed structure of the PSF changes in every image. Figure~\\ref{fig:kecklgs} provides one representation of this PSF variability, plotting histograms of the fractional RMS deviations in $K$-band FWHM and Strehl ratios. Typically, each of our brown dwarf datasets constitutes a series of 6--12~images, each with a integration of about 1~min and total elapsed time of about 15--30~min. Over these time scales, the plotted histograms show significant FWHM and Strehl variations. This is obviously a challenge for science programs requiring a stable PSF. \\begin{figure}[t] \\vskip -0.3in \\begin{center} \\begin{tabular}{c} \\hskip -0.75in \\includegraphics[width=3.75in,angle=90]{make-figure-orbit-skyplot.labels.fancy_color.1orbit.ps} \\hskip -1.8in \\includegraphics[width=3.75in,angle=90]{make-figure-orbit-sep.ps} \\end{tabular} \\end{center} \\vskip -6ex \\caption[fig:2m1534-orbit] { \\label{fig:2m1534-orbit} First dynamical mass determination for a binary T~dwarf, the T5.0+T5.5 binary 2MASS~J1534-2952AB (Ref.\\ \\citenum{liu08-2m1534orbit}). Six epochs of data from Keck LGS are combined with \\HST\\ archival data to form a dataset from 2000-2008 that spans about 50\\% of the 15-year orbital period. Through careful analysis of the binary images with PSF fitting, the Keck LGS data achieve sub-milliarcsecond relative astrometry of the two components. The total mass is $59\\pm3~\\Mjup$. This is the first field binary for which both components are directly confirmed to be below the stellar/substellar limit ($\\lesssim$75~\\Mjup). This is also the coolest and lowest mass binary with a dynamical mass determination to date. The left plot shows the relative astrometry and best-fitting orbit. The right plot shows the separation measurements compared to the best fitting orbit (solid line) and two alternative orbital periods of similar total mass (dashed and dotted lines). The right bottom sub-panel shows the residuals between the observations and the predictions from the orbits.} \\end{figure} Despite the notable PSF fluctuations, high precision science can be with Keck LGS data. For example, we have been using LGS to carefully monitor the orbits of binary brown dwarfs in order to determine their dynamical masses and thereby test theoretical models. Despite the hundreds of brown dwarfs that have been identified and had their spectrophotometric properties characterized, direct measurements of their most fundamental property, namely their mass, are sorely lacking. Typical estimated orbit periods are a decade a more, but many of the binaries were discovered $\\gtrsim$5~years ago with \\HST\\ and thus multi-epoch LGS followup is well-suited to dynamical mass determinations. Figure~\\ref{fig:2m1534-orbit} shows an example of what is possible, presenting the orbit of the T5+T5.5 binary 2MASS~J1534-2952AB by Liu, Dupuy and Ireland (Ref.\\ \\citenum{liu08-2m1534orbit}). This is the first dynamical mass determination for a binary T~dwarf, the coolest and least luminous class of brown dwarfs. There are two stars within 6\\arcsec\\ of \\twomassbin\\ which we use as PSFs to deblend the light of the two binary components. Through extensive Monte Carlo testing of different deblending methods applied to simulated images of binary stars constructed from the single PSFs, we are able to measure the relative position and orientation of the components to 0.3--0.7~mas and 0.2--0.7\\degs\\ RMS at $K$-band. The precision of the best measurements is such that atypical sources of uncertainty need to be considered, including: \\begin{enumerate} \\item The calibration of the instrumental plate scale and orientation can introduce additional uncertainty, since these values for the Keck facility near-IR camera are known to about 1~part in $10^{-3}$ and 0.1\\degs, respectively. \\item The differing spectral types of the two components means that the light from each is subjected to a slightly different amount of differential chromatic refraction (DCR). Given the sky orientation of the binary (just about North-South), the declination of the target ($\\delta=-29\\degs$ meaning the smallest possible airmass for Keck observing is 1.55), and the fact that we observe it near transit (for best AO performance), the DCR causes the separation of the binary to appear slightly smaller at $H$ and $K$-bands and slightly larger at $J$-band compared to the true position as would be observed at zenith. The amplitude of the DCR effect is about 0.3~mas, much smaller than the measurement errors at most (but not all) of the Keck epochs. However, the effect is a systematic one so we correct the relative astrometry of the two components based on synthetic photometry from T~dwarf spectra. (Since T~dwarfs have suppressed flux at the longward portion of the $K$-band, the DCR is smaller than it would be for L~dwarfs observed at the same airmass.) \\end{enumerate} \\noindent The quality of the Keck LGS astrometry equals or exceeds that available for the same target from \\HST, though at the very smallest separations ($\\approx$FWHM) \\HST\\ data still have an advantage since the very stable \\HST\\ PSF can be simulated to high fidelity\\cite{1993ASPC...52..536K} and then fitted to tight binary images (\\eg, Ref.\\ \\citenum{liu08-2m1534orbit}). However, LGS AO provides the necessary long-term platform for synoptic monitoring of visual binaries, especially since the required amount of observing time at each epoch is relatively modest but many epochs are needed. This in contrast to \\HST\\ where target acquisition is slow and monitoring a populous sample over many epochs is telescope time-intensive. \\subsection{Review of Published Keck Performance} To further summarize the current capability, {\\bf Table~3} provides a summary of achieved Keck LGS performance as presented in the published science papers. Obviously this is a heterogeneous assemblage from both the standpoint of the observations and the subsequent analysis, but it does provide ``ground truth'' of the quality of on-sky data that has proven suitable for publication. We consider several types of measurements that readily lend themselves to quantitative uncertainties (as opposed to more difficult measurements to quantify such as, \\eg, morphological studies of resolved objects): \\begin{itemize} \\item {\\em Relative photometry:} the flux of a science target relative to other sources in the same image of known brightness (\\eg, from 2MASS); \\item {\\em Absolute photometry:} the absolute flux of a science target as directly measured from the images of the target and then compared to a photometric calibrator star observed contemporaneously (but not simultaneously); \\item {\\em Relative astrometry:} the location of sources in an LGS dataset relative to each other. \\item {\\em Absolute astrometry:} the location of sources in a LGS dataset as referenced to absolute astrometry tied to an external dataset (\\eg, \\HST). \\item {\\em Binary (relative) photometry:} the relative flux ratio of a binary; \\item {\\em Binary (relative) astrometry:} the separation and position angle of a binary; \\item {\\em Crowded field astrometry:} the positions and/or proper motions for a field with many sources (\\eg, star clusters and resolved nearby galaxies); \\item {\\em Crowded field (relative) photometry:} the same as relative photometry above, except for a field with many sources; \\item {\\em Crowded field (absolute) photometry:} the same as absolute photometry above, except for a field with many sources. \\end{itemize} \\noindent {\\bf Table~3} shows that high quality measurements have been achieved with LGS, especially for relative measurements. In the published papers, the most measurements are available for relative photometry and astrometry of low-mass binaries, since this is a class of science that readily benefits from LGS." }, "0807/0807.0128_arXiv.txt": { "abstract": "We present detailed analysis of three globular cluster X-ray sources in the XMM-Newton extended survey of M31. The X-ray counterpart to the M31 globular cluster Bo\\thinspace 45 (XBo\\thinspace 45) was observed with XMM-Newton on 2006 December 26. Its combined pn+MOS 0.3--10 keV lightcurve was seen to vary by $\\sim$10\\%, and its 0.3--7.0 keV emission spectrum was well described by an absorbed power law with photon index 1.44$\\pm$0.12. Its variability and emission is characteristic of low mass X-ray binaries (LMXBs) in the low-hard state, whether the accretor is a neutron star or black hole. Such behaviour is typically observed at luminosities $\\la$10\\% Eddington. However, XBo\\thinspace 45 exhibited this behaviour at an unabsorbed, 0.3--10 keV luminosity of { 2.5$\\pm$0.2$\\times 10^{38}$} erg s$^{-1}$, or{ $\\sim$140\\%} Eddington for a 1.4 $M_{\\odot}$ neutron star accreting hydrogen. Hence, we identify XBo\\thinspace 45 as a new candidate black hole LMXB. { XBo\\thinspace 45 appears to have been consistently bright for $\\sim$30 years, consistent with theoretical prediction for a globular cluster black hole binary formed via tidal capture}. Bo\\thinspace 375 was observed in the 2007, January 2 XMM-Newton observation, and has a two-component spectrum that is typical for a bright neutron star LMXB. Bo\\thinspace 135 was observed in the same field as Bo\\thinspace 45, and could contain either a black hole or neutron star. ", "introduction": "Of the 14 bright low mass X-ray binaries (LMXBs) associated with Galactic globular clusters (GCs), 13 are known to contain neutron stars, and the other one probably does as well \\citep[see e.g.][ and references within]{wa01,int04}. However, black hole binaries have been seen in extragalactic GCs, proving that they form in such environments. For example, \\citet{mac07} discovered the first GC black hole binary in the giant elliptical galaxy NGC 4472. { The lack of known GC black hole binaries has been a subject of great interest for theoretical modellers. \\citet{kal04} show that black hole binaries formed through exchange interactions should have duty cycles of $\\sim$0.001, consistent with the absence of GC black hole binaries at the time. They also predict that black hole binaries formed through tidal capture of a main sequence star would be bright, persistent X-ray sources, and infer from their absence that tidal capture probably disrupts the main sequence star. However, neither of these results preclude the discovery of new GC black hole binaries. } The bright X-ray sources of M31 have been studied for over 25 years, with Einstein \\citep{tf91}, ROSAT \\citep{S97,S01}, Chandra \\citep[see e.g.][]{dis02, kaa02,will04} and XMM-Newton \\citep[see e.g.][]{shi01,trud05,pfh05,lsg08}. \\citet{dis02} conducted a Chandra survey of selected regions of M31, and found that most of their bright X-ray sources were associated with GCs, with $\\sim$10\\% of GC sources exhibiting 0.5--7 keV luminosities $\\ga$10$^{38}$ erg s$^{-1}$. They also showed the M31 GC population to be significantly different from that of the Milky Way, as $\\sim$30\\% of the M31 GC X-ray sources exhibited luminosities $>$10$^{37}$ erg s$^{-1}$, whereas only one out of 11 Galactic GC X-ray sources exceeded 10$^{37}$ erg s$^{-1}$ \\citep{verb95}. These results contradict those of \\citet{S01}, who concluded that the GC X-ray populations of the Milky Way and M31 were similar. \\citet{dis02} also compared the optical properties of M31 GCs with and without X-ray sources, as well as M31 GCs and Milky Way GCs with X-ray sources. They compared optical colours and apparent magnitudes, radial velocities, metallicities and colour excesses, as well as core radii. They found the only possibly significant difference to be in their luminosities: the median luminosity of M31 GCs with X-ray sources was $\\sim$0.55 magnitudes higher than that of M31 GCs without. Similarly, M31 X-ray GCs were slightly brighter in V than Milky Way X-ray GCs. \\citet{dis02} { speculated} that this might associate X-ray sources with higher GC masses. We report on three bright X-ray sources in M31, all associated with populous, old globular clusters: Bo\\thinspace 45, Bo\\thinspace 135 and Bo\\thinspace 375. We will argue that Bo\\thinspace 45 contains a good black hole (BH) candidate, while Bo\\thinspace 375 is likely to contain a neutron star (NS). Bo\\thinspace 135 is more likely to contain a NS, but we cannot rule out a BH accretor. We first review the emission and variability of LMXBs in their various states in Section~\\ref{bhs}, and then discuss known properties of our targets and their host clusters in Section~\\ref{gc}. We describe the observation and analysis in Section~\\ref{obs}, and our results in Section~\\ref{res}. We finally present our discussion and conclusions in Section~\\ref{dis}. ", "conclusions": "\\label{dis} We have examined the emission spectra and time variability of three X-ray sources associated with GCs in M31, using the 2006 December 26 and 2007 January 2 XMM-Newton observations. The emission of XBo\\thinspace 45 is well described by a pure power law with photon index $\\sim$1.4, and is highly variable. This is consistent with a NS or BH LMXB in the low state \\citep{vdk94}, but is not consistent with a BH LMXB in the high state or steep power law state \\citep[and references within]{mr03}, or a NS LMXB emitting at $>$10$^{38}$ erg s$^{-1}$ \\citep{wsp88,cbc01,bcb03}. { \\citet{sg98} calculated a distance to M31 of 784 kpc, $\\pm$13 kpc of statistical error, $\\pm$17 kpc of systematic error. Combining these distance uncertainties adds further uncertainties in the source luminosities of $\\pm$5\\%. Therefore, the 0.3--10 keV luminosity range for XBo\\thinspace 45 is 2.5$\\pm$0.2$\\times$10$^{38}$ erg s$^{-1}$, or 140$\\pm$10\\% of the Eddington limit for a 1.4 M$_{\\odot}$ neutron star primary. However, several LMXBs have been found to contain neutron stars with mass as high as $\\sim$2.1 M$_{\\odot}$ \\citep[see e.g.][]{nice05,ozel06}; XBo\\thinspace 45 has a 0.3--10 keV luminosity of $\\sim$80\\% Eddington for such systems. Since \\citet{glad07} showed that transitions in neutron star LMXBs occur at $\\la$10\\% Eddington, we consider XBo\\thinspace 45 to exhibit low state behaviour at an apparent luminosity too high for a neutron star. We therefore identify the accretor in XBo\\thinspace 45 as a BH candidate. We note that XBo\\thinspace 45 has been observed several times by the Einstein and ROSAT observatories over the last $\\sim30$ years, varying in luminosity only by a factor of $\\sim$2. This behaviour is consistent with that predicted for a GC BH binary formed by tidal capture \\citep{kal04}. In contrast to XBo\\thinspace 45,} the observed two component emission of XBo\\thinspace 375 is consistent with a bright NS LMXB, but not a low state NS or BH LMXB, nor a BH in the high or steep power law states. Hence, we classify XBo\\thinspace 375 as a NS LMXB. Finally, the emission spectrum of XBo\\thinspace 135 is consistent with a pure power law with photon index $\\sim$1.6, but the fit is significantly improved by adding a blackbody component. Hence, XBo\\thinspace 135 is consistent with a NS or BH LMXB, and deeper observation is required for further classification. { Of the thirteen bright X-ray sources in Galactic GCs, twelve have exhibited X-ray bursts, confiming their natures as NS LMXBs \\citep[see ][ for a review]{vl06}. \\citet{ph05} conducted a survey of XMM-Newton observations of X-ray sources associated with M31 GCs, looking for X-ray bursts. They found simultaneous pn and MOS detections of bursts in two sources, and several burst candidates that were detected in the pn only. Such bursts would be identifiable in the lightcurves of our target sources. However, X-ray bursts are thought to be forbidden at luminosities $>$ 50\\% Eddington \\citep{lpt93}, hence we do not expect bursts from XBo\\thinspace 135 or XBo\\thinspace 375. } XBo\\thinspace 45 exhibits low-state behaviour at luminosities $\\sim$10 times higher than expected for 1.4 M$_{\\odot}$ NS LMXBs; XBo\\thinspace 135 may also. A composite of two or three bright X-ray sources would likely result in a different spectrum to that observed in XBo\\thinspace 45; hence, if XBo\\thinspace 45 were a composite, it would be more likely to be made from { $\\sim$10} NS LMXBs in the low state. \\citet{dis02} calculated the probability for a GC to contain multiple bright X-ray sources: they argue that if $p$ is the probability of finding one bright X-ray source in a GC, then Poisson statistics dictate that probability for two bright X-ray sources should be $p$/2, while the probability for 3 bright X-ray sources should be $p^{2}$/6. For a $p$ $\\sim$0.1--0.2, \\citet{dis02} predict 3--5 GCs with 2 X-ray sources and less than 1 with 3 X-ray sources in M31. It is therefore very unlikely that XBo\\thinspace 45 combines the emission of $\\sim$10 low-state neutron star LMXBs. It is however possible that the emission from XBo\\thinspace 45 is anisotropic. It exhibits low-state spectra at luminosities $\\sim$10--20 times higher than expected for neutron star LMXBs; hence it could simply be beamed by a factor of 10--20. In this case, the emission would be restricted to a small solid angle; one might expect to observe such beaming in $\\sim$5--10\\% of randomly oriented systems. \\citet{tp04} modeled the spectra of 43 M31 GCs. If we loosely class a low-state spectrum as a power law with $\\Gamma$ $<$1.7, then \\citet{tp04} found 20 GCs consistent with low-state spectra. Five of those, XBo\\thinspace 5, XBo\\thinspace 82, XBo\\thinspace 135, XBo\\thinspace 153 and XBo\\thinspace 386 exhibited apparent luminosities $>$10$^{38}$ erg s$^{-1}$. If we include XBo\\thinspace 45, then 6 out of 21 GC systems consistent with low-state spectra (including XBo\\thinspace 45), have measured luminosities $>$10$^{38}$, i.e. $\\sim$30\\%, a significantly larger fraction than expected from beaming. We note that the host cluster Bo\\thinspace 45, which contains a BH LMXB candidate, is significantly larger than the cluster Bo\\thinspace 375 { (see Section 3.2)}, which we think contains a NS LMXB. \\citet{dis02} describe Bo\\thinspace 375 as not at all unusual, with parameters close to the median of M31 GCs. This suggests that Bo\\thinspace 45 (and also Bo\\thinspace 135) are particularly massive, and may therefore be more prone to forming BH LMXBs. Therefore we conclude that, unlike the Milky Way, at least one GC in M31 is likely to contain a black hole binary." }, "0807/0807.4558_arXiv.txt": { "abstract": "We present abundances of C, N, O, F, Na, and Fe in six giant stars of the tidally disrupted globular cluster NGC 6712. The abundances were derived by comparing synthetic spectra with high resolution infrared spectra obtained with the Phoenix spectrograph on the Gemini South telescope. We find large star-to-star abundance variations of the elements C, N, O, F, and Na. NGC 6712 and M4 are the only globular clusters in which F has been measured in more than two stars, and both clusters reveal F abundance variations whose amplitude is comparable to, or exceeds, that of O, a pattern which may be produced in $M$ $\\gtrsim$ 5$M_\\odot$ AGB stars. Within the limited samples, the F abundance in globular clusters is lower than in field and bulge stars at the same metallicity. NGC 6712 and Pal 5 are tidally disrupted globular clusters whose red giant members exhibit O and Na abundance variations not seen in comparable metallicity field stars. Therefore, globular clusters like NGC 6712 and Pal 5 cannot contribute many field stars and/or field stars do not form in environments with chemical enrichment histories like that of NGC 6712 and Pal 5. Although our sample size is small, from the amplitude of the O and Na abundance variations, we infer a large initial cluster mass and tentatively confirm that NGC 6712 was once one of the most massive globular clusters in our Galaxy. ", "introduction": "\\label{sec:intro} The formation and evolution of our Galaxy remains one of the great unanswered questions in modern astronomy. \\citet{els62} suggested formation via the monolithic collapse of a gaseous protocloud on a timescale of 10$^8$ years. \\citet{sz78} challenged this notion by proposing that the halo formed through the accretion of independent fragments over a longer period, 10$^9$ years. These seminal works studied Galactic archaeology using the kinematics and metallicities of stars and globular clusters in the disk and halo. Today, Galaxy formation is discussed within the context of $\\Lambda$CDM cosmology and hierarchical structure formation \\citep{white78,freeman02} with the ongoing accretion of the Sagittarius dwarf galaxy being the most prominent example \\citep{ibata94}. Another mechanism for populating the disk and halo is through the destruction of globular clusters via tidal shocks, two body relaxation etc \\citep{gnedin97}. Although the current mass in globular clusters is small, the initial globular cluster population may have been considerably larger than the present population \\citep{gnedin97}. The globular cluster Palomar 5 exhibits large tidal tails that extend over 10 degrees and contain more mass than the remaining cluster \\citep{odenkirchen01,odenkirchen03}. Therefore, Pal 5 is in the process of being tidally disrupted and is currently contributing stars to the disk and halo. Chemical abundances place strong constraints upon the fraction of halo and disk stars that may come from disrupted globular clusters and/or the types of globular clusters that may populate the disk and halo. Specifically, every well studied Galactic globular cluster exhibits large star-to-star abundance variations for the light elements from C to Al \\citep{smith87,kraft94,gratton04}. Although the amplitude may vary from cluster to cluster, the abundances of C and O are low when N is high, O and Na are anticorrelated as are Mg and Al. Indeed, Str{\\\"o}mgren photometry reveals that every globular cluster has large star-to-star variations in the $c_1$ = $(u-v)-(v-b)$ index at all evolutionary stages \\citep{grundahl00}. The Str{\\\"o}mgren $u$ filter includes the 3360\\AA\\ NH molecular lines, and \\citet{6752nh} recently showed that the N abundances are directly correlated with the $c_1$ index. Therefore, it is likely that all globular clusters possess large N abundance variations at all evolutionary stages. Although hydrogen burning at high temperatures may explain the observed abundance patterns \\citep{langer93,langer95,denissenkov98,karakas03}, the source of the nucleosynthesis and the nature of the pollution mechanism remain unknown. Intermediate-mass ($\\sim$3 to 8$M_{\\odot}$) asymptotic giant branch (AGB) stars were the assumed polluters owing to the mono-metallic nature of most GCs, even though detailed AGB models have so far mostly failed to match the observations \\citep{fenner04,karakas06b}. Nevertheless, these abundance patterns seen in every cluster have rarely, if ever, been observed in field stars to date \\citep{pilachowski96,gratton00b}. \\citet{smith02} conducted a detailed abundance analysis of four bright giant stars in Pal 5 and found variations of O, Na, and Al. (No abundance measurements have been performed upon stars in the tidal tails of Pal 5.) While most stars lost from a tidally disrupted cluster would be main sequence stars, abundance variations of O, Na, and Al have now been identified on the main sequences of globular clusters \\citep{gratton01,cohen05}. Since no radial gradients are associated with the O to Al abundance variations (with the exception of 47 Tucanae [\\citealt{nf79,briley97}]), observations of red giants in the cluster should be equivalent to observing red giants in the tidal tails. That abundance variations of O, Na, and Al are found in Pal 5 suggests that clusters like Pal 5 cannot provide many field stars and/or field stars do not form in environments with chemical enrichment histories similar to Pal 5. Of great interest for our understanding of Galactic and globular cluster formation would be the identification of clusters undergoing tidal disruption in which no light element abundance variations are detected. Of the large sample of globular clusters studied by \\citet{paresce00} using the Hubble Space Telescope, all have mass functions (as inferred from their luminosity functions) which peak at 0.25M$_\\odot$. Not surprisingly, the mass function of Pal 5 is flatter than other clusters revealing significant depletions of low mass stars presumably stripped by the Galactic tidal field \\citep{koch04}. The globular cluster NGC 6712 is a small and sparse globular cluster whose mass function peaks at 0.75M$_\\odot$ instead of 0.25M$_\\odot$ \\citep{demarchi99,andreuzzi01}. That is, NGC 6712 is the only cluster whose mass function decreases with decreasing mass. With an orbit penetrating deep into the bulge, $R_{\\rm pericentric}$ = 0.9 kpc \\citep{dinescu99}, tidal forces have stripped away a substantial fraction of NGC 6712's lower mass stellar population. Calculations suggest that NGC 6712 may have lost up to 99\\% of its original mass \\citep{takahashi00}. All that remains of NGC 6712 is a remnant core of a cluster that was probably once one of the most massive in the Galaxy. The presence of a high luminosity x-ray source and a surprisingly large blue straggler population reinforce the idea that NGC 6712 was once much more massive and concentrated \\citep{paltrinieri01}. Therefore, NGC 6712 has almost certainly contributed stars to the disk and/or halo. Previous abundance analyses of NGC 6712 only considered one post-AGB star \\citep{jasniewicz04,mooney04} whose composition may not reflect the composition of the cluster due to the rich nucleosynthesis occurring in the late phases of stellar evolution. In this paper, we present the first detailed chemical abundance analysis of bright red giant stars in this tidally disrupted globular cluster. ", "conclusions": "\\label{sec:summary} Based on high resolution infrared spectra, we derive abundances of C, N, O, F, Na, and Fe in six giant stars of the tidally disrupted globular cluster NGC 6712. For the elements C, N, O, F, and Na, we find large star-to-star abundance variations and correlations between these elements, a characteristic that NGC 6712 shares with every well studied Galactic globular cluster. This is only the second cluster in which F abundances have been measured in useful numbers of stars and both clusters show F variations whose amplitude is comparable to, or exceeds, that of O. Within the limited data, globular clusters appear to have lower F abundances than field and bulge stars at the same metallicity. Of great interest would be measurements of F in additional stars in $\\omega$ Cen and other globular clusters as well as in larger samples of field stars, with both samples overlapping in metallicity. From the amplitude of the O and Na abundance variations, we tentatively confirm that NGC 6712 was once one of the most massive clusters in our Galaxy. NGC 6712 is a tidally disrupted cluster as revealed through its highly unusual luminosity function. Pal 5 is another tidally disrupted globular cluster. Both NGC 6712 and Pal 5 have almost certainly contributed stars to the disk and halo. Both clusters exhibit large star-to-star abundance variations for light elements, a characteristic which has yet to be identified in field halo stars. Therefore, the light element abundance variations detected in NGC 6712 indicate that clusters like NGC 6712 and Pal 5 have not provided many field stars and/or field stars did not form in environments with chemical enrichment histories like NGC 6712 and Pal 5. As pointed out by \\citet{smith02}, disrupted globular clusters like Pal 5 have lost CN-strong, O-poor, Na-rich, Al-rich stars to the halo field. But where are these stars? Of great interest would be an abundance analysis of stars within the tidal tails of Pal 5 as well as a large-scale dedicated search for O, Na, and Al abundance anomalies in field halo stars." }, "0807/0807.3652_arXiv.txt": { "abstract": "{We continue investigation of the hidden plane-mirror symmetry in the distribution of excursion sets in cosmic microwave background (CMB) temperature anisotropy maps, previously noticed in the three-year data of the Wilkinson microwave anisotropy probe (WMAP), using the WMAP 5 years maps. The symmetry is shown to have higher significance, $\\chi^2 < 1.7$, for low multipoles $\\ell < 5$, while disappearing at larger multipoles, $\\chi^2 > 3.5 $ for $\\ell > 10$. The study of the sum and difference maps of temperature inhomogeneity regions, along with simulated maps, confirms its existence.The properties of these mirroring symmetries are compatible with those produced by the Sachs-Wolfe effect in the presence of an anomalously large component of horizon-size density perturbations, independent of one of the spatial coordinates, and/or a slab-like spatial topology of the Universe.} ", "introduction": "The properties of the CMB temperature anisotropy, as well as its polarization, are among the basic sources of information on cosmological parameters \\cite{DB1,Sp,Kom}. Their tiny features, such as the local spikes in the multipoles power spectrum, deviation from the statistical isotropy, and non-Gaussianity signatures, may all be the result of various fundamental processes having occurred in the early Universe. Among the reported anomalies are the alignment of the principal (Maxwellian) vectors of low multipoles, the north-south power asymmetry, the southern anomalous cold spot, etc (see de Oliveira-Costa et al. 2004; Copi et al. 2004,2007; Schwarz et al. 2004; Eriksen et al. 2004,2007; Cruz et al. 2005; Morales \\& Saez 2008). In the present paper we continue the study of another deviation from statistical isotropy: the hidden partial plane-mirror symmetry in the distribution of CMB temperature fluctuations excursion sets, i.e. of one-connected pixel sets equal to and higher than the given temperature threshold (lower for negative thresholds) previously found in the WMAP 3-years temperature maps \\cite{mirr}. We use the WMAP 5-years data \\cite{WMAP5}, not only to confirm the mirroring effect found in WMAP3 maps while studying the inhomogeneities in the distribution of the excursion sets, but also to reveal its other properties. By inquiring into the dependence of the mirror symmetry on the angular scale, we show that the effect has the highest significance at low multipoles $\\ell <5$ and that it quickly disappears at higher multipoles. Similarly, the study of the sum and difference maps of temperature inhomogeneity regions provides additional insight into the mirroring. Namely, when the sum and difference maps are created via reflection of one of the maps, as it should be for mirrored images, anisotropic properties of excursion sets do survive, while they disappear if the sum map is created without reflection. Difference maps from independent radiometers (A-B) have been used to test the role of scan inhomogeneities and noise \\cite{mirr}. The signal-to-noise ratio for the studied excursion sets is about 4:1 and the excursion sets in (A-B) map do not show any specific property observed in the sum (A+B) map. The negligible role of the noise was also checked using the foreground reduced maps available in http://lambda.gsfc.nasa.gov/product/map/current/. Although contamination of a Galactic or interplanetary origin at these multipoles certainly cannot be excluded, following Gurzadyan et al (2007b), in the last section we discuss which properties of the Universe might be responsible for this effect, if it had a cosmological origin. ", "conclusions": "We have analyzed the scale dependence of partial mirror symmetry in the distribution of excursion sets, as previously found in the WMAP3 maps, with a nearly antipodal location of symmetry centers. Studies used the WMAP5 W-band maps, and the results obtained using WMAP5 and WMAP3 data agree. The centers lie close to one of the $\\ell=3$ multipole Maxwellian vectors, but not close to the sum of multipoles vectors up to $\\ell=8$ \\cite% {mirr}. Also they are close to the ecliptic pole and are nearly orthogonal to the CMB dipole apexes. They are moving towards the Galactic equator with the increase in the temperature threshold interval. This symmetry appears to be a large-angle effect, i.e. it is stronger at low multipoles and it weakens rapidly for larger $\\ell$: its statistical significance is quantified by $\\chi^2 < 1.7$ at $\\ell < 5$ and $\\chi^2 > 3.5$ at $\\ell > 10$. The symmetry was also tested using the following procedure: the sum map of the symmetry regions was obtained first, via rotation over $\\pi$, as is usually the case for mirrored images, and then without such a rotation. The clear mirroring in the first case and its complete absence in the second case make the case for a partial mirror symmetry stronger. Turning to the origin of this symmetry, an unknown interplay of interplanetary and Galactic foregrounds or another unspecified non-cosmological contribution to the low multipoles certainly remains a possibility. If, however, it has a cosmological origin, then a signature of the simplest non trivial, $T^1$, spatial topology of the Universe is among the options, as discussed in Gurzadyan et al (2007b). For this topology, the points with coordinates $z$ and $z+L$ are identified where $z$ is one of the spatial coordinates. Such a model may be also considered as a limiting case of the Universe with compact flat spatial sections having the $T^3$ topology if the identification scales $L_1,L_2$ along two other spatial coordinates are much more than $L$ -- the slab topology (for early papers on a non trivial spatial topology of the Universe, see Zeldovich (1973), Sokolov \\& Schwartsman (1975), Sokolov \\& Starobinsky (1975), Fang \\& Sato (1983)). Note that, one should not expect any mirror symmetry, even a partial one, for comparable topological scales $L\\sim L_2,L_3$. As shown in Starobinsky (1993), for this $T^1$ topology, a large-angle pattern of a CMB temperature anisotropy has just the form (\\ref{split}). The first term on its right hand side has the exact mirror symmetry with respect to the $(x,y)$-plane. It originates from the Sachs-Wolfe effect at the last scattering surface from density perturbations that do not depend on $z$. The second term represents a remaining part of anisotropy and does not have any symmetry at all. However, for $a_0L$ on the order of $R_{\\mathrm{hor}}$ or slightly more, where $a_0=a(t_0)$ is the present scale factor of a Friedmann-Robertson-Walker cosmological model, the latter term should somehow be suppressed since the Sachs-Wolfe contribution to it from the last scattering surface comes from perturbations having wave vectors with $|% \\mathbf{k}|\\ge 2\\pi/L$. That is why one expects the total large-angle pattern of $\\Delta T/T$ to have an \\emph{approximate} mirror symmetry in this case. \\footnote{% Note the other effect worsening the symmetry at large angles (low multipoles): a contribution from the integrated Sachs-Wolfe effect at small redshifts due to a cosmological constant (first calculated in \\cite{KS85}) or dynamical dark energy.} More generically and not connected with a non trivial spatial topology of the Universe, such an approximate mirror symmetry at large angles arises when the large-scale $z$-independent part of density perturbations inside the last scattering surface is anomalously large. In all these cases, the rms amplitude of the first term on the right hand side of Eq. (\\ref{split}) quickly becomes negligible compared to the second one with the growth of $% \\ell$. Weakening of the mirroring symmetry for higher values of $\\ell$ found in Sect. 3 is in a good agreement with this theoretical prediction and may be considered as an additional argument for the reality of the mirroring effect. Until now, searches for the mirroring effect of the form (\\ref{split}) or directly for the $T^1$ topology gave negative results (see e.g. de Oliveira-Costa et al. 1996; de Oliveira-Costa et al. 2004; Cornish et al. 2004), rasing the lower limit on the physical topological scale $a_0L$ to $\\sim R_{\\mathrm{hor}}=14$ Gpc. The numerical value is given for the standard $\\Lambda$CDM cosmological model with $\\Omega_m=0.3,~% \\Omega_{\\Lambda}=0.7$. However, for higher values of $a_0L$ this topology is not excluded. This follows already from the fact that the much more restrictive cubic $T^3$ topology ($L=L_1=L_2$) with $a_0L > R_{hor}$ is still considered as a viable possibility; see the recent papers \\cite{AJLS07,A08}, where an inconclusive evidence for the latter case with $a_0L \\approx 1.15R_{hor}$ is presented. This shows that topological explanation of the partial mirror symmetry investigated in this paper is possible. From our analysis, it is still too early to speak about the value of $L$, since the secure separation of CMB temperature fluctuations into a mirrored and non-mirrored parts needs higher resolution maps. Also, a non-topological (though still cosmological) explanation of such an effect is possible, as pointed out above. In this respect, see the recent papers \\cite{GK2008} where it was shown that voids can act as hyperbolic lenses in a spatially flat Universe, producing specific signatures in CMB temperature fluctuations. Future observational data will help solve these problems. Shortly before submission of this manuscript, a paper appeared \\cite{GE} where CMB statistical anisotropy of an axial type was studied with the preferred axis very close to the one defined by our $CE_N-CE_S$ direction." }, "0807/0807.0717_arXiv.txt": { "abstract": "A line list of vibration-rotation transitions for $^{13}$C substituted HCN is presented. The line list is constructed using known experimental levels where available, calculated levels and {\\it ab initio} line intensities originally calculated for the major isotopologue. Synthetic spectra are generated and compared with observations for cool carbon star WZ Cas. It is suggested that high resolution HCN spectra recorded near 14 $\\mu$m should be particularly sensitive to the $^{13}$C to $^{12}$C ratio. ", "introduction": "\\label{sec:Intro} Carbon giant stars are though to arise from the third dredge up of asymptotic giant branch (AGB) stars, which pollutes the envelope and atmosphere with nuclear processed material from the interior. This increases the abundance of carbon and the $^{13}$C/$^{12}$C ratio to well above the terrestrial level (0.011). In fact the $^{13}$C/$^{12}$C ratio in carbon giants has been measured to be as high as one-third \\citep{ab97}. There are a number line lists or opacity functions available for \\hcn\\ \\citep{Jorgensen1,Jorgensen2,ao98,ha02b,ha06}. However, when considering the contribution to opacity by molecular species such as HCN it may not be sufficient to account only for \\hcn, but also for the isotopologue \\htcn. The calculation of a complete triatomic linelist is computationally expensive, requiring several tens of thousands of CPU hours \\citep{te07}. However, within the Born-Oppenheimer approximation the electronic structure of isotopologues is identical. Thus both the potential energy and electric dipole moment functions are identical for all isotopologues. This implies that the vibration-rotation frequencies and transition intensities are likely to be very similar. For a hetronuclear molecule, such as HCN, the main differences between the spectra of isotopologues are caused by the change in reduced mass. In the harmonic approximation the vibrational contribution to the line frequency is proportional to $\\rho^{-\\frac{1}{2}}$, where $\\rho=m_1m_2/(m_1+m_2)$ is the reduced mass. So the frequency of the CN stretch vibrational mode of \\hcn\\ and \\htcn\\ will differ by the order of 1\\% and the bend and H-C stretch mode by less. In terms of both the observation of \\htcn\\ and the line blanketing of a model atmosphere this shift in line frequency is more important than the small differences in line intensity between comparable lines of \\hcn\\ and \\htcn. It is the aim of this work to compute a set of energy levels and line frequencies for \\htcn\\ and \\htnc. These energy levels will be used in conjunction with the Einstein A coefficients for \\hcn\\ and \\hnc\\ computed by \\citet{ha02b}, to generate a \\htcn/\\htnc\\ linelist. The line frequencies in this \\htcn/\\htnc\\ linelist have been corrected using available laboratory data in order to increase its accuracy. The new linelist has been used to compute synthetic spectra for C-stars, with different $^{13}$C/$^{12}$C ratios using star WZ Cas as a prototype. ", "conclusions": "\\label{sec:conc} We present a new set of {\\em ab initio} rotation-vibration energy levels for \\htcn\\ and \\htnc\\ which were calculated for angular momenta of $J$ = 0, 1, 2, 3, 5, 10, 20, 30, 40, 60 and for both even ($e$) and odd ($f$) parity. The states were computed up to an energy of at least $10000 + B(J(J+1))$~\\cm\\ above the \\htcn\\ ground state, where $B\\sim1.5$~\\cm\\ is the HCN rotational constant. The quantum number assignments were made using a method similar to that described in \\citet{ha06}. The new linelist has been incorporated into our computations of C-rich synthetic spectra. The detailed analysis of the infrared spectra of C-giant star with high $^{13}$C/$^{12}$C ratios ought to take account of \\htcn\\ and \\htnc\\ species. Moreover in many cases HCN spectra probably provides the best chance of determining the $^{13}$C/$^{12}$C ratios in atmospheres of the coolest stars as the CO bands at 2.3 $\\mu$m are usually saturated and other molecular bands are severely blended. Our upgraded opacity sources can be used for the determination of carbon isotopic ratios in atmospheres of carbon stars. The most promising regions are those around 3.6 and 14 $\\mu$m where the the $v_2+v_3$ (bend and CN stretch) combination bands and $v_2$ (bend) fundamental and hot bands of \\hcn\\ and \\htcn\\ molecules are located, respectively. However, to do this requires to use the spectral data of the proper quality. Finally, the use of \\hcn\\ and \\htcn\\ lines for numerical analysis of infrared spectra of evolved stars is restricted by the incompleteness of presently available opacity sets, see \\citet{te07}. Special attention needs be paid to the computation of other lines for polyatomic molecules such as C$_3$, NH$_3$, CH$_4$ and C$_2$H$_2$, and their isotopologues." }, "0807/0807.2462_arXiv.txt": { "abstract": "We carry out a suite of cosmological simulations of modified action $f(R)$ models where cosmic acceleration arises from an alteration of gravity instead of dark energy. These models introduce an extra scalar degree of freedom which enhances the force of gravity below the inverse mass or Compton scale of the scalar. The simulations exhibit the so-called chameleon mechanism, necessary for satisfying local constraints on gravity, where this scale depends on environment, in particular the depth of the local gravitational potential. We find that the chameleon mechanism can substantially suppress the enhancement of power spectrum in the non-linear regime if the background field value is comparable to or smaller than the depth of the gravitational potentials of typical structures. Nonetheless power spectrum enhancements at intermediate scales remain at a measurable level for models even when the expansion history is indistinguishable from a cosmological constant, cold dark matter model. Simple scaling relations that take the linear power spectrum into a non-linear spectrum fail to capture the modifications of $f(R)$ due to the change in collapsed structures, the chameleon mechanism, and the time evolution of the modifications. ", "introduction": "\\label{sec:intro} Cosmic acceleration can arise from either an exotic form of energy with negative pressure or a modification to gravity in the infrared. Self-consistent models for the latter are highly constrained by the stringent tests of gravity in the solar system. Additional propagating degrees of freedom must be suppressed by non-linearities in their equations of motion \\cite{Vainshtein:1972sx,Deffayet:2001uk,Dvali:2006su} in a stable manner \\cite{Dolgov:2003px,Sawicki:2007tf,Seifert:2007fr}. These suppression mechanisms manifest themselves with the formation of non-linear structure in the Universe \\cite{Lue:2004rj,Koyama:2007ih,hu07b}. Understanding the physical content, phenomenology and even the basic viability of such models thus requires cosmological simulations. One possibility that has received much recent attention is the so-called $f(R)$ class of models (see \\cite{Sotiriou:2008rp} and references therein). These models generate acceleration through a replacement of the Einstein-Hilbert action by a function of the Ricci or curvature scalar $R$ \\cite{Caretal03,NojOdi03,Capozziello:2003tk}. They also introduce an extra propagating scalar degree of freedom that acts as an effective fifth force on all forms of matter \\cite{Chi03,Chiba:2006jp}. The range of the force depends non-linearly on the local curvature and can be made to become infinitesimal at high curvature. With an appropriate choice of the function $f(R)$, deep potential regions can trap the field at high curvature leading to a nonlinear ``chameleon mechanism'' \\cite{khoury04a} that suppresses local deviations from ordinary gravity \\cite{Navarro:2006mw,Faulkner:2006ub,hu07a}. Whether or not solar system tests of gravity are satisfied in an $f(R)$ model then depends on the depth of the gravitational potential including the astrophysical and cosmological structure surrounding it \\cite{hu07a}. Likewise observable deviations from ordinary gravity for upcoming dark energy probes such as weak lensing, galaxy clustering and clusters of galaxies depend on the whole history of non-linear structure formation. Previous cosmological simulations ({\\it e.g.}~\\cite{WhiKoc01,Maccio:2003yk,Nusser:2004qu,StaJai06,Laszlo:2007td}) have focused on modifications of the force law with a fixed and density independent range. Such modifications alone are incapable of satisfying local tests of gravity. In a companion paper \\cite{oyaizu08b}, the numerical methodology for solving the non-linear field equation of $f(R)$ gravity was established. In this second paper of the series, we apply this methodology and carry out a suite of cosmological simulations of $f(R)$ models that are chosen to expose the impact of the chameleon mechanism on the power spectrum of the matter and the lensing potential. We begin in \\S \\ref{sec:fr} with a review of non-linear gravitational dynamics in $f(R)$ models and proceed to the simulation results in \\S \\ref{sec:results}. We discuss these results in \\S \\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} We have carried out the first cosmological simulations of $f(R)$ models for cosmic acceleration that exhibit the chameleon mechanism. The chameleon mechanism involves a non-linear field equation for the scalar degree of freedom that suppresses the range of the gravitational force modification or Compton scale in the deep gravitational potential wells of cosmological and astrophysical structure. We have here focused on its impact on the matter power spectrum or equivalently the potential power spectrum relevant for weak lensing surveys. While we have simulated only one particular functional form of $f(R)$, we expect that the qualitative behavior of other models that exhibit a chameleon behavior to yield similar results once scaled to the appropriate Compton scales and field amplitudes. In the absence of the chameleon mechanism, gravitational interactions would have an enhancement of a factor of 4/3 on all scales smaller than the Compton scale in the cosmological background eventually leading to order unity enhancements in the power at high wavenumber. The chameleon mechanism turns on when the depth of the local gravitational potential becomes comparable to the field amplitude in the background. We have shown through otherwise identical simulations of structure with the chameleon mechanism artificially turned off that once the chameleon appears, it causes a substantial reduction of the enhanced power. For example, for a field amplitude of $|f_{R0}|=10^{-6}$ the change in the enhancement of power at $k \\sim 1h^{-1}$Mpc is a factor of four. Even in cases where current cosmological structures do not possess a chameleon $(|f_{R0}| \\simgt 10^{-5})$, there still is an impact on the power spectrum due to evolutionary effects. In $f(R)$ models where the field amplitude decreases with curvature, the chameleon can appear again at high redshift when the building blocks of current structure were assembled. Scaling relations which take the linear power spectrum and map it into the non-linear regime qualitatively misestimate the non-linear power spectrum in several ways. For example, the Smith {\\it et al.}~\\cite{smith03a} prescription assumes that the non-linear power spectrum depends only on the shape of the linear power spectrum near the non-linear scale. This prescription fails to describe both the chameleon mechanism and the change in the structure and abundance of collapsed objects leading to a severe misestimate at high $k$. A halo model can in principle do better to model these effects but simple prescriptions that scale the mass function to the linear variance and leave halo profiles unchanged also do not describe the non-linear effects to sufficient accuracy ({\\it cf.}~\\cite{hu07b}). In the next paper of this series, we intend to study the impact of $f(R)$ modifications on halo properties. \\smallskip \\noindent {\\it Acknowledgments}: We thank N. Dalal, B. Jain, J. Khoury, K. Koyama, A. Kravtsov, A. Upadhye, I. Sawicki, F. Schmidt, J. Tinker, and F. Stabenau for useful conversations. This work was supported in part by the Kavli Institute for Cosmological Physics (KICP) at the University of Chicago through grants NSF PHY-0114422 and NSF PHY-0551142 and an endowment from the Kavli Foundation and its founder Fred Kavli. HO was additionally supported by the NSF grants AST-0239759, AST-0507666, and AST-0708154 at the University of Chicago. WH and ML were additionally supported by U.S.~Dept.\\ of Energy contract DE-FG02-90ER-40560 and WH by the David and Lucile Packard Foundation. Some of the computations used in this work have been performed on the Joint Fermilab - KICP Supercomputing Cluster, supported by grants from Fermilab, Kavli Insititute for Cosmological Physics, and the University of Chicago." }, "0807/0807.2181_arXiv.txt": { "abstract": "Single-field models of inflation are analysed in light of the WMAP five-year data. Assuming instantaneous reheating, we find that modular/new inflation models with small powers in the effective inflaton self-interaction are more strongly constrained than previously. The model with a cubic power lies outside the $2\\sigma$ regime when the number of e-folds is $\\caln \\le 60$. We also find that the predictions for the intermediate model of inflation do not overlap the $1\\sigma$ region regardless of the power of the monomial potential. We analyse a number of ultra-violet, DBI braneworld scenarios involving both wrapped and multiple-brane configurations, where the inflaton kinetic energy is close to the maximum allowed by the warped geometry. In all cases, we find that the parameters of the warped throat are strongly constrained by observations. ", "introduction": "The inflationary scenario postulates that the universe underwent a phase of very rapid, accelerated expansion in its distant past. Observations -- most notably from the anisotropy power spectrum of the Cosmic Microwave Background (CMB) -- have provided strong support for the paradigm. Despite this success, however, the mechanism which drove the inflationary expansion has yet to be identified. Indeed, there remain many viable versions of the scenario. (For reviews, see, e.g., Refs. \\cite{Lyth:1998xn,Liddle:2000cg,Guth:2005zr,Cline:2006hu,HenryTye:2006uv,McAllister:2007bg,Linde:2007fr,Lyth:2007qh}.) In view of this, it is important to constrain, and ultimately rule out, as many models of inflation as possible. In this paper, we consider a wide range of inflationary models that are well motivated from particle physics and unified field theory. We focus on models where the cosmic expansion was driven by a single, self-interacting scalar `inflaton' field, $\\phi$. In general, single-field inflationary models are characterised by the action \\begin{equation} \\label{generalaction} S= \\int d^4x \\sqrt{-g} \\left[ \\frac{\\mpl^2}{2} R + P(\\phi , X) \\right], \\end{equation} where $R$ is the Ricci curvature scalar, the `kinetic function' $P(\\phi , X)$ is a function of the inflaton field and its kinetic energy $X \\equiv - \\frac{1}{2} g^{\\mu\\nu}\\nabla_{\\mu} \\phi \\nabla_{\\nu} \\phi$, and $\\mpl = (8\\pi G)^{-1/2}$ is the reduced Planck mass. In the simplest versions of the scenario, the inflaton has a canonically normalised kinetic energy, where $P=X- V(\\phi )$ and $V(\\phi)$ denotes the inflaton potential. In many higher-dimensional models motivated directly from string/M-theory, however, the function $P (\\phi ,X)$ has a more complicated form. This is the case, for example, in the Dirac-Born-Infeld (DBI) inflationary scenario, where inflation is driven by the propagation of one or more ${\\rm D}$-branes \\cite{DBI}. We compare both canonical and non-canonical models of inflation with the recent observational bounds derived from the combined data of the Wilkinson Microwave Anisotropy Probe (WMAP) \\cite{wmap5}, Baryon Acoustic Oscillations (BAO) \\cite{Percival:2007yw} and Supernovae (SN) surveys \\cite{Riess:2004nr,Astier:2005qq, Riess:2006fw,WoodVasey:2007jb}. In Ref. \\cite{Alabidi:2006qa}, the three-year WMAP data was employed to rule out for the first time a large fraction of viable canonical models at more than $3\\sigma$. We begin by reconsidering this analysis in the light of the WMAP five-year data \\cite{wmap5}. We find that the conclusions of \\cite{Alabidi:2006qa} are confirmed by WMAP5, with the improved bounds on the scalar spectral index strongly constraining models of modular inflation at more than the 2$\\sigma$ level.% We then proceed to consider non-canonical DBI brane inflationary models driven by a wrapped ${\\rm D}5$- or ${\\rm D}7$-brane \\cite{DBI,BECKER,KOB}. We focus on the `relativistic, ultra-violet' version of the scenario, where the brane is moving towards the tipped region of a warped throat with a kinetic energy close to the upper bound imposed by the warpfactor of the higher-dimensional space. We find that independently of the form of the inflaton potential, the geometry of the warped throat must be strongly constrained if such models are to satisfy the improved observational bounds from WMAP5. We also consider extensions of the scenario to multiple-brane configurations \\cite{thomasward,HLTW} and find that the same challenges arise as for the single-brane case. The structure of the paper is as follows. In section II, we summarise the current bounds on the observational parameters. In section III, we investigate models of canonical inflation and determine which models are still viable. In Sections IV and V, we derive new observational limits on wrapped and multi-brane UV DBI inflation. We conclude with a discussion in Section VI. ", "conclusions": "In this paper, we have considered the status of a wide class of single-field inflationary models in the light of the recent WMAP5 data. For the reasonable range of e-fold values aforementioned, $\\caln = 54 \\pm 7$, we conclude that modular/new inflation models with $20$ models poses a challenge to model builders, as discussed in Section III. Moreover, it is significant that independent data sets are now converging towards a preferred value of $n_s$, with only tiny differences between WMAP5 only, ${\\rm WMAP5}+{\\rm SDSS}$ \\cite{Percival:2006gt} and ${\\rm WMAP5}+{\\rm BAO} +{\\rm SN}$ data. This convergence implies that we can be more confident as to what represents a viable model from a phenomenological point of view. We have also shown that the intermediate inflationary scenario can only satisfy the data at $2\\sigma$ for $|\\alpha|\\geq 1.72$, and that the predictions of the model do not overlap the $1\\sigma$ region for any value of $\\alpha$. We emphasize that the above conclusions follow from assuming that the reheating of the universe immediately after inflation was instantaneous. However, the precise duration of the reheating process is not known. In particular, for a quadratic potential, a matter dominated phase may arise if the inflaton oscillates about its potential minimum for an extended interval before decaying. This would alter the number of e-folds that elapsed from the epoch when observable scales first crossed the Hubble radius during inflation to the time when inflation ended. In general, the number of e-folds depends on the effective equation of state after inflation and this would need to be known if a more precise value of $\\cal{N}$ is to be determined. Complementary to our work is that of Ref.~\\cite{Peiris:2008be}, where the authors start from the different stand point of aiming to construct the most general inflationary potential using slow-roll reconstruction with a specific e-fold and reheat temperature prior. Reconstructing the inflationary slow-roll parameters and not the spectral parameters from the data has the advantage of directly probing more fundamental parameters. They locate regions in $\\epsilon,\\eta$ space which are consistent with the WMAP5 data, and this allows to reconstruct the Hubble parameters and therefore (via the Hamilton-Jacobi equations) the inflationary potential. The authors also analyse higher-order parameters, but for the purposes of this work, we only need mention that the simplest fit is consistent with the crucial prior of $\\caln>15$, which we interpret as support for the inflationary paradigm. We have also considered the relativistic, ultra-violet, DBI braneworld scenario, where the inflaton is identified in terms of the radial position of a wrapped ${\\rm D}5$- or ${\\rm D}7$-brane. We found that when the brane's kinetic energy is maximised, new constraints on the parameters of the bulk geometry can be derived from the WMAP5 data in the $(r , n_s )$ plane. Such constraints are independent of the precise form of the inflaton potential and are stronger than existing limits originating from field-theoretic bounds on the tensor-scalar ratio. In all cases, consistency with the data requires a significant reduction in the volume of the base or the discovery of new classes of Calabi-Yau four-folds which allow for larger Euler numbers. We then extended our analysis to a recently proposed relativistic, multi-brane DBI scenario. For the case where the inflaton has a quadratic potential, we found that the predicted value of the spectral index is compatible with observations. However, the ratio $\\Vol/N$ is still strongly bounded from above, as in the single brane models, and consequently the same theoretical problems arise in this case also." }, "0807/0807.4522_arXiv.txt": { "abstract": "Infrared absorption lines of $\\mathrm{H}_{3}^{+}$, including the metastable $R$(3,3)$^l$ line, have been observed toward eight bright infrared sources associated with hot and massive stars located in and between the Galactic Center Cluster and the Quintuplet Cluster 30~pc to the east. The absorption lines with high velocity dispersion arise in the Galaxy's Central Molecular Zone (CMZ) as well as in foreground spiral arms. The temperature and density of the gas in the CMZ, as determined from the relative strengths of the $\\mathrm{H}_{3}^{+}$ lines, are $T$=200~--~300~K and $n\\leq$50~--~200~cm$^{-3}$. The detection of high column densities of $\\mathrm{H}_{3}^{+}$ toward all eight stars implies that this warm and diffuse gaseous environment is widespread in the CMZ. The products of the ionization rate and path length for these sight lines are 1000 and 10 times higher than in dense and diffuse clouds in the Galactic disk, respectively, indicating that the ionization rate, $\\zeta$, is not less than 10$^{-15}$~s$^{-1}$ and that $L$ is at least on the order of 50~pc. The warm and diffuse gas is an important component of the CMZ, in addition to the three previously known gaseous environments: (1) cold molecular clouds observed by radio emission of CO and other molecules, (2) hot ($T~=$~10$^4$~--~10$^6$~K) and highly ionized diffuse gas ($n_e~=$~10~--~100~cm$^{-3}$) seen in radio recombination lines, far infrared atomic lines, and radio-wave scattering, and (3) ultra-hot ($T~=$~10$^7$~--~10$^8$~K) X-ray emitting plasma. Its prevalence significantly changes the understanding of the environment of the CMZ. The sight line toward GC~IRS~3 is unique in showing an additional $\\mathrm{H}_{3}^{+}$ absorption component, which is interpreted as due to either a cloud associated with circumnuclear disk or the ``50~km~s$^{-1}$ cloud'' known from radio observations. An infrared pumping scheme is examined as a mechanism to populate the (3,3) metastable level in this cloud. ", "introduction": "\\subsection{The Central Molecular Zone} The region of the Galaxy within 200~pc of the center, known as the Central Molecular Zone (CMZ), contains 10\\% of the Galactic interstellar molecular mass \\citep[and references therein]{mor96,gen94,mez96}. The interstellar medium (ISM) in the CMZ is exceptional in many other ways. It is highly turbulent, pervaded by an intense magnetic field \\citep[for a review; ][]{mor06}, contains a multitude of massive stars \\citep[e.g.,][]{mon92,cot98,mun06} and super-massive clusters \\citep[e.g.,][for a short review]{nag90,nag95,fig04}, and also is home to numerous highly energetic objects ranging from supernova remnants to cataclysmic variables \\citep[CVs, e.g.,][]{mun03}. The particle densities in its dense molecular clouds are of order 10$^4$~cm$^{-3}$ and higher. The volume filling factor, \\emph{f$_{\\rm v}$}, of dense molecular clouds has previously been estimated to be $\\sim$ 10\\% \\citep{mor96}, but is more likely $\\sim$ 1\\% \\citep{oka05} in view of the observed visual extinction of $\\mathit{A_V}$~=~25~--~40~mag \\citep{cot00}. If dense molecular clouds only occupy 1\\% of the CMZ, what interstellar environments make up the remaining 99\\%? There have been some reports of lower density ``diffuse'' molecular gas in the intercloud region. Detailed analysis of the $\\mathit{J}$ = 1 $\\rightarrow$ 0 emission lines of CO and C$^{18}$O by \\citet{dah98}, and the $\\mathit{J}$~=~2~$\\rightarrow$~1 and $\\mathit{J}$ = 1 $\\rightarrow$ 0 emission lines of CO by \\citet{oka98a} have demonstrated the presence of gas with densities of $\\sim$ 10$^{2.5}$ cm$^{-3}$. Indeed \\citet{oka98a} conclude that CO emission from the Galactic center (GC) is dominated by emission from gas of that density. The volume filling factor of \\emph{f$_{\\rm v}$} $\\sim$ 0.1 claimed by \\citet{mor96} may include such gas. In this paper we present evidence for an even lower density molecular environment in which H$_2$ must still be plentiful but where CO is not abundant. In hindsight, the radio absorption lines with high velocity dispersion of OH \\citep[e. g.][]{bol64}, H$_2$CO \\citep[e. g.][]{pal69}, HCO$^+$ \\citep{lin81}, and CO \\citep{oka98b} may originate in this newly reported environment. A second candidate to fill the intercloud medium is the ultra-hot plasma ($T$ $\\sim 10^7$ -- 10$^8$~K) observed by its diffuse X-ray emission at energies of 0.5 -- 10~keV. The radiation from the plasma is observed in all sightlines to the CMZ \\citep[e.g.,][]{koy89,yam93,koy96}. \\citet{laz98} suggested that the highly ionized regions with \\emph{n$_e$} $\\sim 10$ cm$^{-3}$ and $T$ $\\sim 10^6$~K that are required to explain the $\\lambda^2$ dependent radio-wave scattering toward the GC are located at the photon-dominated interfaces between the molecular clouds and this ultra-hot X-ray emitting plasma. They proposed that high density molecular clouds with a volume filling factor of perhaps \\emph{f$_{\\rm v}$} $\\sim$ 0.1 are surrounded by the scattering electron gas and the rest of the space is filled with this plasma (see their Fig. 9). For an environment such as that described above, which is so dominated by ionized gas, ``Central Plasma Zone\" rather than CMZ would be a more appropriate term for the region. For a variety of reasons the ultra-hot and hot plasma cannot occupy the same regions where molecules abound. How those different categories of gas are distributed and how they coexist in the CMZ is yet to be understood. Observations of a new astrophysical probe of the CMZ, the infrared absorption spectrum of H$_3^+$ \\citep{geb96}, are beginning to shed fresh light on these questions \\citep{bol06}. Earlier measurements of this molecular ion in the Galactic center by \\citet{geb99}, \\citet{got02}, and \\citet{oka05} were made along only a very few sight lines. In this paper we report and analyze observations of seven transitions of H$_3^+$ along eight sight lines, which provide a more comprehensive diagnostic of the extent and nature of the H$_3^+$ - containing gas in the CMZ. \\begin{table*}[bht] \\begin{center} \\tablewidth{\\textwidth} \\scriptsize \\caption{Target list.}\\label{tb1} \\begin{tabular}{lcccccll} \\hline \\hline Source & R.A. & Dec. & $l$ & $b$ &$L$/$L'$& Other Names & Reference \\\\ & (J2000)& (J2000) & [\\arcdeg]& [\\arcdeg]& [mag] & in SIMBAD & 1, 2, 3. \\\\ \\hline GC~IRS~21 \\dots& 17:45:40.2 & $-$29:00:31 &$-$0.0561 &$-$0.0468 & &NAME SGR~A~IRS~21, GCIRS~21 & 1, 2, 3. \\\\ GC~IRS~3 \\dots & 17:45:39.8 & $-$29:00:24 & $-$0.0552&$-$0.0448 & 5.3 &BHA~11, GCIRS~3 & 1, 2, 3. \\\\ GC~IRS~1W \\dots& 17:45:40.4 & $-$29:00:27 &$-$0.0549 &$-$0.0471 & 5.5 &NAME SGR~A~IRS~1W, GCIRS~1W & 1, 2, 3. \\\\ NHS~21 \\dots& 17:42:53.8 & $-$28:51:41 & +0.0992 &$-$0.0553 & 4.62 &qF~577, [NHS93]~21 & 4, 5, 6. \\\\ NHS~22 \\dots& 17:46:05.6 & $-$28:51:41 & +0.1199 &$-$0.0482 & 6.0 & [NHS93]~22 & 4, 5, 6. \\\\ NHS~42 \\dots& 17:46:08.3 & $-$28:49:55 & +0.1481 &$-$0.0426 & 6.05 & qF~578, [NHS93]~42 & 4, 5, 6. \\\\ NHS~25 \\dots& 17:46:15.3 & $-$28:50:04 & +0.1592 &$-$0.0657 & 6.2 &NAME PISTOL STAR, [NHS93]~25& 4, 5, 6. \\\\ GCS~3-2 \\dots & 17:46:14.7 & $-$28:49:41 & +0.1636 &$-$0.0606 & 2.5 & GCS~3-2, [NHS93]~24 & 4, 5, 6. \\\\ \\hline \\end{tabular} \\tablecomments{Reference: 1. \\citet{kra95}, 2. \\citet{ott99}, 3. \\citet{tan06}, 4. \\citet{nag90}, 5. \\citet{nag93}, 6. \\citet{fig99}. } \\end{center} \\end{table*} \\subsection{H$_3^+$ as an Astrophysical Probe} The crucial information obtained from the observations of H$_3^+$ reported here is the product of the ionization rate of molecular hydrogen \\emph{$\\zeta$}, and the aggregate length of the line of sight through the H$_3^+$ - containing clouds, $L$. The simple chemistry of H$_3^+$ allows us to obtain this product from the observed total H$_3^+$ column densities \\citep{oka05,oka06b}, as described below, under the assumptions that (1) there is a steady-state balance between creation and destruction of H$_3^+$ and (2) the dominant supplier of electrons is carbon, all of which is singly ionized. Creation of H$_3^+$ begins with ionization of H$_2$. Because in interstellar regions where H$_2$ is plentiful the ion-neutral reaction H$_2$ + H$_2^+$ $\\rightarrow$ H$_3^+$ + H is rapid, the limit to the formation of H$_3^+$ is the much slower ionization rate of molecular hydrogen, \\[ {\\rm H_2} + {\\rm R} \\rightarrow {\\rm H_2^+} + e + {\\rm R}^{\\prime}. \\] \\noindent The ionizing rays, R, are dominated by cosmic rays in the Galactic disk \\citep[][and references therein]{ind07}, but in the CMZ the ionization by X-rays and EUV radiation may be significant. The production rate per unit volume of H$_3^+$ is thus given by \\emph{$\\zeta$}$n$(H$_2$), where \\emph{$\\zeta$} is the effective ionization rate of H$_2$ including all the aforementioned contributors. In steady state, production of H$_3^+$ balances its destruction, which is dominated by dissociative recombination with electrons, \\[ {\\rm H_3^+} + {\\rm e} \\rightarrow {\\rm 3H~~or~~H_2} + {\\rm H}, \\] \\noindent and one obtains \\citep{geb99} \\[ \\zeta n({\\rm H_2}) = k_{\\rm e}n({\\rm H_3^+})n({\\rm e}), \\] \\noindent where $k_e$ is the electron recombination rate constant, whose temperature dependent value is $\\sim$ 10$^{-7}$ cm$^3$ s$^{-1}$ \\citep{mcc04}. Replacing the electron density with the atomic carbon density after depletion gives \\[ {n({\\rm e})}/{n({\\rm H_2})} = ({2}/{f}) \\left[{n_{\\rm C}}/{n_{\\rm H}}\\right]_{\\rm SV} R_X, \\] \\noindent where $R_X$ = 3 -- 10 is the factor increase in relative carbon abundance at the Galactic center over the solar vicinity, $[n_{\\rm C}/n_{\\rm H}]_{\\rm SV}$ \\citep{sod95}, and $f = 2n({\\rm H_2}) /n_{\\rm H}$ is the fractional density of molecular hydrogen relative to the total number of hydrogen atoms, $n_{\\rm H} = 2n({\\rm H_2}) + n({\\rm H})$. Using $N$(H$_3^+$)~=~$L$$n$(H$_3^+$) because of the constancy of $n$(H$_3^+$) \\citep[e.g.,][]{oka06b}, we obtain the equation central to the analysis in this paper relating the product \\emph{$\\zeta$L} to the total H$_3^+$ column density and other measurables, \\begin{equation} \\zeta L = 2 k_{\\rm e} \\left[{n_{\\rm C}}/{n_{\\rm H}}\\right]_{\\rm SV} R_X N({\\rm H_3^+})_{total}/f. \\end{equation} As discussed by \\citet{oka05}, in the Galactic center the spectrum of H$_3^+$ provides direct information about the temperature and density of the gas. There the relative population of H$_3^+$ in the ($J$, $K$) = (3,3) metastable level, which is 361 K above the lowest (1,1) level (note that $J$ =~0 does not exist in the ground vibrational state), is measurable and is a useful thermometer for temperatures of $\\sim$ 100 -- 1000~K. Observations of the fractional population in the (2,2) unstable level, from which H$_3^+$ decays to the (1,1) level by spontaneous emission in 27 days \\citep{pan86,nea96}, provides a good measure of cloud densities less than several hundred per cm$^{3}$. The lifetime of H$_3^+$ in diffuse clouds is $1/(k_{\\rm e} n_{\\rm e}) \\approx 10^9$~s, which is two orders of magnitude longer than the collisional timescale with molecular hydrogen. The rotational temperature of H$_3^+$ is therefore maintained via collisions with the ambient H$_2$. The thermalization of H$_3^+$ in interstellar clouds has been modeled by \\citet{oka04}, who provide H$_3^+$ population ratios $n$(3,3)/$n$(1,1) and $n$(3,3)/$n$(2,2) plotted as functions of temperature and density (see their Figs.~2-4). ", "conclusions": "% \\subsection{The Ionization Rate} An estimate of the ionization rate, $\\zeta$, in the Galactic center is the most important result of these $\\mathrm{H}_{3}^{+}$ observations. To avoid confusion, we use $\\zeta$(H) and $\\zeta$(H$_2$) for the ionization rates of H and H$_2$, respectively, in the Galactic disk where ionization is mainly due to cosmic-rays. We use $\\zeta$ for the rate in the CMZ where the effect of X-rays and FUV radiation may be significant. $\\zeta$(H) and $\\zeta$(H$_2$) are related approximately as $\\zeta$(H$_2$) = 2$\\zeta$(H) \\citep{dal06}. Early studies of the cosmic ray ionization rate were made in order to understand the physical state of the interstellar medium and its heating mechanisms \\citep{hay61,spi68,fie69}. \\citet{spi68} reported a minimum value of \\emph{$\\zeta_f$}~=~6.8~$\\times$~10$^{-18}$~s$^{-1}$ from an extrapolation of the cosmic ray spectrum observed at Earth, and an upper limit of 1.2 $\\times$ 10$^{-15}$ s$^{-1}$ calculated from an assumed supernova frequency. Their \\emph{$\\zeta_f$} is the total ionization rate for atomic hydrogen including the effect of secondary electrons, written here as \\emph{$\\zeta$}(H). The most recent measurement of \\emph{$\\zeta$}(H), reported by \\citet{web98} using data from the \\emph{Voyager} and \\emph{Pioneer} spacecraft at distances up to 60 AU from the Sun, is \\emph{$\\zeta$}(H)~$\\geq$~(3--4)~$\\times~10^{-17}$~s$^{-1}$, which should replace the lower limit used by \\citet{spi68}. With the advent of the {\\it Copernicus} satellite and its far-ultraviolet spectrometer, which allowed FUV studies of chemical abundances in the diffuse ISM \\citep[for a review; ][]{spi75}, column densities of molecules and radicals such as HD and OH could be measured. The reaction sequences leading to the creation of these two molecules start from H$^+$ and are thus useful for estimating \\emph{$\\zeta$}(H). Values of \\emph{$\\zeta$}(H) determined in this way \\citep{odo74,bla77,har78} are mostly a few times 10$^{-17}$ s$^{-1}$, comparable to the lower limit given above. The comprehensive model of the diffuse interstellar medium by \\citet{dis86} suggested a somewhat higher value of \\emph{$\\zeta$}(H)~$\\approx$~7~$\\times$~10$^{-17}$~s$^{-1}$. However they used an H$_3^+$ dissociative recombination rate constant one thousand times lower than the current value which was popular in some circles at the time. They noted that if the thousand times higher constant were to be used, the value of \\emph{$\\zeta$}(H) would need to be increased by a factor of $\\sim$ 10. The formation and destruction mechanisms for H$_3^+$ are the simplest among all astrophysical probes of the ionization rate, and this makes H$_3^+$, when detectable, a powerful tool for measuring the ionization rate. While the H$_3^+$ column densities observed in dense clouds, \\mbox{(0.4~--~2.3)~$\\times$~10$^{14}$~cm$^{-2}$}, are consistent with \\emph{$\\zeta$}(H$_2$) on the order of 3~$\\times$~10$^{-17}$~s$^{-1}$ \\citep{geb96, mcc99}, the surprisingly high $N$(H$_3^+$) observed in diffuse clouds, comparable to those in dense clouds in spite of their 10 times smaller visual extinction \\citep{mcc98, geb99, mcc02}, require an order of magnitude higher \\emph{$\\zeta$}(H$_2$). Crucial to this conclusion are the laboratory measurements of the dissociative recombination rate constant $k_e$ in Eq.~(1) \\citep{ama88,lar93,lar00,mcc03,mcc04}. Together, \\citet{mcc02} and \\citet{ind07} have carried out a survey of H$_3^+$ in the diffuse interstellar medium toward 29 nearby stars. H$_3^+$ was detected in 14 of these sight lines with column densities of (0.6~--~6.5)~$\\times$~10$^{14}$~cm$^{-2}$. For these diffuse clouds the product $\\zeta$$L$ is in the range (0.5~--~4.4)~$\\times$~10$^4$~cm~s$^{-1}$. Separating $\\zeta$ and $L$ is not straightforward. Based on various methods of estimating $L$, which gave values of 2.2~-~31 pc for different clouds, the cosmic ray ionization rates were found to be in the range \\emph{$\\zeta$}(H$_2$)~=~(1.2~--~7.4)~$\\times$~10$^{-16}$~s$^{-1}$ with an average of 5 $\\times $10$^{-16}$~s$^{-1}$. Note that the primary ionization rate of H, $\\zeta$$_p$, used by \\citet{ind07} is related to $\\zeta$(H$_2$) by 2.3$\\zeta$$_p$ = $\\zeta$(H$_2$) \\citep{gla74}. This has established that the cosmic ray ionization rate in diffuse clouds is an order of magnitude higher than in dense clouds. This result might be understandable in view of attenuation of soft components of cosmic rays in dense clouds \\citep{cra78}, if a large flux of low energy cosmic rays is ubiquitous in the Galaxy \\citep{mcc03}, or if higher energy cosmic rays are trapped by Alfv\\'{e}n waves in diffuse clouds \\citep{pad05}. The discrepancy between the high \\emph{$\\zeta$}(H$_2$) and previous low \\emph{$\\zeta$}(H) values determined from HD and OH has been explained by \\citet{lis03} as due to neutralization of H$^+$ by charge exchange with small grains which had not been considered in the previous studies, in which the neutralization of H$^+$ was assumed to be due to much slower radiative recombination with electrons. As seen in Table~\\ref{tb3}, the total column density of H$_3^+$ in the CMZ toward the eight observed stars ranges over (1.8 -- 6.1) $\\times$ 10$^{15}$ cm$^{-2}$, more than an order of magnitude higher than the column densities in diffuse clouds in the Galactic disk, suggesting even higher values of $\\zeta$ and $L$. Since we are interested in setting the minimum value of both $\\zeta$ and $L$, we use the maximum value of $f$~=~1 and the minimum value of $R_X$ = 3 \\citep{sod95, ari96} in Eq.~(1) and obtain the numerical expression, \\begin{equation} (\\zeta L)_{\\rm min} = 7.4 \\times 10^{-11} {\\rm cm}^3 {\\rm s}^{-1} N({\\rm H}_3^+)_{\\rm total}. \\end{equation} \\noindent To obtain Eq.~(2), we have used $k_e$~=~7.7~$\\times$~10$^{-8}$~cm$^3$~s$^{-1}$ calculated for $T$ = 250 K from the temperature dependence reported by \\citet{mcc04} assuming the same electron temperature as the gas temperature, which may be justified in view of the fast thermalization of electrons with molecules and atoms. We used the C to H ratio in the solar vicinity of 1.6 $\\times$ 10$^{-4}$ given by \\cite{sof04}, which is close to the value of 1.4 $\\times$ $10^{-4}$, determined directly from infrared spectra of CO and H$_2$ toward an embedded young star NGC~2024~IRS~2 \\citep{lac94}, and $1.3 \\times 10^{-4}$, found by \\citet{usu08} with the same technique but toward NGC~7538~IRS~1 in the outer Galaxy. The observed $N$(H$_3^+$)$_{\\rm total}$ yield lower limits for $\\zeta$$L$, ($\\zeta$$L$)$_{\\rm min}$ = (1.3 -- 4.5) $\\times$ 10$^5$ cm s$^{-1}$. Although lower limits, the values of ($\\zeta$$L$)$_{\\rm min}$ are 1000 times higher than values of $\\zeta$$L$ determined for dense clouds \\citep{mcc99} and 10 times higher than values for diffuse clouds in the Galactic disk \\citep{ind07}. We cannot cleanly separate $\\zeta$ and $L$ based on current understanding of the CMZ, but discuss two extreme cases in the following sections. \\begin{table*} \\begin{center} \\tablewidth{\\textwidth} \\tiny \\caption{Column densities of H$_3^+$ in the Galactic center.\\label{tb4}} \\begin{tabular}{l cccc cc cc c cc} \\hline \\hline &\\multicolumn{4}{c}{$N({\\rm H_3^+})_{\\rm Level}$\\tablenotemark{a}} &\\multicolumn{2}{c}{Relative Population\\tablenotemark{a}} & & & & \\\\ &\\multicolumn{4}{c}{-----------------------------------------------------------------------} &\\multicolumn{2}{c}{--------------------------------------------} & & & & & \\\\ & $N(1,1)$ &$N(1,0)$ &$N(3,3)$ &$N(2,2)$ & $N(3,3)/N(1,1)$ &$N(3,3)/N(2,2)$ % & $n(\\rm{H_2})$ & $T(\\rm{H_2})$ & $N({\\rm H_3^+})_{\\rm Total}$ & $(\\zeta L)_{\\rm min}$ & $L$ \\tablenotemark{b}\\\\ Source &[10$^{14}$cm$^{-2}$]&[10$^{14}$cm$^{-2}$]&[10$^{14}$cm$^{-2}$]&[10$^{14}$cm$^{-2}$] & & & [cm$^{-3}$] & [K] & [10$^{14}$cm$^{-2}$] & [$10^3$cm~s$^{-1}$] & [pc]\\\\ \\hline GC~IRS~21\\dots & 28.1$\\pm$12.3 & 9.0$\\pm$ 4.4 & 24.2$\\pm$12.1 & $<$ 8.3 & 0.86$\\pm$0.57 & $>$2.93 & $<$ 125 & 150--450 & 61.3$\\pm$ 17.8 & 451$\\pm$ 131 & 146$\\pm$ 43\\\\ GC~IRS~3 \\dots & 10.8$\\pm$ 2.1 & 3.4$\\pm$ 1.3 & 8.4$\\pm$ 1.9 & $<$ 1.6 & 0.78$\\pm$0.23 & $>$5.36 & $<$ 50 & 225--400 & 22.5$\\pm$ 3.2 & 166$\\pm$ 23 & 54$\\pm$ 8\\\\ GC~IRS~1W\\dots & 18.1$\\pm$ 3.8 & 7.9$\\pm$ 2.4 & 11.7$\\pm$ 3.0 & $<$ 3.3 & 0.65$\\pm$0.21 & $>$3.53 & $<$ 75 & 200--300 & 37.6$\\pm$ 5.4 & 277$\\pm$ 40 & 90$\\pm$ 13\\\\ NHS~21 \\dots & 8.9$\\pm$ 2.2 & 3.7$\\pm$ 1.5 & 5.6$\\pm$ 1.4 & $<$ 2.3 & 0.62$\\pm$0.22 & $>$2.44 & $<$ 125 & 175--275 & 18.2$\\pm$ 3.1 & 134$\\pm$ 22 & 44$\\pm$ 7\\\\ NHS~22 \\dots & 16.9$\\pm$ 5.6 & 7.9$\\pm$ 3.5 & 9.7$\\pm$ 2.7 & $<$ 3.6 & 0.57$\\pm$0.25 & $>$2.73 & $<$ 100 & 150--275 & 34.5$\\pm$ 7.2 & 254$\\pm$ 53 & 82$\\pm$ 17\\\\ NHS~42 \\dots & 17.7$\\pm$ 5.1 & 8.4$\\pm$ 4.1 & 8.6$\\pm$ 4.1 & $<$ 4.0 & 0.49$\\pm$0.27 & $>$2.13 & $<$ 100 & 125--250 & 34.6$\\pm$ 7.7 & 255$\\pm$ 57 & 83$\\pm$ 18\\\\ NHS~25 \\dots & 11.4$\\pm$ 5.9 & 4.0$\\pm$ 3.8 & 7.4$\\pm$ 5.3 & $<$ 3.9 & 0.65$\\pm$0.57 & $>$1.89 & $<$ 175 & 100--300 & 22.8$\\pm$ 8.8 & 168$\\pm$ 64 & 54$\\pm$ 21\\\\ GCS~3-2 \\dots & 17.0$\\pm$ 1.7 & 4.6$\\pm$ 0.8 & 9.8$\\pm$ 1.6 & $<$ 3.0 & 0.57$\\pm$0.11 & $>$3.21 & $<$ 80 & 200--250 & 31.4$\\pm$ 2.5 & 231$\\pm$ 18 & 75$\\pm$ 6\\\\ \\hline GC~IRS~3~(50~km~s$^{-1}$)\\dots & 3.2$\\pm$ 0.5 & 2.1$\\pm$ 0.4 & 4.9$\\pm$ 0.5 & 1.9$\\pm$ 0.6 & 1.55$\\pm$0.29& 2.61$\\pm$0.50& 150--350 & 350--500 & 10.2$\\pm$ 0.8 & 75$\\pm$ 6 & 24$\\pm$ 2\\\\ \\hline \\end{tabular} \\tablenotetext{a}{All column densities are for the CMZ only.} \\tablenotetext{b}{Using $\\zeta$ = 1 $\\times$ 10$^{-15}$~s$^{-1}$.} \\normalsize \\end{center} \\end{table*} \\subsection{The Case of Large $L$ in the CMZ} \\indent \\citet{oka05} favored a large $L$ and hence a high volume filling factor for the newly found warm and diffuse molecular gas. This was not only because the $R$(3,3)$^l$ metastable lines with Large velocity widths and similar velocity profiles had been detected along all eight lines of sight toward sources scattered over 30~pc in projected distance from Sgr~A$^\\ast$, but also because the deduced $T$ and $n$ of the gas were similar in all cases. The latter suggests that the absorbing diffuse material is continuous rather than contained in several separate entities. Further support for this interpretation is that the range of H$_3^+$ absorption velocities in each sight line spans the entire range velocities observed at radio and millimeter wavelengths with large apertures and interpreted as coming from the front side of the CMZ. Extending spectroscopy of H$_3^+$ over more widely spaced sight lines in the CMZ will constitute a more definitive test of the filling factor. A long path length may also be favored over high ionization, in that there are limits to the value of $\\zeta$ for efficient production of H$_3^+$ as discussed in the next subsection. $L$ is unlikely to be significantly larger than the radius of the CMZ, which for the region of interest is $\\sim$ 130 pc, as estimated using Fig.~11b of \\citet{saw04}. Setting $L$~=~100 pc, close to this maximum permissible value, the aforementioned range of values of ($\\zeta$$L$)$_{\\rm min}$ yields the lower limits $\\zeta$$_{min}$ = (0.4 -- 1.5)~$\\times$~10$^{-15}$~s$^{-1}$ for the various sight lines. For lower values of $f$ and higher values of $R_X$, $\\zeta$$_{min}$ will be higher. The large extent of the warm and diffuse molecular gas postulated above conflicts with the previous concept of the gas in the CMZ, represented, for example, in Fig.~9 of \\citet{laz98} where the ultra-hot X-ray emitting plasma fills the CMZ. The warm and diffuse gas revealed in this paper cannot coexist with the ultra-hot plasma because the plasma electrons will be immediately cooled and molecules will be destroyed. Thus in the above model the volume filling factor of the ultra-hot gas would be reduced by a large factor. The large extent of the ultra-hot plasma gas in the CMZ was proposed because of observations of extensive X-ray emission near the GC \\citep[e.g.,][]{koy89,yam93,koy96}. \\citet{sun93} and \\citet{mar93} observed similar X-ray emission, but interpreted it differently in view of the difficulty of confining such a hot gas by the gravitational potential of the Galactic center, and the extraordinarily large energy input required to maintain the high temperature. With the advent of the \\emph{Chandra} X-ray satellite, a great many X-ray point sources have been resolved in the GC. These have been interpreted as explaining all \\citep{wan02}, or a significant fraction \\citep{mun03}, of the observed intense X-rays as being due to stellar sources such as cataclysmic variables and young stellar objects. Additional observational reports also argue against the ultra-hot plasma based on the close correlation between the stellar distribution and the observed intensities of the X-rays in the GC \\citep [e.g.,][]{war06} and in the Galactic ridge \\citep [e.g.,][]{rev06}. However, interpretation of the X-ray emission remains controversial \\citep[e.g.,][]{mun04,moN06,koy07}. Yet another category of high temperature gas observed in the CMZ is the hot gas observed in recombination lines and other phenomena. \\citet{laz98} studied the $\\lambda^2$-dependent hyper-strong radio-wave scattering and free-free radio emissions and absorptions and interpreted the data as evidence for \\emph{T$_e$} $\\sim 10^6$ K and \\emph{n$_e$} $\\sim 10$ cm$^{-3}$ gas with nearly 100\\% surface filling factor toward the CMZ \\citep[see][for a different interpretation]{gol06}. \\citet{rod05} have observed fine structure and recombination lines toward a sample of 18 sources and reported even higher electron densities of 30 -- 100~cm$^{-3}$. \\citet{laz98} inferred that these regions were at interfaces between molecular clouds and the X-ray plasma and proposed (see their Fig.~9) that the high density molecular clouds with a volume filling factor of perhaps \\emph{f$_{\\rm v}$} $\\sim$ 0.1 are surrounded by the scattering electron gas, and the rest of the space is filled with X-ray emitting plasma. Our observations and analysis \\citep[see also][]{oka05} contradict such a picture \\citep{bol06}. \\begin{figure*} \\begin{center} \\includegraphics[angle=-90,width=.8\\textwidth]{f4.eps} \\caption{Temperature and density plotted in filled circles as functions of population ratios $n$(3,3)/$n$(1,1) and $n$(3,3)/$n$(2,2) obtained by inverting the diagram of \\citet{oka04}. Temperature is indicated by thick white lines and density by thin lines. The estimated uncertainties in $n$(3,3)/$n$(1,1) are shown in error bars. Velocity resolved data of GCS~3-2 from \\citet{oka05} is shown in filled squares for comparison. \\label{f4}} \\end{center} \\end{figure*} \\subsection{The Case of High $\\zeta$ in the CMZ} It is usually assumed that the density of high energy ($E > 100$~MeV) cosmic rays is approximately uniform over the Galaxy because of their long mean free paths. The uniform distribution on the sky of diffuse $\\gamma$ ray emission appears to lend support to this \\citep[e.g.][]{die01}. Recent detailed calculations by \\citet{bue05} typically give a 20~\\% variation depending on location and time. However, the situation is very different for low-energy cosmic rays, which are the most important for ionization of the gas; their energy densities can vary greatly depending on proximity to local cosmic sources \\citep{kul71,spi75,ces75} because of their short mean free paths. For example, a proton of 2~MeV, which effectively ionizes hydrogen, travels only 3~pc before losing its energy in a medium with $n_H$ = 100~cm$^{-3}$ \\citep{cra78}. The ionization rate by low-energy cosmic rays could be proportional to the number density of the accelerating sources \\citep[SNRs; e.g.][]{koy95} and therefore have strong local variations. If the energy density of cosmic-rays follows the surface density of OB stars, the cosmic ray ionization rate would be an order of magnitude higher in the Galactic center than in the solar neighborhood \\citep{wol03}. \\citet{yus02} and \\citet{yus07} called for an even higher local enhancement of low-energy cosmic rays in the Galactic center ($\\zeta = 2~\\times~10^{-14}$ s$^{-1}$ to $5~\\times~10^{-13}$s$^{-1}$!) to account for the iron fluorescence line at 6.4~keV. Such values are much larger than those derived in the previous section using H$_3^+$. X-rays and far ultraviolet radiation from stars may also increase $\\zeta$. A factor of 10 or higher increase of $\\zeta$ due to X-ray emission has even been reported in the circumstellar disk of a low mass star \\citep{dot04}. The enhancements could be large for the CMZ, where there are abundant intense emitters of X-rays and EUV radiation. For the above high values of $\\zeta$, $L$ would be quite short if Eq.~(2) is applied. However, there are two arguments against adopting such high values of $\\zeta$. First, as argued earlier, the observed high velocity dispersion of H$_3^+$ indicates that the gas is quite extensive rather than confined in many small clumps. This indicates that there are non-local sources of ionization covering an extensive region. Second, the results of Eqs.~(1) and (2) are based on a simple linear formalism, which is a good approximation for low $\\zeta$. For much higher $\\zeta$ this should be replaced by higher order non-linear formalism in which $N$(H$_3^+$) no longer increases proportionally to $\\zeta$, but tends to saturate. One source of such nonlinearity, revealed by \\citet{lis06,lis07}, is the dissociative recombination of H$_2^+$, H$_2^+$ + e$^-$ $\\rightarrow$ H + H, which competes with the H$_3^+$ production reaction H$_2^+$ + H$_2$ $\\rightarrow$ H$_3^+$ + H. The former reaction, in addition to the dissociative recombination of H$_3^+$, reduces the H$_3^+$ concentration twice. It also increases $n$(H) and decreases $f$, hence the non-linearity. In an example given in the lower figure of Fig.~3 of \\citet{lis06}, a ten times increase of $\\zeta$ from 10$^{-16}$ s$^{-1}$ to 10$^{-15}$ s$^{-1}$ leads to only a fourfold increase of H$_3^+$ (the saturation is more severe for the upper figure). The importance of saturation depends sensitively on the value of $f$, the fractional abundance of H$_2$. Another source of non-linear behavior is the production of H through dissociation of H$_2$ by cosmic rays, X-rays, and EUV radiation. The rates of these processes are comparable to the rate of ionization \\citep{gla73,gla74}. Since the reproduction of H$_2$ on dust grains is slow, H accumulates in the gas and the charge exchange reaction, H$_2^+$ + H $\\rightarrow$ H$_2$ + H$^+$, which has a high rate constant of 6.4~$\\times$~10$^{-10}$~cm$^3$~s$^{-1}$, will compete with the H$_3^+$ producing reaction introducing additional non-linearity. The increase of $n$(H) and resulting reduction of $f$ affect the gas in several other ways, all of which make the increase of $\\zeta$ ineffective in increasing $\\mathrm{H}_{3}^{+}$. While more detailed treatments of these nonlinearities have yet to be worked out, it seems unlikely to us that $\\zeta$ can be larger than $\\sim$ 3 $\\times$ 10$^{-15}$ s$^{-1}$ over extensive portions of the CMZ. Small volumes with higher $\\zeta$ may well exist in the CMZ, but they would not necessarily produce a large fraction of the observed H$_{3}^{+}$. \\subsection{The $+$50~km~s$^{-1}$ absorption toward GC~IRS~3} The sight line toward GC~IRS~3 is special in that at $+$50~km~s$^{-1}$ LSR it has a detectable population of H$_3^+$ in the unstable (2,2)~level in addition to the large populations in the (1,1) and (3,3) levels. Also this is the only case for which the population in the (3,3) metastable level is definitely higher than in the (1,1) ground level (see $N$(3,3)/$N$(1,1) in Table 4). The observable population in the (2,2) level indicates a high density and the larger population in the (3,3) level than in the (1,1) level indicates high temperature environment. Our analysis yields $n$~=~175~--~300 cm$^{-3}$ and $T$~=~350~--~500~K. As seen from Fig.~4, this environment is clearly distinct from those elsewhere. We speculate below on two possible locations for this gas: (1) the Circumnuclear Disk (CND) and, (2) the ``50~km~s$^{-1}$ cloud'' behind the GC. \\subsubsection{Circumnuclear disk} GC~IRS~3 is only 5\\arcsec~(0.2~pc) from Sgr~A$^\\ast$, which is well within the projected span of the Central Cluster \\citep[the maximum transverse separation of bright stars in the cluster is $\\sim$0.65~pc; ][]{vie05}. The central 3~pc of the GC contains hot, high density gas which shows intricate mini-spiral structures \\citep{eke83,lo83} termed the Western Arc, the Northern Arm, the Eastern Arm, and the Bar \\citep{gue87}. The extension of the Western Arc has been associated with a nearly completely circular, Keplerian circumnuclear disk by analysis of HCN radio emission \\citep{gue87,gen87,jac93,chr05} and by other methods; recent direct observations of the disk via its far infrared dust emission by \\citet{lat99} are particularly noteworthy. There also have been attempts to interpret other three mini-spirals as CNDs which are not necessarily elliptical \\citep[][e. g.]{lis85,qui85,lac91,jac93,lis03}. GC~IRS~3 is located in the cavity between the Northern Arm and the Bar, a region of low gas density. H. S. Liszt (private communication) has suggested that the \\ion{H}{1} absorption at $+$50~km~s$^{-1}$ in Fig.~5 of \\citet{lis85} may be due to the same cloud producing the H$_3^+$ absorption in the sightline toward GC~IRS~3. Not only does the velocity match, but also the high velocity gradient of 100~km~s$^{-1}$ arcmin$^{-1}$ found by \\citet{lis85} is consistent with the width of the H$_3^+$ 50~km~s$^{-1}$ absorption feature. However, there are two possible problems with this interpretation. First, it calls for an uncomfortably high cosmic ionization rate in order to account for the observed H$_3^+$ column density in the short path length within the CND. Assuming a path length on the order of 0.5 pc, the observed ($\\zeta$$L$)$_{\\rm min}$ of 7.5 $\\times$ 10$^4$~cm~s$^{-1}$ in Table~\\ref{tb4} gives $\\zeta$~$\\sim$~2~$\\times$~10$^{-14}$~s$^{-1}$, an extremely high value. Such a value is possible \\citep{yus07} in view of the high density of stars and radiation field in the region, but in such conditions the higher order chemistry \\citep{lis06,lis07} discussed in Section 5.3. surely plays a role. Although the detection of H$_3^+$ in the unstable (2,2) level indicates higher density compared with other sight lines, the population of the (3,3) level is still much higher and the distribution is highly non-thermal. The density is definitely much lower than that reported for the CND (10$^4$~--~10$^6$~cm$^{-3}$). This could be due to the fact that GC~IRS~3 is in the northern cavity and is only behind the edge of the CND rather than its densest regions. The second problem is the absence of the $R$(2,2)$^l$ line in the spectrum of GC~IRS~1W. It is remarkable that the $R$(2,2)$^l$ line is detected toward GC~IRS~3, but not toward GC~IRS~1W which is just 8\\farcs4 apart, or 0.31~pc for a common distance of 7.6~kpc \\citep{eis05,nis06}. If both stars belong to the Central Cluster \\citep{bec78,vie05}, it is not likely that they are much more than 0.5~pc apart. Although the source might well be located in the foreground, GC~IRS~1W overlaps with the high density part of the Northern Arm, while GC~IRS~3 is located in the cavity. It is difficult To understand why only GC~IRS~3 shows $R$(2,2)$^l$ absorption, if the absorption occurs within the CND. \\subsubsection{The``50~km~s$^{-1}$ cloud''} The observed velocity of $+$50~km~s$^{-1}$ toward GC~IRS3 immediately suggests that the absorption might arise in the well known ``50~km~s$^{-1}$ cloud'', one of the giant molecular clouds ($> 10^6 M_\\odot$) located near the Galactic center. The ``50~km~s$^{-1}$ cloud'' is sometimes referred to as M$-$0.02$-$0.07, or was called the ``40~km~s$^{-1}$ cloud'' by \\citet{oor77}, and has been discussed in detail by \\citet{bro84}. Many atomic and molecular studies have established that it is located \\emph{behind} the Galactic nucleus, in between the non-thermal radio source Sgr~A East (SNR) and the \\ion{H}{2} region Sgr~A West, which enshrouds the central star cluster \\citep{whi74,gue83,pau96}. \\citet{yus96} argued that the several masers they discovered in the southeast of Sgr~A$^\\ast$ are excited in the shock-heated interface between Sgr~A~East and M$-$0.02$-$0.07, where the expanding supernova remnant hits the giant molecular cloud. \\citet{tsu06} also found clues for an interaction between Sgr~A~East and the ``50~km~s$^{-1}$ cloud'' in the high velocity wings of CS $J=1-0$, although it may not be the case that the ``50~km~s$^{-1}$ cloud'' is the sole cloud that falls in the radial velocity $+$50~km~s$^{-1}$ in the line of sight toward the Galactic center \\citep{zyl92}. \\citet{geb89} found that GC~IRS~3 and IRS~7 show deep absorption lines of the CO fundamental band at 4.7~$\\mu$m at the radial velocity $+$50~km~s$^{-1}$, while GC~IRS~1 and IRS~2 do not. They attributed the $+$50~km~s$^{-1}$ absorption to the CND but also mentioned the possibility that GC~IRS~3 is not a member of the central cluster, but behind the $+$50~km~s$^{-1}$ cloud and thus not in the Central Cluster. The nature of GC~IRS~3 is controversial \\citep{pot08}; it is among the brightest sources in the Central Cluster region only less lumnous than GC~IRS~1 and GC~IRS~10, and much redder than them in the mid-infrared even though it is (apparently) in the cavity. Because of its location and featureless infrared continuum \\citep[e.g. ][]{kra95}, GC~IRS~3 has widely been believed to be a hot massive star heavily obscured by its own dust (similar to the bright Quintuplet sources), in addition to foreground dust. However, new interferometric observation that resolve the circumstellar dust shell have been interpreted by \\citet{pot08} as evidence that the central object is a late-type carbon star, the photospheric spectral features of which are completely obscured. This interpretation is consistent with the apparent lack of ionization of the circumstellar shell, which one might expect if the object lies within the Central Cluster, and indeed the lack of ionized gas near GC~IRS~3 \\citep{roc85}. Although somewhat unlikely due to its proximity on the plane of the sky to the Central Cluster, the placement of GC~IRS~3 well behind the cluster and the ``$+$50~km~s$^{-1}$ cloud'' is otherwise a straightforward assignment and avoids the difficulties associated with the CND hypothesis mentioned in the previous section. \\subsection{Infrared pumping} During the 2005 Royal Society Discussion Meeting on H$_3^+$ \\citep{oka06a}, J. H. Black proposed an infrared pumping mechanism which can produce a non-thermal rotational level distribution of H$_3^+$ if the infrared flux is sufficiently high. The mechanism is illustrated in Fig.~\\ref{f5}. Infrared radiation pumps H$_3^+$ from the fully populated lowest ortho-level, (1,0), to the (2,0) level in the first excited vibrational state via the $R$(1,0) transition. The molecule decays to the (3,0) level in the ground state by spontaneous emission with a lifetime of 25.3~ms. The molecule further decays to the (3,3) metastable level through the forbidden rotational transition (3,0) $\\rightarrow$ (3,3) emitting a far-infrared photon at 49.62 $\\mu$m in 3.75 hours \\citep{pan86,nea96}. Thus, H$_3^+$ molecules are pumped from the (1,0) level to the metastable (3,3) level. \\begin{figure} \\begin{center} \\includegraphics[angle=0,width=.4\\textwidth]{f5.eps} \\caption{Schematic of the infrared pumping mechanism proposed by J. H. Black. H$_3^+$ in the lowest ortho-level (1,0) is excited to the (2,0) level in the first excited vibrational state via the $R$(1,0) transition (red line) by ambient photons. The excited H$_3^+$ spontaneously decays to the (3,0) ground level through the P(3,0) transition (orange line) and then to the metastable (3,3) level (shown in blue). See text for more details. Other transitions used in this paper are also shown. \\label{f5}} \\end{center} \\end{figure} Since the spontaneous emissions are all very fast, the rate determining process for Black's mechanism is the rate of infrared pumping of the (1,0) level of H$_3^+$. We rewrite the excitation rate given in Eq.~(15) of \\citet{dis82} as \\[ W_{R(1,0)} = \\lambda^2A\\phi_\\nu/8\\pi = 1.51 \\times 10^{-7} \\phi_\\nu {\\rm s}^{-1}, \\] \\noindent where $\\phi_\\nu$ is the photon flux of the ambient radiation at the wavelength of the transition in the units of photon cm$^{-2}$ s$^{-1}$ Hz$^{-1}$, and the values of $\\lambda$ = 3.6685~$\\mu$m, the Einstein coefficient $A$ = 98.5~s$^{-1}$, and the dilution factor of 0.286 are used. The dilution factor takes into account the spontaneous emission (2,0) $\\rightarrow$ (1,0) back to the ground level (1,1) with $A$ = 35.5 s$^{-1}$, resulting in a reduction of the pumping efficiency. In order for pumping to be significant, the value of $ W_{R(1,0)}$ has to be comparable to the collision rate, \\[ R = n\\sigma v = n k_L \\sim 2 \\times 10^{-7} {\\rm s}^{-1}, \\] \\noindent where we used a number density of $n$ = 100 cm$^{-3}$ and the temperature independent Langevin rate constant, 2~$\\times$~10$^{-9}$~cm$^3$~s$^{-1}$ for the H$_3^+$ + H$_2$ collision. The energy density of the interstellar radiation field in the solar neighborhood at 4~$\\mu$m has been estimated by \\citet{por06} to be $\\lambda u_\\lambda$ = 0.055 $\\mu$m~eV~cm$^{-3}$~$\\mu$m$^{-1}$ which translates to $\\phi_\\nu$~=~6~$\\times$10$^{-5}$ photons~cm$^{-2}$~s$^{-1}$~Hz$^{-1}$. This is consistent with the value reported by \\citet{mat83} within a factor of 1.5. The corresponding pumping rate is $W_{R(1,0)}$~=~1~$\\times$~10$^{-11}$ s$^{-1}$, about 2$\\times$10$^{4}$ times lower than the collision rate. \\citet{por06} have estimated an average energy density about 20 times higher in the GC. \\citet{lau02} have reported an energy density in the central 120~pc (their Fig.~10) which is another factor of 10 higher. The corresponding pumping rate is still 100 times lower than the collision rate and the infrared pumping is perhaps not very effective for majority of the gas reported in this paper. In the central parsec of the Galaxy, however, the stellar density is $\\sim$ $10^6$~$M_\\odot$~pc$^{-3}$ \\citep{sch07}, including a large number of luminous red giants and at least 80 hot and massive stars \\citep{pau06}. In this environment infrared pumping likely would overwhelm the collisional relaxation of any H$_3^+$ that is present. The effect, however, must be local to the very central region of the CMZ and probably does not seriously alter the population of the (3,3) level beyond a few tens of parsecs from the center, i.e., in the bulk of the CMZ. However, it may well be the best explanation for the strong $R$(3,3)$^l$ absorption line toward GC~IRS~3 at 50~km~s$^{-1}$ discussed in the previous subsection." }, "0807/0807.3628_arXiv.txt": { "abstract": "{Infrared Dark Clouds (IRDCs), condensed regions of the ISM with high column densities, low temperatures and high masses, are suspected sites of star formation. Thousands of IRDCs have already been identified. To date, it has not been resolved whether IRDCs always show star formation activity and, if so, if massive star formation ($\\gtrsim8 M_\\odot$) is the rule or the exception in IRDCs.} {Previous analysis of sub-millimeter cores in the cloud MSXDC G048.65$-$00.29 (G48.65) indicates embedded star formation activity. To characterize this activity in detail, mid-infrared photometry (3--30~$\\mu$m) has been obtained with the \\textit{Spitzer Space Telescope}. This paper analyzes the point sources seen in the 24~$\\mu$m band, combined with counterparts or upper limits at shorter and longer wavelengths.} {Data points in wavelength bands ranging from 1~$\\mu$m up to $850~\\mu$m (\\textit{Spitzer} IRAC, MIPS~24 and 70~$\\mu$m, archival 2MASS data and sub-millimeter counterparts) are used to compare each 24~$\\mu$m source to a set of Spectral Energy Distributions of Young Stellar Object (YSO) models. By assessing the models that fit the data, an attempt is made to identify YSOs as such and determine their evolutionary stages and stellar masses.} {A total of 17 sources are investigated, 13 of which are classified as YSOs, primarily -- but not exclusively -- in an early embedded phase of star formation. The modeled masses of the central stellar objects range from sub-solar to $\\sim$8 $M_\\odot$. Every YSO is at less than 1 pc projected distance from its nearest YSO neighbor.} {IRDC G48.65 is a region of active star formation. We find YSOs in various evolutionary phases, indicating that the star formation in this cloud is not an instantaneous process. The inferred masses of the central objects suggest that this IRDC hosts only low to intermediate mass YSOs and none with masses exceeding $\\sim$8 $M_\\odot$.} ", "introduction": "Infrared dark clouds (IRDCs), discovered independently by \\citet{perault1996} and \\citet{egan1998} as dark patches against the Galactic mid-infrared background, are suspected to be sites of clustered star formation \\citep{rathborne2006,simon2006}. Moreover, the suggestion is raised \\citep[see e.g.][]{menten2005,beuther2007} that IRDCs harbor massive young stars. Morphologies of IRDCs range from globular to more filamentary structures. One of the most important properties of IRDCs is their high column density \\citep[$\\gtrsim10^{22}\\ \\mathrm{cm^{-2}}$, e.g.][]{carey2000}. Dust absorbs mid-infrared radiation and makes the clouds stand out as dark features against the mid-infrared background. Temperatures in IRDCs are $< 25 \\ \\mathrm{K}$ and volume densities can exceed $10^5 \\ \\mathrm{cm^{-3}}$ \\citep{egan1998,carey2000}. Typical size and mass scales are $\\sim$5 pc and $10^3$--$10^4~M_\\odot$ \\citep{simon2006}. These low initial temperatures, high molecular densities and total masses typical for IRDCs are precisely what make them suitable candidates for star forming regions. IRDCs are generally not quiescent: they harbor compact cores of sub-millimeter emission \\citep[e.g.][]{carey2000,ormel2005,rathborne2005}. IRDCs are found primarily in the inner Galaxy and close to the Galactic plane. Thousands are known in the first and fourth quadrants of the Galactic plane \\citep{simon_catalog2006}, i.e., the inner Galaxy. \\citet{frieswijk2007} recently identified the first IRDC in the outer Galaxy (in the absence of a bright mid-infrared background), using 2MASS (Two Micron All Sky Survey) color distributions of background stars, followed up by observations of molecular lines and \\textit{Spitzer Space Telescope} photometry. It has not yet been established whether all IRDCs show active star formation or that some may in fact harbor only starless cores. Moreover, the association of IRDCs to \\emph{massive} ($\\gtrsim8 M_\\odot$) star formation in particular is still a matter of dispute. Studies of star forming regions in general \\citep[e.g.][]{indebetouw2007} and IRDCs in particular \\citep[e.g.][]{beuther2007} prove that star formation activity in a wide stellar mass range and at various phases of evolution can be probed by Spitzer photometry. The IRDC under investigation in this paper is the cloud at Galactic coordinates $(\\ell,b)=(48\\fdg66, -0\\fdg30)$. It has been previously observed by the Mid-course Space Experiment (MSX), the SCUBA instrument on the James Clerk Maxwell Telescope \\citep[JCMT,][]{ormel2005}, and by the JCMT in CO, $^{13}$CO and HCO$^+$ \\citep{shipman2003}. In addition, it is covered by the Galactic Legacy Infrared Mid-Plane Survey Extraordinaire \\citep[GLIMPSE,][]{benjamin2003}. In the dark cloud catalog of \\citet{simon_catalog2006} this cloud has the designation ``MSXDC G048.65$-$00.29\", hereafter simply ``G48.65\". Its distance is determined kinematically from molecular line data and an assumed Galactic rotation curve. G48.65 is found to be at a distance of $\\sim$2.5$\\ \\mathrm{kpc}$ \\citep{ormel2005,simon2006}. Its distance to the Galactic Center is $\\sim$7~$\\mathrm{kpc}$ and it is less than 20 pc away from the mid-plane of the Galaxy. Its total mass is estimated at almost 600\\,$M_\\odot$ within a 2 pc area and the molecular (H$_2$) density is $\\sim$$10^3 \\ \\mathrm{cm^{-3}}$ \\citep{simon2006}. \\citet{ormel2005} identified three distinct emission cores at 450 and 850 $\\mu$m; their modeling indicates the presence of central luminosity sources on the order of $10^2$--$10^3 \\,L_\\odot$ in at least two of these cores. The earliest phases of star formation show emission at $>$30$~\\mu$m as they are embedded in an accreting envelope; the peak wavelength shifts through the mid- and near-infrared to the optical regime as stars evolve toward the main sequence. As mentioned above, sub-millimeter observations indicate ongoing embedded star formation in two or three cores in G48.65 \\citep{ormel2005}. This raises the question whether the star forming activity in G48.65 is limited to the sub-millimeter cores. Additionally, it is realized that observations at shorter wavelengths can improve our understanding of the previously identified star forming cores. This was the motivation to take a closer look in the mid-infrared (roughly 3--30\\,$\\mu$m), where Young Stellar Objects (YSOs) are known to emit the bulk of their energy. This paper describes the interpretation of \\textit{Spitzer Space Telescope} photometry of G48.65, focussing on those sources that show emission at 24\\,$\\mu$m, where the brightest, massive, embedded young stars are expected to be found. Section \\ref{sec:observations} describes the Spitzer data and their reduction, Section \\ref{sec:results} deals with the classification of twenty sources based on the available data points in the near- and mid-infrared and sub-millimeter regimes. Conclusions and a discussion are presented in Sect.~\\ref{sec:conclusions_and_discussion}. ", "conclusions": "\\label{sec:conclusions_and_discussion} \\subsection{Conclusions} \\label{sec:conclusions} Of the 20 sources near IRDC G48.65 that are visible at 24\\,$\\mu$m, a total of 13 are classified as YSOs, seven of which are found to be in Stage I, two are in Stage II, two more are in either Stage I or Stage II, and two are uncertain. While all reliably classified YSOs lie along the IRDC extinction filament, the four sources matched to photosphere models all lie away from the filament. Each YSO generally resides within a projected distance of $<$1\\,$\\mathrm{pc}$ from another YSO. We conclude that G48.65 is an example of a dark cloud environment forming a group of stars. The most important modeled properties of the 24\\,$\\mu$m sources that are classified as YSOs are summarized in Table \\ref{table:allYSOs}. The stellar masses of the modeled objects range from slightly sub-solar to $\\sim$8\\,$M_\\odot$. The inferred evolutionary Stages are predominantly early phases, ranging from early Stage I to late Stage II. This is believed to be caused partly by the selection effect discussed in Sect.~\\ref{sec:MIPSselection}. Therefore, the presented range of evolutionary phases should not be interpreted as a full picture of all star forming activity in this IRDC. The emission peak P1 identified by \\citet{ormel2005} is resolved by Spitzer into two distinct emission cores: S6 and S5. Both objects are found to be in the earliest Stage of star formation. Stellar masses are not well constrained for these sources but are unlikely to be in excess of $8\\,M_\\odot$. The summed total luminosity of the best-fit models for S5 and S6 is $\\sim$$10^2$--$10^3\\,L_\\odot$, consistent with results from modeling by \\citet{ormel2005} based purely on sub-mm observations. The total luminosity of the best fitting model for objects S12 (EP) and S15 (P2) are consistent with values found by \\citet{ormel2005}. \\subsection{Robustness of Results} \\label{sec:discussion} It is important to assess a YSO based on data points in a wavelength regime as extended as possible. While a detection of a certain YSO at sub-millimeter wavelengths seems to be a strong indication that it is in Stage I (S5, S6, S12, S15), a detection by 2MASS in the near-infrared does not rule out a Stage I classification, as sources S2 and S7 show. It may therefore lead to erroneous classification of YSOs if one bases it either on only near-infrared or on only far-infrared/sub-millimeter data. The availability of data points in the 3--30\\,$\\mu$m regime is in some cases enough to constrain the evolutionary stage of a YSO. Source S8 (see Fig.~\\ref{fig:YSOSEDs} for the SED fit) is an example in which the degeneracy of the model fitting is decreased considerably by adding 2MASS data points (see also Sect.~3.3 and Fig.~3 of \\citet{robitaille2007}). \\subsection{Continuous Low-mass Star Formation} The Stage I sources in particular show large uncertainties in stellar mass (see Table~\\ref{table:allYSOs}). This is not surprising considering that, for objects that emit mostly through re-radiation from the accreting envelope, it is of course inherently difficult to model parameters of the obscured central heating source. However, the OH maser survey by \\citet{pandian2007} indicates the absence of OH maser emission in the direction of IRDC G48.65. Since OH masers are commonly associated with high-mass star formation ($\\gtrsim$8\\,$M_\\odot$), this agrees with the fact that we find no massive YSOs in G48.65. Considering the above, we are confident that any YSO associated to G48.65 is unlikely to harbor a central star more massive than 8\\,$M_\\odot$. Based on the range of stellar masses and evolutionary Stages of the YSOs, we conclude that G48.65 has been an active low-mass star forming region for at least $\\sim$$10^6$ years (late Stage II) up to as recent as $\\sim$$10^4$ years ago (early Stage I) and is likely to be in the process of forming still younger stars. Assuming that the IRDC is indeed the birth place of these young stars, the cloud itself must have been stable over a time scale of at least $\\sim$$10^6$ years. This is consistent with modern views on lifetimes of $\\lesssim$$10^6$ years for molecular clouds \\citep[as reviewed in][]{maclow2004} and with lifetimes of $\\lesssim$$10^7$ years for dense cluster forming clouds \\citep[see review by][]{larson2003}. \\subsection{Mid-infrared Selection Effect} % \\label{sec:MIPSselection} The starting point of the study in this paper -- the set of point sources in the MIPS 24\\,$\\mu$m image -- introduces an observational bias toward the younger evolutionary stages. Early stage YSOs are simply brighter in the mid-infrared, since the total mid-infrared luminosity of a YSO generally decreases with time. Especially in a fairly distant region such as IRDC G48.65, at $\\sim$2.5\\,kpc, one must keep in mind that later stages and intrinsically fainter sources are less likely to be detected. This study in particular is biased toward the earlier stages, since it focusses on the 24\\,$\\mu$m objects. It is primarily the relatively cold envelope component, dominant in Stage I objects, that gives a YSO SED its large ``bump\" longward of 20~$\\mu$m. Hence, an analysis of objects near the IRDC that are only visible at wavelengths shortward of 8~$\\mu$m is expected to lead to the identification of more evolved (Stages II and III) and less massive YSOs. The IRAC 8~$\\mu$m image, for example, shows at least a factor 3 more point sources than the 24~$\\mu$m image in the same field of view. Such a study will be essential in getting a complete census of the star formation activity in G48.65, both in terms of time scales and in terms of a stellar mass function. In principle, the set of young stars of various masses presented in this paper could be used to construct an initial mass function (IMF). Extrapolation of such an IMF could provide insight into the absence of massive stars ($>$8\\,$M_\\odot$). However, we refrain from presenting such an exercise here. The first reason for this is the incompleteness of our sample discussed above; a second reason is that the inferred stellar mass of each individual YSO (see Col.~(5) of Table~\\ref{table:allYSOs}) is rather poorly constrained. \\subsection{Outlook} \\label{sec:outlook} In the near future, the 2MASS point source catalog (1--3\\,$\\mu$m) and the GLIMPSE catalog (3--9.5\\,$\\mu$m) will be complemented by the MIPSGAL point source list of the Galactic plane at 24 and 70\\,$\\mu$m. Moreover, the UKIRT Infrared Deep Sky Survey \\citep{lawrence2007} is already partly available to the general community and will eventually cover half of the Galactic plane in the $J$, $H$ and $K$ bands, superseding the sensitivity of 2MASS by roughly 3 magnitudes. This will open up the possibility to investigate YSOs in hundreds of other IRDCs in the Galaxy using an approach similar to the one taken in this paper. This approach has shown that source-by-source investigation of emission cores with 1--30\\,$\\mu$m data can lead to the identification of young stars in various phases of evolution. Perhaps positions that show sub-millimeter emission are the most interesting cases to start with if one is attempting to identify very early stages of star formation. However, even with the combination of data from 2MASS, GLIMPSE and MIPSGAL, covering just the 1--70\\,$\\mu$m regime, it is possible to systematically study star formation activity across the Galactic plane. IRDCs may eventually be divided into groups according to whether they show star formation activity, and if they do, whether they form low-mass stars or high-mass stars. The Stage I sources in Table \\ref{table:allYSOs} are expected to show associated outflows and shocks, traced by e.g. SiO. Detection of such tracers could confirm the star forming activity and reveal kinematic and spatial structure of YSOs. In addition, measurements by the VISIR instrument (8--13\\,$\\mu$m) at VLT will be able to resolve more detailed structure ($<$$10^3\\ \\mathrm{AU}$ at 2.5\\,kpc) of the sources and show envelope and disk components. It may even show that objects that appear as one source in the Spitzer images are in fact multiple YSOs. Source S5, for example, appears slightly elongated in the Spitzer images. In fact, some of the poorer model fits (see Col.~(4) in Table~\\ref{table:allYSOs}) may be explained by the fact that the models do not account for multiple young star-disk-envelope systems inside one telescope beam. The high spatial and spectral resolution and new wavelength regimes of future observatories, such as ALMA, the Herschel Space Observatory \\citep[specifically the Hi-GAL survey,][]{noriega-crespo2008} and the James Webb Space Telescope, will enable more detailed studies of individual objects associated with G48.65 and other (distant) IRDCs." }, "0807/0807.0301_arXiv.txt": { "abstract": "{ Due to its proximity, youth, and solar-like characteristics with a spectral type of K2V, $\\epsilon$ Eri is one of the most extensively studied systems in an extrasolar planet context. Based on radial velocity, astrometry, and studies of the structure of its circumstellar debris disk, at least two planetary companion candidates to $\\epsilon$ Eri have been inferred in the literature ($\\epsilon$ Eri b, $\\epsilon$ Eri c). Some of these methods also hint at additional companions residing in the system. Here we present a new adaptive optics assisted high-contrast imaging approach that takes advantage of the favourable planet spectral energy distribution at 4 $\\mu$m, using narrow-band angular differential imaging to provide an improved contrast at small and intermediate separations from the star. We use this method to search for planets at orbits intermediate between $\\epsilon$ Eri b (3.4 AU) and $\\epsilon$ Eri c (40 AU). The method is described in detail, and important issues related to the detectability of planets such as the age of $\\epsilon$ Eri and constraints from indirect measurements are discussed. The non-detection of companion candidates provides stringent upper limits for the masses of additional planets. Using a combination of the existing dynamic and imaging data, we exclude the presence of any planetary companion more massive than 3 $M_{\\rm jup}$ anywhere in the $\\epsilon$ Eri system. Specifically, with regards to the possible residual linear radial velocity trend, we find that it is unlikely to correspond to a real physical companion if the system is as young as 200 Myr, whereas if it is as old as 800 Myr, there is an allowed semi-major axis range between about 8.5 and 25 AU. ", "introduction": "\\label{sec_intro} With its distance of 3.2 pc and spectral type K2V, $\\epsilon$ Eri is the most nearby solar-like single star. Largely due to this fact, it has been targeted by several attempts to indirectly or directly detect planetary companions. So far, two exoplanet candidates have been reported as companions to $\\epsilon$ Eri -- $\\epsilon$ Eri b (1.55 $M_{\\rm jup}$, $a=3.4$ AU), which was inferred from a periodic radial velocity signal by Hatzes et al. (2000) and subsequently also from astrometric motion of the primary (Benedict et al. 2006), and $\\epsilon$ Eri c (0.1 $M_{\\rm jup}$, 40 AU), which results from an interpretation of observed irregularities in the dust disk surrounding the primary (Quillen \\& Thorndike, 2002). Targeted attempts have been made to directly image both $\\epsilon$ Eri b (Janson et al. 2007) and $\\epsilon$ Eri c (Macintosh et al. 2003, Marengo et al. 2006), but did not yield any detections. $\\epsilon$ Eri has also been observed as a part of large-scale near-infrared adaptive optics surveys, such as Lafreniere et al. (2007), and Biller et al. (2007). Typically, direct imaging surveys from the ground are performed in narrow-band filters around 1.6 $\\mu$m, since atmospheric opacities cause cool objects (giant planets and brown dwarfs) with temperatures below 800 K to emit a large fraction of their H-band flux just shortward of the CH$_4$ absorption feature at 1.6 $\\mu$m, and since this feature also allows for spectral differential imaging (SDI) to enhance the image contrast. However, the primary-to-secondary brightness contrast is generally smaller at longer wavelengths, such that an improved performance can be gained for some types of high-contrast systems. This was demonstrated with an imaging survey in the L'-band of 22 young, nearby stars by Kasper et al. (2007). The brightness difference is even smaller in M-band (see e.g. Hinz et al. 2006), but here the noise contribution of the thermal background is also significantly higher, making it observationally challenging. Here we examine an intermediate approach, where we use a narrow-band filter towards the red end of the L'-band range, which simultaneously allows for both good brightness contrast and observational simplicity. We test this technique on the $\\epsilon$ Eri system to search for planetary companions at separations intermediate between $\\epsilon$ Eri b and $\\epsilon$ Eri c. The paper is organized as follows: In Sect. \\ref{sec_obsstrat}, we describe the basic principles of the observing strategy chosen, and in Sect. \\ref{sec_obs} we describe the actual observations performed, in terms of the techniques used and the observing conditions and corresponding instrument performance. This is followed in Sect. \\ref{sec_datared} by a description of the data reduction, and the results as given in Sect. \\ref{sec_contrast}. Our choice of filter is further motivated in Sect. \\ref{sec_comparison}, where it is also explained over which parameter range the method should be preferable to other common methods in high-contrast imaging. In Sect. \\ref{sec_age}, we discuss the controversial issue of the age of $\\epsilon$ Eri, which has fundamental implications for the detectability of planetary companions in the system. In Sect. \\ref{sec_masslimits} the brightness contrast is translated into a mass contrast. We examine the possible clues for an intermediate companion in Sect. \\ref{sec_dyn}. Finally, we summarize what these issues mean in terms of mass-equivalent detection limits for exoplanet companions to $\\epsilon$ Eri in Sect. \\ref{sec_limits}, and we conclude in Sect. \\ref{sec_conclusion}. ", "conclusions": "\\label{sec_conclusion} We have performed a deep, high-contrast imaging observation of $\\epsilon$ Eri with a new method that allows for detection of lower-mass planets at small separations from bright stars than previously possible. The method is based on ADI in the NB 4.05 $\\mu$m filter, and compares favourably to other methods in the literature. Its advantage is however limited to very bright stars, since the background limit will dominate the contrast limit for faint stars, such that other methods (e.g. SDI+ADI or L'-band ADI) are preferable for such targets. As the final output images show no signs of any convincing companion candidate, we place constraints on any planets that may reside in the system. In particular, we test the hypothesis of the existence of an intermediate orbit planet possibly indicated by radial velocity, Hipparcos $\\Delta \\mu$ astrometry and numerical attempts to reproduce the circumstellar debris disk structure of $\\epsilon$ Eri. In order to emphasize the hypothetical nature of such a companion, it is referred to as $\\epsilon$ Eri x. We find that, due to the uncertain age of the $\\epsilon$ Eri system, which we quote as 200-800 Myr based on a detailed investigation of the existing literature, the presence of $\\epsilon$ Eri x can not be excluded based on our images, but its allowed parameter range can be constrained. In general, a combination of the existing RV data and high-contrast images allows us to conclude that whatever the amount and distribution of planets is in the $\\epsilon$ Eri system, none of them can be more massive than 3 $M_{\\rm jup}$." }, "0807/0807.0137_arXiv.txt": { "abstract": "% We present the results of Spectral Energy Distribution (SED) fitting analysis for Lyman Break Galaxies (LBGs) at $z\\sim5$ in the GOODS-N and its flanking fields. From the SED fitting for $\\sim100$ objects, we found that the stellar masses range from $10^{8}$ to $10^{11}M_{\\odot}$ with a median value of $4\\times10^{9}M_{\\odot}$. By using the large sample of galaxies at $z\\sim5$, we construct the stellar mass function (SMF) with incompleteness corrections. By integrating down to $10^{8}M_{\\odot}$, the cosmic stellar mass density at $z\\sim5$ is calculated to be $7\\times10^{6}M_{\\odot}\\textrm{Mpc}^{-3}$. ", "introduction": "Recent observations show the gradual increase of the stellar mass density of the universe with time \\citep[e.g.,][]{wilkins08}. However, the studies on the stellar mass of galaxies at $z\\ga5$ are restricted because the lack of sufficiently deep mid-infrared data. With the advent of Spitzer, we can access the rest-frame optical wavelength, and thus, the stellar mass of galaxies at $z\\sim5$. In this work, we explore the stellar mass of Lyman Break Galaxies (LBGs) at $z\\sim5$ using the Subaru/S-Cam imaging data and the Spitzer/IRAC data. ", "conclusions": "" }, "0807/0807.2242_arXiv.txt": { "abstract": "} \\nc{\\eab}{ We extend the previously described CMB Gibbs sampling framework to allow for exact Bayesian analysis of anisotropic universe models, and apply this method to the 5-year WMAP temperature observations. This involves adding support for non-diagonal signal covariance matrices, and implementing a general spectral parameter MCMC sampler. As a worked example we apply these techniques to the model recently introduced by Ackerman et al., describing for instance violations of rotational invariance during the inflationary epoch. After verifying the code with simulated data, we analyze the foreground-reduced 5-year WMAP temperature sky maps. For $\\ell \\le 400$ and the W-band data, we find tentative evidence for a preferred direction pointing towards $(l,b) = (110^{\\circ}, 10^{\\circ})$ with an anisotropy amplitude of $g_* = 0.15 \\pm 0.039$. Similar results are obtained from the V-band data [$g_* = 0.10\\pm 0.04$; $(l,b) = (130^{\\circ}, 20^{\\circ})$]. Further, the preferred direction is stable with respect to multipole range, seen independently in both $\\ell=[2,100]$ and $[100,400]$, although at lower statistical significance. We have not yet been able to establish a fully satisfactory explanation for the observations in terms of known systematics, such as non-cosmological foregrounds, correlated noise or asymmetric beams, but stress that further study of all these issues is warranted before a cosmological interpretation can be supported. ", "introduction": "\\label{secintroduction} Since the early 1990's, great advances have been made in the field of data analysis techniques for studying the cosmic microwave background (CMB). Observations of the CMB anisotropies, for instance those made by the Wilkinson Microwave Anisotropy Probe (WMAP) experiment \\citep{bennett:2003, hinshaw:2007}, provides the single most powerful probe in contemporary cosmology. From these, various theoretical universe models may be constrained, and today an effective concordance model based on the inflationary $\\Lambda$CDM framework has been established. The theory of inflation was initially proposed as a solution to the horizon and flatness problem \\citep{guth:1981}. Additionally, it established a highly successful theory for the formation of primordial density perturbations, thus providing the required seeds for the large-scale structures (LSS), later giving rise to the temperature anisotropies in the cosmic microwave background radiation that we observe today \\citep{starobinsky:1980, guth:1981,linde:1982, muhkanov:1981, starobinsky:1982, linde:1983, linde:1994, smoot:1992, ruhl:2003, runyan:2003, scott:2003}. A firm prediction of inflation is that the observed universe should be nearly isotropic on large scales. Yet, recent theoretical studies have demonstrated that anisotropic inflationary models are indeed conceivable \\citep{armendariz:2006, contaldi:2007, pullen:2007, kanno:2008, yokoyama:2008}. Two other examples are those presented by \\citet{ackerman:2007} (ACW) and \\citet{erickcek:2008}. The first model considers violation of rotational invariance in the early universe, while the second model describes the effects on the observed perturbation distribution due to a large-scale curvaton field. The introduction of anisotropic models poses several problems in terms of data analysis. The definition of a proper likelihood function may be non-trivial for a general case, although many models can be described as multivariate Gaussians with non-diagonal covariance matrices. All models mentioned above are examples of this. Yet, even in these relatively simple cases, the numerical evaluation of the likelihood is computationally unfeasible due to the sheer size of the relevant covariance matrix. In the present paper, we extend the previously described CMB Gibbs sampling framework \\citep{jewell:2004, wandelt:2004, eriksen:2004b} to allow for non-diagonal, but sparse, covariance matrices. As currently described in the literature, this framework allows for exact Bayesian analysis of high-resolution CMB data, but only under the assumption of isotropy, i.e., a diagonal CMB covariance matrix. This method has already been applied several times to the WMAP data \\citep{odwyer:2004, eriksen:2007a, eriksen:2007b, eriksen:2008b}, and has been extended to take into account both polarization \\citep{larson:2007} and internal component separation \\citep{eriksen:2008a}. The question of isotropy has received considerable attention during recent years, due to unexpected signatures observed in the WMAP sky maps. These data appear to exhibit several significant and distinct signatures of violation of statistical isotropy. First, \\citet{de Oliveira-Costa:2004} found a striking alignment between the two largest harmonic modes in the temperature anisotropy sky, the quadrupole and the octopole. Second, \\citet{vielva:2004} pointed out the presence of a very large cold spot in the southern Galactic sky, apparently incompatible with $\\Lambda$CDM-based simulations. Finally, \\citet{eriksen:2004a} found a significantly anisotropic distribution of power between two hemispheres. The tools developed in the present paper may be able to constrain specific models relevant for these observations. In particular, we use these methods to estimate the anisotropy parameters in the ACW model from the 5-year WMAP temperature data. The paper is structured as follows: In \\S \\ref{sec:acwmodel}, we review the ACW universe model, and briefly introduce the relevant posterior distribution. Next, we present the method in \\S\\ref{sec:method}, before we apply our tools to simulated data in \\S\\ref{sec:simulations}. In \\S\\ref{sec:wmap} we analyze the five-year WMAP temperature sky maps. Finally, we conclude in \\S\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} We have generalized a previously described CMB Gibbs sampler to allow for exact Bayesian analysis of any anisotropic universe models that predicts a sparse signal harmonic space covariance matrix. This generalization involved incorporation of a sparse matrix library into the existing Gibbs sampling code called ``Commander'', and implementation of a new sampling algorithm for the anisotropy parameters given a sky map, $P(\\theta|\\mathbf{s})$. We then considered a special case of anisotropic universe models, namely the \\citet{ackerman:2007} model which generalizes the primordial power spectrum $P(k)$ to include a dependence on direction, $P(\\mathbf{k})$. Explicit expressions for the resulting covariance matrix is provided in their paper. We implemented support for this model in our codes, and demonstrated and validated the new tools with appropriate simulations. First, we compared the results from the Gibbs sampler with brute-force likelihood evaluations, and then verified that the input parameters were faithfully reproduced in realistic WMAP simulations. Finally, we analyzed the five-year WMAP temperature sky maps, and presented for the first time the WMAP posteriors of the ACW model. The results from this analysis are highly intriguing, but we emphasize that the effect of instrumental systematics, particularly in the form of correlated noise, must be better understood before the findings can be given a cosmologically interpretation. Taken at face value, we find a preferred direction in the W-band WMAP temperature data pointing towards $(l,b) = (110^{\\circ}, 10^{\\circ})$ (Galactic longitude and latitude), with an anisotropy amplitude of $g_* = 0.15\\pm 0.039$, formally corresponding to a $3.8\\sigma$ detection of $g_* \\ne 0$. Similar results for $g_*$ are found for the V-band data, although with a somewhat lower significance ($g_* = 0.10 \\pm 0.04$; $2.5\\sigma$). The preferred direction is very stable with respect to both data set and multipole range. Figure \\ref{figanisotropicmap} illustrates the underlying anisotropic contribution for a simulation with parameters corresponding to the W band posterior. We have not been able to identify a plausible explanation for this effect in terms of known systematics. First, foregrounds do not appear to have much impact on the results, as consistent results are obtained both from foreground-corrected and raw maps. Second, although correlated noise does lead to a signature similar to the ACW model, its amplitude appears too low in the 5-year data. The least well constrained possibility is that of asymmetric beams, for which we lack proper simulations. \\begin{figure*} \\includegraphics[width=170mm]{f12.eps} \\caption{A simulated realization drawn from a Gaussian distribution with zero mean and a covariance matrix given by the anisotropic $\\Delta$ term in the ACW model, computed for an asymmetry amplitude of $g_* = 0.14$ and a preferred direction $(l,b) = (110^{\\circ}, 10^{\\circ})$, marked by red dots. Notice the rotational structure about the preferred direction. The amplitude of the anisotropic component is $\\sim\\pm 15\\mu\\textrm{K}$, or $\\sim3$\\% of the isotropic component.} \\label{figanisotropicmap} \\end{figure*} While this particular signature certainly is highly intriguing, we would like to point out that the main purpose of this paper is the demonstration of a general framework for analyzing anisotropic signal models. This is useful both for studying particular universe models (e.g., the ACW model), but also for understanding systematic effects (e.g., correlated noise) in a given data set. We therefore believe that these methods may be useful in a wide range of applications, only some of which have been demonstrated in this paper." }, "0807/0807.2060_arXiv.txt": { "abstract": "We report on a systematic investigation of the \\hb\\ and \\feii\\ emission lines in a sample of 568 quasars within $z < 0.8$ selected from the Sloan Digital Sky Survey. The conventional broad \\hb\\ emission line can be decomposed into two components---one with intermediate velocity width and another with very broad width. The velocity shift and equivalent width of the intermediate-width component do not correlate with those of the very broad component of \\hb, but its velocity shift and width do resemble \\feii. Moreover, the width of the very broad component is roughly 2.5 times that of the intermediate-width component. These characteristics strongly suggest the existence of an intermediate-line region, whose kinematics seem to be dominated by infall, located at the outer portion of the broad-line region. ", "introduction": "The geometry and kinematics of the broad-line region (BLR) in active galactic nuclei (AGNs) have been studied for about three decades but the details are far from well understood. It is widely accepted that the BLR is stratified: high-ionization lines originate from small radii and low-ionization lines arise further out \\citep{collin88}. This stratification picture is supported by the results of reverberation mapping, which show that lines of different ionization have different lags (e.g., \\citealt{peterson99}). The dependence of the systemic velocities of the emission lines on ionization \\citep[e.g.,][]{gaskell82,sulentic00a,richards02,shang07} suggests that the BLR may originate from a wind and a disk \\citep[e.g.,][and references therein]{leighly04a,leighly04b}. However, the profiles of the broad emission lines often contain multiple velocity components, suggesting that the structure of the BLR may be more complex than can be described by a simple stratification or wind $+$ disk model. It has been known that the broad \\hb\\ line profiles are generally not well described by a single Gaussian. Two Gaussians \\citep[e.g.,][]{netzer07} or a Gauss-Hermite function \\citep[e.g.,][]{salviander07} are often used. Additionally, the profiles show great diversity from object to object. Sources with narrower \\hb\\ lines tend to have stronger line wings, while those with broader \\hb\\ lines are dominated by the line core \\citep[e.g.,][]{sulentic02}. Some sources have asymmetric and shifted \\hb\\ profiles, suggesting that the BLR has a structure more complex than a single virialized component. The \\hb\\ profile of OQ 208, for example, has an additional redshifted \\hb\\ component of intermediate width that closely resembles the kinematics of \\feii\\ \\citep{marziani93}. Additional evidence that the BLR contains two or more kinematically distinct components comes from differential variability between the line core and wing \\citep[e.g.,][]{ferland90,peterson00}. The existence of an intermediate-line region and a very broad-line region for \\hb\\ emission was suggested by some previous studies \\citep[e.g.,][]{brotherton96,sulentic00b}. In a recent spectral decomposition of a large sample of quasars selected from the Sloan Digital Sky Survey (SDSS), Hu et al. (2008, hereinafter Paper I) find that the majority of quasars show \\feii\\ emission that is both redshifted and narrower than \\hb. Moreover, the magnitude of the \\feii\\ redshift correlates inversely with the Eddington ratio. These characteristics suggest that \\feii\\ originates from an exterior portion of the BLR, whose dynamics may be dominated by infall. These findings offer fresh insights into the structure of the BLR. In light of the trends associated with \\feii\\ emission, we expect that a portion of the \\hb-emitting gas may be related to an inflowing component too. In this Letter, we systematically study the \\hb\\ profiles of SDSS quasars to try to answer two questions: do all quasars have an intermediate-width \\hb\\ component similar to OQ 208, and, if so, is this component also associate with the \\feii\\ emission? We show that the conventional broad \\hb\\ line actually consists of two kinematically linked components, one of which originates from the same region that emits \\feii. ", "conclusions": "\\begin{figure} \\includegraphics[angle=-90,width=0.47\\textwidth]{f3.eps} \\caption{\\footnotesize Correlations between ({\\it a}) \\feii\\ and \\hbi\\ shifts and ({\\it b}) \\feii\\ and \\hbi\\ widths. The solid diagonal lines denote that \\feii\\ and \\hbi\\ have the same shifts and widths. The contours in panel ({\\it a}) show the density of the data points.} \\label{fig-hbife} \\end{figure} \\begin{figure} \\includegraphics[angle=-90,width=0.47\\textwidth]{f4.eps} \\caption{\\footnotesize Plots of ({\\it a}) \\feii\\ vs. \\hbvb\\ shifts and ({\\it b}) \\feii\\ vs. \\hbvb\\ widths. The two lines have different shifts and widths.} \\label{fig-hbvbfe} \\end{figure} We find that \\hbi\\ and \\feii\\ emission have similar kinematics. The similarity can be seen not only in individual sources, but also statistically for the whole sample. From examples {\\it a} to {\\it c} in Figure \\ref{fig-example}, the \\hb\\ profiles change progressively while the \\feii\\ shifts become lower and lower. Figure \\ref{fig-hbife}{\\it a} shows a strong correlation between \\hbi\\ and \\feii\\ shifts. Pearson's correlation coefficient $r_{\\rm P}$ is 0.22, and the probability $P$ of a chance correlation is $< 1\\times10^{-5}$. \\hbi\\ and \\feii\\ have approximately the same shifts. The widths of \\hbi\\ and \\feii\\ are also well correlated and roughly equal (Fig. \\ref{fig-hbife}{\\it b}); $r_{\\rm P}$ = 0.48 and $P < 1\\times10^{-5}$. Except for some sources to the upper left of the solid lines whose errors are large, the majority of sources follow the relation that \\hbi\\ and \\feii\\ have the same shifts and widths. By contrast, \\feii\\ and \\hbvb\\ have different shifts and widths (Fig. \\ref{fig-hbvbfe}). We also find that the kinematic connection between \\hbi\\ and \\hbvb\\ is complicated. No correlation between \\hbi\\ and \\hbvb\\ shifts is seen (Fig. \\ref{fig-hbihbvb}{\\it a}), but the \\hbi\\ and \\hbvb\\ widths are strongly linked, such that FWHM(\\hbvb) $\\approx$ 2.5 FWHM(\\hbi) (Fig. \\ref{fig-hbihbvb}{\\it b}). \\begin{figure} \\includegraphics[angle=-90,width=0.47\\textwidth]{f5.eps} \\caption{\\footnotesize Plots of ({\\it a}) \\hbvb\\ vs. \\hbi\\ shifts and ({\\it b}) \\hbvb\\ vs. \\hbi\\ widths. The two components have different shifts and widths. The solid line in panel ({\\it b}) denotes FWHM(\\hbvb) $= 2.5$ FWHM(\\hbi).} \\label{fig-hbihbvb} \\end{figure} The observational results described above suggest a scenario in which the conventional BLR consists of two components---an intermediate-line region (ILR) and a very broad-line region (VBLR). If both regions are virialized, so that $R \\propto v^{-2}$, then the ILR is about $2.5^2$ or 6.25 times farther from the center than the VBLR. Because of its redshift, the kinematics of the ILR may be dominated by infall. \\hb\\ emission emerges from both the ILR and VBLR, while most of the \\feii\\ emission comes from the ILR. The ILR and VBLR defined in this paper are essentially similar to those described in \\citet{corbin95}, \\citet{brotherton96}, \\citet{sulentic00b}, and \\citet{zhu08} but differ from those in \\citet{brotherton94} or \\citet{sulentic99}, which refer to the \\civ-emitting region. Note that the ILR and VBLR of the \\hb-emitting region are distinct from those of the \\civ-emitting region because the \\civ\\ ILR usually has the systemic redshift while the \\civ\\ VBLR shows a blueshift. \\begin{figure} \\includegraphics[angle=-90,width=0.47\\textwidth]{f6.eps} \\caption{\\footnotesize Plots of ({\\it a}) \\hbvb\\ vs. \\hbi\\ EWs and ({\\it b}) \\feii\\ vs. \\hbi\\ EWs. No correlations are found.} \\label{fig-ew} \\end{figure} The lack of correlation between the EWs of \\hbi\\ and \\hbvb\\ (Fig. \\ref{fig-ew}{\\it a}) strongly suggests that the two components are emitted from different regions. If both are photoionized, they must have different covering factors. The relative strength between \\hbi\\ and \\hbvb\\ determines the final \\hb\\ profiles. In Figure \\ref{fig-hbwfp}, sources broader than $\\sim$5000 \\kms\\ can be fitted well using one Gaussian. This trend can be interpreted in our two-component BLR scenario. It is well known that sources with broad \\hb\\ tend to have weak \\feii/\\hb\\ \\citep[e.g.,][]{bg92,sulentic00a}. The ILR is weak in these systems because \\feii\\ is weak. Their profiles show little deviation from a single Gaussian under the typical S/N level of SDSS spectra. The composite spectra of sources with large \\feii\\ redshifts, on the other hand, do show red asymmetry in the \\hb\\ profiles (see Fig. 13 of Paper I). The variation in the relative strength of the \\hbvb\\ and \\hbi\\ components in different sources reflects the competition between the two components, although they apparently do so in such a manner that their kinematics remained coupled. It is of interest to note that the strengths of \\hbi\\ and \\feii\\ are not correlated (Fig. \\ref{fig-ew}{\\it b}). The wide range of \\feii/\\hbi\\ ratios reflects either the complexity of the excitation mechanism of \\feii\\ emission \\citep[e.g.,][and references therein]{baldwin04} or large variations in quasar metallicities \\citep[e.g.,][]{netzer07}." }, "0807/0807.0854_arXiv.txt": { "abstract": "In this geometrical approach to gravitational lensing theory, we apply the Gauss-Bonnet theorem to the optical metric of a lens, modelled as a static, spherically symmetric, perfect non-relativistic fluid, in the weak deflection limit. We find that the focusing of the light rays emerges here as a topological effect, and we introduce a new method to calculate the deflection angle from the Gaussian curvature of the optical metric. As examples, the Schwarzschild lens, the Plummer sphere and the singular isothermal sphere are discussed within this framework. ", "introduction": "The deflection of light by gravitational fields has been studied with great interest in astrophysics as well as in theoretical physics. Fundamental properties such as Fermat's principle for Lorentzian manifolds, conditions on image multiplicity and caustics in spacetime have been discussed in a fully relativistic setting (e.g., see \\cite{perlick} and references therein). In the astrophysical context, however, an impulse approximation with piecewise straight light rays in flat space has proven useful since deflection angles on cosmological scales are very small (for a comprehensive introduction, see e.g. \\cite{schneider} or \\cite{straumann} pp 272--95 and references therein). Despite their different premisses, both treatments have yielded mathematically interesting, general properties which depend on topology. In particular, image counting theorems like the odd number theorem have been established with different versions of Morse theory both in the spacetime lensing and impulse approximation frameworks \\cite{mckenzie}. In this article, we would like to present another approach to gravitational lensing theory which emphasizes global properties. Specifically, we consider the astrophysically relevant weak deflection limit not in the impulse approximation but treat light rays as spatial geodesics of the optical metric, and use the Gauss-Bonnet theorem. This approach has previously been applied to lensing by cosmic strings \\cite{gibbons}, and we extend it here to static, spherically symmetric bodies of a perfect fluid as simple models for galaxies acting as gravitational lenses. It turns out that the focusing of light rays is, from this point of view, essentially a topological effect. Hence we find, rather surprisingly, that the deflection angle can be calculated by integrating the Gaussian curvature of the optical metric outwards from the light ray, in contrast to the usual description in terms of the mass enclosed within the impact parameter of the light ray. To illustrate, we discuss how this works for three well-known models, namely the Schwarzschild lens, the Plummer sphere and the singular isothermal sphere. The structure of this article is therefore as follows. In section 2 we give a brief review of the Gauss-Bonnet theorem and introduce two constructions which will be used to investigate the lensing geometry. The optical metric and its Gaussian curvature for static, spherically symmetric systems of a perfect fluid is discussed in section 3, followed by the application to the three lens models mentioned above in section 4. With regard to conventions, we use metric signature $(-,+,+,+)$, Latin and Greek indices for space and spacetime coordinates, respectively, and set the speed of light $c=1$. $G$ denotes the gravitational constant as usual. ", "conclusions": "In this article, we have introduced a geometrical approach to gravitational lensing theory different from the spacetime and impulse approximation treatments. By applying the Gauss-Bonnet theorem to the optical metric, whose geodesics are the spatial light rays, we found that the focusing of light rays can be regarded as a topological effect. We also gave a new expression (\\ref{gb3}) to calculate the deflection angle by integrating the Gaussian curvature of the optical metric outwards from the light ray, in the weak deflection limit. The lens models considered were given by static, spherically symmetric bodies of a non-relativistic, perfect fluid, and we discussed as examples the Schwarzschild lens, the Plummer sphere and the singular isothermal sphere. It would therefore be interesting to see whether this approach could also be extended and be fruitfully applied to lenses without spherical symmetry, where images are no longer collinear with the lens centre, or to the relativistic strong deflection limit. \\ack MCW would like to thank Claude Warnick for useful discussions, and the Science and Technology Facilities Council, UK, for funding." }, "0807/0807.0123_arXiv.txt": { "abstract": "% We briefly summarize our findings from the unbiased surveys for $z$$\\sim$5 LBGs based on Subaru/Suprime-Cam and follow-up optical spectroscopy. ", "introduction": " ", "conclusions": "" }, "0807/0807.3316_arXiv.txt": { "abstract": "We describe a sample of low-mass Seyfert 2 galaxies selected from the Sloan Digital Sky Survey, having a median absolute magnitude of $M_g = -19.0$ mag. These galaxies are Type 2 counterparts to the Seyfert 1 galaxies with intermediate-mass black holes identified by Greene \\& Ho (2004). Spectra obtained with the \\emph{Echellette Spectrograph and Imager} at the Keck Observatory are used to determine the central stellar velocity dispersions and to examine the emission-line properties. Overall, the stellar velocity dispersions are low ($\\sim40-90$ \\kms), and we find 12 objects having $\\sigmastar < 60$ \\kms, a range where very few Seyfert 2 galaxies were previously known. The sample follows the correlation between stellar velocity dispersion and FWHM([\\ion{O}{3}]) seen in more massive Seyfert galaxies, indicating that the narrow-line FWHM values are largely determined by virial motion of gas in the central regions of the host galaxies, but the [\\ion{O}{3}] emission lines exhibit a higher incidence of redward asymmetries and double-peaked profiles than what is found in typical Seyfert samples. Using estimates of the black hole masses and AGN bolometric luminosities, we find that these galaxies are typically radiating at a high fraction of their Eddington rate, with a median $\\lbol/\\ledd = 0.4$. We identify one galaxy, SDSS J110912.40+612346.7, as a Type 2 analog of the nearby dwarf Seyfert 1 galaxy NGC 4395, with a nearly identical narrow-line spectrum and a dwarf spiral host of only $M_g = -16.8$ mag. The close similarities between these two objects suggest that the obscuring torus of AGN unification models may persist even at the lowest luminosities seen among Seyfert galaxies, below $\\lbol = 10^{41}$ ergs s\\per. Spectropolarimetry observations of four objects do not reveal any evidence for polarized broad-line emission, but SDSS J110912.40+612346.7 has a continuum polarization significantly in excess of the expected Galactic foreground polarization, possibly indicative of scattered light from a hidden nucleus. Forthcoming observations of this sample, including X-ray and mid-infrared spectroscopy, can provide new tests of the obscuring torus model for active galaxies at low luminosities. ", "introduction": "The majority of active galactic nuclei (AGNs) are found in giant galaxies with substantial bulges, and there is a dramatic drop in the AGN fraction for late Hubble types \\citep{hfs97} and for host galaxies with stellar masses below $\\sim10^{10}$ \\msun\\ \\citep{kau03agn}. This trend is consistent with the generally accepted scenario in which black hole growth and bulge growth are closely coupled, as expected from the correlations between black hole mass and bulge mass \\citep{kr95}, and between black hole mass and stellar velocity dispersion \\citep[the \\msigma\\ relation;][]{fm00, geb00}. The low AGN fraction in late-type and low-mass galaxies arises from a combination of factors. Perhaps most important, the black hole occupation fraction in low-mass galaxies is apparently below unity, as demonstrated by the stellar-dynamical non-detections of central black holes in the Local Group galaxies M33 \\citep{geb01, mfj01} and NGC 205 \\citep{val05}. For those late-type galaxies that do contain a central black hole, shallow central gravitational potential wells may lead to a low efficiency for fueling black hole accretion. A low-mass black hole ($\\lesssim10^6$ \\msun), even if radiating at its Eddington rate, cannot produce a very luminous AGN, and such objects can only be readily detected as AGNs if they are fairly nearby. In optical surveys, dust extinction and blending with circumnuclear star-forming regions can further hinder the detection of low-luminosity active nuclei in late-type spirals. If the \\msigma\\ relation continues toward low masses, then a black hole of $10^6$ \\msun\\ would correspond to a host galaxy velocity dispersion of $\\sigmastar\\approx60$ \\kms\\ \\citep{tre02}. Thus, to explore the intermediate-mass regime for black holes, it is particularly interesting to search for AGN host galaxies having $\\sigmastar<60$ \\kms. Such objects are known to exist, but they are rare. Prior to the Sloan Digital Sky Survey (SDSS), there were only two nearby examples of dwarf galaxies known to contain Seyfert 1 nuclei: POX 52, a dE galaxy with $\\sigmastar=36\\pm5$ \\kms\\ \\citep{ksb87, bar04}, and NGC 4395, an Sd-type spiral with a central velocity dispersion of $\\sigmastar\\lesssim30$ \\kms\\ \\citep{fs89, fh03}. The availability of SDSS has made it possible for the first time to search systematically for more examples of AGNs with low-mass black holes and small central velocity dispersions. Greene \\& Ho (2004; hereinafter GH04) carried out the first such survey using Data Release 1 of the SDSS, identifying 19 Seyfert 1 galaxies as likely candidates for having $\\mbh < 10^6$ \\msun. Although host galaxy properties were not considered in the sample selection, the hosts for this sample turned out to be galaxies of relatively low luminosity, on average about 1 mag fainter than \\lstar. \\citet{bgh05} measured stellar velocity dispersions for many of the galaxies in this sample and found that they are also low ($\\sim35-80$ \\kms) and fall close to the local \\msigma\\ relation extrapolated toward lower masses. Most previous searches for low-mass AGNs have concentrated on Type 1 (broad-lined) objects (Greene \\& Ho 2004, 2007; Dong \\etal\\ 2006) because the broad-line widths and AGN continuum luminosity can be used to estimate the black hole mass. However, at low luminosities, most Seyferts are Type 2 objects \\citep{hfs97, hao05b}, and a full determination of the demographics of AGNs with low-mass black holes must take into account the Type 2 population. Furthermore, without the glare of a bright AGN point source, the host galaxies of Seyfert 2s are more easily studied. This can make it possible to examine the central stellar populations and star formation histories of Seyfert 2 host galaxies in a level of detail that would be impossible for Type 1 AGNs \\citep{kau03agn, hec04}. Since searches for broad-lined AGNs require both the flux and the width of a broad emission line (such as \\hal) to be above some detection threshold, surveys for Type 2 AGNs based on measurements of narrow emission lines can potentially probe lower AGN luminosities, and therefore may be sensitive to AGNs in smaller host galaxies, or containing black holes of lower mass than those that can be found in Seyfert 1 surveys. Optical spectroscopic surveys such as the Palomar survey \\citep{hfs97} have previously identified some examples of very low-luminosity Seyfert 2 nuclei in late-type galaxies, such as the Sc galaxy NGC 1058, which has a central velocity dispersion of only $31\\pm6$ \\kms\\ \\citep{bhs02}. Other strategies for finding obscured AGNs, such as mid-infrared searches for high-ionization coronal line emission from obscured active nuclei \\citep{sat07}, have contributed further evidence for black holes in late-type spirals. Since stellar-dynamical searches for low-mass black holes are limited to very nearby galaxies (within a few Mpc at best), AGN surveys remain the best way to examine the demographics of these objects, and to determine the properties of the host galaxies in which low-mass black holes form and grow. This paper describes new observations of a set of Seyfert 2 galaxies selected from Data Release 2 of the Sloan Digital Sky Survey \\citep[SDSS DR2]{aba04} to have host galaxies with relatively low luminosities and small central velocity dispersions. We select galaxies spectroscopically identified as having Seyfert 2 nuclei (rather than LINER or AGN/starburst transition or composite types), in order to focus on objects having nuclear emission-line spectra that are clearly and unambiguously dominated by an accretion-powered AGN. These galaxies are in many respects Type 2 counterparts of the low-mass Seyfert 1 sample found by GH04, although the two samples are selected by very different criteria. Most of the objects in both of these samples are not small or faint enough to be considered dwarf galaxies, but the objects described here are sub-\\lstar\\ galaxies that occupy ranges in mass, luminosity, and central velocity dispersion where very few Type 2 AGNs were known before SDSS. We present measurements of the properties of these galaxies from high-resolution spectra taken at the Keck Observatory. The small-aperture, high-resolution Keck spectra make it possible to confirm the AGN classification of these galaxies, to detect weak emission-line features and search for faint broad-line components, to measure stellar velocity dispersions, and to fully resolve the profiles of the narrow emission lines and measure accurate linewidths. We discuss the properties of these galaxies and the Seyfert 1s from the GH04 sample in the context of unified models of AGNs. We also describe a search for polarized broad-line and continuum emission in four of these low-mass Seyfert 2 galaxies. In this paper, distance-dependent quantities are calculated assuming $H_0 = 72$ km s\\per\\ Mpc\\per, $\\Omega_m$ = 0.3, and $\\Omega_\\Lambda = 0.7$. ", "conclusions": "We have presented an initial, exploratory study of the properties of Seyfert 2 galaxies with sub-\\lstar\\ luminosities selected from the Sloan Digital Sky Survey. Our measurements reveal very low central stellar velocity dispersions (\\sigmastar $<60$ \\kms) in 12 objects; these are among the smallest velocity dispersions found in any AGN host galaxies, and imply that these galaxies contain some of the least massive black holes in any known AGNs. We have also identified one of the very few known examples of a high-excitation Seyfert 2 nucleus in a late-type, dwarf spiral host galaxy similar to NGC 4395. The correlations between [\\ion{O}{3}] linewidth and \\sigmastar\\ established for higher-mass Seyferts continue to hold in this low-mass regime, while the low-mass Seyfert 2 galaxies exhibit an unusually high incidence of peculiarities (such as redward asymmetries and double-peaked profiles) in their forbidden emission lines. Using rough estimates of black hole mass and bolometric luminosity, we find that the median value of $\\lbol/\\ledd$ for this sample is 0.4, indicating that many of these are objects in which the black hole is undergoing a major growth phase. Future work on these objects will include \\emph{XMM-Newton} observations and \\emph{Spitzer} mid-infrared spectroscopy, in order to search for evidence of nuclear obscuration and to test whether the obscuring torus of AGN unified models is present in this low-mass, low-luminosity regime." }, "0807/0807.1914_arXiv.txt": { "abstract": "Absent any indirect tests on the thermal history of the Universe prior to the formation of light nuclear elements, it is legitimate to investigate situations where, before nucleosyntheis, the sound speed of the plasma was larger than $c/\\sqrt{3}$, at most equalling the speed of light $c$. In this plausible extension of the current cosmological paradigm, hereby dubbed Tensor-$\\Lambda$CDM (i.e. T$\\Lambda$CDM) scenario, high-frequency gravitons are copiously produced. Without conflicting with the bounds on the tensor to scalar ratio stemming from the combined analysis of the three standard cosmological data sets (i.e. cosmic microwave background anisotropies, large-scale structure and supenovae), the spectral energy density of the relic gravitons in the T$\\Lambda$CDM scenario can be potentially observable by wide-band interferometers (in their advanced version) operating in a frequency window which ranges between few Hz and few kHz. ", "introduction": " ", "conclusions": "" }, "0807/0807.3911_arXiv.txt": { "abstract": "This paper explores the mapping between the observable properties of a stellar halo in phase- and abundance-space and the parent galaxy's accretion history in terms of the characteristic epoch of accretion and mass and orbits of progenitor objects. The study utilizes a suite of eleven stellar halo models constructed within the context of a standard $\\Lambda$CDM cosmology. The results demonstrate that coordinate-space studies are sensitive to the recent (0-8 Gyears ago) merger histories of galaxies (this timescale corresponds to the last few to tens of percent of mass accretion for a Milky-Way-type galaxy). Specifically, the {\\it frequency, sky coverage} and {\\it fraction of stars} in substructures in the stellar halo as a function of surface brightness are indicators of the importance of recent merging and of the luminosity function of infalling dwarfs. The {\\it morphology} of features serves as a guide to the orbital distribution of those dwarfs. Constraints on the earlier merger history ($>$ 8 Gyears ago) can be gleaned from the abundance patterns in halo stars: within our models, dramatic differences in the dominant epoch of accretion or luminosity function of progenitor objects leave clear signatures in the [$\\alpha$/Fe] and [Fe/H] distributions of the stellar halo --- halos dominated by very early accretion have higher average [$\\alpha$/Fe], while those dominated by high luminosity satellites have higher [Fe/H]. This intuition can be applied to reconstruct much about the merger histories of nearby galaxies from current and future data sets. ", "introduction": "Phase- and abundance-space substructure in the stellar distribution around a galaxy is commonly interpreted as a natural outcome of the process of hierarchical structure formation, where large galaxies are built in part from the accretion of dwarf galaxies. Numerous previous studies have looked at how to understand individual debris features around galaxies in terms of the properties of the progenitor dwarfs \\citep[e.g.][]{johnston98,helmi99a,johnston01,law05,fardal06,warnick08}. More generally, we might ask: to what extent can the merger history of a galaxy be reconstructed from its surrounding substructure?; and what could you learn about the nature of merger histories in our Universe by examining the outskirts of a large sample of galaxies to very low surface brightness? These questions have yet to be addressed beyond using simple analytic estimates for the expected scalings in tidal debris \\citep{bullock01,johnston01}. Motivation for answering these questions in more depth has been amply provided in the last decade by observations which have revealed abundant substructure in the spatial distribution around both the Milky Way and Andromeda galaxies, already seen at a few tens of kpc from their centers, and dominant at 100kpc \\citep[e.g.][and see \\S \\ref{current.sec} for more examples]{ivezic00,newberg02,newberg03,ferguson02,majewski03,belokurov06,ibata07}. The ubiquity of such substructure has become apparent because of dramatic increases in the sky-coverage of halo surveys, the number of tracers (and hence surface brightness to which such surveys are sensitive) and the distances from the parent galaxies which have been probed. In contrast, only slightly more than a decade ago the classical picture of stellar halos around galaxies was that the stars in them were smoothly distributed in phase-space --- a picture informed by observations of the RR Lyrae and globular cluster distribution around the Milky Way using samples of a few dozen to hundreds of objects \\citep[e.g.][]{wetterer96}. In our own Galaxy, even halo stars in the Solar neighborhood have been shown to be clumped once their full phase-space coordinates are known and their orbital distribution is considered \\citep{helmi99b,morrison08}. Satellite accretion is the preferred explanation for apparent phase-space substructure. Debris from the destruction of a satellite will phase-mix along the progenitor's orbit over time to fill the full volume of the original orbit in coordinate space \\citep{johnston98}, while becoming locally colder at each point in velocity-space \\citep[in order to satisfy Louiville's theorem --- see][for a rigorous description]{helmi99a}. The small range in orbital angular momenta and energy is largely conserved during any gradual evolution of the parent galaxy potential, although the average values can evolve \\citep{penarrubia06}. \\citep[See also][for a discussion in a more violent context.]{knebe05,warnick08} Putting these results for individual accretion events together leads to a spectrum of lumps in phase-space, as has been demonstrated in composite studies of halo distributions \\citep{helmi00,bullock01,johnston01,helmi03,bullock05,delucia08}. Similarly, substructure is now becoming apparent in metallicity distributions and abundance patterns of halo stars. So far, these studies have relied on using coordinates in both phase- and abundance-space to identify these structures \\citep[e.g.][]{majewski96,navarro04,helmi06,ibata07,gilbert08}. However, abundance space has the advantage over phase-space of preserving a complete memory of a star's initial conditions which cannot be destroyed by subsequent mixing or scattering. In principle,``chemically tagging'' stars could lead to associations that are not apparent in phase-space alone \\citep{blandhawthorn03}. Indeed, it has already been demonstrated that the distinct merging time and mass scales of stellar halo, satellite and dominant substructure progenitors means that stars in each of these systems should have distinct patterns in $\\alpha$-element/metallicity space \\citep{robertson05,font06a,font06b,font08}, which implies that distinctions between lower-contrast systems could be possible in higher dimensions. These results also suggest that abundance distributions in a stellar halo should reflect its accretion history \\citep[as has already been shown for metallicity distributions:][]{renda05,font06b}. This paper takes a first step towards addressing how to broadly relate substructures apparent today to a galaxy's past by exploring how the observed properties of eleven stellar halo models built entirely from accretion events within the context of a $\\Lambda$CDM universe reflect their accretion histories. Further work will build on the intuition developed here to define statistical measures of substructure that are tuned to be sensitive to the epoch of accretion and mass and orbit type of progenitor satellites. The modeling methods are described in \\S2. In \\S 3 the 1515 simulated accretion events from all eleven model halos are first analyzed individually in order to characterize how the {\\it observed} properties of debris (i.e. spatial and velocity scales, morphology, surface brightness and abundance patterns) are related to the {\\it intrinsic} properties of the progenitor satellite (i.e. accretion time, luminosity and orbit). The intuition developed in \\S 3 is then applied in \\S 4 to understanding how the {\\it observed} properties of stellar halos (i.e. frequency, morphology, surface brightness and stellar populations of non-uniform features) arise from their unique formation histories. These ideas are illustrated by contrasting our eleven ``standard'' (i.e. built within a $\\Lambda$CDM context) halos with four model halos built from accretion histories that have been artificially constrained to be dominated by ancient/recent and high/low luminosity events, as well as two more built from events predominantly on radial/circular orbits. The results are reviewed in \\S5, and subsequently applied to ``observations'' of our model halos in order to demonstrate in principle how such data might be interpreted. In \\S6 we discuss applications of these ideas to both current and future data sets. We summarize our conclusions in \\S7. ", "conclusions": "In this paper we have explored the connection between the merger history of a galaxy and the present day structure of its stellar halo in phase- and abundance-space. We have found that studies of substructure in the surface-density and/or density distribution in stellar halos are sensitive to the recent ($<$ 8 Gyears ago) merging history of a galaxy. These substructures can tell us about the mass, orbit and accretion time of progenitors, while the proportion of mass in the smooth component reflects the importance of early events (whether in the form of direct accretions from minor or major mergers or {\\it in situ} formation from a heated disk or monolithic collapse). For a Milky-Way-type galaxy, built within the context of $\\Lambda$CDM, we expect of order few to tens of percent of the halo to be in the form of substructure, for those substructures to be increasingly dominant at large radii and for the inner halo to be relatively smooth. These expectations are broadly consistent with the current data for the Milky Way and Andromeda galaxies . Our analysis was based on relating the results of a subjective morphological classification of debris morphology from individual accretion events to the properties of the progenitor satellite. As our sample of galaxies and substructures within galaxies grows, the challenging nature of reaching very low surface brightness, and the multiple dependencies of substructure characteristics mean that interpretations of individual features in terms of progenitor properties may not always be unique. Hence, for specific comparisons of data and observations, it makes sense to recast our results in terms of statistical measures of substructure rather than individual interpretations. For example, \\citet{bell08} report on a preliminary analysis of the SDSS halo turnoff-star data which quantifies the level of substructure via the dispersion in counts around a smooth background. The same analysis applied to the model stellar halos presented in this paper suggests a degree of substructure similar to the observations. More generally, these statistics need to be designed to be sensitive to differences in accretion history, armed with the knowledge developed here that: (i) the epoch of galaxy formation sets the percentage of the stellar halo contained in phase-space substructure (and its average [$\\alpha$/Fe]); (ii) the number and mass scale of recently accreted objects set the number, angular scales (and mean metallicities) of substructure; and (iii) the orbit type of progenitors set the morphology of substructure. In contrast to coordinate-space, signatures in abundance-space are not subject to mixing over time so should last indefinitely. This promises a way to look further back time \\citep{blandhawthorn03}. Our own results demonstrate that variations in [$\\alpha$/Fe] and [Fe/H] among the smooth components of different stellar halos reflect variations is the dominant epoch of accretion and masses of progenitor objects respectively. Moving to higher dimensions in abundance space should allow even more precise interpretations." }, "0807/0807.1255_arXiv.txt": { "abstract": "The algorithm of the ensemble pulsar time based on the optimal Wiener filtration method has been constructed. This algorithm allows the separation of the contributions to the post-fit pulsar timing residuals of the atomic clock and pulsar itself. Filters were designed with the use of the cross- and autocovariance functions of the timing residuals. The method has been applied to the timing data of millisecond pulsars PSR B1855+09 and PSR B1937+21 and allowed the filtering out of the atomic scale component from the pulsar data. Direct comparison of the terrestrial time TT(BIPM06) and the ensemble pulsar time PT$_{\\rm ens}$ revealed that fractional instability of TT(BIPM06)--PT$_{\\rm ens}$ is equal to $\\sigma_z=(0.8\\pm 1.9)\\cdot 10^{-15}$. Based on the $\\sigma_z$ statistics of TT(BIPM06)--PT$_{\\rm ens}$ a new limit of the energy density of the gravitational wave background was calculated to be equal to $\\Omega_g h^2 \\sim 3\\cdot 10^{-9}$. ", "introduction": "The discovery of pulsars in 1967 \\cite{Hewish68} showed clearly that their rotational stability allowed them to be used as astronomical clocks. This became even more obvious after discovery of the millisecond pulsar PSR B1937+21 in 1982 \\cite{Backer82}. Now a typical accuracy of measuring the time of arrivals (TOA) of millisecond pulsar pulses comprises a few microseconds and even hundreds of nanoseconds for some pulsars. For the observation interval in the order $10^8$ seconds this accuracy produces a fractional instability of $10^{-15}$, which is comparable to the fractional instability of atomic clocks. Such a high stability cannot but used for time metrology and time keeping. There are several papers considering applicability of stability of pulsar rotation to time scales. The paper \\cite{Guinot88} presents principles of the establishment of TT (Terrestrial Time) with the main conclusion that one cannot rely on the single atomic standard before authorised confirmation and, for pulsar timing, one should use the most accurate realisations of terrestrial time TT(BIPMXX) (Bureau International des Poids et Mesures). The paper of Ilyasov et. al. \\cite{ilyasov89} describes the principles of a pulsar time scale, definition of \"pulsar second\" is presented. Guinot \\& Petit (1991) show that, because of the unknown pulsar period and period derivative, rotation of millisecond pulsars is only useful for investigations of time scale stability a posteriori and with long data spans. The papers \\cite{ilyasov96}, \\cite{kopeikin97}, \\cite{rodin97}, \\cite{ilyasov98} suggest a binary pulsar time (BPT) scale based on the motion of a pulsar in a binary system with theoretical expressions for variations in rotational and binary parameters depending on the observational interval. The main conclusion is that BPT at short intervals is less stable than the conventional pulsar time scale (PT), but at a longer period of observation ($10^2\\div 10^3$ years) the fractional instability of BPT may be as accurate as $10^{-14}$. The paper of Petit \\& Tavella \\cite{petit96} presents an algorithm of a group pulsar time scale and some ideas about the stability of BPT. The paper of Foster \\& Backer \\cite{foster90} presents a polynomial approach for describing clock \\& ephemeris errors and the influence of gravitational waves passing through the Solar system. In this paper the author presents a method of obtaining corrections of the atomic time scale relative to PT from pulsar timing observations. The basic idea of the method was published earlier in the paper \\cite{rodin06}. In Sect.~\\ref{sect2}, the main formulae of pulsar timing are described. Sect.~\\ref{sect3} contains theoretical algorithm of Wiener filtering. Sect.~\\ref{sect4} presents the results of numerical simulation, i.e. recovery of harmonic signal from noisy data by Wiener optimal filter and weighted average algorithm. The latter one is used similarly to the paper of Petit \\& Tavella \\cite{petit96}. Sect.~\\ref{sect5} describes an application of the algorithm to real timing data of pulsars PSR B1855+09 and PSR B1937+21 \\cite{kaspi94}. ", "conclusions": "An algorithm of forming of ensemble pulsar time scale based on the method of the optimal Wiener filtering is presented. The basic idea of the algorithm consists in the use of optimal filter for removal of additive noise from the timing data {\\it before} construction of the weighted average ensemble time scale. Such a filtering approach offers an advantage over the weighted average algorithm since it utilises additional statistical information about common signal in the form of its covariance function or power spectrum. Since timing data are always available relative to a definite time scale, for separation of the pulsar and clock contributions one need to use observations from a few pulsars (minimum two) relative to the same time scale. Such approach allows estimation of the signal covariance function (power spectrum) by averaging all cross-covariance functions or power cross-spectra of the original data. The availability of two scale differences UTC -- TT and UTC -- PT has resulted in the long awaited possibility of comparing the ultimate terrestrial time scale TT and extraterrestrial ensemble pulsar time scale PT of comparable accuracy. The fractional instability of the terrestrial time scale TT relative to PT and their high correlation have demonstrated that PT scale can be successfully used for monitoring the long-term variations of TT." }, "0807/0807.3066_arXiv.txt": { "abstract": "When the universe was about 10 $\\mu $seconds old, a first order cosmological quark - hadron phase transition occurred at a critical temperature of around 200 MeV. In this work, we study the quark-hadron phase transition in the context of brane-world cosmologies, in which our Universe is a three-brane embedded in a five-dimensional bulk, and within an effective model of QCD. We analyze the evolution of the physical quantities, relevant for the physical description of the early universe, namely, the energy density, temperature and scale factor, before, during, and after the phase transition. To study the cosmological dynamics and evolution we use both analytical and numerical methods. In particular, due to the high energy density in the early Universe, we consider in detail the specific brane world model case of neglecting the terms linearly proportional to the energy density with respect to the quadratic terms. A small brane tension and a high value of the dark radiation term tend to decrease the effective temperature of the quark-gluon plasma and of the hadronic fluid, respectively, and to significantly accelerate the transition to a pure hadronic phase. By assuming that the phase transition may be described by an effective nucleation theory, we also consider the case where the Universe evolved through a mixed phase with a small initial supercooling and monotonically growing hadronic bubbles. ", "introduction": "The possibility that our $4D$ universe may be viewed as a brane hypersurface embedded in a higher dimensional bulk space has attracted considerable interest lately. A scenario with an infinite fifth dimension in the presence of a brane can generate a theory of gravity which mimics purely four-dimensional gravity, both with respect to the classical gravitational potential and with respect to gravitational radiation \\cite{RS99a}. The gravitational self-couplings are not significantly modified in this model. This result has been obtained from the study of a single 3-brane embedded in five dimensions, with the $5D$ metric given by $ds^{2}=e^{-f(y)}\\eta _{\\mu \\nu }dx^{\\mu }dx^{\\nu }+dy^{2}$, which can produce a large hierarchy between the scale of particle physics and gravity due to the appearance of the warp factor \\cite{RS99a}. Even if the fifth dimension is uncompactified, standard $4D$ gravity is reproduced on the brane. In contrast to the compactified case, this follows because the near-brane geometry traps the massless graviton. Hence this model allows the presence of large or even infinite non-compact extra dimensions. Our brane is identified to a domain wall in a 5-dimensional anti-de Sitter space-time. The effective gravitational field equations on the brane-world, in which all the matter forces except gravity are confined on the 3-brane in a 5-dimensional space-time with $Z_2$-symmetry have been obtained, by using an elegant geometric approach, in \\cite{SMS00,SSM00}. The correct signature for gravity is provided by the brane with positive tension. If the bulk space-time is exactly anti-de Sitter, then generically the matter on the brane is required to be spatially homogeneous. The electric part of the 5-dimensional Weyl tensor $E_{IJ}$ gives the leading order corrections to the conventional Einstein equations on the brane. This implies a modification of the basic equations describing the cosmological and astrophysical dynamics, which has been extensively considered \\cite{all2}. For reviews of the dynamics and geometry of the brane Universes, as well as for the discussions of the cosmological implications see \\cite{Ma01}. According to standard cosmology, as it expanded and cooled, the early Universe is expected to have undergone a series of symmetry-breaking phase transitions, at which topological defects may have formed. Phase transitions are labelled first or second order, according to whether the position of the vacuum state in field space changes discontinuously or continuously, as the critical temperature is crossed. A first order phase transition proceeds by bubble nucleation and expansion. When at least $(4-n)$ of these bubble collide, for $n=0,1,2$, an $n$-dimensional topological defect may form in the region between them \\cite{Kajantie:1986hq}. Recent lattice QCD calculations for two quark flavors suggest that QCD makes a transition at a temperature of $T_{c}\\sim 150$ MeV \\cite{TaBo07}. This phase transition, which may have occurred in the early Universe, could lead to the formation of relic quark-gluon plasma objects, which still survive today. A first order quark-hadron phase transition in the expanding Universe can be described generically as follows \\cite{Kajantie:1986hq}. As the color deconfined quark-gluon plasma cools below the critical temperature $T_{c}\\approx 150$ MeV, it becomes energetically favorable to form color confined hadrons (primarily pions and a tiny amount of neutrons and protons, due to the conserved net baryon number). However, the new phase does not show up immediately. As is characteristic for a first order phase transition, some supercooling is needed to overcome the energy expense of forming the surface of the bubble and the new hadron phase. When a hadron bubble is nucleated, latent heat is released, and a spherical shock wave expands into the surrounding supercooled quark-gluon plasma. This reheats the plasma to the critical temperature, preventing further nucleation in a region passed by one or more shock fronts. Generally, bubble growth is described by deflagrations, with a shock front preceding the actual transition front. The nucleation stops when the whole Universe has reheated to $T_{c}$. This part of the phase transition passes very fast, in about $0.05$ $\\mu $s, during which the cosmic expansion is totally negligible. After that, the hadron bubbles grow at the expense of the quark phase and eventually percolate or coalesce. The transition ends when all quark-gluon plasma has been converted to hadrons, neglecting possible quark nugget production. The physics of the quark-hadron phase transition, as well as the cosmological implications of this process have been extensively discussed in the framework of general relativistic cosmology in \\cite{IgKaKuLa94} - \\cite{IgSc01} In the context of brane-world scenarios, the Friedmann equation contains deviations to the $4D$ case, which results in an increased expansion in early times. This in general has important effects on the cosmological evolution, and in particular on cosmological phase transitions. First order phase transitions have also been considered, in the framework of the brane-world model, in \\cite{DaVe01}. Due to the effects coming from the extra-dimensions, phase transitions require a higher nucleation rate to complete, and baryogenesis and particle abundances could be suppressed. The evolution of topological defects is also affected, but the increased expansion cannot solve the monopole and domain wall problems. In this work, we consider the quark-hadron phase transition in the brane-world scenario. By assuming that the phase transition is of the first order, we study in detail the evolution of the relevant cosmological parameters (energy density, temperature, scale factor, etc) of the quark-gluon and hadron phases, and the phase transition itself. It is important to emphasize that in the early universe the energy density is high, so that one may neglect the terms linearly proportional to the energy density with respect to the quadratic terms. This is carried out in detail in this work. An important parameter to describe the phase transition is the hadron fraction, whose time evolution describes the conversion process. We also consider the effect of the dark radiation term on the phase transition. The brane world effects (the quadratic corrections to the matter energy-momentum tensor, described by the numerical value of the brane tension, and the dark radiation term) lead to an overall decrease of the temperature of the very early universe, and accelerate the transition to the pure hadronic phase. This paper is organized in the following manner. In Section \\ref{SecII}, we briefly outline, for self-completeness and self-consistency, the field equations in brane-world models and the basics of the quark-hadron phase transition. In the latter, we lay down the equations of state and the relevant physical quantities that are analyzed in the remaining Sections. In Section \\ref{SecIII} we analyze in detail the quark-hadron phase transition. In Section \\ref{SecIV}, we consider bubble nucleation in the brane-world scenario, by assuming that the phase transition may be described by an effective nucleation theory. We discuss and summarize our results in Section \\ref{SecV}. In the present paper we use a system of units so that the speed of light is $c=1$. ", "conclusions": "\\label{SecV} In the context of brane-world scenarios, in the high density cosmological phase the Friedmann equation contains deviations to the $4D$ case, which imposes fundamental phenomenological consequences on the cosmological evolution, and in particular on the cosmological phase transitions. In this work, we have analyze the evolution of the relevant physical quantities, namely, the energy density, temperature and scale factor before, during and after the phase transition. It is important to emphasize that in the early universe the energy density is extremely high, so that one can neglect the terms linearly proportional to the energy density with respect to the quadratic terms. This approximation has been considered in detail throughout this work. In the high density regime the Hubble function is proportional to the energy density of the cosmological matter, which drastically affects the cosmological dynamics of the universe. Moreover, in the early universe the dark radiation term, the projection of the Weyl tensor from the bulk, which appears in form of a radiation-like term in the field equations, may also play an important role. The magnitude of the brane world effects can be characterized by the numerical value of the brane tension $\\lambda $. In the limit $\\lambda \\to \\infty$ we recover standard general relativity \\cite{Ma01}. On the other hand, the study of the quark-hadron phase transition is also very important from an observational point of view, since the inhomogeneities generated at the QCD phase transition might have a noticeable effect on nucleosynthesis \\cite {IgKaKuLa94a}. By fully taking into account the brane world effects we have found that the temperature evolution of the universe in the brane world scenario is different from the idealized standard FRW model. The temperature of the early universe in the quark phase is smaller in the brane world scenario, as one can see from Fig.~\\ref{FIG1}, where the long dashed curve corresponds (approximately) to the general relativistic limit. Hence a small value of the brane tension would significantly reduce the temperature of the quark-gluon plasma, and accelerate the phase transition to the hadronic era. An increase of the dark radiation term for fixed bag constant $B$ and brane tension $\\lambda $ gives the same effect, as one can see from Fig.~\\ref{FIG2}. Once the quark-hadron phase transition starts, the hadron fraction $h$ is again strongly dependent on the brane tension. For small values of $\\lambda $, $h(t)$ is much higher than in the standard general relativity, as one can see from Fig.~\\ref{FIG3}. The increase of the dark radiation on the brane also strongly accelerates the formation of the hadronic phase (see Fig.~\\ref{FIG4}), and decreases the time interval necessary for the transition. A small brane tension and a high energy density of the dark radiation also tend to reduce the temperature of the hadronic fluid. Of course, the temperature evolution also depends upon the relativistic degrees of freedom in the equation of state and upon the equation of state. In addition to this, by assuming that the phase transition may be described by an effective nucleation theory, we have also considered the case where the Universe evolved through a mixed phase with a small initial supercooling, and monotonically growing hadronic bubbles. It was shown that at a certain temperature below the critical value, an enormous amount of hadronic bubbles are nucleated, which grow explosively to transform all the plasma into hadrons, indicating that the small supercooling phase transition is very rapid with respect to the simple first order phase transition. Many details of the QCD phase transition are not yet conclusively understood. Even the order of transition is still a matter of debate. An advance in the understanding of the numerical values of the QCD coupling constants would be very helpful in obtaining accurate cosmological conclusions. Such an advance may also provide a powerful method for testing on a cosmological scale the theoretical predictions of the brane world models and the possible existence of the extra-dimensions." }, "0807/0807.4470_arXiv.txt": { "abstract": "{ During the study of a large set of late-type stellar X-ray sources, we discovered a large fraction of multiple systems.} { In this paper we investigate the orbital elements and kinematic properties of three new spectroscopic triple systems as well as spectral types and astrophysical parameters ($T_{\\rm eff}$, $\\log g$, $v\\sin i$, $\\log N$(Li)) of their components.} { We conducted follow-up optical observations, both photometric and spectroscopic at high resolution, of these systems. We used a synthetic approach and the cross-correlation method to derive most of the stellar parameters.} { We estimated reliable radial velocities and deduced the orbital elements of the inner binaries. The comparison of the observed spectra with synthetic composite ones, obtained as the weighted sum of three spectra of non-active reference stars, allowed us to determine the stellar parameters for each component of these systems. We found all are only composed of main sequence stars.} { These three systems are certainly stable hierarchical triples composed of short-period inner binaries plus a tertiary component in a long-period orbit. From their kinematics and/or Lithium content, these systems result to be fairly young.} ", "introduction": "\\label{sec:Intro} Binary and multiple stars are very important astrophysical laboratories. In particular, spectro-photometric and spectro-astrometric binaries offer the unique opportunity to determine, with a high level of accuracy, the basic stellar parameters (mass, radius, and effective temperature) to study stellar structure and evolution. However, the formation and evolution of binary stars are still debated subjects \\citep[e.g.,][]{ZM2001}. Especially, a still unsolved problem is the formation of close binaries with main sequence components separated by few solar radii that in the proto-stellar phase should have been in contact. In the last years, to answer many of these open questions, relevant observational and theoretical efforts are being done to improve continuously the statistics of binary systems with different periods, mass ratios, etc. \\citep[e.g.,][]{Tok06}. \\begin{table*} \\caption[RasTyc sources]{Main data of the three {\\it RasTyc} sources from the literature.} \\begin{center} \\begin{tabular}{clccccrrcc} \\hline \\hline \\noalign{\\medskip} \\textsl{RasTyc} Name & Name & $\\alpha$ (2000) & $\\delta$ (2000) & V$_T^{\\,\\,\\rm a}$ & $\\pi^{\\,\\,\\rm b}$ & $\\mu_{\\alpha}^{\\,\\,\\rm a}$~~~ & $\\mu_{\\delta}^{\\,\\,\\rm a}$~~~ & X-ray source & Counts \\\\ & & (h m s) & ($\\degr ~\\arcmin ~\\arcsec$) & (mag) & (mas) & \\multicolumn{2}{c}{(mas\\,yr$^{-1}$)} & 1RXS & (ct\\,s$^{-1}$) \\\\ \\noalign{\\medskip} \\hline \\noalign{\\medskip} RasTyc\\,0524+6739 & \\object{BD+67\\,381} & 05 24 53.2 & +67 39 39 & 9.065 & 7.8$\\pm$9.3\t& $-0.5$~~ & 26.7~~ & J052454.0+673939 & 4.21$\\times 10^{-1}$ \\\\ RasTyc\\,1828+3506 & \\object{BD+35\\,3261} & 18 28 50.3 & +35 06 34 & 9.049 & 12.2$\\pm$8.2\t& 12.3~~ & $-3.5$~~ & J182849.7+350637 & 6.24$\\times 10^{-2}$ \\\\ RasTyc\\,2034+8253 & \\object{BD+82\\,622} & 20 34 27.5 & +82 53 35 & 9.730 & ---\t\t& 61.5~~ & 35.5~~ & J203426.2+825334 & 3.75$\\times 10^{-1}$\\\\ \\hline \\noalign{\\medskip} \\end{tabular} \\begin{tabular}{cc} $^{\\rm a}$ $V$ magnitude and proper motions from the TYCHO-2 catalog \\citep{Hog00}; & $^{\\rm b}$ Parallax from the TYCHO-1 catalog \\citep{Hipp} \\\\ \\end{tabular} \\end{center} \\label{Tab:SourcesRasTyc} \\end{table*} \\onlfig{1}{ \\begin{figure*}[!th] \\centering \\includegraphics[width=9.0cm]{Evolution_Ha_RXJ0524.eps} \\includegraphics[width=9.0cm]{Evolution_Li_RXJ0524.eps} \\includegraphics[width=9.0cm]{Evolution_Ha_RXJ1828.eps} \\includegraphics[width=9.0cm]{Evolution_Li_RXJ1828.eps} \\includegraphics[width=9.0cm]{Evolution_Ha_RXJ2034.eps} \\includegraphics[width=9.0cm]{Evolution_Li_RXJ2034.eps} \\caption{\\label{fig:EvolutionSpectraAll} High resolution spectra of RasTyc\\,0524+6739 (top panels), RasTyc\\,1828+3506 (middle panels) and RasTyc\\,2034+8253 (lower panels) acquired with the A{\\sc urelie} spectrograph at the 152-cm telescope of the OHP both in the H$\\alpha$ (left panels) and the Lithium spectral regions (right panels). The laboratory wavelengths of the \\ion{Fe}{i}\\,$\\lambda$6546.2 and the H$\\alpha$ lines as well as those of the \\ion{Li}{i}\\,$\\lambda$6707.8, and the \\ion{Ca}{i}\\,$\\lambda$6717.7 are marked with vertical dashed lines in the H$\\alpha$ and Lithium spectral regions, respectively.} \\end{figure*} } Close binaries containing at least one late-type component, such as RS~CVn and BY~Dra systems, are objects with the strongest magnetic activity (starspots, plages, flares) induced by a dynamo action in the sub-photospheric convection zone. Their strong activity is mainly due to their very fast rotation (spin-orbit synchronization by tidal forces) and to proximity effects. X-ray sky surveys performed in recent years have allowed to identify thousands of active late-type stars in the field and in open clusters. Follow-up observations of the optical counterparts of X-ray sources have led to discover very young stars far from the typical birth sites, i.e. open clusters and stars forming regions \\citep[e.g.,][]{Wichmann03, Zickgraf05, Torres2006, Guillout2008}, as well as to detect several spectroscopic binaries \\citep[e.g.,][]{Wichmann03b, Frasca2006}. The knowledge of the incidence of binaries and multiple systems in X-ray selected samples of active stars is extremely important to study the recent local star formation history. One of the largest ($\\sim14\\,000$ active stars) and most comprehensive set of stellar X-ray sources in the field is the so-called \\textsl{RasTyc} sample, which is the result of the cross-correlation of the ROSAT All-Sky Survey (RASS) with the TYCHO catalog \\citep{Guillout1999}. We began to analyze a representative sub-sample of the \\textsl{RasTyc} population in the northern hemisphere \\citep{Guillout2008} to obtain some reliable statistics about the \\textsl{RasTyc} stellar characteristics. For this purpose, we led campaigns of high-resolution spectroscopic observations, with the E{\\sc lodie} \\'echelle spectrograph at the 193-cm telescope and the A{\\sc urelie} spectrograph at the 152-cm telescope of the {\\it Observatoire de Haute Provence} (OHP). For all the sources, we performed a detailed analysis of the cross-correlation function (CCF) and found that single-lined (SB1), double-lined (SB2), and triple-lined (SB3) spectroscopic systems altogether account for more than 35\\,\\% of the sample. In particular, at least 10 sources are clearly identified as triple systems. Our aim is to determine the orbital and physical parameters of these systems. For this reason, we have monitored these new multiple systems, both photometrically and spectroscopically, with the 91-cm telescope of the {\\it Osservatorio Astrofisico di Catania} (OAC). The majority of the triple systems studied so far are nearby objects and look as visual binaries where one componentis a SB2 system. Here we study three of such newly discovered SB3 late-type systems for which we obtained enough data. A detailed analysis of the entire sample of stellar X-ray sources and of the multiple systems discovered so far is necessary for drawing statistically significant conclusions. Nevertheless, the properties of these three systems can give us some insights into the typical composition of triple systems among stellar X-ray sources. The paper is organized as follows. We summarize briefly the observations and their reduction in Sect.~\\ref{sec:ObsMet}. The determination of radial velocity and physical parameters ($v\\sin i$'s, Lithium abundances, etc.) are shown in Sect.~\\ref{sec:ObsMet} as well. The photometric modulation, the spectral composition, and the age estimate are discussed in Sect.~\\ref{sec:Discussion}. The conclusions are outlined in Sect.~\\ref{sec:Conclusions}. ", "conclusions": "\\label{sec:Conclusions} This paper is devoted to the analysis of three new triple systems discovered in the \\textsl{RasTyc} sample of stellar X-ray sources. Their spectroscopic and photometric data allow us to conclude that they are almost certainly stable hierarchical triple systems composed of short-period inner binaries plus a tertiary component in a long-period orbit. The orbital periods of the inner binaries range from 3.5 to 7.6 days and the orbits are practically circular. From the high-resolution spectra we also found the spectral composition and the astrophysical parameters of the components that turn out to be all G-K main sequence stars. In all cases, the components of the inner binaries have nearly the same masses, spectral types, and luminosities. From their kinematics and/or Lithium content, these systems result to be fairly young. RasTyc\\,2034+8253 is the only system in which the \\ion{Li}{i}\\,$\\lambda$6707.8 line is strong enough to be clearly visible in the spectra of all the three components and suggests an age in the range $100-300$\\,Myr. It is a possible new member of the IC\\,2391 supercluster. For the remaining systems, the membership to young moving groups is rather uncertain. Our spectroscopic survey has revealed that multiple systems represent a large fraction of the \\textsl{RasTyc} sources. However, a detailed analysis is absolutely necessary for drawing statistically significant conclusions. Since \\textsl{RasTyc} objects are relatively nearby, the discovery and the study of new triple systems, such as those presented in the present paper, can contribute to a better understanding of the formation and the evolution of close binaries and multiple systems in the solar neighborhood." }, "0807/0807.0473_arXiv.txt": { "abstract": "NGC 2770 has been the host of three supernovae of Type Ib during the last 10 years, SN 1999eh, SN 2007uy and SN 2008D. SN 2008D attracted special attention due to the serendipitous discovery of an associated X-ray transient. In this paper, we study the properties of NGC 2770 and specifically the three SN sites to investigate whether this galaxy is in any way peculiar to cause a high frequency of SNe Ib. We model the global SED of the galaxy from broadband data and derive a star-formation and SN rate comparable to the values of the Milky Way. We further study the galaxy using longslit spectroscopy covering the major axis and the three SN sites. From the spectroscopic study we find subsolar metallicities for the SN sites, a high extinction and a moderate star-formation rate. In a high resolution spectrum, we also detect diffuse interstellar bands in the line-of-sight towards SN 2008. A comparison of NGC 2770 to the global properties of a galaxy sample with high SN occurance ($\\geq$ 3 SN in the last 100 years) suggests that NGC 2770 is not particularly destined to produce such an enhancement of observed SNe observed. Its properties are also very different from gamma-ray burst host galaxies. Statistical considerations on SN Ib detection rates give a probability of $\\sim$ 1.5\\% to find a galaxy with three Ib SNe detected in 10 years. The high number of rare Ib SNe in this galaxy is therefore likely to be a coincidence rather than special properties of the galaxy itself. NGC 2770 has a small irregular companion, NGC 2770B, which is highly starforming, has a very low mass and one of the lowest metallicities detected in the nearby universe as derived from longslit spectroscopy. In the most metal poor part, we even detect Wolf-Rayet features, against the current models of WR stars which require high metallicities. ", "introduction": "Massive stars end their lives in various ways, as governed by their mass, composition, angular momentum and whether or not they are interacting with a companion star \\citep[e.g.][]{Heger02}. The most massive stars loose their outer layers through winds whose strength strongly depend on the metallicity of the star \\citep[e.g.][and references therein]{Crowther02}. These stripped stars explode as supernovae (SNe) Type Ib (SN Ib) if they have lost their hydrogen envelope or as SNe Type Ic (SN Ic) if they have also lost their He envelope. If the star is also rapidly rotating, it is believed that the star might even produce a Gamma-Ray Burst (GRB) \\citep{MacFadyen99, Woosley06}. SNe Ib/c are much rarer than SNe II, which are produced by stars with masses of 8 to 40 M$_\\odot$, whereas SNe Ib/cs are assumed to require zero age main sequence (ZAMS) masses of $\\gtrsim$ 35 M$_\\odot$ for non-rotating stars \\citep{Woosley02}. GRBs are even less common than SNe Ic which is in line with the general picture that GRB progenitors require some special conditions to produce a GRB. Only for SNe II has it been possible to identify the progenitor star - which turned out to be yellow, red or blue giants with masses of around 8 -- 20 M$_\\odot$, e.g. SN 1987A \\citep{Gilmozzi87}, SN 2002ov \\citep{Li07}, SN 2004A \\citep{Hendry06} and SN 2006gl \\citep{Gal-Yam07}. Furthermore, several detections have been made using preexplosion imaging from the {\\it HST} archive \\citep[e.g.][]{Smartt03}. SNe Ib/c have so far evaded an identification with a progenitor star \\citep{Maund05, Crockett07, Crockett08}. This is also the case for GRBs which most often occur at distances that do not even allow us to resolve their host galaxies. When the progenitor cannot be directly identified, studying the environment can provide important information about what kind of star exploded. Spatially resolved photometric studies of the sites of different types of SNe show that SNe Type II trace the light of their host galaxies while GRBs are more concentrated towards bright regions \\citep{Fruchter06} and that SNe Type Ib, Ic and GRBs appear to be differently distributed within their hosts \\citep{Kelly07}. Spectroscopic investigations show that the broadlined SNe Ic which are not connected to GRBs are found at sites with higher metallicities than those that are connected to GRBs \\citep{Modjaz08a}. A spatially resolved study of the host of one long-duration GRB which was not connected to a SN also showed a low metallicity in comparison to the rest of the host galaxy \\citep{Thoene08}. Long duration GRBs have been found to be associated with broadlined SNe Ic. There are four spectroscopically confirmed cases so far, namely GRB\\,980425 \\citep{Galama98}, GRB\\,030329 \\citep{Hjorth03, Matheson03, Stanek03}, GRB\\,031203 \\citep{Cobb04, Thomson04, Malesani04, Gal-Yam04} and GRB\\,060218 \\citep{Pian06,Sollerman06, Modjaz06, Mirabal06}. For all other nearby long GRBs up to 2006 where a SN could have been observed, additional light from the SN component was found in the late time lightcurves \\cite[e.g.,][]{Zeh04}. GRBs therefore offer a unique opportunity to observe a SN from the very onset of the explosion, as indicated by the prompt emission from the GRB. The connection between GRBs and broadlined SNe Ic, however, had to be revised in 2006 when two long-duration GRBs were found not to show any sign of an associated SN \\citep{Fynbo06, Gehrels06, DellaValle06, Gal-Yam06, Ofek07}. One important supernova that recently caught the attention of the community is SN 2008D which was associated with an X-ray transient (XT) \\citep{Soderberg08}. This supernova was serendipitously detected while the XRT instrument onboard the {\\emph Swift} satellite \\citep{Gehrels04} observed another supernova, SN 2007uy, in the same galaxy. Whether this prompt XT originated from a weak GRB-like event or if it was due to the shock breakout from the star is still under debate \\citep{Soderberg08, Modjaz08b, Malesani08, Xu08}. The X-ray emission, however, is clearly associated with the onset of the explosion, and consequently SN 2008D was one of the earliest observed SNe. The SN spectrum evolved from a smooth spectrum with small undulations characteristic of a high-velocity ejecta into a typical SN Ib \\citep{Soderberg08, Modjaz08b, Malesani08}. Also the host galaxy of this supernova, NGC 2770, has attracted some attention as it produced three SNe within the last 9 years, SNe 1999eh, 2007uy and 2008D. Intriguingly, all three were stripped envelope core-collapse SNe Ib. In this paper, we present a study on the local properties at the SN sites as well as of other regions in NGC 2770 with longslit spectroscopy and compare the host itself with a sample of other supernova producing galaxies. We want to investigate if the occurrence of three recent SNe Ib can be explained by some physical properties of the host galaxy. In Section 2 we present the observations and data reduction of the spectra, Section 3 studies the global properties of NGC 2770 and properties derived from modeling the spectral energy distribution (SED). In Section 4 we analyze the different regions in the host in terms of metallicity, extinction and star formation rate (SFR). Section 5 compares NGC 2770 to a sample of other nearby galaxies with several recent SNe and discuss the probability to find three SNe Ib in a galaxy within nine years. Finally, Section 6 investigates the properties of the companion galaxy NGC 2770B. Throughout the paper we use a cosmology with H$_0$=70 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_\\Lambda$=0.7 and $\\Omega_\\mathrm{m}$=0.3. At z = 0.007 this corresponds to 0.13 kpc per arcsecond and we use a distance of 27 Mpc to NGC 2700. ", "conclusions": "We have investigated the properties of the three SN sites in NGC 2770, the host of three SNe Ib, and the other regions in the host along 4 longslit positions. Previous observations in all wavelengths from UV to radio allow us to fit the SED of NGC 2770 and derive a range of global properties from it. We then set NGC 2770 in context to a sample of galaxies with frequent SN occurrence (3 or more SNe detected) and also compute the probability to detect three SNe Ib in a galaxy within only 10 years as it was the case for NGC 2770. From this analyses, we then conclude the following: \\begin{itemize} \\item NGC 2770 has global properties similar to the MW, even though it has a higher number of SNe observed. Its SFR and SNR are around the average of other nearby spiral galaxies and our sample of frequent SN galaxies. \\item The only outstanding property of NGC 2770 is its high HI mass which indicates a large reservoir for forming stars. \\item The metallicities at the SN sites in NGC 2770 are around 0.5 solar which is similar to the values observed for nearby GRB sites, but lower than for broadline SN Ic sites. \\item Almost half of the galaxies with SNe II and Ib/c only have one type of SN which is likely connected to the age of the dominant stellar population. Galaxies producing SNe Ib/cs have a higher deVaucouleurs number than those producing SN II. \\item SN and GRB hosts seem to be somewhat different in terms of SFRs and masses. \\item The probability to detect 3 SNe Ib in a galaxy is 0.6 to 1.5 \\% assuming 10,000 monitored galaxies. \\end{itemize} It therefore seems to be likely that observing 3 SNe Ib in NGC 2770 was only a chance coincidence. NGC 2770 is by no means a special galaxy that would be predestined to produce only stripped-envelope SNe. In fact its properties are not typical for galaxies with (frequent) SN Ib/c occurrence. However, it might also imply that the local properties at the SN sites are more important, at least in some galaxies, to produce a certain type of SN than its global properties." }, "0807/0807.0229_arXiv.txt": { "abstract": "Measuring the integrated stellar halo light around galaxies is very challenging. The surface brightness of these haloes are expected to be many magnitudes below dark sky and the central brightness of the galaxy. Here I show that in some of the recent literature the effect of very extended Point Spread Function (PSF) tails on the measurements of halo light has been underestimated; especially in the case of edge-on disc galaxies. The detection of a halo along the minor axis of an edge-on galaxy in the Hubble Ultra Deep Field can largely be explained by scattered galaxy light. Similarly, depending on filter and the shape one assumes for the uncertain extended PSF, 20 to 80 per cent of the halo light found along the minor axis of scaled and stacked Sloan Digital Sky Survey (SDSS) edge-on galaxy images can be explained by scattered galaxy light. Scattered light also significantly contributes to the anomalous halo colours of stacked SDSS images. The scattered light fraction decreases when looking in the quadrants away from the minor axis. The remaining excess light is well modelled with a S\\'ersic profile halo with shape parameters based on star count halo detections of nearby galaxies. Even though the contribution from PSF scattered light does not fully remove the need for extended components around these edge-on galaxies, it will be very challenging to make accurate halo light shape and colour measurements from integrated light without very careful PSF measurements and scattered light modelling. ", "introduction": "In recent years we have begun to appreciate the importance of galaxy stellar envelopes as tracers of the hierarchical galaxy formation process. Hierarchical models in a $\\Lambda$CDM context predict that the stellar envelopes around galaxies are created from many disrupted satellites, where size, shape, amount of substructure, and metallicity of the envelope principally depend on the primordial power spectrum, the reionisation epoch, the star formation history of accreted dwarfs, and the total dark matter mass of the host galaxy \\citep[e.g.,][]{BekChi05,BulJoh05,AbaNav06,PurBul07}. The discovery of the Sagittarius Dwarf currently being disrupted by the Milky Way \\citep{Iba94} and the highly structured envelope around M31 \\citep[e.g.,][]{FerIrw02,Iba07} has given much credence to the hierarchical model. While these observations were performed using resolved stars, many measurements of galaxy haloes have been attempted using integrated light \\citep[e.g.,][]{MorBor94,Fry99,WuBur02,ZibWhi04,ZibFer04}. These observations are very difficult, as the halo light is typically at least a factor of 10$^{4}$ below sky level, and therefore careful attention has to be paid to flat fielding and sky subtraction. Here I investigate another effect that has sometimes been underestimated in integrated light studies: the effect of scattered light in extreme PSF tails when examining edge-on galaxies. The effect of convolving a PSF with a spherical light distribution can be fairly well estimated. For instance, with an elliptical galaxy the light in the central region will be dispersed by the generally broader PSF shape. Further out, the light distribution is nearly unaffected as the PSF shape is steeper than the slowly declining galaxy profile. By over-plotting the PSF on the measured light distribution normalised to the same central brightness, one gets a fairly good impression of which radii are affected by light convolution as modelled by the PSF. For edge-on galaxies the procedure is not as simple as the luminosity profile perpendicular to the disc can be steeper than the outer tails of the PSF. Close to the midplane of the galaxy the vertical light distribution will still be modified by a point-source-like convolution; however, farther away we are less dominated by the local light. Instead the light scattered from the whole disc --- not just the central core--- significantly contributes to the measured distribution at large scale heights. To estimate this contribution at more than 2 scale lengths above the disc (about 6--15 scale heights) one should not use a PSF profile normalised to the central surface brightness of the edge-on galaxy, but instead normalised to the total brightness of the galaxy. Calculating the actual scattered light contribution at any point is obviously best determined by convolving a 2D model of the intrinsic light distribution with a full 2D PSF. Another problem arises when studying galaxy haloes from stacked galaxy images, scaled to a common size, in order to reach fainter levels. One can estimate the PSF of the stack by combining appropriate stellar images from each field, scaled by the same factor as the galaxy in the frame. However, for a typical sample selection the distribution of scaling is not symmetric and highly biased toward the scales near the selection limit of the sample. The distribution of scaled PSF images is therefore strongly skewed, meaning that combining the PSFs using median or mean values gives very different results. Assuming that the galaxies are very similar after scaling, no such skewed distribution is present and median and mean combining should give very similar results. \\begin{figure*} \\includegraphics[width=\\linewidth,trim=1.5cm 0.5cm 1.5cm 0.5cm,clip=true]{udf_profs.eps} \\caption{Minor (left) and major (right) axis surface brightness profiles of the edge-on HUDF galaxy. The data points are derived from \\citet{ZibFer04}. The profiles for the different bands are coded in different colours and symbols and are offset in order to avoid confusion, as indicated by the legend. The input galaxy model is shown by black dashed lines for the different passbands. Thin solid lines indicate crosscuts through the $B$ and $z$-band PSFs (arbitrarily normalised), whilst thick lines show the convolved galaxy model using the same colours for the different passbands as used for the data points. \\label{UDFprofs} } \\end{figure*} In this research note I investigate a few cases in the literature where the effects of PSF convolution with edge-on galaxies has been underestimated. I will also show that the colour measurements of the envelopes around galaxies can be strongly affected by scattered light. In particular, in Section~\\ref{HUDF} I will show the effect of PSF convolution on the light distribution around an edge-on galaxy in the Hubble Ultra Deep field \\citep[HUDF;][]{BecSti06} as measured by \\citep{ZibFer04}. In Section~\\ref{SDSS} I estimate the scattered light effect on the stacked SDSS image analysis of \\citet{ZibWhi04}. Finally, in Section~\\ref{Concl} I summarise my conclusions and give a few recommendations. ", "conclusions": "\\label{Concl} Measuring the exact halo light distribution around edge-on disc light from the integrated light distribution is very challenging. In addition to careful flat fielding and removal of fore- and background sources, particular attention has to be paid to the effect of scattered light from the central disc. Specifically, for edge-on galaxies it is not sufficient to compare the galaxy minor axis light profile with the PSF profile; at large heights above the disc scattered light is contributed from the whole galaxy, not just the centre, and a full convolution of a 2D model is needed to assess the scattered light contamination. Basic modelling suggests that the minor axis may not be the optimum place to detect a halo, but for a flattened halo one should look in the quadrants away from the major and minor axis where scattered light contribution from the disc is smaller relative to the halo. I find that the extended component seen along the minor axis of an HUDF edge-on galaxy can almost fully be explained by scattered light. The observations of \\citet{ZibFer04} for this galaxy do show an extended component along the major axis that cannot be explained by scattered light. Similarly, the halo detection around stacked SDSS edge-on galaxies reported by \\citet{ZibWhi04} has a significant scattered light contamination, especially along the minor axis and in the $i$-band. Depending on the assumed best method to create a stacked PSF the contamination along the minor axis in the $i$-band can amount to 80\\% of the light, in the $g$ and $r$-bands 50\\%. The contamination decreases with other PSF assumptions and going to one of the quadrants, but the PSF uncertainties make it hard to derive accurate stellar halo properties. The excess light seen on top of the disk only model is consistent with a S\\'ersic law halo with parameters typical of those found by the GHOSTS survey \\citep[][de Jong et al., in preparation]{deJRad07}. The anomalous minor axis halo colours reported by \\citet{ZibFer04} and \\citet{ZibWhi04} are consistent with originating from scattered light contamination. In particular, the colour gradients reported by \\citet{ZibWhi04} can be modelled by an extended PSF as calculated from the average of the observed scaled SDSS PSFs. There seems to be no need to invoke extreme stellar population to explain these colours as proposed by \\citet{ZacBer06}. The expected contribution from a typical GHOSTS halo is so small in the $i$-band that it has only a small effect on the predicted colour gradient. There have been other reports of anomalous galaxy stellar halo colours. \\citet{LeqFor96,LeqCom98} report thick disk or halo {\\em BVI}-band colours of edge-on galaxy NGC\\,5907 that are only consistent with colours of elliptical galaxies, i.e.\\ an old, metal rich stellar population. In this case the errors are unlikely to be due to scattered light; the thick disk is detected at about a factor 10$^3$ below the central brightness at about 80 arcsec above the midplane, while the PSF is already 10$^7$ below peak brightness at 16 arcsec radius. Not even a factor 10 enhancement of the scattered light effect due to the edge-on disc configuration can make up that difference. Similarly, the red haloes reported around Blue Compact Galaxies \\citep[see e.g., ][and references therein]{ZacBer06} are not caused by scattered light. The galaxies are too large compared to the PSF and the haloes show substructure different from the central galaxy that is unlikely to be due to the substructure in the PSF. Unfortunately, the examples of \\citet{ZibFer04} and \\citet{ZibWhi04} are close to the limit where scattered light is no longer significant and slightly larger objects would have avoided problems. The HUDF does not contain a larger edge-on galaxy that would render scattered light unimportant. However, the SDSS study could now be repeated on a sample of larger galaxies as the area surveyed by SDSS has nearly tripled since the \\citet{ZibWhi04} study. Ideally, such a study would use only a small range of scale sizes to avoid large spatial scaling corrections and uncertainties in the effective PSF. Taking about a two times as large an isophotal selection limit would probably suffice, especially in the $g$ and $r$-bands. While measuring halo properties from integrated light remains plagued with large uncertainties, we do have a technique that allows us to accurately measure the stellar envelopes around galaxies. By performing star counts of resolved stellar populations we can measure equivalent surface brightnesses to very faint limits as spectacularly demonstrated for M31 and M33 \\citep[e.g.,][]{Iba07}. These observations are not affected by scattered light and only minimally by flat fielding errors. The main limitations of this method lie in fore- and background contaminating objects and low number statistics (too few brights stars at very low surface brightnesses/densities). Very few massive, highly inclined galaxies can be studied with ground-based resolution, but HST allows a much larger sample of galaxies to be studied. Indeed, preliminary results from the GHOSTS survey show conclusively that most large galaxies do have very extended stellar envelopes, with envelope size likely depending on galaxy mass and bulge-to-disc ratio \\citep[][de Jong et al., in preparation]{deJRad07}." }, "0807/0807.1533_arXiv.txt": { "abstract": "We describe a new program for determining photometric redshifts, dubbed EAZY. The program is optimized for cases where spectroscopic redshifts are not available, or only available for a biased subset of the galaxies. The code combines features from various existing codes: it can fit linear combinations of templates, it includes optional flux- and redshift-based priors, and its user interface is modeled on the popular HYPERZ code. A novel feature is that the default template set, as well as the default functional forms of the priors, are not based on (usually highly biased) spectroscopic samples, but on semi-analytical models. Furthermore, template mismatch is addressed by a novel rest-frame template error function. This function gives different wavelength regions different weights, and ensures that the formal redshift uncertainties are realistic. We introduce a redshift quality parameter, $Q_z$, that provides a robust estimate of the reliability of the photometric redshift estimate. Despite the fact that EAZY is not \"trained\" on spectroscopic samples, the code (with default parameters) performs very well on existing public datasets. For $K$-selected samples in CDF-South and other deep fields we find a $1\\sigma$ scatter in $\\Delta z/(1+z)$ of 0.034, and we provide updated photometric redshift catalogs for the FIRES, MUSYC, and FIREWORKS surveys. ", "introduction": "Accurate redshifts of distant galaxies are crucial for nearly all of observational cosmology. Whereas extensive spectroscopy with multi-object spectrographs on 8-10m class telescopes has yielded redshifts for thousands, and in some cases tens of thousands, of galaxies \\citep[e.g.][]{steidel:03,deep2,vvds}, these galaxies tend to be relatively bright at optical wavelengths. For galaxies fainter than $R\\sim 25$ we rely almost exclusively on photometric redshifts, derived from fitting template spectra to broad- or medium-band photometry \\citep[e.g.][]{lanzetta:96,wolf:03,franx:03,mobasher:04,drory:05}. This situation is not likely to change, even with the advent of efficient spectrographs with very wide fields \\citep[such as WFMOS;][]{wfmos}, multi-object capabilities in the near-infrared \\citep[e.g. MOIRCS;][]{moircs}, or larger telescopes. The signal-to-noise ratio (S/N) per resolution element in the continuum decreases with spectral resolution as ${\\rm S}/{\\rm N} \\propto R^{-0.5}$ for a given exposure time. Therefore, the required integration time to maintain a given S/N per resolution element increases linearly with the spectral resolution, quite independent of the details of the telescope and instruments. As a typical set of broad band filters corresponds to $R\\sim 5$ and typical faint object spectrographs have $R\\sim 1000$, spectroscopy is about two orders of magnitude more time consuming than photometry for a given telescope size. A notable exception is spectroscopy of emission line objects, which can be extremely efficient at faint magnitudes. The methodology for determining photometric redshifts using the template-fitting approach is essentially straightforward: the photometric data are compared to synthetic photometry for a large range of template spectra and redshifts, and the most likely redshift follows from a statistical analysis of the differences between observed and synthetic data. Several codes exist that perform this task, each employing its own techniques for creating the synthetic photometry and interpreting the residuals in the redshift -- template plane. Popular examples include HYPERZ \\citep{hyperz}, \\textsc{ImpZ} \\citep{impz}, and Le PHARE\\footnote{\\texttt{http://www.oamp.fr/people/arnouts/LE\\_PHARE.html}} (Arnouts \\& Ilbert), which do a straightforward $\\chi^2$ minimization; GREGZ\\footnote{Greg Rudnick did not name his code; the name GREGZ is used for convenience in the present paper.} \\citep{rudnick:01,rudnick:03}, which allows linear combinations of templates and uses Monte Carlo methods to determine the redshift uncertainties; BPZ \\citep{bpz}, which uses Bayesian statistics allowing the use of priors; and ZEBRA \\citep{zebra} and \\texttt{kcorrect} \\citep[][hereafter BR07]{blanton:07}, which include (distinct) iterative template-optimization routines that make use of the extensive spectroscopic databases of the zCOSMOS \\citep{lilly:07} and Sloane Digital Sky Survey \\citep[SDSS; ][]{york:00} projects, respectively. For obvious reasons photometric redshifts benefit from having high quality photometry in many bandpasses and from sampling strong continuum features in the observed wavelength region (such as a Lyman or Balmer break), irrespective of the methodology. However, given a set of objects with good quality photometry, the aspect that is of paramount importance for obtaining reliable photometric redshifts is the selection of the template set (see Feldmann et al.\\ 2006, \\S\\ref{s:template_set}). \\cite{zebra} obtain very good results by iteratively adapting the templates, minimizing the systematic differences between the best fitting templates and the actual galaxy photometry. This approach not only reduces the random uncertainty in the photometric redshifts but can also eliminate systematic effects in certain redshift ranges \\citep[see][]{zebra}. The disadvantage of this optimization is that its effects can only be assessed when a large sample of galaxies with spectroscopic redshifts is available, and when this sample is a random subset of the entire photometric sample. This assumption may be valid in the case of zCOSMOS, but this is generally not the case in studies of galaxy samples which are significantly fainter than the spectroscopic limit. In this paper we describe a new photometric redshift code which was written specifically for samples with incomplete and/or biased spectroscopic information (such as, for example, faint $K$-selected samples). Rather than minimizing the scatter in the relation between photometric and spectroscopic redshift using the spectroscopic sample as a training set, a user-defined template error function is introduced to account for wavelength-dependent template mismatch. The code combines features from various existing programs: the possibility of fitting linear combinations of templates (as done in GREGZ), the use of priors (as first done in BPZ), and a user-friendly interface based on HYPERZ. The default template set and the redshift-magnitude priors are derived from semi-analytical models. These models are, of course, only an approximation of reality, but their ``perfect'' completeness down to very faint magnitudes outweighs their imperfect representation of the real Universe. The outline of this paper is as follows. In \\S\\ref{s:implementation} we describe the implementation of the code, including the optimized template set and redshift priors derived from semi-analytical models and the template error function derived from the GOODS-CDFS photometric catalog. In \\S\\ref{s:application} we test the code on a combined photometric catalog from a variety of deep multi-wavelength surveys and compare the photometric redshifts to spectroscopic redshifts of nearly 2000 galaxies at $0 < z < 4$. In \\S\\ref{s:reliability} we discuss the reliability of the photometric redshift estimates and provide cautionary examples for relying solely on spectroscopic samples to estimate the photometric redshift quality. Finally, in \\S\\ref{s:summary} we summarize the features and performance of the photometric redshift code and discuss future avenues for improvement. ", "conclusions": "" }, "0807/0807.0791_arXiv.txt": { "abstract": "% We review VLBI observations of supernovae over the last quarter century and discuss the prospect of imaging future supernovae with space VLBI in the context of VSOP-2. From thousands of discovered supernovae, most of them at cosmological distances, $\\sim$50 have been detected at radio wavelengths, most of them in relatively nearby galaxies. All of the radio supernovae are Type II or Ib/c, which originate from the explosion of massive progenitor stars. Of these, 12 were observed with VLBI and four of them, SN 1979C, SN 1986J, SN 1993J, and SN 1987A, could be imaged in detail, the former three with VLBI. In addition, supernovae or young supernova remnants were discovered at radio wavelengths in highly dust-obscured galaxies, such as M82, Arp 299, and Arp 220, and some of them could also be imaged in detail. Four of the supernovae so far observed were sufficiently bright to be detectable with VSOP-2. With VSOP-2 the expansion of supernovae can be monitored and investiated with unsurpassed angular resolution, starting as early as the time of the supernova's transition from its opaque to transparent stage. Such studies can reveal, in a movie, the aftermath of a supernova explosion shortly after shock breakout. ", "introduction": "A supernova (SN), the explosion of a star, is one of the most energetic single events in the universe. Thousands of optical SNe are now known but most of them are at cosmological distances. Only $\\sim$50 have been detected at radio wavelengths and 12 observed with VLBI, the most distant at almost 100 Mpc. In addition, several SNe and supernova remnants (SNRs) were discovered at radio wavelengths in dust-obscured galaxies, such as M82, Arp 299, and Arp 220, and observed with VLBI. All of the SNe observed with VLBI are thought to have resulted from the core collapse of a massive progenitor and emit synchrotron radiation generated from the electrons accelerated in the region where the shock front interacts with the circumstellar medium (CSM) left over from the wind of the mass-losing progenitor. In addition, some of the radiation may come from the environment of the stellar corpse, a neutron star or a black hole left over from the explosion. During the early stage of the expansion of a radio SN, the radio lightcurve rises quickly because of the decreasing optical depth due to synchrotron self-absorption or to thermal absorption in the ionized CSM along the line of sight. It reaches its peak, first at high frequencies later at lower frequencies, when the optical depth decreases below unity, and it declines thereafter during the optically thin stage of the SN while the SN is expanding \\citep[e.g.,][]{Chevalier1982a}. For a typical shock front expansion velocity of 10,000 km s$^{-1}$, the angular size of a SN at a distance of 4 Mpc expands at a rate of 1 mas yr$^{-1}$. With global VLBI and a wavelength of 4 cm, an angular resolution of $\\sim$0.6 mas can be obtained, allowing detailed investigations of the expanding SN and the making of a movie. Global VLBI with VSOP-2 promises to increase the angular resolution at this wavelength to 0.2 mas and to a correspondingly higher resolution at shorter wavelengths, enabling us to witness the aftermath of the explosion shortly after shock breakout. ", "conclusions": "" }, "0807/0807.0758_arXiv.txt": { "abstract": "Models of magnetospheric accretion on to classical T Tauri stars often assume that stellar magnetic fields are simple dipoles. Recently published surface magnetograms of BP~Tau and V2129~Oph have shown, however, that their fields are more complex. The magnetic field of V2129~Oph was found to be predominantly octupolar. For BP~Tau the magnetic energy was shared mainly between the dipole and octupole field components, with the dipole component being almost four times as strong as that of V2129~Oph. From the published surface maps of the photospheric magnetic fields we extrapolate the coronal fields of both stars, and compare the resulting field structures with that of a dipole. We consider different models where the disc is truncated at, or well-within, the Keplerian corotation radius. We find that although the structure of the surface magnetic field is particularly complex for both stars, the geometry of the larger scale field, along which accretion is occurring, is somewhat simpler. However, the larger scale field is distorted close to the star by the stronger field regions, with the net effect being that the fractional open flux through the stellar surface is less than would be expected with a dipole magnetic field model. Finally, we estimate the disc truncation radius, assuming that this occurs where the magnetic torque from the stellar magnetosphere is comparable to the viscous torque in the disc. ", "introduction": "Classical T Tauri stars (cTTSs) are young solar analogs which are accreting material from circumstellar discs. Many observations are consistent with the scenario of the stellar field truncating the inner disc and channelling gas onto discrete regions of the stellar surface. The shapes of near-IR spectral energy distributions (e.g. \\citealt{rob07}), and the kinematics of CO lines formed in the disc \\citep*{naj03}, are consistent with the disc having been truncated at a distance of a few stellar radii. Average surface fields of order a kG have been detected on a number of cTTSs \\citep{joh07}, which models suggest will be strong enough to disrupt the inner disc \\citep{kon91}. The detection of inverse P-Cygni profiles, with widths of several hundred kms$^{-1}$, can also be explained by material essentially free-falling along the field lines of the stellar magnetosphere from the location of the inner disc \\citep{edw94}. Furthermore, the excess continuum emission (veiling) in the optical and UV is likely to arise from shocks at the base of accretion funnels, with emission lines with high excitation potentials, e.g. HeI 5876{\\AA }, forming mainly in such regions \\citep{ber01}. The magnetic interaction between the stellar field and the disc may also have important consequences for the formation of planets. Simulations by \\citet{rom06}, and analytic work by \\citet*{lin96} and \\citet{fle08}, suggest that the inner disc hole, cleared by the star-disc interaction, may provide a natural barrier that decreases the rate of inward migration of forming planets. There is also evidence that the large scale magnetosphere may directly disrupt the inner disc, producing warps which in some systems cross the observer's line-of-sight to the star (e.g. AA Tau, \\citealt{bou07}). Indeed 3D MHD simulations have demonstrated that complicated warping effects in the disc truncation region may be common for T Tauri stars \\citep{rom03,rom04}. The star-disc interaction may also explain the slower rotation of cTTSs compared to the typically older weak-line T Tauri stars, whose discs have largely dispersed (see e.g. the review by \\citealt{bou07IAU}). Accretion of material from the inner disc would act to spin-up the central star in the absence of some physical mechanism to remove angular momentum from the system. Various magnetospheric accretion models have been developed to explain this mechanism, which differ in their assumed location of the inner disc, the details of how angular momentum is removed, and how the magnetic field topology controls both accretion and outflows (\\citealt{kon91}; \\citealt{col93}; \\citealt{shu94}; \\citealt*{fer00}; \\citealt*{fer06}; \\citealt*{kuk03}; \\citealt{mat05a,mat05b,mat08a,mat08b}; \\citealt*{lon05}; \\citealt{bes08}). Although observations indicate that the magnetic field topologies of cTTSs are complex (e.g. see the discussion in \\citealt{gre06a}), the majority of accretion models simply assume that the star has a strong dipolar magnetic field. Recently however a few studies have dropped this assumption, with \\citet{gre05,gre06a} and \\citet{jar06} being the first to consider how magnetic fields with a realistic degree of complexity would affect the accretion process. Other field geometries have since been considered by \\citet*{lon07,lon08} who present the results of a 3D MHD simulation of accretion to stars with composite fields that consist of dipole and quadrupole components tilted at various angles relative to the stellar rotation axis. It is only recently, however, that instrumentation has advanced to the stage where it is possible to map the magnetic topologies of classical T Tauri stars. \\citet{don07,don08a} have recently published magnetic surface maps derived from Zeeman-Doppler imaging of the young solar-analogs V2129~Oph and BP~Tau, using data from the ESPaDOnS and NARVAL spectropolarimeters at the Canada-France-Hawai'i Telescope and T\\'elescope Bernard Lyot respectively. Using the pre-main sequence evolutionary model of \\citet*{sie00}, \\citet{don07} determined that V2129~Oph, a high-mass T Tauri star of $1.35\\,{\\rm M}_{\\odot}$, had developed a radiative core. In contrast, BP~Tau, at $0.7\\,{\\rm M}_{\\odot}$, should be completely convective \\citep{don08a}. As the internal structure, and therefore the magnetic field generation process, is different in both V2129~Oph and BP~Tau, it is of particular interest to compare the structure and properties of their magnetic fields, as well as comparing both to a dipole. In \\S2 we summarise the stellar parameters and magnetic field measurements of BP~Tau and V2129~Oph. In \\S3 we describe how three-dimensional coronal fields can be extrapolated from Zeeman-Doppler maps of photospheric magnetic fields, and compare the resulting field topologies with a large scale dipole. In \\S4 and \\S5 we discuss the structure of the open stellar, and accreting, field, while \\S6 contains our conclusions. ", "conclusions": "By extrapolating the fields of two T Tauri stars from observationally derived surface magnetograms we have compared the resulting magnetic field topologies with a simple dipole. We have found that although the surface magnetic fields of both stars are particularly complex, the larger scale field is simpler and more well ordered. This is consistent with previously published spectropolarimetric studies of accreting T Tauri stars, which suggested that bundles of accreting field lines connect to the stellar surface in single polarity regions, and were therefore likely well-ordered (e.g. \\citealt{val04}). The larger scale field of BP~Tau in particular is closely matched to a dipole magnetic field, although V2129~Oph still shows departures from a dipole field even on the largest scales. However, for both stars, the footpoint separation at the stellar surface of the largest scale field lines is greater than for a purely dipolar magnetic field. In other words as the largest scale field lines arrive at the stellar surface, their structure is influenced by the much stronger field regions. The effect of this is that the largest scale field connects to the star at higher latitudes than would be expected when using a dipole magnetic field model. Thus by considering complex magnetic field topologies we find that the fractional open flux through the stellar surface is less than would be expected from dipole field models. The reduction in the fractional open flux is mainly attributable to the larger closed flux for the complex fields, which arises due to the numerous small magnetic loops close to the stellar surface. For both V2129~Oph and BP~Tau there is less total open flux than for a dipole field, however, the difference is small enough (see Fig. \\ref{fluxes}) that there is unlikely to be any significant implications for the amount of angular momentum that can be carried away from such systems by a stellar wind \\citep{mat05b,mat08a,mat08b}. However, more spectropolarimetric data for many different T Tauri stars will be required to confirm this. If the extent of stellar magnetospheres are limited to well within corotation then accretion may proceed along the more complex regions of the stellar surface field. The result of this would be the formation of many low-latitude accretion hotspots \\citep{gre05,gre06a}. There is at least one star which shows evidence for such equatorial spots, GQ~Lup \\citep{bro07}. However, for V2129~Oph and BP~Tau the spectroscopic data of \\citet{don07,don08a} clearly showed that the bulk of accreting material is carried into high latitude regions of the stellar surface. In order to explain such high latitude hotspots, \\citet{don07} and \\citet{jar08} demonstrated that for V2129~Oph the source surface must be set to at least 6.0$\\,{R_{\\ast}}$, while for BP~Tau the stellar magnetosphere must extend to at least 4$\\,{R_{\\ast}}$ \\citep{don08a}. Selecting the location of the source surface sets the structure of the stellar magnetic field, although the inner disc may still be truncated closer to the stellar surface. Using the assumption that inner disc is located at the point where the torque due to viscous processes in the disc is comparable to the magnetic torque due to the stellar magnetosphere we estimated the disc truncation radius using the extrapolated fields. For BP~Tau we found that the inner disc would be truncated close to the radius predicted for a dipole field, whereas for V2129~Oph, the inner disc is likely to be closer to the star. However, with large uncertainties in the assumed mass accretion rate (which influences the torque due to viscous processes in the disc), and with the static field structures considered here, such calculations are only of qualitative value. Further 3D MHD simulations, such as those already performed by, for example \\citet{lon08}, are required in order to model how magnetic fields with an observed degree of complexity can disrupt and influence the structure of planet forming discs, and will represent a major future development in this field. Although the radius of the source surface is essentially a free parameter of our model, the observed hotspot locations provide constraints. Once the source surface has been selected, however, this then also sets the structure of the X-ray emitting surface field, allowing predictions to be made regarding the stellar X-ray emission properties \\citep{jar08}. Such predictions will allow direct testing of our field extrapolation model with future independently obtained X-ray observations. A further test of the model will be the simulation of accretion related emission line profiles, and are currently being undertaken. V2129~Oph, despite its young age, is massive enough to have developed a radiative core. In contrast to this BP~Tau is likely to be completely convective. Although the field of BP~Tau is more complex than a dipole, it is much simpler than that of V2129~Oph. The differences in the field structure of both stars likely reflect the different internal structure and the process by which their magnetic fields are generated and maintained. For stars with radiative cores the magnetic field is likely generated in the shear layer between the core and the outer convective envelope. In the absence of such a shear layer, it is difficult to explain how fully convective stars can generate and maintain large-scale almost axisymmetric kilo-Gauss fields, although recent theoretical models suggest that this is possible \\citep{bro08}. With only three magnetic surface maps available to date, V2129~Oph and two of BP~Tau \\citep{don07,don08a}, it is not possible to determine whether this is a general feature of all T Tauri stars, however it is similar to what it found for low mass main sequence stars \\citep{don08b}. More magnetic surface maps of T Tauri stars with varying stellar parameters are now required to test these ideas fully, and are an essential requirement to advance our understanding of stellar magnetism on the pre-main sequence." }, "0807/0807.2088_arXiv.txt": { "abstract": "We present a new survey for pulsars in the error boxes of the low-latitude EGRET sources 3EG J1027$-$5817, 3EG J1800$-$2338 and 3EG J1810$-$1032. Although all of these sources have been covered by previous pulsar surveys, the recent discovery of the young, energetic pulsar PSR~J1410$-$6132 at 6.7~GHz has shown that pulsars of this type can be hidden from low frequency surveys. Using an observing frequency of 3.1~GHz we discovered a 91-ms pulsar, PSR J1028$-$5819, which observations made at the Parkes telescope and the Australia Telescope Compact Array have shown to be young and energetic. We believe this pulsar is likely to be powering the unidentified EGRET source 3EG J1027$-$5817. Like other energetic pulsars, PSR J1028$-$5819 is highly linearly polarised, but astonishingly has a pulse duty cycle of only 0.4\\%, one of the smallest in the entire pulsar catalogue. ", "introduction": " ", "conclusions": "" }, "0807/0807.0334_arXiv.txt": { "abstract": " ", "introduction": "\\noindent Evolution of galaxies is directed both by external and internal factors. Internal factors include mainly dynamical instabilities and depend on mass and angular momentum of a galaxy. External factors are determined by the galaxy environments -- other galaxies, intergalactic medium -- and can be either gravitational or hydrodynamical mechanisms. Galactic groups are the best place to study the external factors of galaxy evolution. Many of them have X-ray halo of hot gas; this high-pressure intergalactic medium can strip the own gas reservoirs of spiral galaxies and transform them into lenticulars. Gravitational interactions are provided by the tight neighborhood. Galaxy velocity dispersions inside the groups are not so large, of order of the intrinsic stellar velocity dispersions of early-type galaxies, and cannot prevent development of tidal effects. Numerical simulations proved that tidal interactions affect not only external parts of galaxies but also stimulate bar formation in the centers which in turn re-distribute matter along the radius. Secular evolution can even change the morphological type of a galaxy by causing a growth of a bulge by resulting in gas inflow and subsequent star formation in the center; a whole class of bulges, named `pseudobulges', may be formed rather recently by such a mechanism (Kormendy and Kennicutt 2004). If several galaxies are settled within similar environments in a center of a group and mutually interact gravitationally, it seems probable that the re-shaping of their structures and formation of new central stellar components should be synchronous. Observationally, it means that the stellar ages would be the same, and the spatial star distributions would be similar. If observations reveal synchronous evolution of the central parts of the group galaxies, it would be an argument in favour of dominance of external mechanisms; if no -- we should put attention to variance of internal conditions in the galaxies under consideration. We have already studied 5 nearby groups with the Multi-Pupil Fiber Spectrograph (MPFS) of the Russian 6m telescope; in each group we have observed central parts of 2--3 centrally located galaxies. In Leo I (Sil'chenko et al. 2003a) and in the groups around NGC~5576 (Sil'chenko et al. 2002) and NGC~3169 (Sil'chenko and Afanasiev 2006) we have found similar properties of the stellar populations in the circumnuclear stellar disks: they all have been formed rather recently, 1--3 Gyr ago; and in 3 dominant galaxies of Leo I they are almost co-spatial. In the Leo Triplet, on the contrary, the stellar ages and kinematics in the centers of NGC~3623 and NGC~3627 were quite different (Afanasiev and Sil'chenko 2005), and we concluded that the galaxies of the Triplet had met recently, not earlier than 1 Gyr ago. The only group with the X-ray halo among those studied by us is Leo II (Afanasiev and Sil'chenko 2007); its two central galaxies, NGC~3607 and NGC~3608, have both old kinematically decoupled stellar structures, but their mean stellar ages are different, 10 and 6 Gyr respectively. In this paper we present results of our study of stellar populations in galaxies of another massive group with X-ray hot gas -- the group of NGC~80. A group of galaxies around the giant lenticular galaxy NGC 80 is a rich and massive one. Ramella et al. (2002) in their catalog of galactic groups mentioned 13 members within 2 magnitudes from the brightest one, Dell'Antonio et al. (1994) estimated the number of galaxies in the group as 21, and Mahdavi et al. (2004), with more data, -- even as 45. The X-ray luminosity of the group, according to Mahdavi et al. (2000), is $\\log L_x = 42.56 \\pm 0.09$ ($L_x$ is expressed in erg/s); it is rather high for a group. The center of the X-ray brightness distribution coincides practically with NGC 80, the most luminous galaxy in the group. According to Mahdavi et al. (2004), the velocity dispersion of the galaxies in the NGC 80 group is 336 km/s and the systemic velocity is $5771 \\pm 48$ km/s; Dell'Antonio et al. (1994) rejected some non-member galaxies and after that obtained 246 km/s and $5663 \\pm 51$ km/s, respectively. The central galaxy of the group, the S0 NGC 80, has its own line-of-sight velocity of 5698 km/s which is close to the systemic velocity of the group; it is consistent with NGC 80 being a dynamical center of the group. Interestingly, another giant galaxy, a E0 NGC 83, having the same luminosity as NGC 80 and being located very nearby the group center, has a line-of-sight velocity exceeding the systemic group velocity by more than 500 km/s. One more luminous galaxy in the group is a spiral galaxy NGC 93. Bothun and Schommer (1983) observed NGC 93 at the 21 cm, in the neutral-hidrogen emission line, and they found that the galaxy is very massive, with the rotation velocity of 317 km/s, and `anemic', that means very red and with the low $M_{HI}/L_B$ ratio -- just as massive spiral galaxies of the Virgo cluster. Spiral galaxies in clusters are thought to be `anemic' because of effect of the surrounding hot intergalactic medium -- the most probable cause is the stripping of neutral hydrogen from the outer disks of galaxies by ram pressure. Since the group of NGC 80 is rich by hot gas, NGC 93 may be also stripped in such a manner. We have made integral-field spectroscopic observations of the central parts of the most luminous galaxies in the group, NGC 80, NGC 83, and NGC 93, as well as of E-galaxy NGC 79 and S0-galaxies NGC 86, IC 1541, and IC 1548; the stellar population properties and stellar and gaseous (if the gas is present) kinematics have been studied. The galaxies studied are distributed from the very center (NGC 80 and NGC 83) to the most peripheric parts, at 0.5 Mpc from the group center (IC 1548 and IC 1541). Their global parameters taken from the NED and HYPERLEDA are given in the Table 1. {\\normalsize \\begin{table*} \\caption[ ] {Global parameters of the galaxies studied} \\begin{flushleft} \\begin{tabular}{lccccccc} \\hline\\noalign{\\smallskip} NGC(IC) & 80 & 83 & 79 & 93 & 86 & (1541) & (1548) \\\\ Morph. type (LEDA$^1$) & SA$0-$ & E & E & S? Sab & Sa & S0 & S0 \\\\ $D_{25},\\, ^{\\prime}$ (LEDA) & 1.82 & 1.29 & 0.81 & 1.35 & 0.76 & 0.76 & 0.69 \\\\ $B_T^0$ (LEDA) & 12.98 & 13.22 & 14.66 & 13.66 & 14.35 & 15.15 & 15.20\\\\ $M_B$ (LEDA) & --21.6 & --21.6 & --19.8 & --20.8 & --20.2 & --19.5 & --19.4 \\\\ $(B-V)_e$ (LEDA) & 1.07 & 1.12 & -- & 1.16 & -- & -- & -- \\\\ $(U-B)_e$ (LEDA) & 0.66 & 0.60 & -- & 0.61 & -- & -- & -- \\\\ $V_r$, km/s (NED$^2$) & 5698 & 6227 & 5485 & 5380 & 5591 & 5926 & 5746 \\\\ Distance, Mpc & \\multicolumn{7}{c}{77 (Mahdavi \\& Geller 2004)} \\\\ $PA_{phot}$ (LEDA) & $4.5^{\\circ}$ & $112^{\\circ}$ & $163^{\\circ}$ & $49^{\\circ}$ & $8^{\\circ}$ & $36^{\\circ}$ & $78^{\\circ}$ \\\\ $\\sigma _*$, km/s (LEDA) & 260 & 262 & 201 & -- & -- & -- & 154 \\\\ \\hline \\multicolumn{8}{l}{$^1$\\rule{0pt}{11pt}\\footnotesize NASA/IPAC Extragalactic Database}\\\\ \\multicolumn{8}{l}{$^2$\\rule{0pt}{11pt}\\footnotesize Lyon-Meudon Extragalactic Database} \\end{tabular} \\end{flushleft} \\end{table*} } ", "conclusions": "\\noindent We have studied 7 early-type (E--Sab) galaxies which are members of the massive X-ray galaxy group NGC 80, by means of 2D spectroscopy at the 6m telescope. We have searched for consequences of their synchronous secular evolution. Five of seven galaxies have old bulges, with the mean stellar ages of 10--15 Gyr. Concerning the central galaxy of the group, NGC~80, we have found earlier (Sil'chenko et al. 2003b) that it has circumnuclear structures of intermediate age, namely, the nucleus and the ring with the radius of about 2 kpc, but they are not younger than 5 Gyr. The `anemic' giant spiral galaxy NGC~93, which neutral gas has been probably stripped away partly by ram pressure of the hot intragroup medium at the moment of galaxy infall, has the mean stellar age of the chemically distinct nucleus of 4 Gyr. If to relate its nuclear star formation burst having produced the chemically distinct nucleus with the galaxy infall into the group, we would obtain that the moment of the main group gathering occured about 4--5 Gyr ago. However there are some early-type galaxies in this group that possess rather young stellar populations. The S0-galaxy IC~1548 reveals clear signatures of the recent central star formation burst: the mean stellar age in the bulge is 3 Gyr, and that in the nucleus is 1.5 Gyr. In the same galaxy we have found a circumnuclear polar gaseous ring which warps smoothly into the counterrotating gaseous disk coplanar with the stellar one in more outer parts. Probably, IC~1548 suffered strong tidal interaction with the nearby late-type spiral galaxy and perhaps even accreted gas from this neighbor; once acquiring the gas, IC 1548 had to be provoked to the central star formation by the same tidal interaction. The giant E0 galaxy NGC~83 possesses a compact massive stellar-gaseous disk with the radius of some 2 kpc, very fast rotating, with the noticeable current star formation. Perhaps, all these properties are consequences of merging a spiral or an irregular gas-rich galaxy, so called `minor merger'. It is interesting that the line-of-sight systemic velocity of NGC~83, a giant galaxy in the center of the group, exceeds far the systemic LOS velocity of the group, by more than two group velocity dispersion values. Only a few nearby group members -- NGC 81, NGC 85, PGC 1327 -- have LOS velocities close to that of NGC~83. The majority of the group members have LOS velocities that differ from that of NGC~80 by no more than 200--300 km/s. We may suggest that the poor subgroup of NGC~83 has been accreted by the group of NGC~80 only recently. Was NGC~83 an elliptical galaxy before the accretion event? Did its central gas belong to NGC~83 before the subgroup accretion, or now it is a result of merging? Several variants of answers are possible. \\medskip We thank Dr. A. V. Moiseev of SAO RAS for supporting the observations at the 6m telescope. The 6m telescope is operated under the financial support of Science Ministry of Russia (registration number 01-43). During the data analysis we have used the Lyon-Meudon Extragalactic Database (LEDA) supplied by the LEDA team at the CRAL-Observatoire de Lyon (France) and 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. Our study is supported by the grant of the Russian Foundation for Basic Researches no. 07-02-00229a \\clearpage" }, "0807/0807.2331_arXiv.txt": { "abstract": "We analyse the signatures of brane inflation models with moduli stabilisation. These are hybrid inflation models with a non-trivial field-space metric which can induce complex trajectories for the fields during inflation. This in turn could lead to observable features on the power spectrum of the CMB fluctuations through departures from near scale invariance or the presence of isocurvature modes. We look specifically at multi-brane models in which the volume modulus also evolves. We find that the signatures are highly sensitive to the actual trajectories in field space, but their amplitudes are too small to be observable even for future high precision CMB experiments. ", "introduction": "The early universe gives us the best possibility to test string theory models. Brane-anti-brane \\cite{Dvali:1998pa,Burgess:2001fx,GarciaBellido:2001ky,Jones:2002cv} and modular inflation models \\cite{BlancoPillado:2004ns,Conlon:2005jm} give predictions that are largely compatible with the present cosmological data. Therefore, in order to differentiate between the various models we have to look at their predictions beyond the usual set of experimentally determined parameters (the amplitude of the density perturbations, the spectral index and the running of the spectral index). With the advent of high precision cosmological Cosmic Microwave Background (CMB) and Large Scale Structure (LSS) data the possibility of characterising the initial perturbation spectrum beyond these simplest spectral parameters is now a reality. Efforts to reconstruct the initial perturbation spectrum with model-independent 'inversion' techniques are already under way (see e.g. \\cite{reconstruct}) and have yielded some tantalizing hints of structure in the spectrum. There is a lack of power on the largest scales \\cite{quadrupole} and some indications of oscillations on intermediate scale \\cite{oscillations}. Arguably, the statistical significance of these effects are not sufficiently strong to motivate a departure from the simplest phenomenological models of single-field, slow roll inflation but an accurate analysis of observable features in theoretically driven scenarios is certainly warranted. A generic feature of models of inflation within warped compactification mechanisms is the presence of many extra degrees of freedom in addition to the field driving inflation. Depending on the masses and couplings of the extra fields their presence can induce non-trivial trajectories in the field space. This can result in observable effects such as radically broken scale invariance, presence of isocurvature (or entropy) perturbations and enhanced non-Gaussianity of the perturbations close to points of high acceleration in the trajectories \\cite{ng}. A non-vanishing isocurvature contribution to the total initial perturbations is a particularly interesting possibility since tight limits on any such contribution will be available when CMB polarization measurements becomes precise. The polarization data will also help break fundamental degeneracies in our ability to accurately constrain any broken scale invariance beyond a simple running of the spectral index. In this work we specifically look at multi-brane models \\cite{Cline:2005ty}. These models are very promising since they naturally avoid a number of significant fine-tuning problems present in the simplest brane inflation models. Firstly they naturally produce a larger number of $e$-folds via the assisted inflation mechanism \\cite{Liddle:1998jc}, and secondly, they offer the possibility that the inflaton potential is flattened dynamically. In these models, all branes are initially initially in a local minimum and as branes tunnel through the potential barrier and annihilate with anti-branes the barrier is lowered dynamically until it completely disappears. The potential becomes monotonic and very close to flat once a sufficient number of branes-anti-brane pairs have annihilated and the remaining branes will roll down the potential before annihilating. Quantum fluctuations will cause the rolling branes to start from slightly different initial positions and therefore follow different trajectories. We study the signatures that these models give which may be detected through cosmological observations. All models under consideration feature moduli stabilisation for both the shape \\cite{Giddings:2001yu} and volume \\cite{Kachru:2003aw} moduli and anti-branes needed to lift the anti-deSitter vacuum generated by the stabilising mechanism to a deSitter one during inflation. The bulk contains a warped throat, the anti-branes being located at the bottom of it, and the mobile branes roll down this throat while inflation takes place. Specifically, we follow the evolution of a number of perturbation modes for a set of scalar fields (in this case the positions of the mobile branes) coupled to gravity both inside and outside the Hubble horizon. The non-trivial shape of the potential and K\\\"ahler metric leads to a residual evolution of the scalar fields after horizon crossing. We evolve the perturbations of the scalars coupled to gravity using the Mukhanov-Sasaki \\cite{Mukhanov:1990me,Sasaki} variables and decompose the perturbations into adiabatic and entropy ones. We finally compute the separate spectra for the adiabatic and entropy perturbations and we find that while the adiabatic perturbations are insensitive to the trajectories followed by the many inflaton fields, the entropy ones are highly dependent on the trajectory. However, their amplitude is much too small to lead to observable features in the measured CMB spectrum. The study performed here applies to a larger class of models, namely the ``inflection point'' models \\cite{Baumann:2007np}, where the mass of the inflaton(s) is large, except for a very small region around an inflection point of the potential seen as a function of one inflaton field at a time (keeping all other fields constant). Most of the $e$-folds are coming from this small region of the field space where the inflaton trajectory can be very well approximated by a straight line therefore leading to a very small amplitude for the density perturbations. ", "conclusions": "We analysed the possible signatures that brane-inflation models could give, that may be detectable in observations of the CMB spectrum. To do so we derived the evolution equation for the perturbations of the inflatons and metric and numerically evolved a large number of modes starting inside the horizon, all the way to the end of inflation. We then separated the fluctuations into adiabatic and entropy ones and derived their respective spectra. We found that the adiabatic spectrum is insensitive to the inflaton trajectories and has the has the same shape as was determined in earlier work \\cite{Burgess:2004kv,Cline:2005ty} using the Sasaki-Stewart formalism \\cite{Sasaki:1995aw}. The entropy one is highly sensitive to the inflaton trajectories but its amplitude is too small to be observable. We also offered an analytical argument that in a large class of models where most of the $e$-folds come from a small region around an inflection point of the potential, the inflaton trajectory is almost a straight line and therefore the entropy perturbations are expected to be small. Thus multi-brane models solve a number of fine-tuning problems but due to their $e$-foldings coming from a reduced section of the field trajectory they display less phenomenology than single field models. This argument would lead us to expect that other potentially observable effects such as non-Gaussianities would also have a low amplitude in these models. Typically large values of the non-Gaussian $f_{\\rm NL}$ parameter are produced during periods of high acceleration in the field trajectory which are not present in these models \\cite{ng}. Interestingly, the old models of brane inflation using branes oriented at angles and toroidal compactifications \\cite{Jones:2002cv,Burgess:2001fx,GarciaBellido:2001ky,Shandera:2003gx} have highly non-trivial inflaton trajectories as a result of the brane interaction inside a compact manifold. This leaves open the possibility that more complicated models, like multi-throat models \\cite{Iizuka:2004ct} will feature non-trivial inflaton trajectories and therefore observable features in the power spectrum and non-Gaussianity. As we have shown the models have no particular problem in reproducing the observed tilt in the power spectrum but as expected in models that probe short trajectories, the running is small. If the observed running is confirmed to be at the current best-fit value of $dn_s/d\\ln k \\approx -0.05$ with increasing significance in future measurements it may turn out to be the strongest constraint on these models. As with most multi-field inflation models a detection of a gravitational wave background would also be in conflict with the models since the tensor perturbation are not boosted by the number of fields contributing to inflating the universe." }, "0807/0807.4983_arXiv.txt": { "abstract": "{ Previous surveys for \\HI\\ 21-cm absorption in $z > 0.1$ radio galaxies and quasars yield a $\\approx40$\\% detection rate, which is attributed to unified schemes of active galactic nuclei (AGN). In this paradigm absorption is only witnessed in (close to) type-2 objects, where the central obscuration is viewed (nearly) edge-on and thus absorbs the rest frame 1420 MHz emission along our sight-line. However, we find this mix of detections and non-detections to only apply at low redshift ($z < 1$): From a sensitive survey of eight $z\\gapp3$ radio sources we find no 21-cm absorption, indicating a low abundance of cold neutral gas in (the sight-lines searched in) these objects. Analysing the spectral energy distributions of these sources, we find that our high redshift selection introduces a bias where our sample consists exclusively of quasars with ultra-violet luminosities in excess of $L_{\\rm UV}\\sim10^{23}$ W Hz$^{-1}$. This may suggest that we have selected a class of particularly UV bright type-1 objects. Whatever the cause, it must also be invoked to explain the non-detections in an equal number of $z < 0.7$ sources, where we find, for the first time, the same exclusive non-detections at $L_{\\rm UV}\\gapp10^{23}$ W Hz$^{-1}$. These objects also turn out to be quasars and, from these exclusive high UV luminosity--21-cm non-detections, it is apparent that orientation effects alone cannot account for the mix of 21-cm detections and non-detections at any redshift. ", "introduction": "Observations of the redshifted 21-cm transition of neutral hydrogen (\\HI) provide a powerful probe of the nature and contents of the early Universe. For example, this line can be used to: \\begin{enumerate} \\item Measure the baryonic content of the young Universe, when, before its consumption by star formation, neutral gas outweighed the stars. \\item Probe the evolution of large-scale structure and galaxy formation. At high redshifts, interactions occur more frequently and \\HI\\ observations provide an indispensable means to studying the dynamics of galactic mergence and accretion. \\item Provide a lower limit the time of the epoch of reionisation, when neutral hydrogen collapsed forming the first galaxies and igniting the first stars. \\item Measure any changes in the values of the fundamental constants of nature: Comparing the redshifted frequencies of the 21-cm transition with those of metal-ion and molecular lines against laboratory values, can in principle yield measures of various combinations of fundamental constants at large look-back times (see \\citealt{cdk04}). \\end{enumerate} However, redshifted \\HI\\ 21-cm absorption systems are currently very rare with only 67 known at $z\\gapp0.1$ (see table 1 of \\citealt{cww+08}). We have therefore embarked upon a large survey to search for this transition in both systems intervening the lines-of-sight to background radio sources \\citep{cmp+03,ctp+07} and within the sources themselves (i.e. for absorption ``associated'' with the host, \\citealt{cwm+06,cww+08}). From a recent survey for associated absorption in $z = 2.9 - 3.8$ radio sources with the Giant Metrewave Radio Telescope (GMRT), 21-cm was not detected in any of the eight targets for which good data were obtained. We discuss the reasons for this here. ", "conclusions": "Our observations are discussed in detail in \\citet{cwm+06,cww+08} and in Fig. \\ref{N-z} we show the derived 21-cm line strengths, indicating the upper limits of our and the previous surveys. \\begin{figure*}[t!] \\resizebox{\\hsize}{!}{\\includegraphics[angle=-90,clip=true]{N-z-shrink.eps}} \\caption{\\footnotesize The scaled velocity integrated optical depth of the \\HI\\ line ($1.823\\times10^{18}.\\int \\tau dv$) versus the host redshift for the published $z\\gapp0.1$ searches for associated 21-cm absorption. The filled symbols represent the 21-cm detections and the unfilled symbols the non-detections, with stars designating quasars and circles galaxies. The hatched region shows the range of our recent survey. The detections in this range are from \\cite{ubc91,mcm98} and the two $z_{\\rm em}>5$ non-detections are from \\cite{cwh+07}.} \\label{N-z} \\end{figure*} As seen from the figure, our limits are comparable to the vast majority of detections (which are primarily at low redshift)\\footnote{Note that most of the low redshift non-detections have been searched sufficiently deeply to detect 21-cm in most of the known absorbers.}. Currently, the known mix of detections and non-detections (until our survey, mostly at $z\\lapp1$) are attributed to unified schemes of active galactic nuclei (AGN), where the observed properties of the active galaxy or quasar are due to the orientation at which the active nucleus is viewed (see \\citealt{ant93,up95}): In type-1 objects, the AGN is viewed directly and in type-2 objects through a large column of dense obscuring gas, thus giving rise to the 21-cm absorption by the cold neutral gas located along our line-of-sight (e.g. \\citealt{jm94,cb95}, Fig. \\ref{agn}). \\begin{figure*}[t!] \\begin{center} % \\includegraphics[scale=0.32]{agn3-rot.eps} \\end{center} % \\caption{\\footnotesize Schematic showing the lines-of-sight to an AGN, where according to unified schemes the object type depends upon the orientation at which the nucleus is observed: In type-1 objects we view the AGN directly and in type-2s this is obscured by a circumnuclear torus of dense neutral gas. The 21-cm absorption detection shown is in PKS 1555--140, a known type-2 object, and the non-detection is in PKS 2300--189, a known type-1 object (the spectra are taken from \\citealt{cwm+06}, where the ordinate is the flux density in Jy and the abscissa the observed frequency in GHz). } \\label{agn} \\end{figure*} However, for our sample we obtain exclusive non-detections, where, according to unified schemes, we may expect the $\\approx40$\\% (31 out of 73) mix seen at $z\\lapp1$\\footnote{Perhaps more, since the density of \\HI\\ at $z \\sim 3$ is expected to be higher than it is presently (e.g. \\citealt{psm+01}).}. With two detections already obtained at $z = 2.64$ and $3.40$ (Fig. \\ref{N-z}), it is clear that our high redhsift selection (alone, at least) cannot be responsible for the lack of 21-cm absorption in our targets. If we consider the ultra-violet ($\\lambda\\sim1000$ \\AA) luminosities, however (which we have estimated from the available optical and near-infrared photometry, \\citealt{cww+08}), we see that all our targets have luminosities of $L_{\\rm UV}\\gapp10^{23}$ W Hz$^{-1}$ (Fig. \\ref{lum-z}). \\begin{figure*}[t!] \\resizebox{\\hsize}{!}{\\includegraphics[angle=-90,clip=true]{lum-z-shrink.eps}} \\caption{\\footnotesize As per Fig. \\ref{N-z}, but with the calculated ultra-violet luminosities on the ordinate. } \\label{lum-z} \\end{figure*} Furthermore, although there is a roughly equal mix of detections (33 objects) and non-detections (36 objects) at $L_{\\rm UV}\\lapp10^{23}$ W Hz$^{-1}$, the exclusive non-detections at high UV luminosities relation is also seen for the low redshift sources. We therefore suspect that the high UV fluxes are ionising the neutral gas, rendering this undetectable through the 21-cm transition of hydrogen. This is the first time such a correlation has been noted and, if indeed the case, perhaps exclusive 21-cm non-detections should have been expected for our sample, which at luminosity distances of 26 to 34 Gpc are severely flux limited. That is, at such large distances only the most UV luminous sources are known\\footnote{The quasar frame UV flux being redshifted into the optical band in these cases.}. It should be noted, however, that our main selection criterion was choosing radio-loud sources with $B\\gapp19$. This introduces a luminosity ``ceiling'' of $L_{\\rm UV}\\lapp3\\times10^{24}$ W Hz$^{-1}$, thus causing us to select the {\\it dimmer} ultra-violet sources known at these redshifts. With hindsight, a high redshift selection yielding high luminosity objects undergoing a large degree of ionisation may be expected, although, what is surprising is the discovery of the eight 21-cm non-detections at low redshift (all at $z\\leq0.7$, Fig. \\ref{lum-z}) for which the UV luminosity also exceeds $L_{\\rm UV}\\sim10^{23}$ W Hz$^{-1}$. These objects could be understood in terms of unified schemes (Fig. \\ref{agn}), where we may expect high UV luminosities to go hand-in-hand with 21-cm non-detections (i.e. all are type-1 objects), if it were not for the non-detections at $L_{\\rm UV}\\lapp10^{23}$ W Hz$^{-1}$ (in fact down to $L_{\\rm UV}\\approx4\\times10^{18}$ W Hz$^{-1}$). Again, this is the first time that such a segregation has been noted and suggests that the $L_{\\rm UV}\\gapp10^{23}$ W Hz$^{-1}$ targets (which are all flagged as quasars) are different from their low luminosity counterparts in which 21-cm also remains undetected (which comprise of a mix of galaxies and quasars)." }, "0807/0807.1274_arXiv.txt": { "abstract": "{Recent observations of two black hole candidates (GX 339-4 and J1753.5-0127) in the low-hard state ($L_{\\rm{X}}/L_{\\rm{Edd}} \\simeq 0.003-0.05$) suggest the presence of a cool accretion disk very close to the innermost stable orbit of the black hole. This runs counter to models of the low-hard state in which the cool disk is truncated at a much larger radius. We study the interaction between a moderately truncated disk and a hot inner flow. Ion-bombardment heats the surface of the disk in the overlap region between a two-temperature advection-dominated accretion flow and a standard accretion disk, producing a hot ($kT_{\\rm{e}}\\simeq 70$keV) layer on the surface of the cool disk. The hard X-ray flux from this layer heats the inner parts of the underlying cool disk, producing a soft X-ray excess. Together with interstellar absorption these effects mimic the thermal spectrum from a disk extending to the last stable orbit. The results show that soft excesses in the low-hard state are a natural feature of truncated disk models. ", "introduction": "The geometry of the low-luminosity (``low-hard'') state of Galactic Black Hole Candidates (GBHC), in which the spectrum is dominated by a power law X-ray flux extending to high energies, has been an open question for several decades. While it is generally believed that the power law spectrum is formed by inverse Compton scattering, there is no consensus about the geometry of the flow, source of seed photons or energy distribution for the Comptonizing electrons. Broadly speaking, there are two classes of model to explain the spectrum in the low-hard state. The first is the ``corona'' model, in which the disk remains untruncated or nearly untruncated at luminosities $L_{\\rm{X}} \\simeq 10^{-3} L_{\\rm{Edd}}$. The hard power law spectrum comes from a hot and patchy corona (perhaps powered by magnetic flares \\citep{1998MNRAS.299L..15D, 1999ApJ...510L.123B, 2001MNRAS.321..549M}) on top of the disk, while the surrounding region is bombarded with high energy photons, producing the observed reflection and Fe-K fluorescence components. In the alternate, ``truncated disk'' model the thin disk is truncated at some distance from the black hole and the inner region is filled with a hot, radiatively-inefficient flow, which produces the hard spectrum. The reflection spectrum and Fe-K fluorescence is then produced by the interaction of the hard X-rays with the inner part of the truncated disk, or in some cool outflow moving away from the disk. For a recent discussion of the low-hard state see sect. 4 of \\cite{2007A&ARv..15....1D}. In theory, the presence or absence of a cool disk should be confirmable through direct detection of a soft X-ray blackbody component at low energies. In practice however, this is made difficult by the fact that at low accretion rates the temperature of even an untruncated disk will drop from about 1-2keV in the high soft state to $\\sim$ 0.1-0.3keV, which puts it out of the range of most X-ray detectors. Additionally, the effects of interstellar absorption become very strong at around 0.1keV, so that detecting a soft excess and accurately measuring its parameters will depend somewhat on how accurately the interstellar absorption can be determined. Even with these challenges, a soft excess in the low-hard state has previously been reported in several sources. The first was Cyg X-1 \\citep{1995A&A...302L...5B, 2001ApJ...547.1024D}, although its association with an accretion disk is complicated by the fact that Cyg X-1 is a high mass X-ray binary accreting from a wind. This question was also the focus of two recent papers, \\cite{2006ApJ...652L.113M, 2006ApJ...653..525M}, in which the authors studied long-exposure {\\it{XMM-Newton}} spectra of two different GBHCs, SWIFT J1753.5-0127 and GX 339-4, at low luminosities ($L_{\\rm{X}}/L_{\\rm{Edd}} \\sim 0.003-0.05$). Soft excesses at similar luminosities in these two sources have also been reported in \\cite{2007MNRAS.378..182R} (J1753.5-0127) and \\cite{2008arXiv0802.3357T} (GX 339-4). Since these two observations, there have also been observations of soft excesses in several other sources. \\cite{2007ApJ...666.1129R} made several observations of the soft component of XTE J1817-330 with $Swift$ during the outburst decline of that source down to a luminosity of $L_{\\rm{X}}/L_{\\rm{Edd}} \\sim 0.001$, while a soft component in GRO J1655-40 has been reported by both \\cite{2006MNRAS.365.1203B} and \\cite{2008PASJ...60S..69T} using different telescopes. To interpret the soft excesses in SWIFT J1753.5-0127 and GX 339-4, \\cite{2006ApJ...652L.113M, 2006ApJ...653..525M} fit the data with a several {\\it XSPEC} models, trying various black-body disk models and simple hard X-ray components (both a power law and various Comptonization models). In GX 339-4 a broad Fe-K line was also observed and fit with a relativistically broadened reflection model. Using blackbody models for a standard accreting disk, the authors found disks with maximum temperatures of {\\it{kT}} $\\sim$ 0.2-0.4 keV, and inner radii consistent with the innermost stable circular orbit of a black hole. At the inferred low accretion rates in the hard state, a disk extending to the last stable orbit would produce a soft X-ray component with peak close to the cutoff due to interstellar absorption. Unless an accurate independent measure of the interstellar absorption column is available, spectral fitting procedures cannot reliably distinguish between a thermal peak at $kT = 0.3$ keV with one interstellar absorption column and a cooler component with a lower energy component cutoff by a slightly higher interstellar absorption column. For energetic reasons the hard X-ray component which dominates the luminosity in the hard state must originate near the black hole, the same region as the proposed cool disk. Some form of interaction of hard X-rays with the cool disk must take place, and this implies that the isolated cool disk models used as `components' in fits to observed spectra are unrealistic. In fact, most models for the hard X-ray component include some prescription for the reprocessing of hard into soft radiation, whether these be truncated disks or extended disk models. As shown by Haardt and Maraschi (1991), such models generically produce a similar energy flux in soft and hard X-rays. A strong soft component is thus a natural consequence in truncated as well as extended disk models for the hard state. The main difference in a truncated disk model is that the soft flux originates from a larger surface area and consequently has a lower temperature, putting its spectral peak below 0.5 keV. After interstellar absorption the soft component has a peak around 0.5 keV that can be mistaken for an apparent thermal peak with the temperature of a disk near the last stable orbit. In this paper we examine this with a more quantitative model for truncated disks. At the inner edge of a truncated disk the accretion flow must change in nature from a relatively cool, thin disk into a much hotter, vertically-extended inner flow. There will thus necessarily be some interaction between the two, either through radiation (e.g. \\citet{1991ApJ...380L..51H}) or matter exchange \\citep{1997LNP...487...67S}, or both. Our goal is to determine whether such a model could reproduce the soft spectral components reported by \\cite{2006ApJ...652L.113M,2006ApJ...653..525M}. We will find that disks truncated at 15--20 Schwarzschild radii can in fact produce soft components of the observed strength and shape. An alternative model investigating re-condensation fron am ADAF is considered by \\cite{2008arXiv0807.3402T}. ", "conclusions": "\\label{sec:conclusions} From energetic considerations, the hard spectra observed in the low-hard state of LMXBs must be produced by hot ($kT_{e}\\sim$ 100 keV) matter in the inner regions surrounding the black hole. If there is also a much cooler disk present, there will necessarily be some degree of interaction between the two components, and the disk will be somewhat heated by irradiation from the hot Comptonizing component. The fits reported in \\cite{2006ApJ...652L.113M} and \\cite{2006ApJ...653..525M} neglect this interaction by fitting the disk and hard component separately. In this paper we have shown that incorporating the effects of this interaction heats the inner regions of a moderately truncated disk so that, when coupled with the effects of interstellar absorption, the size of the soft excess matches observations. Our work also highlights the potential pitfalls of using simple power law or analytic Comptonization fits at low energies, which can provide significant deviations in the soft X-rays, thus changing the shape and intensity of the observed soft excess. In the case of GX 339-4, our model predicts an Fe-K component of comparable strength to that observed, although we did not do a detailed comparison. However, work by others has suggested that part of the broadening in the Fe-K line that was observed for GX 339-4 can be attributed to a large outflow, and detailed models of Fe-K fluorescence in galactic black holes show lines that are much broader than is found in AGN models (and which are normally used to fit spectra). The model we have envisioned presents several opportunities for further improvement, in order to better constrain the introduced fitting parameters, $C$ and $\\eta$. The spectrum from the hot ring and ADAF are particularly uncertain, and dependent on a more detailed model for the radiative transfer through this region, as well as the source and number of seed photons (which will set the electron temperature in both regions). The model's global accretion rate (which is limited by the rate at which the hot layer spills over into the hot ring and then evaporates into the ADAF) is also very low, although this can be increased if the viscosity in the warm layer can be increased, perhaps as a result of accretion through an ordered magnetic field. CD'A acknowledges financial support from the National Sciences and Engineering Research Council of Canada." }, "0807/0807.2985_arXiv.txt": { "abstract": "We consider black holes resulting from binary black hole mergers. By fitting to numerical results we construct analytic formulas that predict the mass and spin of the final black hole. Our formulas are valid for arbitrary initial spins and mass ratios and agree well with available numerical simulations. We use our spin formula in the context of two common merger scenarios for supermassive galactic black holes. We consider the case of isotropically distributed initial spin orientations (when no surrounding matter is present) and also the case when matter closely aligns the spins with the orbital angular momentum. The spin magnitude of black holes resulting from successive generations of mergers (with symmetric mass ratio $\\eta$) has a mean of $1.73\\eta + 0.28$ in the isotropic case and $0.94$ for the closely aligned case. ", "introduction": " ", "conclusions": "" }, "0807/0807.2511_arXiv.txt": { "abstract": "We study the ejecta chemistry of a zero-metallicity progenitor, massive, supernova using a novel approach based on chemical kinetics. Species considered span the range of simple, di-atomic molecules such as CO or SiO to more complex species involved in dust nucleation processes. We describe their formation from the gas phase including all possible relevant chemical processes and apply it to the ejecta of a primordial 170 \\Ms~supernova. Two ejecta cases are explored: full mixing of the heavy elements, and a stratified ejecta reflecting the progenitor nucleosynthesis. Penetration of hydrogen from the progenitor envelope is considered. We show that molecules form very efficiently in the ejecta of primordial supernovae whatever the level of mixing and account for 13 to 34\\% of the total progenitor mass, equivalent to 21 to 57 \\Ms~of the ejecta material in molecular form. The chemical nature of molecules depends on mixing of heavy elements and hydrogen in the ejecta. Species produced include O$_2$, CO, CO$_2$, SiS, SO, SiO and H$_2$. Consequently, molecules can be used as observational tracers of supernova mixing after explosion. We conclude that primordial massive supernovae are the first molecule providers to the early universe. ", "introduction": "Large amounts of dust have been conjectured to explain the reddening of background quasars and damped Ly$\\alpha$ systems in the early universe (Pettini et al. 1994, Pei \\& Fall 1995). Possible dust makers could be primordial, very massive stars exploding as supernovae, for the time scales at redshifts $\\ge$ 6 imply short stellar evolution times, thus excluding low-mass, evolved stars. Dust formation in such massive objects has been studied (Nosawa et al. 2003, Schneider et al. 2004) using a classical nucleation theory and excluding details on nucleation processes of dust clusters from the gas phase. However, it is now well accepted that dust forms in other evolved stellar environments under non-equilibirum conditions close to those encountered in laboratory dust condensation experiments where chemical kinetics commands dust nucleation from the gas phase (Donn \\& Nuth 1985, Cherchneff et al. 1992). In supernovae like in other environments, the nucleation will take place via the formation of a molecular phase in the ejecta. Detection of CO, SiO and H$_3^+$ in SN1987A, a core-collapse supernova in the Large Magellanic Cloud (Spyromilio et al. 1988, Roche et al. 1991, Miller et al. 1992), triggered a few theoretical studies on modeling CO and SiO observations (Petuchowski et al. 1989, Lepp et al. 1990, Liu \\& Dalgarno 1994, Gearhart et al. 1999). Those models often used incomplete chemical networks at steady state. No complete chemical description of a primordial, massive supernova ejecta has been so far attempted. We present here the first results of such an endeavor and consider as a surrogate a primordial massive supernova (hereafter PMSN) with zero-metallicity 170 \\Ms~progenitor. We study the formation of chemical species, some of which being dust precursors, in its ejecta. We show that chemistry is usually not at steady state and fosters the formation of complex molecules. Up to one third of the ejecta is found to be in molecular form depending on the level of mixing after explosion. ", "conclusions": "Molecular abundances with respect to the total gas number density are displayed in Figure 2 for the fully-mixed case. Here, we consider two sub-cases: the extreme case where the entire H envelope is microscopically mixed to the He core (hereafter referred as H-rich case), and a case for which only 1\\% of the H envelope mixes with the He-core (i.e. H-poor case). In both cases several molecules do form but differences exist among the two situations. For the H-rich case, the dominant ejected species at 1000 days are H$_2$, O$_2$, SO, CO$_2$, and N$_2$. Their corresponding masses are respectively 19.3 \\Ms, 15.3 \\Ms, 13.0 \\Ms, 8.4 \\Ms, and 1.0 \\Ms, thereby implying a total molecular mass of $\\sim$ 57 \\Ms~equivalent to 34\\% of the PMSN progenitor mass. SiO is abundant up to t = 550 days, but is rapidly depleted due to silica/quartz precursor formation. Inspection of Figure 2 shows that, chemically speaking, the steady state assumption does not hold for molecular formation and destruction at t $\\le$ 1000 days. At early times, the dominant chemical processes at play are neutral-neutral and RA reactions. The OH radical is a key species to molecular formation. Destruction occurs mainly via He$^+$ attack. For t$\\ge$ 600 days, neutral-neutral processes without activation barrier and ion-molecule reactions are dominant. As for dust precursors, (SiO$_2$)$_2$, a ring precursor to silica/quartz nucleation, forms in large amount -- 33.3 \\Ms, as early as 500 days after explosion, followed by AlO, the gas phase precusor to corumdum (Al$_2$O$_3$) -- 0.04~\\Ms~at 530 days, and finally (FeO)$_2$ -- 0.001~\\Ms~at 750 days. For the H-poor case, the dominant molecular species are SiO - 15.8~\\Ms, CO - 4.6~\\Ms~and CO$_2$ - 1~\\Ms, resulting in a total molecular content of 21.5~\\Ms, or 12.6\\% of the progenitor mass. When H is absent, the formation chemistry is powered at early times by RA reactions as opposed to neutral-neutral reactions involving OH. The RA forming SiO has a reaction rate 50 times greater than that for O$_2$ formation, resulting in large amounts of SiO compared to O$_2$. Molecular formation by neutral-neutral processes without activation barrier is postponed to later times, as seen in Figure 2. A smaller molecular content is thus produced due to lower gas temperature and density. As for dust, (SiO$_2$)$_2$ again is the dominant precursor to form at level up to 10.7~\\Ms, while (MgO)$_2$, AlO and (FeO)$_2$ contents are negligible. The above dust nucleation sequences are different from those of existing classical nucleation studies which rely on condensation temperature analysis for solids (Todini \\& Ferrrara 2001, Nozawa et al. 2003, Schneider at al. 2004). Such studies predict the condensation of corundum, forsterite (Mg$_2$SiO$_4$), quartz (SiO$_2$), amorphous carbon and magnetite (Fe$_3$O$_4$). The discrepancy resides in the approach used: in the present model, molecules form from the gas simultaneously to dust precursors, thereby depleting some of the available elements from the gas phase. Dust precursor formation is then commanded by chemical kinetics at play in an altered gas phase whose chemical composition differs drastically from the initial elemental composition. The absence of carbon dust precursors in our fully-mixed case illustrates this point. Indeed, carbon is locked up in CO and CO$_2$ despite the inclusion of Compton electron destruction reactions (Clayton et al. 1999), and is not available for further build-up of C-rich molecules and amorphous carbon precursors. Hence chemistry is acting as a bottleneck to dust nucleation. If we assume that all gas-phase precursors formed in our fully-mixed ejecta are included into dust during condensation, we get an upper limit for freshly formed dust of roughly 33.4~\\Ms~in the H-rich case and 10.7~\\Ms~in the H-poor case, equivalent to $\\sim 19\\%$ and 6\\% of the PMSN progenitor mass, respectively. However, such an assumption overestimates the dust content as dust condensation efficiencies in pyrolysis or flame experiments in the laboratory are usually less than one, with a residual population of dust precursors in the gas phase (J{\\\"a}ger et al. 2006). Furthermore, our H-rich case is somehow extreme and the total H envelope mixing with the He core unlikely, though H mixing was observed in SN1987A and should occur in PISNe as well. Nozawa et al. (2003) and Schneider et al. (2004) find respectively $\\sim$ 33 \\Ms~and $\\sim$ 39 \\Ms~of dust for the fully-mixed He core of a 170 \\Ms~progenitor mass PISN and similar ejecta thermodynamics than that in the present study. These results are to be compared to the 10.7 \\Ms~upper limit derived for our H-poor case. Our two mixed cases highlight trends in dust formation scenarios and content in PMSNe and our limits for freshly formed dust are a few \\% of the progenitor mass only. This is certainly lower than existing predicted dust amounts. For the unmixed ejecta, molecules and their derived mass are listed in Table 2. Again, molecules do form efficiently and account for $\\sim$ 42~\\Ms, equivalent to 25\\% of the PMSN progenitor mass. Their chemical nature now traces their location in the He core and the amount of H mixing from the progenitor envelope. In Zone 1, devoid of oxygen, SiS forms from reaction between atomic Si and S$_2$ and traces a Si/S-rich gas. In Zone 4, we allowed for 18 \\% of total mass hydrogen penetration in the C/He--rich layer as in Nosawa et al. (2003), with subsequent formation of H$_2$, C$_2$H$_2$, and CO. Carbon monoxide is preferably formed over CO$_2$ in this layer as CO$_2$ is rapidly destroyed by reactions with ions like C$^+$ and He$^+$ to re-form CO. Thus, CO presence in the ejecta indicates a C/O ratio greater than 1, a large free carbon atom content, and some mixing with He and H. This result is supported by the simultaneous detection of CO and amorphous carbon dust in SN1987A, implying carbon-rich inhomogeneities in the ejecta. In Zone 2, CO$_2$ dominates over CO, because of the rapid CO conversion to CO$_2$ via reaction with O$_2$, which is very abundant in those regions where the C/O ratio is less than 1. CO$_2$ is thus a tracer of oxygen-rich ejecta regions whereas the molecular composition of Zone 4 is typical of a C/H/He-rich environment. Similar chemical processes and species are encountered in the inner shocked winds of O-rich and C-rich, evolved low-mass stars, where the gas parameters resemble those of SN ejecta (Cherchneff 2006). As for dust molecular precursors, we find that Zone 1 produces $\\sim$ 1.6~\\Ms~of (Si)$_4$ and $\\sim$ 1.3~\\Ms~of (FeS)$_2$, while Zone 2 produces 3~\\Ms~of (SiO$_2$)$_3$. No dust precursors are formed in Zone 3, and Zone 4 produces 0.5~\\Ms~of C$_3$. Again, grain types and nucleation sequences are different from the study of Nosawa et al. (2003). An upper limit to dust production in the unmixed case is $\\sim$ 6.5 \\Ms~or $\\sim$ 3.8$\\%$ of the total PMSN progenitor mass, which is much less than the value of $\\sim$ 18 \\% derived by Nozawa et al. for their 170 \\Ms~PISN unmixed case. This points again to the crucial role of chemical kinetics as bottleneck to dust formation. Chemically speaking, we notice that Zone 4 forms carbon dust precursors via pure carbon chains like C$_3$ despite the large quantities of C$_2$H$_2$ available and the aromatic formation reactions included in our chemical network. Aromatic chemical pathways to carbon dust are inhibited even in the presence of hydrogen because of the high temperatures encountered in the ejecta up to 600 days after explosion, and the competitive RA pathways to carbon chains combined to large amounts of free C atoms. This result may be specific to supernova ejecta and is supported by the non-detection of polycyclic aromatic hydrocarbon infrared emission lines in SN1987A spectra (Wooden et al. 1993). We conclude that Pop III, massive supernovae are efficient molecule providers to the early universe, with 13\\% to 34$\\%$ of their ejecta in molecular form, corresponding to 22 \\Ms~to 57~\\Ms~of molecular material released to the local, pristine gas after explosion. Formed species depend on mixing in the ejecta and include O$_2$, CO$_2$, SO, CO, SiS, OH, C$_2$H$_2$ and H$_2$. H mixing boosts molecular formation at early times via neutral-neutral processes, resulting in a large molecular component in the ejecta. In a more general context, molecules could be used as observational tracers of mixing in nearby core-collapse supernovae. The substantial amounts of molecular material imply some impact on the local gas cooling if cooling time scales are comparable to those for the ejecta adiabatic expansion. It was recently shown that some dust grains could survive the passage of the PMSN reverse shock some 10$^4$ years after explosion (Nozawa et al. 2007, Bianchi \\& Schneider 2007). The survival of molecules in cool, dense, inhomogeneities passing the reverse shock must then be studied. At later times, it is conjectured that Pop.~II.5 stars can form in the PMSN dense shell if other cooling than that of H$_2$ is provided (MacKey et al. 2003, Salvaterra et al. 2004). If surviving the reverse shock, our predicted molecules could provide part of the necessary cooling and their subsequent impact on second generation star formation needs further investigation. Finally, under present chemical kinetic conditions, our results show that PMSNe do form dust efficiently as the amounts of dust precursors produced are significant. However, the present study points to lower dust contents formed in PMSNe than those predicted by existing, classical nucleation studies." }, "0807/0807.4930_arXiv.txt": { "abstract": "We study the phenomenology of hybrid scenarios of neutrino dark energy, where in addition to a so-called Mass Varying Neutrino (MaVaN) sector a cosmological constant (from a false vacuum) is driving the accelerated expansion of the universe today. For general power law potentials we calculate the effective equation of state parameter $w_{eff}(z)$ in terms of the neutrino mass scale. Due to the interaction of the dark energy field (``acceleron'') with the neutrino sector, $w_{eff}(z)$ is predicted to become smaller than $-1$ for $z>0$, which could be tested in future cosmological observations. For the considered scenarios, the neutrino mass scale additionally determines which fraction of the dark energy is dynamical, and which originates from the ``cosmological constant like'' vacuum energy of the false vacuum. On the other hand, the field value of the ``acceleron'' field today as well as the masses of the right-handed neutrinos, which appear in the seesaw-type mechanism for small neutrino masses, are not fixed. This, in principle, allows to realise hybrid scenarios of neutrino dark energy with a ``high-scale'' seesaw where the right-handed neutrino masses are close to the GUT scale. We also comment on how MaVaN Hybrid Scenarios with ``high-scale'' seesaw might help to resolve stability problems of dark energy models with non-relativistic neutrinos. ", "introduction": "The evidence for the existence of a dark sector in the universe has made the present era of cosmology fascinating and challenging. At present, the microscopic natures of dark energy and dark matter are still an open question, with both components only probed gravitationally \\cite{Colless:2001gk,Riess:1998cb,Perlmutter:1998np,Spergel:2006hy}. Regarding dark matter, there are various particle physics candidates which may belong to the class of Weakly Interacting Massive Particles (WIMPS) or which may interact only gravitationally. On the other hand, for particle physics explanations of the observed dark energy, the main fundamental question is whether dark energy is a cosmological constant or a dynamical field. While in the former case the equation of state parameter $w$ of dark energy is constant and equal to $-1$, in the latter case it is a function of redshift $z$ and in general differs from $-1$, providing a way to distinguish both fundamental types of dark energy experimentally. Such a deviation of $w$ from $-1$ will be searched for in various future surveys, studying supernovae, baryonic acoustic oscillations, weak gravitational lensing effects, galaxy clusters, or other techniques \\cite{Albrecht:2006um}. One obstacle to relate dark energy more explicitly to particle physics is the fact that the amount of the observed energy density $\\rho_{DE}^{1/4} \\sim 0.003$ eV is {\\em much} smaller than the ``theoretical expectation''. In this paper we will not address this ``cosmological constant'' problem, i.e.\\ the suppression/cancellation of the various generic contributions to dark energy which are too large, but assume that it is resolved by some other mechanism. Our work is motivated by the intriguing observation that the energy density of dark energy is close to another very small scale in particle physics, namely the one of neutrino masses. The discovery of flavour conversion of neutrinos from various sources, interpreted within the framework of neutrino oscillations, points to two mass eigenvalues of the light neutrinos above about $0.01$ eV and $0.05$ eV \\cite{Mohapatra:2005wg}, while searches for neutrino masses from Tritium $\\beta$-decay and neutrinoless double $\\beta$-decay yield an upper bound for each mass eigenvalue of roughly $0.5$ eV \\cite{Mohapatra:2005wg}. The proximity of the scale of dark energy to that of neutrino masses has inspired the proposal of the so-called Mass Varying Neutrino (MaVaN) scenario \\cite{Fardon:2003eh}. In this scenario, the relic neutrino density contributes to the effective potential for the dynamical field which is called ``acceleron''. One consequence is that neutrino masses vary on cosmological time scales and, since the relation between neutrino masses and $\\rho_{DE}^{1/4}$ remains valid in earlier cosmic epochs, also dark energy behaves differently from a cosmological constant. Compared to other dynamical models of dark energy (e.g.\\ quintessence \\cite{Wetterich:1987fm,Peebles:1987ek,Caldwell:1997ii}), in the MaVaN scenario one does not have to choose a mass for the dynamical field of the order of the present Hubble parameter $H_0$. In fact it can be much larger, even of the order of the neutrino mass scale, due to the stabilising effect of the contribution of the relic neutrino density to the potential. As a result, the field adiabatically tracks the minimum of the effective potential whose evolution is controlled by the time evolution of neutrino number density. The MaVaN models have been investigated in many studies, e.g.\\ regarding possible experimental signatures and constraints \\cite{Kaplan:2004dq,Zurek:2004vd,Li:2004tq,Barger:2005mn,Schwetz:2005fy,Ringwald:2006ks, Cirelli:2005sg, Weiner:2005ac, Brookfield:2005bz,Bernardini:2008pn} as well as model-building issues \\cite{Fardon:2005wc, Takahashi:2005kw, Ma:2006mr, Takahashi:2007ru, Bjaelde:2008yd}. It has turned out that while there are many interesting possible signatures and attractive features, the scenario is tightly constrained by the requirement of consistency with late time structure formation. In particular, it has been pointed out \\cite{Afshordi:2005ym} that if neutrinos are non-relativistic today, due to fifth force effects the neutrinos would cluster at late time and finally form ``neutrino nuggets'' which would spoil the dark energy behavior of the neutrino fluid. The goal of this paper is to investigate the phenomenology of a generalised scenario which we will refer to as MaVaN Hybrid Scenario, where in addition to a MaVaN sector a cosmological constant (from a false vacuum) is driving the accelerated expansion of the universe today. As we will show, for a generalised power law potential in the MaVaN sector the effective equation of state parameter $w_{eff}(z)$ as well as the fraction to which the dark energy is of dynamical nature is determined by the neutrino mass scale. We will also analyse the possibility to realise neutrino dark energy with a ``high-scale'' seesaw mechanism, where the right-handed neutrino masses are close to the GUT scale, and comment on how such models might help to suppress the formation of ``neutrino nuggets'' and resolve stability problems of MaVaN models with non-relativistic neutrinos. For the main part of the paper we will focus on neutrinos that are non-relativistic today. The paper is organised as follows: In Section 2 we will review the basics of the MaVaN scenario. Section 3 contains the definition of the phenomenological framework and the analysis is performed in Section 4. In Section 5 we discuss the possibility of realising neutrino dark energy with a ``high-scale'' seesaw mechanism and in Section 6 we comment on how this might help to solve stability issues for non-relativistic neutrinos. Section 7 contains our Summary and Conclusions. ", "conclusions": "Motivated by the intriguing proximity of the energy density of dark energy and the neutrino mass scale we have studied the phenomenology of hybrid scenarios of neutrino dark energy, where in addition to a so-called Mass Varying Neutrino (MaVaN) sector, a cosmological constant (from a false vacuum) is driving the accelerated expansion of the universe today. Within the generalised framework we have focused on phenomenological issues such as on the connection to the neutrino mass scale and on its consequences for the dynamical nature of dark energy. We have therefore calculated the effective equation of state parameter $w_{eff}(z)$ in the MaVaN Hybrid Scenario where the effective potential for the dynamical real scalar field (the ``acceleron'' field $A$) has the following form \\begin{equation} V(A)_{eff} = M^{4- \\alpha}A^{\\alpha} + \\frac{\\rho_{\\nu}^{(0)}}{a^3} \\left(\\frac{A_0}{ A}\\right)^{\\beta} + V_{0} \\;. \\end{equation} We found that, for the case of a power law potential in the MaVaN sector, $w_{eff}(z)$ is determined by the neutrino mass scale and by the parameters $\\alpha$ and $\\beta$ (c.f.\\ Eq.~(\\ref{Eq:weff_mnu})). Due to the interactions of the dark energy field with the neutrino sector, $w_{eff}(z)$ is predicted to become smaller than $-1$ for increasing $z>0$ (c.f.\\ Fig.~\\ref{weffective_plot}), which could be tested in future cosmological observations. For the considered scenarios, we have also calculated how the neutrino mass scale determines which fraction of the dark energy is dynamical, and which originates from the ``cosmological constant like'' vacuum energy. In particular, for the case of a mass term potential and a standard seesaw relation (i.e.\\ $\\alpha=2$ and $\\beta=1$) we found that compatibility with the terrestrial neutrino mass bounds requires a large contribution of constant vacuum energy (c.f.\\ Eq.~(\\ref{Eq:MnuvsGamma})). Another interesting question, which we have investigated in the MaVaN Hybrid Scenario with power law potentials is whether it is possible to realise neutrino dark energy with a ``high-scale'' seesaw mechanism, where the right-handed neutrino masses are close to the GUT scale. We found that the field value of the ``acceleron'' field as well as the masses of the right-handed neutrinos can indeed be large and in principle a hybrid scenario of neutrino dark energy might be realised with a ``high-scale'' seesaw. We have also commented on how the Hybrid MaVaN Scenarios with ``high-scale'' seesaw might allow to suppress the formation of ``neutrino nuggets'' and resolve stability problems of dark energy models with non-relativistic neutrinos. In summary, we have found that the considered hybrid scenarios of neutrino dark energy have several attractive features, in particular the close connection to the neutrino mass scale and an effective dark energy equation of state parameter $w_{eff}(z)$ which depends only on the parameters $\\alpha$ and $\\beta$ of the potential and on the neutrino mass scale. The prediction that $w_{eff}(z) < -1$ for $z>0$ provides a ``smoking gun'' signal for such interacting dark energy scenarios, which could be observed in future surveys. Issues which are still open, and which are left for further studies, are whether a MaVaN Hybrid Scenario with ``high-scale'' seesaw can indeed solve the stability problems of conventional MaVaN scenarios with non-relativistic neutrinos and whether a consistent model can be constructed where this is realised." }, "0807/0807.1724_arXiv.txt": { "abstract": "Achieving maximum scientific results from the overwhelming volume of astronomical data to be acquired over the next few decades will demand novel, fully automatic methods of data analysis. Artificial intelligence approaches hold great promise in contributing to this goal. Here we apply neural network learning technology to the specific domain of eclipsing binary (EB) stars, of which only some hundreds have been rigorously analyzed, but whose numbers will reach millions in a decade. Well-analyzed EBs are a prime source of astrophysical information whose growth rate is at present limited by the need for human interaction with each EB data-set, principally in determining a starting solution for subsequent rigorous analysis. We describe the artificial neural network (ANN) approach which is able to surmount this human bottleneck and permit EB-based astrophysical information to keep pace with future data rates. The ANN, following training on a sample of 33,235 model light curves, outputs a set of approximate model parameters ($T_2/T_1$, $(R_1+R_2)/a$, $e \\sin \\omega$, $e \\cos \\omega$, and $\\sin i$) for each input light curve data-set. The whole sample is processed in just a few seconds on a single 2GHz CPU. The obtained parameters can then be readily passed to sophisticated modeling engines. We also describe a novel method \\emph{polyfit} for pre-processing observational light curves before inputting their data to the ANN and present the results and analysis of testing the approach on synthetic data and on real data including fifty binaries from the Catalog and Atlas of Eclipsing Binaries (CALEB) database and 2580 light curves from OGLE survey data. The success rate, defined by less than a 10\\% error in the network output parameter values, is approximately 90\\% for the OGLE sample and close to 100\\% for the CALEB sample -- sufficient for a reliable statistical analysis. The code is made available to the public. Our approach is applicable to EB light curves of all classes; this first paper in the Eclipsing Binaries via Artificial Intelligence (\\ebaints) series focuses on detached EBs, which is the class most challenging for this approach. ", "introduction": "Over the past decade advances in observational technologies, computers and eclipsing binary (EB) analysis codes (e.g., WD code: \\citet{wd1971,wd1993,wd2007}) have enabled the accumulation of an impressive sample of fundamental stellar data. Careful analysis of EB light curves has produced fundamental stellar properties, tests of stellar evolution theories, accurate distances within the Galaxy and to external galaxies, as well as providing tests of stellar structure models and general relativity \\citep[see e.g.][]{guinan2007}. Despite the importance of these astrophysical results, only some hundreds of EBs have been subject to the requisite analysis and the cumulative results populate the astrophysical parameter space sparsely. Current standard practice requires significant human interaction with EB light curve data, particularly in the initial stages of analysis, which defines a \"rate-determining step\" in the process generating these astrophysical results. By 2020, the observational bounty from ground- and space-based programs such as Optical Gravitational Lensing Experiment \\citep[OGLE;][]{udalski1997}, Exp\\'erience de Recherche d'Objets Sombres \\citep[EROS;][]{pd1998}, All Sky Automated Survey \\citep[ASAS;][]{pojmanski2002}, Pan-Starrs \\citep{kaiser2002}, Large Synoptic Survey Telescope \\citep[LSST;][]{tyson2002}, and their space counterparts Hipparcos \\citep{perryman1997}, Kepler \\citep{borucki2004} and Gaia \\citep{perryman2001} will include millions of new EB light curves, even catching some EBs in fleeting stages of stellar evolution. Powerful and mature codes for light curve analysis stand ready to mine this enormous and rich vein of new astrophysical information, and pioneering steps in automating these tools have already been taken by \\citet{devor2005,tamuz2006,mazeh2006} (see \\citet{prsa2007} for a concise overview of current automated techniques). A key issue is to efficiently determine an initial set of model parameters for every light curve as input to the automated analysis process. Starting values have been conventionally supplied by the analyst/astronomer using expert knowledge typically guided by checks with light curve synthesis codes, a time-intensive process that is certainly out of the question for the coming fire-hose of EB data. Approaches to automating light curve solutions have taken various forms to date. \\citet{ww2001,ww2002}, in their work to establish the best distance indicators among detached and semi-detached binaries in the Small Magellanic Cloud, obtained starting parameters for the rigorous WD model by comparing each candidate light curve with a set of template model light curves, sending the best match to an automated version of the WD differential corrector program DC. This could be computationally prohibitive to apply to the expected large future data-sets, even if clever pruning reduced the number of comparison templates. Employing less rigorous physical models, of course, is one approach to computational efficiency. \\citet{devor2005} provides a critical discussion of an automated pipeline employing a simple model of spherical stars without tidal or reflection physics, whose starting values are obtained from an initial guess and then refined using a downhill simplex method with simulated annealing. \\citet{tamuz2006} employ the EBOP ellipsoidal model \\citep{popper1981}. Using this engine, they arrive at initial solutions after a combination of grid search, gradient descent and geometrical analysis of the LC. The challenge is therefore to gain the advantages of a sophisticated model yet keep processing time limited. The Eclipsing Binaries via Artificial Intelligence (\\ebaints) project introduces artificial neural networks (ANNs) that are trained on the rigorous WD physical model and are computationally extremely efficient, towards fully automating the solution process for EBs. In this first paper in the \\ebai series we describe the basic ANN concepts and procedures for applying ANNs to detached EBs. We present results of applying the trained ANN to a set of 10,000 synthetic detached EBs, to 50 detached real world binaries from the Catalog and AtLas of Eclipsing Binaries (CALEB\\footnote{CALEB is maintained by D.~Bradstreet at Eastern University -- see {\\tt http://caleb.eastern.edu}.}), and to the set of 2580 OGLE LMC binaries \\citep{wyrzykowski2003} classified as detached. Subsequent papers will deal similarly with semi-detached systems and overcontact EBs, and address automated light curve classification. ", "conclusions": "ANNs have proved successful in problems of classification, real-time control, data mining and many other tasks in a variety of scientific and technical applications. This paper demonstrates their utility in parameter estimation for detached binaries, even in the face of parameter degeneracy, the correlation of certain pairs of parameters; e.g., the compensating effect of increasing a system's sum of radii vs. increasing its inclination. A task-optimal ANN topology and learning rate was determined and the network was trained on the rigorous WD model. The trained network was tested on a previously unseen test set of 10,000 WD model light curves, yielding model parameters to better than 90\\% accuracy for 90\\% of the test set. In tests with 50 real-world detached binary systems from the CALEB database, the ANN alone achieved sufficiently good fits for 22 systems that differential corrections were not needed. Those requiring differential corrections all converged to produce good quality fits, confirming the utility of the ANN-produced starting parameters. Overall, the CALEB test produced a 100\\% success rate for the ANN. We have shown that a suitable ANN can be successfully trained on a sophisticated EB model, and that the trained network produces quite satisfactory approximate light curve solutions with high computational efficiency. In addition to these advantages, ANNs have favorable properties of interpolation; they are well-behaved in regions around their multi-dimensional training points. Thus the network will interpolate reasonable starting parameters for light curves with asymmetry or spots, and in fact for any (detached) light curve that it has not seen before. It is quite remarkable that the network is able to memorize to such high quality all the light curve solutions in the space of detached binaries with only 8240 numbers ($201\\times40 + 40 \\times 5$ weights). Clearly, this network has sufficient information storage capacity, but it is worth noting that the information capacity corresponding to a given network topology is only known in heuristic approximation \\citep{neelakanta1994}. This confirms that much is yet to be learned about ANNs, and in particular it explains the necessity to explore network topologies and learning rates for their appropriateness to the given problem. Such exploration, in addition to the ability to rapidly train many iterations, further emphasizes the need for speed in the learning phase. This work was greatly facilitated by two developments. The \\emph{polyfit} algorithm produced excellent analytic approximations to real-world light curves sampled at (often quite) unequal phase intervals, where splines or other interpolating methods behave badly \\citep{emery2001}, making equal-phase interpolated data points available for input to the ANN. This algorithm has been found to be an excellent interpolator for EBs of all classes, as well as, for example, Cepheid variables. Since well-behaved interpolation is a problem common to many fields, \\emph{polyfit} could well see further application beyond astronomy. The scheme for parallelization of ANN training over multiple processors was key to quickly exploring topologies and training parameters, and churning through a half-million training iterations in days instead of weeks. This scheme had the beneficial side effect of lending simulated annealing behavior to the training phase, improving the search for the global minimum. The source code for \\emph{polyfit} and ANN is released under the GNU General Public License, which grants users the right to freely use, distribute and modify the code. The programs may be downloaded from the IDP project homepage: {\\tt http://www.eclipsingbinaries.org} or \\phoebe homepage: {\\tt http://phoebe.fiz.uni-lj.si}. This paper focuses on detached EBs. Systems with large proximity effects which are fully treated in the WD model will be discussed in a subsequent paper. We will thus be able to apply ANNs in a unified, automated approach to provide starting solutions for eclipsing binary light curves of all classes." }, "0807/0807.2216.txt": { "abstract": "We perform a detailed photometric analysis (bulge-disk-bar decomposition and Concentration-Asymmetry-Clumpiness - CAS parametrization) for a well defined sample of isolated galaxies, extracted from the Catalog of Isolated Galaxies \\citep{karachentseva73} and reevaluated morphologically in the context of the AMIGA project (\\textbf{A}nalysis of the interstellar \\textbf{M}edium of \\textbf{I}solated \\textbf{GA}laxies). We focus on Sb-Sc morphological types, as they are the most representative population among the isolated spiral galaxies. Our analysis yields a large number of important galactic parameters and various correlation plots are used to seek relationships that might shed light on the processes involved in determining those parameters. Assuming that the bulge S\\'{e}rsic index and/or Bulge/Total luminosity ratios are reasonable diagnostics for pseudo- versus classical bulges, we conclude that the majority of late-type isolated disk galaxies likely host pseudobulges rather than classical bulges. Our parametrization of galactic bulges and disks suggests that the properties of the pseudobulges are strongly connected to those of the disks. This may indicate that pseudobulges are formed through internal processes within the disks (i.e. secular evolution) and that bars may play an important role in their formation. Although the sample under investigation covers a narrow morphological range, a clear separation between Sb and Sbc-Sc types is observed in various measures, e.g. the former are redder, brighter, have larger disks and larger bars, more luminous bulges, are more concentrated, more symmetric and clumpier than the latter. A comparison with samples of spiral galaxies (within the same morphological range) selected without isolation criteria reveals that the isolated galaxies tend to host larger bars, are more symmetric, less concentrated and less clumpy. ", "introduction": "The properties of galaxies and their evolution are thought to be strongly related to their environment. The empirical quantification of environmental influence (``nurture'') on morphology, structure, nuclear activity, star formation properties, etc. requires a robust definition of a sample of galaxies that are minimally perturbed by other galaxies. Such a sample could serve as a ``pure nature'' baseline. In this sense, perhaps the best compilation of isolated galaxies available at this time is the Catalog of Isolated Galaxies \\citep[CIG;][]{karachentseva73}. Both the size (n=1050 galaxies) and the restrictive isolation criteria in the catalog contribute to its statistical value. The definition of isolation requires that, for a galaxy of diameter D, there is no companion/neighbor with a diameter d in the range D/4 to 4D within a distance of 20d. The isolation criteria used to construct the CIG suggest that a typical galaxy of 25 kpc diameter has not been visited by a similar mass perturber in the past $\\sim$ 3Gyr \\citep[assuming a typical field velocity of $\\sim$ 150 km s$^{-1}$;][]{verdes05}. Thus, the evolution of such isolated galaxies is mostly driven by internal processes and to a much lesser degree by environment, at least for the last $\\sim$ 3Gyr of their existence. A recent morphological reevaluation of the CIG galaxies in the context of the AMIGA project (\\textbf{A}nalysis of the interstellar \\textbf{M}edium of \\textbf{I}solated \\textbf{GA}laxies) revealed that the bulk ($\\sim$ 63\\%) show morphological types in the range Sb-Sc \\citep{sulentic06}. In this study we present the results of a photometric characterization for a representative subsample of n $\\sim$ 100 CIG galaxies classified as Sb-Sc in this latter reference. We perform multicomponent decomposition (bulge/disk/bar) using the BUDDA code\\footnote{http://www.mpa-garching.mpg.de/~dimitri/budda.html} \\citep{deSouza04}. Additionally, we evaluate CAS parameters Concentration(C)-Asymmetry(A)-Clumpinesss(S) \\citep[e.g.][]{conselice00,bershady00,conselice03,taylor07}. Assembling a set of parameters combining a model-dependent description of the main components of galaxies (BUDDA) with global structural measures (CAS) could provide valuable hints into the formation and evolution of galaxies. This is the first attempt to date to present a detailed examination of this kind (bulge-disk-bar decomposition combined with CAS parameters) for a well defined sample of isolated galaxies. This study is an integral part of the AMIGA project, which is a dedicated multiwavelength study of the revised CIG catalog. The goal of AMIGA is to quantify the fundamental properties of a statistically meaningful sample of isolated galaxies which can then be used as a baseline for comparison and for estimation of the effects of environment in other less isolated samples of galaxies. The CIG catalog has recently been reevaluated in terms of galaxy positions \\citep{leon03}, isolation \\citep{verley07a,verley07b} and morphology \\citep{sulentic06}. A series of studies were produced in the context of the AMIGA project: 1) an optical characterization of the refined sample \\citep{verdes05}, 2) an analysis of mid- and far-infrared properties \\citep{lisenfeld07}, 3) a study of the neutral CO and HI gas \\citep{espada05,espada06}, 4) radio continuum emission \\citep{leon08} and 5) nuclear activity \\citep{sabater08}. Another recent study used a subsample of isolated AMIGA galaxies to investigate the role of bars in star formation processes \\citep[e.g.][]{verley07c}. Our present study offers a detailed photometric analysis of a representative sample of the core AMIGA population of Sb-Sc morphological types. We should note that all data produced within the AMIGA project are periodically updated and made publicly available at http://www.iaa.es/AMIGA.html. Theoretical models and numerical simulations exploring the formation and evolution of galaxies rely on empirical results that could separate and quantify the relative roles of internal secular processes (that develop on time scales much longer than the galaxy formation/collapse process itself) and slow or fast external perturbations (environment) in defining the structural properties of galaxies. In this sense, our present study has a twofold importance: a) it explores a representative and well defined sample of \\textit{the most isolated galaxies} in the local Universe and b) provides an \\textit{extensive photometric structural analysis} of these galaxies. Our main goal is to identify potential scaling relations and correlations: i) between parameters describing the same structural component (bulge, disk or bar), ii) between components, iii) between components and global properties of the galaxy (morphological type, color, luminosity, concentration, asymmetry, clumpiness, etc.). With such correlations available one could explore for example the nature of bulges in isolated galaxies and how they are formed, the role of bars (if any) in the formation/evolution of bulges, whether the isolated spiral galaxies are different relative to spirals in richer environments in terms of global properties and/or in terms of properties of their components (bulge, bar, disks). This paper is organized as follows: \\S~2 presents the selection and basic properties of the sample, \\S~3 offers a concise view on data reduction, \\S~4 and \\S~5 present the results of BUDDA decomposition analysis and CAS parametrization, respectively. \\S~6 combines various measures obtained from the BUDDA code with CAS parameters. \\S~7 is dedicated to discussion and conclusions. Throughout the paper we use H$_{o}$ = 75 km s$^{-1}$ Mpc$^{-1}$. ", "conclusions": "We have presented a detailed structural analysis for a well defined sample of $\\sim$ 100 late-type isolated galaxies. If a {\\it bona fide} isolated (pure ``nature'') population of galaxies exists then our previous work \\citep{sulentic06} suggests that it is dominated by systems with spiral morphology ($\\sim$ 84\\%) with the bulk in the range Sb-Sc (63\\%). We assume that the galaxies we investigate here are best described as ``minimal nurture and maximal nature'' systems because they are as isolated as individual galaxies can be. This hypothesis does not imply that these isolated galaxies have undergone no merger activity since their epoch of formation but rather that major mergers are probably absent from their past $\\sim$ 3Gyr history. We do note that the AMIGA sample includes 14\\% early-type galaxies and those are systems of such low luminosity as to suspect little or no major merger activity over their entire history \\citep{sulentic06}. One might reasonably expect the tightest correlations between various intrinsic properties from a sample of isolated galaxies, where it is assumed that nurture (i.e. interactions) would increase the scatter (e.g. UBV-colors; \\citealt{larson78}). The strength of this study is manifold: the large size of the sample, the uniformity of the SDSS data, the robustness of the BUDDA code and the stringent isolation criteria underlying the definition of the parent AMIGA sample. In this study we have retained subjective morphological classifications and investigate morphological type dependence of various properties even though the typical range is narrow. This narrowness coupled with our ``nurture-free'' assumption raises the possibility that Hubble type T=4$\\pm$1 may represent the seed population for all spiral galaxies. \\subsection{Pseudobulges in Isolated Galaxies} We present evidence favoring the hypothesis that most or all late-type isolated galaxies host pseudobulges (\\S\\S~4.1) rather than classical bulges: A. A large majority of our isolated systems host relatively ``unevolved'' bulge structures (as hypothesized by \\citealt{hunt04}); most S\\'{e}rsic indices (n$_{bulge}$) are smaller than 2.0-2.5 (see Table 5 and Figure 4) with the largest concentration around n$_{bulge}$ $\\sim$ 1.3-1.4. Such bulges are probably not as relaxed as larger bulges in earlier spiral types. They are likely dominated by rotation unlike higher S\\'{e}rsic index bulges (for a detailed discussion on this subject see section 4.6 in \\citealt{kormendy04}, section 4.2 in \\citealt{laurikainen07} and references therein). B. We observe a large range of effective surface brightness $\\mu_{e}$ for a rather narrow range of r$_{e}$ (these two last parameters defining in part the fundamental plane) - see Figure 5a. The locus occupied by the bulges of our Sb-Sc galaxies in this plane is similar to that of disky bulges of galaxies at the end of dissipative collapse \\citep{capaccioli92}. The lack of correlation between $\\mu_{e}$ and r$_{e}$ supports the case of ``pseudobulges'' for isolated spiral galaxies in our sample. As pointed out in \\citet{macarthur03}, these results support an ``iceberg'' scenario, i.e. late-type spiral bulges are ``more deeply embedded in their host galaxy disk than earlier type bulges''. This idea is further complemented by the fact that the size of the bulge (r$_{e}$) scales with the scalelength of the disk (Figure 10a) \\citep[see also e.g.][]{khosroshahi00,macarthur03,mendez08,fisher08}. We observe a larger dispersion in r$_{e}$/h$_{R}$ for larger Bulge/Total luminosity ratios (Figure 10b). However, in Figures 10b-c one can clearly see that Sbc-Sc galaxies do show a clear increasing trend for r$_{e}$/h$_{R}$ and r$_{e}$/a$_{25}^{i}$ with Bulge/Total and Bulge/Disk luminosity ratio, respectively. In contrast, Sb galaxies appear detached from the Sbc-Sc population. Assuming that the bulge S\\'{e}rsic index and/or Bulge/Total luminosity ratios are reasonable discriminators of pseudo- versus classical bulges (\\S\\S~4.1), then amongst our sample Sb galaxies have the greatest chance of hosting classical bulges. Thus, in Figures 10b-c we may have yet another indication that the pseudobulges and the galactic disks are clearly connected, while the classical bulges do not show similar scaling relations. Some studies (e.g. \\citealt{thomas06}) argue that ``secular evolution through the disk and the phenomenon of pseudobulge formation are most likely restricted to spirals of types Sc and later''. Our results (but see also \\citealt{laurikainen07}) find a large fraction of pseudobulges among spiral types earlier than Sc (see \\S\\S~4.1). This may be telling us that the formation of pseudobulges does not appear exclusively restricted to Sc types or later. Our results suggest that if one considers only morphological types later than Sc, one may identify an almost pure pseudobulge population of galaxies. A fundamental question mentioned earlier is whether the isolated Sb-Sc spiral galaxies constitute the seed population of unnurtured spirals? If so then isolated galaxies might be expected to host a pure pseudobulge population. In this context Sb types in our sample have the greatest chance of bulge building via nurture and may involve a mixed classical and psuedobulge population. In Figure 4 it is interesting that a linear correlation emerges only for galaxies of Sb type which bridges the classical and pseudobulges domains. Alternatively, the trend may be telling us that all/most Sb galaxies contain a real (classical) bulge. This would suggest that some large bulges are natural or that all Sb spirals in the sample are a product of nurture. The latter interpretation is disfavored by the extreme isolation of our sample. \\subsection{The Role of Bars in the Formation of Pseudobulges} The results of the present study could set constraints for various galaxy formation and evolution models. Two important galaxy formation scenarios have been proposed and advocated: 1) spheroidal component (bulge) forms prior to the disk component in a monolithic collapse or via early mergers (so called ``inside out'' formation, e.g. \\citealt{eggen62,baugh96,kauffmann96,vandenBosch98,cole00,merlin06}) and 2) bulges form after the disk component as a result of secular dynamics/evolution driven by a disk instability \\citep[e.g.][]{courteau96,zhang04} possibly triggered by external satellite accretion \\citep[e.g][]{aguerri01,eliche06}. The former mechanism may be dominant for elliptical galaxies and in early spiral galaxies with large bulges (as they all appear to share similar properties and scaling relations within the fundamental plane; e.g. \\citealt{kormendy85,djorgovski87,faber87}). The latter mechanism may be more plausible for late type spiral systems \\citep[e.g.][]{carollo99,hunt04,debattista04}, as they largely harbor pseudobulges. Some authors proposed that bulges of late type spiral galaxies are formed primarily through secular evolution of bars \\citep[e.g.][]{kormendy79,kormendy93,norman96,hasan98,fathi03,kormendy04,athanassoula05,jogee05,debattista06}. Others have suggested that bars can help the process of ``pseudobulge'' formation (making it faster and more efficient), but is not a necessary requirement for that process \\citep[e.g.][and references therein]{laurikainen07}. Bars can transport gas inward \\citep[e.g.][]{sakamoto99, sheth05} potentially contributing to the formation of a bulge. On the other hand it has been proposed that even without a bar the stellar disk component could be redistributed due to a secular torque action \\citep[e.g.][]{zhang07}. We find a larger fraction of barred galaxies among Sb types relative to Sbc-Sc types (\\S\\S~4.2). Sb galaxies also appear to host the largest bars (Table 6a) within the morphological sequence Sb-Sbc-Sc. If bars are assumed as necessary precursors of all pseudobulges, then the smaller bars in later type galaxies ``dissolve'' more efficiently in the process of bulge formation. It is interesting to mention that for Sb and Sbc types in our sample of isolated galaxies we find systematically larger values of the index n$_{bulge}$ for barred galaxies compared to the non-barred galaxies (Tables 6a-b). The difference almost vanishes for Sc barred and non-barred. \\citet{laurikainen07} report a rather opposite result for Sb type (see their Figure 3). If n$_{bulge}$ is one of the empirical discriminators between classical and pseudobulges then any connection with the presence/absence of bars merits further attention. In this context it is relevant to review our Figures 8a-b. We note the ``disappearance'' of objects with n$_{bulge}$ above 1.7 for non-barred galaxies (Figure 8b in contrast to Figure 8a). We tested whether this may be caused by the resolution limitation in SDSS images, thus the BUDDA code's inability to identify the presence of a bar. First, we analyzed the distribution of n$_{bulge}$ values of non-barred galaxies with V$_{R}$ lower and higher than the median V$_{R}$ of the full non-barred sample ($\\sim$ 5700 km s$^{-1}$), respectively. We found no significant difference. Secondly, for our galaxies, the typical seeing FWHM is better than 1\\arcsec, with very few cases at 1.5\\arcsec. Considering the most extreme case, for a galaxy showing V$_{R}$ $\\simeq$ 10,000 km s$^{-1}$ a 1.5\\arcsec seeing would translate into a spatial resolution of $\\sim$ 1.0 kpc, which is well within the capability of the BUDDA code to provide reliable structural measures \\citep{gadotti08}. The scarcity of non-barred galaxies with n$_{bulge}$ above 1.7 is consistent with the scenario that bars could transform by dissolution into pseudobulges. The presence of bars may influence the degree of relaxation of bulges in the sense that n$_{bulge}$ decreases from Sb through Sc only for barred galaxies, but not for non-barred spirals (\\S\\S~4.2). The formation and lifetime of bars may be sensitive to environment \\citep[e.g.][]{gerin90}. It has been suggested that bars in early type spiral galaxies are formed by tidal interactions with other galaxies and those in late types have intrinsic origin \\citep{noguchi96}. The connection bars-environment may be different for early and late type spirals \\citep{noguchi96, noguchi00}, being proposed a ``bimodality'' of bars in this sense. Moreover, numerical simulations have shown that for Sb-Sc galaxies bars are transient features and dissolve progressively in $\\sim$ 1-2 Gyr \\citep{bournaud05}. As we pointed out earlier, the AMIGA/CIG isolated galaxies have been basically nurture-free for at least a comparable time. We find that $\\sim$ 50-60\\% of our present sample are barred galaxies. The conclusion here could be that the bars we observe in these late type isolated spiral galaxies have been likely renewed or reformed through internal processes and not by external accretion or interactions \\citep[e.g.][]{block02,berentzen04}. It is also interesting to mention we find that the largest bars lie in disks with the lowest central surface brightness $\\mu_{o}$ (Figure 9b). This is consistent with the idea that bars build up from the material in the central parts of disks and they are products of secular dynamical evolution within the disk. We find that our isolated galaxies tend to host large bars\\footnote{However, one must be aware that there is no standard definition for the length of a bar in a galaxy \\citep{erwin05} and scaling parameters for galaxy components may be sensitive to the filter that is used for photometry.}. Our Figure 6 shows that most bar radii are clustered in the range 2-6 kpc. \\citealt{erwin05} (based on \\citealt{martin95}) reports typical bar sizes in the range 1-3 kpc (B-band) for morphological types Sb-Sc having absolute magnitudes similar to our sample. A more recent study \\citep{marinova07} presents a characterization of bars in optical (B-band) and near-IR (H-band) for the OSUBSGS sample of galaxies. In order to compare the bar sizes with their estimates, we restrict their sample to Sb-Sc morphological range, based on the RC3 catalog \\citep{deVaucouleurs91}. The OSUBSGS-based sample has a similar distribution of absolute magnitudes as our sample. In terms of l$_{bar}$, our sample of barred galaxies (n=48) is characterized by a mean $\\sim$ 5.0 kpc and a median $\\sim$ 4.8 kpc. For The OSUBSGS sample of n=49 barred galaxies, the mean and median values (H-band) of l$_{bar}$ $\\sim$ 3.8 kpc and $\\sim$ 3.4 kpc, respectively. The conclusion is that the size of bars may be related to the environment, isolation favoring larger bars. Moreover, this conclusion seems to be consistent with reports that the disk scalelength h$_{R}$ of spiral galaxies in rich environments is typically smaller than that of field (i.e. isolated) galaxies \\citep[e.g.][]{aguerri04}. We find that the bar size scales with the disk scalelength h$_{R}$ (our Figure 9; see also \\citealt{laine02}). In extreme environments (e.g. compact groups) spiral galaxies tend to lose their disk components by dissolution into a stellar halo. The size of the disks in what we assumed were initially late type spiral galaxies in Seyfert's Sextet for example (estimated by the last concentric isophote) is less than 10 kpc diameter, comparable to the smallest disks in our present sample \\citep{durbala08}. Comparing our $l_{bar}/a_{25}^{i}$ estimates with the similar quantities reported in \\citet{erwin05} we observe the same declining trend from Sb through Sc; for our sample we do obtain systematically larger $l_{bar}/a_{25}^{i}$ ratios relative to that study, although we note that \\citet{erwin05} measures are based on B-band data from \\citet{martin95}. \\subsection{CAS Structural Measures in Isolated Galaxies} The minimal environmental influence on AMIGA/CIG galaxies investigated here is revealed also by an analysis of the structural properties in terms of CAS parameters (\\S~5). Due to the narrow morphological range represented in our sample of isolated galaxies, any attempt at comparison with other studies must be cautiously explored. Nonetheless, the size of the sample examined in the present study allows a meaningful comparison of the 96 galaxies as a whole (i.e. the full set of Sb-Sc galaxies) with galaxies of same morphological types selected without isolation constraints. Table 11 offers such a comparison with the subsample of Sb-Sc galaxies (n=49) examined in \\citet{conselice03}, extracted from the \\citet{frei96} sample, assumed representative for the population of nearby normal galaxies. The general conclusion is that the isolated galaxies are less concentrated, less asymmetric and less clumpy than other galaxies of same morphological type selected without isolation criteria. Thus, we may have clear indications of environmental influence on the structure of galaxies. This may be telling us that the formation of large central concentrations and large clumps within disks are disfavored in the absence of comparable sized neighbors. \\subsection{Describing the Morphological Classification} Although our study involves a narrow range of morphological types all plots that involve exclusively bulge measures show clear morphological separation (e.g. Figure 5). When we combine disk measures (e.g. Figure 7a), the morphological segregation is less clear or absent suggesting some commonality among disk properties over the Sb-Sc range. Thus, it appears that the morphological separation may be associated with a change in the luminosity profile of bulges as indicated by their S\\'{e}rsic indices. \\citet{hunt04} proposed that spiral galaxies may begin with low bulge S\\'{e}rsic index. As they age they change into structurally more evolved systems (toward n$_{bulge}$ = 4 or higher) also characterized by higher surface brightness (see Figures 5a, 12d) and an increased absolute magnitude (see Figures 5b-d). However, \\citet{carollo99} argue that pseudobulges cannot evolve into denser r$^{1/4}$ (i.e., n$_{bulge}$=4) bulges just by repeated cycles of bar formation/disruption. At the same time one should keep in mind that the bulge S\\'{e}rsic index is associated with rather large uncertainties (\\S\\S~4.1), which complicates its use for a quantitative morphological classification \\citep{gadotti08}. It appears that the concentration index C, the Bulge/Total (or Bulge/Disk) luminosity ratio and the bulge S\\'{e}rsic index are relevant parameters when one describes the morphological sequence of spiral galaxies from earlier to later types. Nonetheless, Figure 12 suggests that the morphological diversity of spiral galaxies is deeply connected to the structure of their bulges. The concentration index C is not a good tracer of Bulge/Total ratio for Bulge/Total $>$ 0.1. This is true because the bulge light is no longer concentrated within the radius that includes 20\\% of the total light (section \\S~3; see also \\citealt{graham01b}). It is not obvious why the bulge surface brightness shows a plateau in its trend versus C (Figure 12d). However, one may speculate that the fact that some of the Sb galaxies curve away from the main trend (described largely by Sbc and Sc types) toward larger C values could be due to a different type of bulges they host. \\subsection{Final Remarks} This present study could be complemented by an extension of a similar type of analysis to the whole set of isolated spiral galaxies, which would include the whole sequence of Hubble morphological types. This would provide a more general and a more clear picture on the morphological type dependence of various structural properties and scaling relations presented and discussed here. Measures of bulge colors and kinematics would both provide strong tests of our hypothesis that most isolated spirals involve pseudobulges. Another complementary approach is a Fourier analysis of our images, which would provide a quantitative description of the spiral structure, intimately connected to galactic morphology as well. This is part of an ongoing project we are working on at this time and the results will be presented in a future paper." }, "0807/0807.1348_arXiv.txt": { "abstract": "{ We present the results from an ESO/VLT campaign aimed at studying the afterglow properties of the short/hard gamma ray burst GRB\\,070707. Observations were carried out at ten different epochs from $\\sim 0.5$ to $\\sim 80$ days after the event. The optical flux decayed steeply with a power-law decay index greater than 3, later levelling off at $R\\sim 27.3$~mag; this is likely the emission level of the host galaxy, the faintest yet detected for a short GRB. Spectroscopic observations did not reveal any line features/edges that could unambiguously pinpoint the GRB redshift, but set a limit $z < 3.6$. In the range of allowed redshifts, the host has a low luminosity, comparable to that of long-duration GRBs. The existence of such faint host galaxies suggests caution when associating short GRBs with bright, offset galaxies, where the true host might just be too dim for detection. The steepness of the decay of the optical afterglow of GRB\\,070707 challenges external shock models for the optical afterglow of short/hard GRBs. We argue that this behaviour might results from prolonged activity of the central engine or require alternative scenarios.} ", "introduction": "Gamma-ray bursts (GRBs) are among the most powerful explosions in the universe. They are revealed in the hard X-ray/soft gamma-ray band and are followed in many cases by a fading afterglow observable from radio to X-ray wavelengths. GRBs are empirically classified in two groups \\citep{Maz81, Nor84, Kou93,Tav96}: short GRBs last less than 2\\,s and have a hard spectrum; long GRBs have longer durations (typically tens to hundreds seconds) and somewhat softer spectra. The emergence of a typical supernova (SN) spectrum superposed on the rapidly decaying non-thermal afterglow weeks after the events and the association with blue, highly star-forming galaxies provided strong evidence that a significant fraction of long GRBs originates in the gravitational collapse of massive stars (but see \\citealt{MDV06, Fyn06, Gal06a}). Short GRBs are revealed less frequently than long GRBs \\citep[they comprise about 1/4 and 1/10 of the BATSE and \\textit{Swift} samples, respectively;][]{Kou93,Ber07}; moreover their afterglows are weaker and, thus, more difficult to detect and follow up. These are among the reasons why the origin of short GRBs is still under debate, despite the important progresses made in the \\textit{Swift} era. The tight upper limits on any associated SN \\citep{Hjo05,Cov06,Kann08} as well as the association with a broad variety of Hubble types hosts, from elliptical \\citep{Berger05} to moderate star forming galaxies \\citep[e.g.][]{Cov06, Fox06, Berger08}, rules out the core-collapse mechanism as the main channel for short-GRBs production and strongly suggest that the explosion mechanism and/or progenitors of short GRBs are different from those of long GRBs \\citep[for a recent review, see][]{LeRa07}. The leading model for short GRBs involves the merging of a system composed of two collapsed objects, a double neutron star (DNS) or a black hole/neutron star binary. In those systems that evolve out of massive stars that were born in a binary system (we term these ``primordial binaries''), the delay between formation and merging is dominated by the gravitational wave inspiral time, ranging from tens of Myr to a few Gyr \\citep{PeBe02}, strongly dependent on the initial system separation. Short GRBs that result from them are expected to: (a) have a redshift distribution which broadly follows that of star formation and (b) drift away in some cases from the star-forming regions in which they were born, and merge outside, or in the outskirts, of galaxies \\citep{Bel02}. A fraction of two collapsed object binaries may also form dynamically through binary exchange interactions in the core of globular clusters \\citep{Gri06}. For such a formation mechanism, the delay between star-formation and merging is driven by the cluster core collapse time, which is comparable to the Hubble time \\citep{Hop06}. Therefore, short GRBs originating from dynamically formed double collapsed object binaries should go off at lower redshifts than short GRBs from primordial binaries \\citep{GuPi05,GuPi06,Gal06,Hop06,Salva08}. In another scenario, a fraction of the short GRBs is due to hyperflares from soft gamma-ray repeaters in the local universe (distances up to $\\sim 100$\\,Mpc; \\citealt{Hur05,Tan05,Fre07,Maz08}). The above summary emphasises that redshift determination, GRB position relative to the host galaxy and properties of the host galaxy are all crucial pieces of information for understanding short GRBs and discriminating among different models \\citep{Bel06}. One key issue in the study of short GRBs is the secure identification of the host galaxy. Indeed, several short GRBs afterglows have been detected only in X-rays and thus localized with a precision of a few arcseconds (whereas sub-arcesecond localizations are required to unambiguously reveal their host). The possibility that a sizeable fraction of short GRBs lie outside the light of their hosts, as predicted for both primordial and dynamically formed NS-NS/BH binaries, further complicates the identification process. Indeed some proposed short GRB associations with bright, nearby galaxies, based on some angular separation, might result from by-chance alignment. So far, a dozen short GRBs were localized with sub-arcsecond precision and host galaxies were firmly detected with small offset. Only in a few cases (e.g. \\object{GRB\\,061201}: \\citealt{Str07}; \\object{GRB\\,080503}: \\citealt{Perley08}) no host galaxy was found down to $R \\geq 26$--28 after the optical afterglow had faded. For other short GRBs at unknown redshift, a putative host galaxy was proposed with magnitude $R \\sim 23$--26. Spectroscopy of the brightest four of these galaxies indicates that they lie at $0.4 \\leq z \\leq 1.1$. A comparison with field galaxy magnitudes suggests that the rest of the sample lies at $z \\geq 1$ \\citep{Ber07b}. The unambiguously localized hosts are both early- and late-type galaxies, with very different star formation rates and masses \\citep{Nak07}. The association with early-type galaxies has provided clear evidence that at least a fraction of the short GRBs have progenitors related to an older stellar population than that of long GRBs, as expected from the compact binary system models. The nature of the progenitors of short GRBs that go off in star forming galaxies is still under debate. The possibility that short GRBs comprise different subclasses cannot be confirmed yet but neither excluded. In this paper we present the results from an extensive campaign aimed at studying the optical afterglow of \\object{GRB\\,070707}. This campaign monitored the decay of the optical afterglow evolution from about 0.45\\,days to more than one month after the burst. This is one of the best optical light curves for a short/hard GRB so far obtained. In Sect.\\,\\ref{sec:grb} we report on previous results on \\object{GRB\\,070707}. In Sect.\\,\\ref{sec:obs} our observations and data analyses are discussed and in Sect.\\,\\ref{sec:res} our results are presented. In Sect.\\,\\ref{sec:disc} we discuss our findings. \\begin{table*} \\caption{VLT observation log for GRB\\,070707. Magnitudes are not corrected for Galactic absorption. Upper limits are given at $3\\sigma$ confidence level.} \\centering \\begin{tabular}{llrcccc} \\hline Mean time & Exposure time & $t-t_0$ & Seeing & Instrument & Magnitude & Filter / Grism \\\\ (UT) & (s) & (days) & ($\\arcsec$) & & & \\\\ \\hline 2007 Jul 08.12988 & $10 \\times 120$ & 0.45742 & 0.9 & VLT/FORS1 & $23.05 \\pm 0.02$ & $R$ \\\\ 2007 Jul 09.07984 & $10 \\times 120$ & 1.40738 & 1.0 & VLT/FORS1 & $23.86 \\pm 0.05$ & $R$ \\\\ 2007 Jul 09.23005 & $ 1 \\times 180$ & 1.55759 & 1.0 & VLT/FORS1 & $24.07 \\pm 0.12$ & $R$ \\\\ 2007 Jul 10.14491 & $10 \\times 120$ & 2.47245 & 0.7 & VLT/FORS1 & $25.33 \\pm 0.08$ & $R$ \\\\ 2007 Jul 10.21523 & $20 \\times 3 \\times 30$ & 2.54277 & 0.7 & VLT/ISAAC & $> 23.6 $ & $J$ \\\\ 2007 Jul 11.13697 & $20 \\times 180$ & 3.46721 & 0.5 & VLT/FORS1 & $26.62 \\pm 0.18$ & $R$ \\\\ 2007 Jul 12.17370 & $17 \\times 300$ & 4.50124 & 0.8 & VLT/FORS1 & $26.81 \\pm 0.22$ & $R$ \\\\ 2007 Jul 15.21785 & $18 \\times 300$ & 7.54539 & 0.6 & VLT/FORS1 & $27.39 \\pm 0.22$ & $R$ \\\\ 2007 Jul 19.12752 & $20 \\times 300$ & 11.45506 & 0.6 & VLT/FORS1 & $27.21 \\pm 0.20$ & $R$ \\\\ 2007 Aug 15.07151 & $20 \\times 300$ & 38.38355 & 0.6 & VLT/FORS1 & $27.15 \\pm 0.29$ & $R$ \\\\ 2007 Sep 28.03125 & $20 \\times 3 \\times 60$ & 81.95974 & 0.5 & VLT/NACO & $> 22.7$ & $K$ \\\\ \\hline 2007 Jul 09.32508 & $5 \\times 2400$ & 1.65262 & 1.0 & VLT/FORS1 & --- & $300V+$GG375 \\\\ 2007 Jul 13.15960 & $2 \\times 900$ & 5.48714 & 0.8 & VLT/FORS1 & --- & $300V+$GG375 \\\\ \\hline \\end{tabular} \\label{tab_log1} \\end{table*} ", "conclusions": "\\label{sec:disc} \\begin{figure} \\includegraphics[width=\\columnwidth]{Amati_2.ps} \\caption{Location of GRB\\,070707 in the plane peak energy vs. isotropic-equivalent energy. The thick solid line shows the position of GRB\\,070707 as a function of redshift, with the diamonds indicating specific values discussed in this paper. Filled circles represent long-duration GRBs (from \\citealt{Amati08}), and the diagonal lines indicate the best-fit Amati relation (dashed) and the 2$\\sigma$ contours (dotted). Empty squares indicate other short-duration events with known redshift and peak energy \\citep{Ama06,Gole06,Ohno07,Gole07b}.} \\label{Amati} \\end{figure} As with a number of other short GRBs, the nearly unconstrained redshift of GRB\\,070707 remains an important limiting factor. In the peak energy vs. isotropic-equivalent energy diagram (Fig. \\ref{Amati}), \\object{GRB\\,070707} should lie on the thick solid curve marked with selected values of the redshift. For the range of possible redshifts, the position of \\object{GRB\\,070707} would be in broad agreement with that of other short GRBs with known redshift and peak energy, which notoriously do not follow the so-called ``Amati-relation'' of long GRBs \\citep{Ama06}. Whichever its redshift, the host of \\object{GRB\\,070707} ($R \\sim 27.3$) is the faintest ever detected for a short GRB, with a magnitude comparable to that of long GRB hosts at high redshift (e.g. \\citealt{Wain07,Fruchter06}). In past works \\citep[e.g.][]{Blo07,Str07,Lev07}, there has been discussion about the possibility to find short GRBs not spatially coincident with a host galaxy. In some cases, nearby, bright objects were proposed to be the GRB host based just on angular proximity. The case of GRB\\,070707 shows that caution is needed. Without deep VLT images, it might have been tempting to associate GRB\\,070707 with the closeby galaxy G6, which is certainly not the GRB host. With such a large magnitude contrast between the afterglow and the host, many more galaxies of short GRBs may be just fainter than the fluxes probed by shallower exposures. While this is consistent with the suggestion that a sizeable fraction of short GRBs reside at redshift larger than $z \\sim 1$ \\citep{Ber07b}, this may also indicate that some short GRBs go off inside low-luminosity galaxies at low redshift. The sparse data on the host galaxy of GRB\\,070707 do not allow for a detailed analysis. At the maximum allowed redshift $z = 3.6$, the host would have an absolute magnitude $M_{\\rm AB} = -18.6$ (at rest-frame wavelength $\\lambda \\approx 1500$~\\AA). This is more that 2 magnitudes fainter than the Schechter luminosity at this redshift ($M^*_{\\rm AB} = -20.7$; \\citealt{Gabasch04}). We can thus firmly set $L < (1/6)L_*$ at any redshift, implying that the host was intrinsically faint. Our late-time NIR upper limit ($K > 22.7$) implies $R - K < 4.3$ and hence can rule out a bright, red host, such as an extremely red object (ERO) or a moderate-redshift elliptical. GRB\\,070707 hence confirms that short GRBs can explode inside faint and possibly extremely faint systems. Short GRB hosts indeed exhibit a wide range of luminosities. Our spectrum did not show any clear absorption feature, which is quite possibly due to the low signal-to-noise and/or limited covered wavelength range. It is however interesting to note that a featureless spectrum has been reported also for the afterglow of the short GRB\\,061201 \\citep{Str07}. If confirmed by further, better-quality observations, this fact may provide hints for a lower density of short GRB environments compared to those of long-duration events. Further information on the GRB progenitor comes from the optical light curve of the GRB\\,070707 afterglow, which decayed very steeply starting $\\sim 1.5$~days from the burst. The power law decay index $\\alpha$ is constrained by our fits to be steeper than 3, or may even be exponential. Assuming that the optical afterglow emission came from the forward shock, as commonly supposed in the fireball model for GRB afterglows, the steepening of the optical light curve could be interpreted as a jet break. However the index measured for GRB\\,070707 is too steep to be explained in terms of jetted emission. In fact this would require $\\alpha_2 = p$ \\citep{Rhoads99}, where $p$ is the electron energy distribution index. Such a steep decay could be marginally consistent with the post-break phase only by adopting a very soft electron energy distribution ($p > 3$). Such large $p$ values have never been found in GRB afterglows, both from theoretical and empirical investigations (e.g. \\citealt{Pana05,She06,Tag06, Kann06}). In particular, values of $p < 3$ are inferred based on afterglow spectra, in which case the analysis is more robust than when relying on the temporal behaviour. Moreover, the magnitude of the steepening from the pre- to the post-break decay would be too pronounced for a jet-break interpretation in GRB\\,070707. The steepness of the observed decay might be difficult to reconcile with forward shock emission even just for causality reasons. The fastest possible decay is the so-called high-latitude emission, which occurs when the fireball emission stops abruptly and the observers see photons coming from the wings of the emitting surface. In this case, $\\alpha = 2 + \\beta$, where $\\beta$ is the afterglow spectral index \\citep{Kumar00}. The observed upper limit in the $J$ band (Table\\,\\ref{tab_log1}) allows us to derive $R-J < 2.1$ on Jul 10.2 UT, which corresponds to $\\beta < 1.95$. This would still be consistent with the high-latitude interpration. A stronger limit on $\\beta$ can however be inferred by using the X-ray data. Assuming a synchrotron spectrum, the optical spectral index can never be softer than the optical-to-X-ray spectral index $\\beta_{\\rm OX}$. For GRB\\,070707, the observed X-ray flux \\citep{Bea07b} implies $\\beta_{\\rm OX} = 0.75^{+0.13}_{-0.09}$. High-latitude emission from a source with such a spectrum cannot decay faster than $t^{-2.75}$, hence effectively ruling out such possibility for GRB\\,070707. A viable alternative to explain the steep decay of GRB\\,070707 is that of a long-lived central engine. In this case, as discussed by several authors \\citep[e.g.][]{Zhang06}, the decay index should be computed after setting the zero time $t_0$ to the end of the extended activity phase. The intrisic decay slope would therefore be shallower than the observed value, eliminating the causality problem. The very steep decay of \\object{GRB\\,070707} thus would provide further evidence that the inner engine powering short GRBs is working for a much longer time than the observed gamma-ray emission. The optical light curve of \\object{GRB\\,070707} might be produced, for example, by a large flare, as also argued for \\object{GRB\\,050724} \\citep{Barthelmy05,Malesani07}. Finally, we mention that the optical light curve of GRB\\,070707 could also be explained within the context of cannonball model (\\citealt{Dado08a}; \\citealt{Dado08b} and references therein). The emitting region in this case has angular dimension well below $1/\\Gamma$, and therefore the causality constraints are much more relaxed." }, "0807/0807.3667_arXiv.txt": { "abstract": "Absorption lines of the $^{12}$CO and $^{13}$CO molecular bands ($\\Delta v$ = 2) at 2.29 -- 2.45 micron are modelled in spectrum of Arcturus (K2III). We compute a grid of model atmospheres and synthetic spectra for giant of Teff = 4300, log g = 1.5, and the elemental abundances of Peterson et al. (1993), but abundances of carbon, oxygen and the carbon isotopic ratio, \\CC are varied in our computations. The computed spectra are fitted to the observed spectrum of Arcturus from the atlas of Hinkle et al. (1995). The best fit to observed spectrum is achieved for log N(C) = -3.78 $\\pm$ 0.1, \\CC = 8 $\\pm$ 1. A dependence of the determined \\CC vs. log N(C) and log N(O) in atmospheres of red giants is discussed. ", "introduction": " ", "conclusions": "" }, "0807/0807.3445_arXiv.txt": { "abstract": "We study the $f(R)$ theory of gravity using metric approach. In particular we investigate the recently proposed model by Hu-Sawicki, Appleby $-$ Battye and Starobinsky. In this model, the cosmological constant is zero in flat space time. The model passes both the Solar system and the laboratory tests. But the model parameters need to be fine tuned to avoid the finite time singularity recently pointed in the literature. We check the concordance of this model with the $H(z)$ and baryon acoustic oscillation data. We find that the model resembles the $\\Lambda$CDM at high redshift. However, for some parameter values there are variations in the expansion history of the universe at low redshift. ", "introduction": "\\vskip 0.5cm \\noindent It is remarkable that different data sets of complementary nature such as supernovae, baryon oscillations, galaxy clustering, microwave back ground and weak lensing all taken together strongly support the late time acceleration of universe. In the standard lore, one assumes that the history of universe is described by the general relativity (GR). The late time acceleration can easily be captured in this frame work by introducing a scalar field with large negative pressure known as {\\it dark energy} \\cite{review}. In view of the fine tuning problem, the scalar field models, specially those with tracker like solutions, are more attractive compared to the models based on cosmological constant. At present, observations are not in a position to reject or to establish the dark energy metamorphosis. A host of scalar field models have been investigated in the literature. The scalar field models can fit the data but lack the predictive power. It then becomes important to seek the support of these models from a fundamental theory of high energy physics. \\vskip 0.5cm \\noindent One can question the standard lore on fundamental grounds. We know that gravity is modified at small distance scales; it is quite possible that it is modified at large scales too where it has never been confronted with observations directly. It is therefore perfectly legitimate to investigate the possibility of late time acceleration due to modification of Einstein-Hilbert action. It is tempting to study the string curvature corrections to Einstein gravity amongst which the Gauss-Bonnet correction enjoys special status. A large number of papers are devoted to the cosmological implications of string curvature corrected gravity\\cite{fr,NOS,KM06,TS,CTS,Neupane,Cal,Sami06,Annalen,Sanyal}. These models suffer from several problems. Most of these models do not include tracker like solution and those which do are heavily constrained by the thermal history of universe. For instance, the Gauss-Bonnet theory with dynamical dilaton might cause transition from matter scaling regime to late time acceleration allowing to alleviate the fine tuning and coincidence problems. However, it is difficult to reconcile this model with nucleosynthesis\\cite{TS,KM06}. Another possibility of large scale modification is provided by non-locally corrected gravity which typically involves inverse of d'Alembertian of Ricci scalar. The non-local construct might mimic dark energy; the model poses technical difficulties and there has been a little progress in this direction\\cite{Wood}. The large scale modification may also arise in extra dimensional theories like DGP model which contains self accelerating brane. Apart from the theoretical problems, this model is heavily constrained by observation. \\vskip 0.5cm \\noindent On purely phenomenological grounds, one could seek a modification of Einstein gravity by replacing the Ricci scalar by $f(R)$. The $f(R)$ gravity models have been extensively investigated in past five years\\cite{Nojiri:2003ft,review1,FRB,FRB1}. The $f(R)$ gravity theories giving rise to cosmological constant in low curvature regime are plagued with instabilities and on observational grounds they are not distinguished from cosmological constant. The recently introduced models of $f(R)$ gravity by Hu-Sawicki and Starobinsky (referred as HSS models hereafter) with disappearing cosmological constant\\cite{HS,star} have given rise to new hopes for a viable cosmological model within the framework of modified gravity ($f(R)$ gravity model with similar properties is proposed in Ref.\\cite{ABTT}). These models contain Minkowski space time as a solution in the low curvature regime which is an unstable solution. In high curvature regime these models reduce to cosmological constant. Both the first and the second derivatives of $f(R)$ with respect to R are positive. The positivity of the first derivative ensures that the scalar degree of freedom, a characteristic of any f(R) theory, is not tachyonic where as the positivity of second derivative tells us that graviton is not ghost thereby guaranteing the stability. In Starobinsky parametrization\\cite{star}, $f(R)$ is given by, $f(R) = R + \\Lambda \\left \\lbrack \\left ( 1 + R^2/R_0^2 \\right )^{-n} - 1 \\right]$. The HSS models can evade solar physics constraints provided that the model parameters are chosen properly. An important observation has recently been made by by Appleby $-$ Battye and Forolov\\cite{AppBatt,frolov} (see also \\cite{Nojiri:2008fk}). The minimum of scalaron potential which corresponds to dark energy can be very near to $\\phi=0$ or equivalently $R=\\infty$. As pointed out in Ref.\\cite{Tsujikawa}, the minimum should be near the origin for solar constraints to be evaded. Hence, it becomes most likely that we hit the singularity if the parameters are not fine tuned. \\vskip 0.5cm \\noindent In order to check whether the $f(R)$ gravity theory is cosmological viable or not, it is necessary that this theory must be compatible with the observations. In this work, we study the Starobinsky model using the data from the recent observations which include $H(z)$ , Hubble parameter at various red-shifts and the Baryon Acoustic Oscillation (BAO) peak from Sloan Digital Sky Survey (SDSS). \\vskip 0.5cm \\noindent This paper is organised as follows. In Section II, we describe the general properties of the model highlighting the fine tuning problem. The Friedmann equation and the special cases of trace equation is studied in Section III. The cosmological constraints from the recent observations are described in Section IV. Finally, Section V contains the results and discussions. ", "conclusions": "\\no The observed late time acceleration of the universe is one of the major unsolved problem in the cosmology. It may hint at the breakdown of Einstein GR . This has lead to modification of the Einstein theory of gravity. One of the attractive possibility to modify this theory is to replace the Ricci scalar $R$ with the generic function $f(R)$ in the Hilbert action. \\vskip 0.5 cm \\no $f(R)$ theories are usually studied by two methods: {\\it metric} and {\\it Palatini} approach. The Palatini approach is explored extensively in the literature both theoretically and observationally \\cite{pal}( see the other ref's given in these papers). This formulation gives second order differential field equation which can explain the late time behavior of the universe. \\vskip 0.5 cm \\no The metric approach of $f(R)$ theory leads to a fourth order non linear differential equation in terms of scale factor. This equation is difficult to solve both analytically and numerically even for the special cases. Therefore, not much observational tests have been performed on $f(R)$ theories based on metric formulation. Our work is an attempt to check the concordance of $f(R)$ theory of gravity using the metric approach with some of the cosmological observations. In particular, we have explored the Starobinsky model in which function $f(R)$ is analytic, satisfying the condition $ f(0) = 0$. This model also passes the Solar system and laboratory tests successfully for large values of $n$. In this work we study the $f(R)$ model for $n= 1$ and $n=2$. \\vskip 0.5 cm \\noindent In order to study the cosmological viability of this theory, we investigate this class of model with two observational tests. The first method is based on the Hubble parameter versus redshift data, $H(z)$. The Hubble parameter is related to the differential age of the universe through this form $$ H(z) = - \\frac{1}{1 +z} \\frac{dz}{dt}. $$ By estimating the $dt/dz$, one can obtain directly the Hubble parameter, $H(z)$ at different redshifts. The H(z) data has one major advantage, unlike in the standard candle approach (SNe Ia): the Hubble function is not integrated over. The other important feature of this test is that differential ages are less sensitive to systematic errors as compared to the absolute ages \\cite{svj}. This observational $H(z)$ have been used earlier also to constrain various other dark energy models \\cite{other}. \\vskip 0.5 cm \\noindent In this work, we check the compatibility of $f(R)$ model with $n =1$ and $n=2$ with the $H(z)$ data. Since $H_0$ is a nuisance parameter, we marginalise over $H_0$. We further combine the results obtained from $H(z)$ data set with the BAO data. To perform the joint test, we define the quantity: \\begin{equation} \\chi^2_{\\mathrm{total}} = \\chi^2_{\\mathrm{H(z)}} + \\chi^2_{\\mathrm{BAO}}. \\end{equation} \\no The results are given in the Table 1. It appears that $n = 2$ is favored by the observations. This fact is also in agreement with solar and laboratory test. The other important conclusions are as follow: \\begin{itemize} \\item \\no The variation of $H(z)$ with redshift becomes independent of $n$ after the redshift around $ z =1.8$, see Fig. 7. Before this redshift ( $ z < 1.8$), there is a small difference between the behavior of $H(z)$ with $z$ for $ n =1$ and $n =2$ models. This variation is still within the error bars of the data points. At higher redshift ( i.e when $R >> R_0$) we recover the standard $\\Lambda$CDM universe for $\\lambda =2$, for all the values of $n$. Hence the thermal history of the universe is correctly reproduced by this model. \\item \\no The expansion history for this model with $ n = 2, \\lambda = 2$ matches exactly with the standard cosmological model ($\\Lambda$CDM) with equation of state parameter $\\omega = -1, \\Omega_m = 0.3$ and $\\Omega_{\\Lambda} = 0.7$ (see fig. 7). The variation in parameters $\\Omega_m$ and $\\Omega_{\\Lambda}$ (keeping $\\Omega_{m0} + \\Omega_{\\Lambda} = 1$) leads to very small changes in values of $\\lambda$, keeping it close to 2. \\item \\no As shown in Table 1, it seems that the present observational data used in this work prefer $n = 2$ over $n = 1$, as indicated by $\\chi^{2}$ per degree of freedom, $\\chi^2_{\\nu}$. However, $ n=1$ cannot be ruled out. \\item \\no As stated earlier, the algebraic expressions become cumbersome for larger values of $n$. It is quite possible that observational tests discussed in this paper may be compatible with values of $n$ larger than two which are also consistent with solar and laboratory tests. However, it should be emphasized that for a given value of $\\lambda$, the minimum of the scalaron potential gets closer to zero for larger values of $n$ making the model vulnerable to singularity, see Fig. 3. \\end{itemize} \\vskip 0.5cm \\no Due to large error bars in the data sets used, the variation of the expansion history of the universe studied in this paper can easily be accommodated by the observations, thereby making the models close to $\\Lambda$CDM at the background level. \\no At present the sample of $H(z)$ data is too small and the error bars are large. In the future, large amount of precise $H(z)$ data is expected to become available. This will not only reveal the fine features of the expansion history of the universe but also tightly constrain the cosmological parameters. \\vskip 0.5cm \\no There is a further need to explore this model with time based observational tests, like age of the universe and high redshift objects. It is known that the evolution of the age of the universe with the redshift vary from model to model. So it is possible that the model of the universe which are able to reproduce the total age of the universe at $z= 0$, may not accommodate the objects at high redshift\\cite{dj}. Therefore the time based observational test may play the key role in short-listing the viable dark energy model in the near future. It is also expected that the study of matter power perturbations would allow to distinguish this model from the standard cosmology." }, "0807/0807.4579_arXiv.txt": { "abstract": "We present 1.4 and 2.5 GHz Australia Telescope Compact Array (ATCA) observations of the galaxy cluster A3158 (z=0.0597) which is located within the central part of the Horologium-Reticulum Supercluster (HRS). Spectroscopic data for the central part of the HRS suggest that A3158 is in a dynamically important position within the supercluster and that it is moving toward the double cluster system A3125/A3128 which marks the centre of the HRS. A total of 110 radio galaxies are detected in a 35 arcminute radius about the cluster at 1.4 GHz, of which 30 are also detected at 2.5GHz. We examine the source counts and compute the Radio Luminosity Function (RLF) at 1.4 GHz from the subset of 88 sources found within the full-width half-power area of the ATCA beam. Comparison of the source counts in the area over the background, as computed by Prandoni et al.~(2001), shows some evidence of an excess of galaxies with L$_{1.4 GHz} \\leq 2 \\times 10^{22} W Hz^{-1}$. This result seems to indicate a star forming population and is a result similar to that found recently by Owen et al.~(2005) for the merging cluster A2125. In addition we find that the radio luminosity function for early-type galaxies (E and S0) below log P$_{1.4} \\sim$ 22.5 is lower than that found for a composite cluster environment (Ledlow \\& Owen, 1996) but is similar to the early-type RLF for clusters in the centre of the Shapley Super cluster (Venturi et al.~ 2000) which are believed to be in the latter stages of merging. This result implies that the cores of superclusters are environments where radio emission, particularly resultant from AGN, is suppressed in the later stages of merging. Thus, radio observations of clusters might be sensitive indicators of the precise merger stage of the cluster but more observational evidence is still required to establish this trend. ", "introduction": "Observational and numerical studies suggest that dynamical interactions are frequent occurrences in clusters of galaxies and that these events are the most energetic in the recent history of the Universe. Matter is believed to stream along preferred axes (so-called filaments and sheets) often accelerating in-falling galaxies to velocities of the order of 10$^3$ km s$^{-1}$. Rich clusters form at the intersection of the filament and sheets, and larger mass congregations, such as superclusters, are often detected as a collection of several clusters and filaments. These dynamical processes cause shocks, turbulence and bulk flows along the filamentary axes resulting in disturbance of the intracluster medium (ICM). Such perturbations are frequently detected with X-ray observations which show a variety of features including sharp temperature jumps, distorted X-ray isophotes and regions of enhanced pressure and entropy. At radio wavelengths, cluster mergers are thought to generate large areas of diffuse synchrotron emission either as a central, unpolarised halo or one or occasionally two peripheral ``radio relics''. Cluster ``weather'' and bulk flows have also been cited as causes of the bending of the jets in head-tail galaxies (Burns et al.~1998, Johnston-Hollitt et al.~2004), although recent results suggest that more local conditions might play a more important role in their formation (Mao et al.~2008). \\begin{figure*} \\begin{center} \\includegraphics {mjh_fig1.eps} \\caption{Plot of all spectroscopic redshifts available in the region between A3158 (shown on the left, marked with a circle) and the A3125/A3128 complex on the right, again with the clusters denoted by large circles. Filled circles represent objects which have velocities in the range of 17425 km s$^{-1}$ and 18575 km s$^{-1}$ which encompasses those in the bridge investigated by Fleenor (2006).} \\label{bridge} \\end{center} \\end{figure*} In addition to the large-area phenomena discussed above, the dynamical history of galaxy clusters should also be imprinted on the constituent galaxies. In particular, increased secondary star formation either observed in the optical (Burns et al.~1994; Caldwell \\& Rose 1997) or through increased low-power radio emission (Owen et al.~1999) is one possible signature of the history of a dynamical cluster. However, the balance between mergers driving gas towards the centre of the galaxies and hence feeding the central AGN to induce a starburst (Bekki 1999) and an increase in the ram pressure which would strip gas from galaxies and hence curtail the feeding of the AGN (e.g. Balogh et al.~1998; Fujita et al.~1999) still needs to be addressed observationally. To date, few studies have been conducted with both sufficient sensitivity in the radio to probe the low-powered radio sources, and good optical observations to confirm cluster membership. The most notable progress to date has been made on sections of the Shapley Supercluster with several areas analysed in both the radio and optical, including radio observations of A3556 (Venturi et al.~1997), A3571 (Venturi et al. 2002), the A3528 complex (Venturi et al.~2001) and the A3558 complex (Venturi et al.~2001; Miller 2005) and its outskirts (Giacintucci et al.~2004, Miller 2005). In addition to clusters in superclusters, the host radio population for several individual clusters have also been examined including A2255 (Miller \\& Owen, 2005), A2256 (Miller et al.~2003), A2125 (Owen et al.~2005) and A2111 (Miller et al.~2006). However, at present, those observational studies that have been undertaken give incomplete results with both instances in which effects are seen on the radio population, and cases where they are not observed. Sometimes neighbouring clusters in the same supercluster show markedly different results in terms of radio population. This may well be due to the radio galaxy population only being affected in certain phases of dynamical activity (Venturi et al.~2002) however, the exact point at which a merger might affect the radio population is, as yet, unclear. Nevertheless, large-area observations of superclusters, such as Shapley, present ideal datasets with which to investigate the way in which environment affects the galaxy population. In this paper, which will be the first in a series, we present a study of the radio population of A3158 which lies within the kinematic core of the massive Horologium-Reticulum Supercluster (HRS). The HRS spans around 180 square degrees on the sky (Fleenor et al.~2006), contains upwards of 20 galaxy clusters (Zucca et al.~1993), is comprised of at least two major filaments (Einasto et al.~2003), and is the second largest mass concentration in the local 300 Mpc (Hudson et al.~1999). A3158 lies adjacent to the so-called center of the HRS, defined by the double cluster complex A3125/A3128 (Rose et al.~2002). The paper will be laid out as follows: Section 2 will discuss A3158 in detail; Section 3 will give details of the radio observations of the cluster; Section 4 will present the radio source population and in Sections 5 and 6 we will discuss the resultant radio source counts and radio luminosity functions; Sections 7 \\& 8 will present the discussion and conclusions. Throughout this paper we will assume H$_0$=71 km s$^{-1}$ Mpc$^{-1}$ which at the average redshift of A3158 (z=0.0597) means that 1 arcsecond $\\sim$ 1.1 kpc. \\begin{table} \\begin{center} \\label{optp} \\caption{Known cluster parameters for A3158. Column 1 and 2 give the parameter and the most recent value from the literature. The reference is given in column 3 where 1 means Havlen \\& Quintana~(1978), 2 is Ebeling et al.~(1996), 3 is Lokas et al.~(2006) and 4 is Fleenor et al.~(2006).} \\begin{tabular}{lll} \\hline Parameter & Value & Reference\\\\ \\hline redshift & 0.0597 & 3\\\\ ellipticity & 0.3 & 1\\\\ X-ray Luminosity & 5.31 $\\times 10^{44}$ erg s$^{-1}$ & 2\\\\ velocity & 17,910 kms$^{-1}$ & 4\\\\ Redshift n=145 & 0.0597 & 3\\\\ velocity disp. (n=145)& 970$\\pm$57 kms$^{-1}$ & 3 \\\\ Viral Mass & $15.4^{+7.6}_{-5.4} 10^{14}$M$_{\\sun}$ & 3\\\\ \\hline \\end{tabular} \\end{center} \\end{table} ", "conclusions": "We present the the first of a series of papers investigating radio emission in the Horologium-Reticulum Supercluster. In this paper we present radio imaging at 1.4 and 2.5 GHz of the cluster A3158, from which we detect 109 radio sources of which 64 are found to have optical counterparts. By considering a reduced subsample of 87 sources within a $23\\farcm05$ radius, over which the primary beam attenuation of the ATCA is negligible, we are able to investigate the statistical properties of radio emission in the cluster. Our results can be summarised as follows. (1) The radio source counts in A3158 are unlikely to be consistent with the background source counts and in particular there appears to be an excess of galaxies with S$_{1.4} \\leq$ 3.96 mJy; a KS test of the entire sample against the background only returns a $\\sim$3\\% - 8\\% chance that the distributions are drawn from the same population, while a KS test over the sample with S$_{1.4} \\leq$ 3.96 mJy gives a much higher chance ($\\sim$60\\%) that the distributions are drawn from the same population suggesting that the low-powered excess is the significant difference. (2) The percentage of radio galaxies associated with the cluster as determined by spectroscopic redshifts is highly correlated to radio flux/power which is consistent with the excess source counts being the result of low-powered radio galaxies associated with A3158; in addition colours of these galaxies suggest they are blue which is likely to indicate starformation as a driving mechanism. (3) The alignment of the candidate starforming galaxies associated with the increased source counts follows the bridge of galaxies connecting A3158 to the A3125/A3128 complex suggestive of merger-induced formation. (4) The radio luminosity function for early-type galaxies is significantly lower than for other composite cluster environments and the field; this result is similar to the results for the late merger in A3558 obtained by Venturi et al.~(2000). Thus, if we can use the radio galaxy population as an indicator of merger history it is likely that A3158 is observed in a late merger state with a significant excess of low-powered, blue galaxies aligned along the axis connecting this cluster to the A3125/A3128 complex. These galaxies are likely the result of merger-induced starformation." }, "0807/0807.4323_arXiv.txt": { "abstract": "Astrometric and radial-velocity planet detections track very similar motions, and one generally expects that the statistical properties of the detections would also be similar after they are scaled to the signal-to-noise ratio of the underlying observations. I show that this expectation is realized for periods small compared to the duration of the experiment $P/T\\ll 1$, but not when $P/T\\ga 1$. At longer periods, the fact that models of astrometric observations must take account of an extra nuisance parameter causes the mass error to begin deteriorating at $P/T\\sim 0.8$, as compared to $P/T\\sim 1.0$ for RV. Moreover, the deterioration is much less graceful. This qualitative difference carries over to the more complicated case in which the planet is being monitored in the presence of a distant companion that generates an approximately uniform acceleration. The period errors begin deteriorating somewhat earlier in all cases, but the situation is qualitatively similar to that of the mass errors. These results imply that to preserve astrometric discovery space at the longest accessible orbits (which nominally have the lowest-mass sensitivity) requires supplementary observations to identify or rule out distant companions that could contribute quasi-uniform acceleration. ", "introduction": "\\label{sec:intro}} Astrometric and radial-velocity (RV) planet detections are, from a mathematical standpoint, extremely similar. In each, one models 1-dimensional projections of Kepler orbits and attempts to fit Kepler parameters. In the limit of circular orbits, the planet signature is encoded in the simple form \\begin{equation} F(t;a_1,a_2,a_3) = a_1\\sin(a_2 t + a_3), \\label{eqn:basicform} \\end{equation} where \\begin{equation} a_1 = \\alpha \\quad {\\rm (astrometry)},\\qquad a_1 = K \\quad {\\rm (RV)}, \\label{eqn:a1eq} \\end{equation} are the astrometric and velocity semi-amplitudes for the two cases, and where \\begin{equation} a_2 = {2\\pi\\over P}, \\quad a_3 = \\phi \\qquad {\\rm (astrometry + RV)} \\label{eqn:a23eq} \\end{equation} designate the period $P$ and phase $\\phi$, respectively. One then goes on to combine information about the star's mass and (for astrometry) its distance, to infer the planet mass $m$ (astrometry) or $m\\sin i$ (RV), where $i$ is the inclination. In the astrometric case, there are of course two such equations, one for each direction in the plane of the sky, whose ratio gives $\\cos i$, and so permit one to break the $m\\sin i$ degeneracy that plagues the intrinsically 1-dimensional RV measurement. Because the form of equation (\\ref{eqn:basicform}) is essentially identical in the two cases, it is generally assumed that the error properties, i.e., the relation between the measurement errors and the derived-parameter errors, is also the same. Of course, it is well known that the parameter errors have a different dependence on semimajor axis $a$, system distance $D$, etc. Most notably, with other parameters held fixed, astrometric sensitivity increases linearly with $a$ whereas RV sensitivity declines as $a^{-1/2}$. But here I am referring to something else. The differences just mentioned all impact the final result because they change the characteristic signal-to-noise ratio of the experiment, \\begin{equation} {\\rm SNR}\\equiv {a_1\\over\\sigma}\\sqrt{N}, \\label{eqn:snrdef} \\end{equation} where $N$ is the number of measurements and $\\sigma$ is the error in each measurement (assumed for simplicity to be all the same). Here I will show that the astrometric and RV measurements have substantially different error properties even when SNR is identical. ", "conclusions": "\\label{sec:discussion}} The results derived here are primarily of interest in regard to future astrometric missions such as GAIA and SIM. The first point is that planets with periods that are even slightly longer than the mission cannot be reliably detected unless they are many times more massive than the nominal thresholds of detection. And second, if one is forced to allow for uniform acceleration, then the same statement applies to $P/T\\ga 0.7$. Hence, to preserve the discovery space at the longest accessible periods, it is critically important to identify or rule out distant perturbers by supplementary data, such as longer-term RV observations to look for large, distant planetary or stellar companions, or perhaps AO observations to find stellar companions. This should be relatively straightforward for SIM, with its limited number of high-precision targets, but may be more difficult for GAIA, which is a survey instrument." }, "0807/0807.3029_arXiv.txt": { "abstract": "We study an extension of the MSSM with a singlet supermultiplet $S$ with coupling $SH_1H_2$ in order to solve the $\\mu$ problem as in the NMSSM, and right-handed neutrino supermultiplets $N$ with couplings $SNN$ in order to generate dynamically electroweak-scale Majorana masses. We show how in this model a purely right-handed sneutrino can be a viable candidate for cold dark matter in the Universe. Through the direct coupling to the singlet, the sneutrino can not only be thermal relic dark matter but also have a large enough scattering cross section with nuclei to detect it directly in near future, in contrast with most other right-handed sneutrino dark matter models. ", "introduction": "Weakly interacting massive particles (WIMPs) are among the best motivated candidates for explaining the cold dark matter in the Universe. WIMPs appear in many interesting extensions of the standard model providing new physics at the TeV scale. Such is the case of supersymmetric models, in which imposing a discrete symmetry (R-parity) to avoid rapid proton decay renders the lightest supersymmetric particle (LSP) stable and thus a good dark matter (DM) candidate. The minimal supersymmetric extension of the standard model (MSSM) provides two natural candidates for WIMPs, the neutralino and the (left-handed) sneutrino. The neutralino is a popular and extensively studied possibility. On the contrary, the left-handed sneutrino in the MSSM~\\cite{Ibanez} is not a viable dark matter candidate. Given its sizable coupling to the $Z$ boson, they either annihilate too rapidly, resulting in a very small relic abundance, or give rise to a large scattering cross section off nucleons and are excluded by direct DM searches~\\cite{Falk:1994es} (notice however that the inclusion of a lepton number violating operator can reduce the detection cross section~\\cite{Hall:1997ah}). However, there is a strong motivation to consider an extension of the MSSM, the fact that neutrino oscillations imply tiny but non-vanishing neutrino masses. These can be obtained introducing right-handed neutrino superfields. Several models have been proposed to revive sneutrino DM by reducing its coupling with Z-boson. This can be achieved by introducing a mixture of left- and right-handed sneutrino~\\cite{ArkaniHamed:2000bq,Arina:2007tm,valle}, or by considering a purely right-handed sneutrino~\\cite{Asaka:2005cn,Gopalakrishna:2006kr,McDonald:2006if,Lee:2007mt}. In the former, a significant left-right mixture is realized by adopting some particular supersymmetry breaking with a large trilinear term~\\cite{ArkaniHamed:2000bq}. Such a mechanism is not available in the standard supergravity mediated supersymmetry breaking, where trilinear terms are proportional to the small neutrino Yukawa couplings. Recently, another realization of large mixing was pointed out~\\cite{valle} by abandoning the canonical see-saw formula for neutrino masses. On the other hand, pure right-handed sneutrinos cannot be thermal relics, since their coupling to ordinary matter is extremely reduced by the neutrino Yukawa coupling~\\cite{Asaka:2005cn,Gopalakrishna:2006kr,McDonald:2006if}. Furthermore, these would be unobservable in direct detection experiments. Another possibility to obtain the correct thermal relic density would consist of coupling the right-handed sneutrino to the observable sector, e.g., via an extension of the gauge \\cite{Lee:2007mt} or Higgs \\cite{pilaftsis,dp} sectors There is one more motivation to consider another extension of the MSSM, % the so-called ``$\\mu$ problem''~\\cite{Kim:1983dt}. The superpotential in the MSSM contains a bilinear term, $\\mu H_1 H_2$. Successful radiative electroweak symmetry breaking (REWSB) requires $\\mu$ of the order of the electroweak scale. The next-to-minimal supersymmetric standard model (NMSSM) offers a simple solution % by introducing a singlet superfield $S$ and promoting the bilinear term to a trilinear coupling $\\l S H_1 H_2$. After REWSB, % $S$ develops a vacuum expectation value (VEV) of order of the electroweak scale thereby providing an effective $\\mu$ term, $\\mu=\\lambda \\langle S\\rangle$. Furthermore, the NMSSM also alleviates the ``little hierarchy problem'' of the Higgs sector in the MSSM~\\cite{BasteroGil:2000bw} and has an attractive phenomenology, featuring light Higgses and interesting consequences for neutralino DM \\cite{Cerdeno:2004xw}. Although the $Z_3$ symmetry of the NMSSM may give rise to a cosmological domain wall problem, this can be avoided with the inclusion of non-renormalisable operators. Motivated by the above two issues, we study an extension of the MSSM where singlet scalar superfields are included % \\cite{ko99,pilaftsis}. A singlet $S$ in order to solve the $\\mu$ problem as in the NMSSM (and which accounts for extra Higgs and neutralino states) and right-handed neutrinos $N$ to obtain non-vanishing neutrino Majorana masses with the canonical, but low scale, see-saw mechanism. Terms of the type $SNN$ in the superpotential can generate dynamically Majorana masses through the VEV of the singlet $S$. In addition, the presence of right-handed sneutrinos, $\\tilde N$, with a weak scale mass provides a new possible DM candidate within the WIMP category. In this letter we analyse the properties of right-handed sneutrinos, showing that not only they can be thermally produced in sufficient amount to account for the DM in the Universe because of the direct coupling between $S$ and $N$, but also that their elastic scattering cross section off nuclei is large enough to allow their detection in future experiments. ", "conclusions": "We propose the right-handed sneutrino as a viable thermal DM candidate in an extension of the MSSM where the singlet superfields, $S$ and $N$, are included to solve the $\\mu$ problem and account for neutrino masses. A direct coupling between $S$ and $N$ provides a sufficiently large annihilation cross section for the right-handed sneutrino, as well as a detection cross section in the range of future direct DM searches. DGC was supported by the program ``Juan de la Cierva''. CM was supported by the MEC project FPA2006-05423 and the EU program MRTN-CT-2004-503369. DGC and CM were also supported by the MEC project FPA2006-01105, and the EU network MRTN-CT-2006-035863. OS was supported by the MEC project FPA 2004-02015. We thank the ENTApP Network of the ILIAS project RII3-CT-2004-506222 and the project HEPHACOS P-ESP-00346 of the Comunidad de Madrid." }, "0807/0807.2733_arXiv.txt": { "abstract": "s{ M\\,87 is the first extragalactic source detected in the Very High Energy (VHE; $E > 100$\\,GeV) $\\gamma$-ray domain that is not a blazar, its large scale jet not being aligned to the line of sight. Slight modification of standard emission models of TeV blazars allows to account for the $\\gamma$-ray spectra obtained with H.E.S.S. We present a multi-blob synchrotron self-Compton model taking explicitly into account large viewing angles and moderate values of the Lorentz factor as inferred from MHD simulations of jet formation. Predictions of the VHE emission for the nearby radiogalaxy Cen\\,A and an interpretation of the broadband radiation of M\\,87 are presented. } ", "introduction": " ", "conclusions": "Our scenario shows the ability to extend the interpretation of TeV blazars by leptonic models to active galactic nuclei with intermediate beaming which jet is not strictly aligned with the line of sight. The {\\it GLAST} mission and H.E.S.S.\\,II project, and CTA in a near future, will allow to probe the sub-TeV spectral region, which is still poorly known and yet essential to constrain the inverse Compton bump. This is required to distinguish between leptonic or hadronic origin of the high energy emission of the jet in these objects." }, "0807/0807.2443_arXiv.txt": { "abstract": "We present and discuss the bounds from the energy conditions on a general $f(R)$ functional form in the framework of metric variational approach. As a concrete application of the energy conditions to locally homogeneous and isotropic $f(R)-$cosmology, the recent estimated values of the deceleration and jerk parameters are used to examine the bounds from the weak energy condition on the free parameter of the family of $f(R)=\\sqrt{R^2 - R_0^2}\\;$ gravity theory. ", "introduction": "The observed late-time acceleration of the Universe poses a great challenge to modern cosmology, which may be the result of unknown physical processes involving either modifications of gravitation theory or the existence of new fields in high energy physics. This latter route is most commonly used, however, following the former, an attractive approach to this problem, known as $f(R)-$gravity,\\cite{Kerner} examines the possibility of modifying Einstein's general relativity (GR) by adding terms proportional to powers of the Ricci scalar $R$ to the Einstein-Hilbert Lagrangian (see also Refs.~\\refcite{Francaviglia} for recent reviews). Although these theories provide an alternative way to explain the observed cosmic acceleration without dark energy, the freedom in the choice of different functional forms of $f(R)$ gives rise to the problem of how to constrain on theoretical and/or observational grounds, the many possible $f(R)-$gravity theories. Theoretical limits have long been discussed in Refs.~\\refcite{Amendola}, while only recently observational constraints from several cosmological datasets have been explored for testing the viability of these theories.\\cite{Fabiocc} Additional constraints to $f(R)$ theories may also arise by imposing the so-called energy conditions.\\cite{Kung} It is well known that these conditions, initially formulated in GR context,\\cite{Hawking} have been used in different contexts to derive general results that hold for a variety of physical situations.\\cite{EC} More recently, several authors have employed the GR classical energy conditions to investigate cosmological issues such as the phantom fields\\cite{Santos} and the expansion history of the universe.\\cite{Nilza} While they are well founded in the context of GR, one has to be cautious when using the energy conditions in a more general framework, such as the $f(R)-$gravity. In this regard, in a recent work Santos \\emph{et al.}\\cite{SARC} have used Raychaudhuri's equation along with the requirement that gravity is attractive, to derive the energy conditions for a general $f(R)-$gravity in the metric formulation. They have shown that, although similar, the energy conditions differ from their formulation in GR context. In this work, we use estimated values of the deceleration and jerk parameters to examine the bounds from these newly derived $f(R)-$energy-conditions on the one-parameter family of a recently proposed $f(R)-$gravity.\\cite{Baghram-a,Baghram-b} ", "conclusions": "" }, "0807/0807.1056_arXiv.txt": { "abstract": "The GMRT reionization effort aims to map out the large scale structure of the Universe during the epoch of reionization (EoR). Removal of polarized Galactic emission is a difficult part of any 21 cm EoR program, and we present new upper limits to diffuse polarized foregrounds at 150 MHz. We find no high significance evidence of polarized emission in our observed field at mid galactic latitude (J2000 08h26m+26). We find an upper limit on the 2-dimensional angular power spectrum of diffuse polarized foregrounds of $[l^2C_l/2\\pi]^{1/2}<3\\,$K in frequency bins of width $\\delta\\nu=1\\,$MHz at $3000.03h\\,$Mpc$^{-1}$, $k<0.1h\\,$Mpc$^{-1}$. This can be compared to the expected EoR signal in total intensity of $[k^3P(k)/2\\pi^2]^{1/2}\\sim 10\\,$mK. We find polarized structure is substantially weaker than suggested by extrapolation from higher frequency observations, so the new low upper limits reported here reduce the anticipated impact of these foregrounds on EoR experiments. We discuss Faraday beam and depth depolarization models and compare predictions of these models to our data. We report on a new technique for polarization calibration using pulsars, as well as a new technique to remove broadband radio frequency interference. Our data indicate that, on the edges of the main beam at GMRT, polarization squint creates $\\sim$ 3\\% leakage of unpolarized power into polarized maps at zero rotation measure. Ionospheric rotation was largely stable during these solar minimum night time observations. ", "introduction": "A current frontier in observational and theoretical cosmology is the search for and understanding of large scale structure during the epoch of reionization (EoR). Detection of reionization at $z\\sim 10$ would allow study of the emergence of the first abundant luminous objects. Such observations holds the promise of examination of astrophysical and cosmological processes at epochs as early as a few hundred million years after the start of the expansion of the Universe. The WMAP satellite has measured polarization in the Cosmic Microwave Background (CMB) at large angular scales. This polarization is believed to arise from Thomson scattering of the CMB photons near the EoR \\citep{2008arXiv0803.0593N,2008arXiv0803.0547K}. The observed optical depth $\\tau\\sim 0.089\\pm 0.016$ corresponds to an instantaneous reionization redshift of $z_{\\rm reion}=10.8\\pm 1.4$. One way to study the reionization transition in detail is by imaging redshifted 21cm emission. At redshifts above the EoR transition the gas is neutral and is predicted to glow with about a 25 mK sky brightness temperature. After reionization is complete this glow is absent. At redshifts close to the transition a patchy sky is expected. Simulations \\citep{2008MNRAS.tmp...77I} suggest that with existing telescopes a measurement near 150 MHz may allow for statistical detection of $\\sim$ 20 Mpc patchiness in the neutral hydrogen. This detection would pin down the reionization redshift and begin the process of more detailed study of the transition. The 21 cm EoR signal, in the 10 milliKelvin range, is much smaller than the expected $\\sim 10$K patchiness due to Galactic synchrotron emission and extragalactic foregrounds. One of the most difficult challenges of 21 cm EoR astronomy is the removal of these bright foregrounds. The proposed strategy makes use of the very different spatial and spectral character of the two components of sky structure. Since the received frequency of the 21 cm line emission encodes the distance of the source, the EoR signal consists of a series of source screens that are largely independent. For example, if one detects a patchy layer of cosmic material at 150 MHz due to the EoR transition, another patchy layer must be present at 151 MHz, but these two images come from distinct radial shells and the bright and dim patches in the two images are not expected to be aligned on the sky. The characteristic size of the ionization structures near the epoch of ionized-cell overlap is approximately 10 $h^{-1}$ Mpc, which corresponds to 5 arc minutes in angular size, or about 1 MHz in radial distance for the redshifted 21cm line at $z \\sim 9$. In contrast, the foreground emission is highly correlated from one frequency to the next, since most radio-bright structures emit with a very smooth synchrotron spectrum. To remove the foregrounds some type of frequency-differencing technique will be needed. At frequencies above 300 MHz many radio sources with a synchrotron spectrum have been shown to be linearly polarized, and this raises a concern for the frequency differencing scheme. In propagation through the interstellar medium the polarization rotates on the sky because of the Faraday effect. In the polarization basis of the telescope this means that each linear polarization component can oscillate with frequency. If the oscillation period falls close to the frequency separation used for the foreground subtraction, incomplete removal of Galactic synchrotron emission may leave a residual that masks the EoR signal. One way to avoid the Faraday rotation problem is to measure total intensity, described by the Stokes I parameter, using a instrument that rejects the linearly polarized Stokes components Q and U. However, all real instruments have various sources of leakage between these components. To design these instruments and plan future EoR observations one needs information on the polarized sky brightness in the EoR band near 150 MHz. Recent studies at 350 MHz \\citep{2006AN....327..487D} report that the polarized component has much more structure on arc minute angular scales than the total intensity. Polarized structure amounting to several Kelvin has been reported. Scaling using a synchrotron spectral index of 2.6 \\citep{2008arXiv0802.1525D}, the polarized structure could be tens of Kelvin in the EoR band. This means that instrumental leakage from Q/U to I could severely limit the ability to remove Galactic emission. This paper places new constraints on polarized sky brightness in the EoR range of frequencies near 150 MHz. We find the level of polarized sky brightness is well below that expected by extrapolation from higher frequencies. We discuss possible mechanisms which would lead to this strong depolarization, and suggest future techniques to differentiate between them. ", "conclusions": "The GMRT EoR data has resulted in the strongest upper limits to polarized emission at 150 MHz to date. The observed polarization is an order of magnitude smaller than had been expected from higher frequency observations. We note that comparison was not made on the same fields. We estimated the expected beam and depth polarization effects, but these are model dependent and poorly constrained by the data. Within the uncertainties, either mechanism could account for the lack of observed polarization. If beam depolarization were the cause, it would imply a significant small scale magnetic field with structure at $l>10,000$. Others have reported depth depolarization at similar RM$\\lambda^2$, so depth depolarization may play a role here as well. When compared to the sensitivities required for EoR observations, the upper bounds we report are still two orders of magnitude larger than the sensitivities required. The measured polarization leakage at GMRT is less than 10\\%, so a modest leakage correction, accurate to 10\\%, should make EoR observations feasible. The low polarized emission we report is promising for the search for Reionization." }, "0807/0807.3053_arXiv.txt": { "abstract": "{The chronology of bulge and disk formation is a major unsolved issue in galaxy formation, which impacts on our global understanding of the Hubble sequence.} {We analyse colours of the nuclear regions of intermediate redshift disk galaxies, with the aim of obtaining empirical information of relative ages of bulges and disks at $0.1 < z < 1.3$.} {We work with an apparent-diameter limited parent sample of 248 galaxies from the HST Groth Strip Survey. We apply a conservative criterion to identify bulges and potential precursors of present-day bulges based on nuclear surface brightness excess above the exponential profile of the outer parts and select a sample of 56 galaxies with measurable bulges. We measure bulge colours on wedge profiles opening on the semi-minor axis least affected by dust in the disk, and compare them to disk, and global galaxy colours.} {For $60\\%$ of galaxies with bulges, the rest-frame nuclear colour distribution shows a red sequence that is well fit by passive evolution models of various ages, while the remainder $40\\%$ scatters towards bluer colours. In contrast, galaxies without central brightness excess show typical colours of star forming population and lack a red sequence. We also see that, as in the local Universe, most of the minor axis colour profiles are negative (bluer outward), and fairly gentle, indicating that nuclear colours are not distinctly different from disk colours. This is corroborated when comparing nuclear, global and disk colours: these show strong correlations, for any value of the central brightness prominence of the bulge. No major differences are found between the low and high inclination samples, both for the bulge and non-bulge samples.} % {Comparison with synthetic models of red sequence bulge colours suggests that such red bulges have stopped forming stars at an epoch earlier than $\\sim 1$ Gyr before the observation. The correlation between nuclear and disk colours and the small colour gradients hints at an intertwined star formation history for bulges and disks: probably, most of our red bulges formed in a process in which truncation of star formation in the bulge did not destroy the disk.} ", "introduction": "\\label{sec:introduction} Three reasons make bulge population ages useful for galaxy formation studies. In all inside-out scenarios, the bulge regions should contain information on the earliest phases of star formation in galaxies. Also, when compared to ages of elliptical galaxies, bulge ages should shed light on whether bulges and ellipticals share a formation history as proposed in CDM-based semi-analytical models \\citep{Baugh96,Kauffmann96bul}. Finally, bulge population ages, when compared to ages of the disks they live in, should provide clues on whether bulges formed before disks, again as assumed in current CDM-based models, or after disks, as proposed in secular evolution models \\citep{Kormendy04}. The current consensus emerging from local Universe studies points to bulges being uniformly old ($\\sim$10 Gyr) in early- to intermediate-type galaxies. \\citet{Peletier99} show that optical-NIR colours are indistinguishable from those of ellipticals in Coma, suggesting similar ages of at most 2 Gyr younger than Coma ellipticals. Independently, from stellar colour-magnitude diagrams rather than integrated colours, for the Milky Way bulge \\citet{Zoccali03} infer homogeneous stellar ages similar to those of globular clusters, or $\\ga$10 Gyr. Bulges of later-type spirals have signs of harboring younger populations, and ongoing star formation; \\citep[e.g.][]{Peletier99,Carollo01}. Studies of local galaxies have also shown that redder bulges live in redder disks \\citep{Balcells94}, i.e., colour-wise, bulges are more similar to their parent disks than to each other \\citep{Peletier96}. This suggests common evolution, although not necessarily that one forms from the other. Because the bulge ages just mentioned do not appear to be much older than the oldest stars in nearby mature disks (about 10~Gyr for the MW disk), the relative chronology of disk and bulge formation appears difficult to determine from the fossil record in nearby galaxies. Studying galaxies at cosmologically-significant look-back times offers a means to date galaxian components seen at epochs closer to that of their formation, and may provide better time resolution to determine age differences between bulges and ellipticals, and between bulges and disks. This approach has been followed by several groups in recent years, with conflicting conclusions. \\citet{Ellis01} and \\citet{Menanteau01} find a prevalence ($30 - 50\\%$) of young field spheroids at $z \\sim 1$ in the two HDFs. In contrast, \\citet{Koo05} emphasized that $85\\%$ of luminous ($M_B < -19$) field bulges at redshifts $z \\sim 0.8$ in the Groth strip are nearly as red ($U - B \\sim 0.50$) as local E/SO's, a result that suggests both an old metal-rich dominant population and a 'drizzling' contribution of star formation to the global colours. \\citet{Koo05HDF} compared bulge colours for field and cluster galaxies, concluding that bulge colours are identical for field and cluster galaxies. \\citet{MacArthur07}, who isolated bulges for a sample of 137 spiral galaxies within the redshift range $0.1 1$) concentrate strongly at 'very red' colours, $(U-B) > 0.25$, defining a red sequence, which fits well passive evolution models of different ages. The remainder $40\\%$ span over a bluer and wider colour range. \\item Galaxies without central brightness excess ($\\eta < 1$) show bluer colours and lack a red sequence. Galaxies show a colour bimodality that is related to the central brigthness excess, in the sense that more concentrated galaxies are redder. \\item Nuclear and global colours are very similar to each other: redder bulges are embedded in redder galaxies. Disk colours also correlate with bulge colours. As a result, the colour difference between bulges and disks is smaller than that between bulges of different galaxies. Colour profiles are smooth, and do not show a discontinuity at the bulge-disk boundary. \\item The correlations between nuclear and disk regions are fulfilled for all types of galaxies, independently of the central brightness excess and the colour. The colour profiles are equally smooth in the samples with ($\\eta > 1$) and without ($\\eta > 1$) central concentration. \\item A major merger origin for bulges, as proposed by others, needs to be able to account for the similarities in bulge and disk colours. Such similarities lend support to models with coeval growth of bulges and disks, and to bulge formation models that do not destroy a pre-existing disk. \\end{enumerate}" }, "0807/0807.1326_arXiv.txt": { "abstract": "As an alternative explanation of the dimming of distant supernovae it has recently been advocated that we live in a special place in the Universe near the centre of a large void described by a Lema\\^itre-Tolman-Bondi (LTB) metric. The Universe is no longer homogeneous and isotropic and the apparent late time acceleration is actually a consequence of spatial gradients in the metric. If we did not live close to the centre of the void, we would have observed a Cosmic Microwave Background (CMB) dipole much larger than that allowed by observations. Hence, until now it has been argued, for the model to be consistent with observations, that by coincidence we happen to live very close to the centre of the void or we are moving towards it. However, even if we are at the centre of the void, we can observe distant galaxy clusters, which are off-centre. In their frame of reference there should be a large CMB dipole, which manifests itself observationally for us as a kinematic Sunyaev-Zeldovich (kSZ) effect. kSZ observations give far stronger constraints on the LTB model compared to other observational probes such as Type Ia Supernovae, the CMB, and baryon acoustic oscillations. We show that current observations of only 9 clusters with large error bars already rule out LTB models with void sizes greater than $\\sim 1.5$ Gpc and a significant underdensity, and that near future kSZ surveys like the Atacama Cosmology Telescope (ACT), South Pole Telescope (SPT), APEX telescope, or the Planck satellite will be able to strongly rule out or confirm LTB models with giga parsec sized voids. On the other hand, if the LTB model is confirmed by observations, a kSZ survey gives a unique possibility of directly reconstructing the expansion rate and underdensity profile of the void. ", "introduction": "Distant supernovae appear dimmer than expected in a matter-dominated homogeneous and isotropic FRW universe. The currently favoured explanation of this dimming is the late time acceleration of the universe due to a mysterious energy component that acts like a repulsive force. The nature of the so-called Dark Energy responsible for the apparent acceleration is completely unknown. Observations seem to suggest that it is similar to Einstein's cosmological constant, but there is inconclusive evidence. There has been a tremendous effort in the last few years to try to pin down deviations from a cosmological constant, e.g.~with deep galaxy catalogues like 2dFGRS~\\cite{2dFGRS} and SDSS~\\cite{SDSS}, and extensive supernovae surveys like ESSENCE~\\cite{ESSENCE}, SNLS~\\cite{SNLS}, and SDSS-SN~\\cite{SDSS-SN}, and many more are planned for the near future e.g.~DES~\\cite{DES}, PAU~\\cite{PAU,Benitez2008}, BOSS~\\cite{BOSS} and JDEM~\\cite{JDEM}. In the meantime, our realisation that the universe around us is far from homogeneous and isotropic has triggered the study of alternatives to this mysterious energy. Since the end of the nineties it has been suggested by various groups~\\cite{Hellaby:1998,Celerier:1999hp,Tomita:2000jj,Moffat:2005yx, Garfinkle:2006sb,Enqvist:2007vb,Mattsson:2007tj,Wiltshire:2007zj,GBH:2008} that an isotropic but inhomogeneous Lema\\^itre-Tolman-Bondi universe could also induce an apparent dimming of the light of distant supernovae, in this case due to local spatial gradients in the expansion rate and matter density, rather than due to late acceleration. It is just a matter of interpretation which mechanism is responsible for the dimming of the light we receive from those supernovae. Certainly the homogeneous and isotropic FRW model is more appealing from a philosophical point of view, but so was the static universe and we had to abandon it when the recession of galaxies was discovered at the beginning of last century. There is nothing wrong or inconsistent with the possibility that we live close to the centre of a giga parsec scale void. Such a void may indeed have been observed as the CMB cold spot~\\cite{Cruz:2006sv,Cruz:2006fy,Cruz:2008sb} and smaller voids have been seen in the local galaxy distribution \\cite{Frith:2003,Granett:2008}. The size and depth of the distant observed voids, i.e.~$r_0 \\sim 2$ Gpc and $\\Omega_M \\sim 0.2$ within a flat Einstein-de Sitter universe, seems to be consistent with that in which we may happen to live~\\cite{Tully:2007tp}, and could account for the supernovae dimming, together with the observed baryon acoustic oscillations and CMB acoustic peaks, the age of the universe, local rate of expansion, etc.~\\cite{GBH:2008}. Moreover, according to the theory of eternal inflation~\\cite{Linde:1993xx}, rare fluctuations at the Planck boundary may be responsible for the non-perturbative amplification of local inhomogeneities in the metric, which would look like local voids in the matter distribution~\\cite{Linde:1994gy}. In the eternal inflation approach one assumes to be a typical observer whose local patch comes directly from a rare fluctuation within an inflationary domain at the Planck scale. Moreover, since it is a rare fluctuation it should be highly spherically symmetric, and the theory predicts that we should live close to the centre of such a void~\\cite{Linde:1994gy}. The size and depth of those voids depends on the theory of inflation on very large scales and may be a probe (perhaps the only probe) of the global structure of the universe. In fact, observations suggest that if there is such a large void, we should live close to the centre, otherwise our anisotropic position in the void would be seen as a large dipole in the CMB. Of course, we do observe a dipole, but it is normally assumed to be due to the combined gravitational pull of the Virgo cluster, and the Shapley super cluster. There is always the possibility that we live off-centre and we are moving towards the centre of the void, so that the two effects are partially cancelled, giving rise to the observed dipole. However, such a coincidence could not happen for all galaxies in the void and, in general, clusters that are off-centred should see, in their frame of reference, a large CMB dipole. Such a dipole would manifest itself observationally for us as an apparent kinematic Sunyaev-Zeldovich effect for the given cluster. It is the purpose of this paper to study the very strong constraints that present observations of the kSZ effect already put on the LTB void models, and predict how near future observations from kSZ surveys like ACT or SPT will strongly rule out (or confirm) LTB models with giga parsec sized voids. In section~2 we describe the general LTB void models, giving the corresponding Einstein-Friedmann equations, as well as parameterisations of their solutions. In a subsection we describe the GBH constrained model, where we assume the Big Bang is homogenous, and thus the model depends on a single function, the inhomogeneous matter ratio $\\Omega_M(r)$. In section~3 we study the induced dipole for off-centred clusters and compute the size of the analogue velocity of those clusters depending on the parameters of the void model. In section~4 we analyse present observations and give constraints on the model from current observations. In section~5 we then explore the prospects that future experiments like ACT, SPT and Planck will provide for strongly constraining or even ruling out LTB models of the universe. Finally, in section~6 we give our conclusions. ", "conclusions": "As an alternative explanation of the dimming of distant supernovae it has recently been advocated that we live in a special place in the Universe near the centre of a large void. The universe is no longer homogeneous and isotropic and the apparent late time acceleration is actually a consequence of spatial gradients in the metric. If we did not live close to the centre of the void, we would have observed a CMB dipole much larger than that allowed by observations. Hence, until now it has been argued, for the model to be consistent with observations, that by coincidence we happen to live very close to the centre of the void. However, even if we are at the centre of the void, we can observe distant galaxy clusters, which are off-centre. In their frame of reference there should be a large CMB dipole, which manifests itself observationally for us as a kinematic Sunyaev-Zeldovich effect. In this paper we have studied the induced dipole for off-centred clusters due to the different trajectories of photons from the last scattering surface, and computed the size of the corresponding apparent velocity of those clusters with respect to the rest frame of the CMB LSS, depending on the parameters of the void model. We then analysed the present observations of the kSZ effect in a handful of clusters and gave very strong constraints on the size of the void in LTB models. In fact, for our specific constrained-GBH model the bounds are in conflict with other constraints from supernovae, baryon acoustic oscillations and CMB, at least at the 3-$\\sigma$ level, see Fig.~\\ref{fig:otherdata}, and therefore we conclude that the constrained-GBH void model is practically ruled out if the current interpretation of kSZ observations are correct. At present, kSZ data leave small voids ($r_0 \\lesssim 800$ Mpc) unconstrained, independently of the inner density contrast ($\\Omega_{\\rm in}$) used, simply because the radius is so small that the void does not impact the cluster with the lowest redshift in the sample, see Table~1. In order to put limits on small voids with large density contrasts, or sudden transitions, we would need to use local large scale structure data that can be checked for homogeneity, or alternatively low redshift supernovae. Current kSZ observations are limited in numbers, and are still not precise enough to make a single positive detection of the peculiar velocity of a cluster at the 2-$\\sigma$ level, see Fig.~\\ref{fig:obs}. Hence, one cannot rule out that they are plagued by large systematic and/or random errors. This is about to change in the coming years, and we have made predictions of how systematic near-future observations from kSZ surveys like ACT or SPT could strongly rule out all LTB models with giga parsec sized voids. In the case that the LTB models were confirmed, the average apparent velocity profile as a function of distance would give a direct handle on the density and expansion rate of the void. This unique relationship makes by far kSZ observations the most powerful data for constraining LTB models. If the sensitivity of experiments like ACT, SPT and Planck are as planned, and the present systematic errors are under control, it is expected that large voids will definitely be ruled out at many sigma. If these experiments yield 10 (100) well observed clusters we could reasonably expect to rule out voids of 800 (500) Mpc radius at the 3-$\\sigma$ level. We hope to come back to this severe challenge to LTB models and redo the analysis with the new data in the near future." }, "0807/0807.1921_arXiv.txt": { "abstract": "The Guitar Nebula is an H$\\alpha$ nebula produced by the interaction of the relativistic wind of a very fast pulsar, PSR B2224+65, with the interstellar medium. It consists of a ram-pressure confined bow shock near its head and a series of semi-circular bubbles further behind, the two largest of which form the body of the Guitar. We present a scenario in which this peculiar morphology is due to instabilities in the back flow from the pulsar bow shock. From simulations, these back flows appear similar to jets and their kinetic energy is a large fraction of the total energy in the pulsar's relativistic wind. We suggest that, like jets, these flows become unstable some distance down-stream, leading to rapid dissipation of the kinetic energy into heat, and the formation of an expanding bubble. We show that in this scenario the sizes, velocities, and surface brightnesses of the bubbles depend mostly on observables, and that they match roughly what is seen for the Guitar. Similar instabilities may account for features seen in other bow shocks. ", "introduction": "\\label{sec:intro} Many pulsars travel at high speed, and the collision between their relativistic winds and the interstellar medium leads to the formation of bow shocks. These shocks are observed most readily at X-ray and optical wavelengths: the shocked relativistic wind will emit mostly synchrotron radiation, while the shocked interstellar medium will emit -- if it is partially neutral -- copious H$\\alpha$ emission (for a review, see \\citealt{gaens06}). Arguably the most spectacular bow shock is the Guitar Nebula, made by one of the fastest pulsars known, PSR~B2224+65 \\citep*{cordrm93}. This H$\\alpha$ nebula has, as the name implies, a guitar-like shape, with a bright head, a faint neck, and a body consisting of two larger bubbles (see \\Fref{guitar}). \\cite{cordrm93} suggest this morphology might reflect variations in either the pulsar energy injection rate or the interstellar medium density. In this {\\em Letter}, we investigate whether instead the peculiar morphology could be due to instabilities in the jet-like flow of pulsar effluvium away from the bow shock. Fast back flows are a natural consequence of bow shocks: the pulsar wind is greatly heated at the shock, which, for the usual case where cooling is slow, leads to a high pressure that drives a flow in the only direction available, to the back. From simulations (e.g., \\citealt{buccadz05}), the flows seem similar to jets, being well-collimated and fast, and seem to carry most of the pulsar wind energy. Jet-like flows are indeed seen in X-ray observations, which also show that only a small fraction of the energy is radiated (for a review, \\citealt{kargp08}). So far, the simulations have not extended far to the back, but if the back flow is similar to a jet, one might expect it to become unstable further downstream. From simulations of jets (e.g., \\citealt{bodo+98}), this instability would lead to mixing with the ambient medium and rapid dissipation of the kinetic energy into heat. The simulations have not followed what happens beyond this initial mixing, but it seems plausible that the material would expand rapidly and drive a bubble. If so, it might initially expand faster than the pulsar motion, and gain further energy from the jet. With time, however, it will slow down, and once the pulsar has moved sufficiently far ahead, the jet will become so long that it becomes unstable before reaching the bubble, and a new bubble will be formed. We suggest the body of the guitar is made up of two such bubbles, while another one has just started to form near the head. In \\Sref{model}, we describe our model in more detail, and in \\Sref{guitar} we compare it with the properties of the Guitar Nebula, finding qualitative agreement. In \\Sref{discussion}, we discuss implications as well as ways in which our model could be tested. ", "conclusions": "We found that we could roughly reproduce the Guitar Nebula assuming the jet-like back flow from the pulsar bow shock becomes unstable and dissipates rapidly, causing expanding bubbles. If this were to happen generally, one might expect other sources with jets or bow shocks to show Guitar-like bubbles, yet none appear to be known. For jet sources, this may not be surprising: many jets are denser than the medium they move through, and hence more stable, and disruptions that do occur may be difficult to distinguish from, e.g., changes in jet orientation. For other bow shocks, the absence of bubbles may partly be a selection effect: most have much larger stand-off radii than the Guitar, and hence any bubbles would be at correspondingly larger distances, where they might be missed, especially as they would be fainter than the bow shock (or even invisible if the expansion velocity became too low or if radiative effects became important; both perhaps relevant especially for stellar wind bow shocks). The one possible exception is PSR B0740$-$28, which has a H$\\alpha$ bow shock with a relatively small stand-off radius of $\\theta_0=1\\farcs0$ as well as ``shoulders'' further behind \\citep{jonesg02}. If related to an instability, one infers $15\\lesssim\\lambda/\\theta_0\\lesssim60$, of the same order as we see for the Guitar. It would be interesting to obtain deeper images further behind the bow shock. In some pulsar bow shocks, the shocked pulsar wind is observed directly, by its synchrotron emission (for an overview, \\citealt{kargp08}). For many, including the Guitar \\citep{huib07}, emission is seen only close to the pulsar, likely at the pulsar wind termination shock. Some, however, have much longer tails. The longest belongs to the ``Mouse,'' associated with PSR J1747$-$2958. This nebula, with $\\theta_0\\approx\\!0\\farcs75$, shows a bulbous structure $\\sim\\!1\\farcm5$ behind the pulsar (the Mouse's body), but also a smooth, straight tail of $12\\arcmin$, without a clear end \\citep{gaen+04}. Scaling with the stand-off distances, one might identify the Mouse's body with the equivalent of the Guitar's head bubble. The long tail has a size equivalent to the bottom bubble, but, apart from changes in polarisation, shows little structure \\citep{yuseb87}. This would seem inconsistent with any bubbles being formed, and thus is puzzling in the context of our model. For two other pulsar bow shocks with long tails, the observations match expectations better. For PSR J1509$-$5850, with $\\theta_0\\approx0\\farcs5$, the X-ray tail extends for $\\gtrsim\\!5\\farcm6$ and shows clear structure, with a change in brightness at $1\\farcm3$, a kink at~$3\\arcmin$, and a bright radio spot coincident with its end point \\citep{huib07b,karg+08}. Comparing with the large bubbles in the Guitar, the typical length scale of $\\sim\\!1\\farcm5$ for the knots and kinks is about a factor 3 larger, roughly consistent with the ratio of the stand-off distances. For PSR B1929+10, with $\\theta_0\\simeq2\\farcs3$, the tail extends up to $10\\arcmin$ and again shows substantial structure, with brightenings at $\\sim\\!2\\arcmin$ and $\\sim\\!5\\arcmin$, the latter coincident with a radio feature \\citep{beck+06,misapg07}. Again scaling with the stand-off radii, the $5\\arcmin$ feature could be similar to the head bubble in the Guitar. Overall, we conclude that our model of instabilities in a bow shock back flow roughly reproduces observations of the Guitar Nebula, without the need to appeal to variations in the density of the ambient medium, nor to energy sources beyond what is expected to be carried by the back flow. It also seems consistent with what is seen in other pulsar bow shocks. The model could be tested further both with observations and simulations. Observationally, one test would be to measure the expansion velocities in the Guitar bubbles, either by determining proper motions, or by spectroscopy (from the broad component of the H$\\alpha$ profile, as done for non-radiative shocks in supernova remnants; \\citealt{raym91}). Given the observed H$\\alpha$ surface brightness, this would allow one to estimate the ambient density, which should be similar to that at the location of the bow shock in our model, but substantially lower if the bubbles reflect density variations \\citep{cordrm93,chatc04}. Simulations of bow shocks that extend to larger scales might show whether instabilities in fact lead to bubbles or rather to more continuous structure, or whether perhaps the process is sufficiently stochastic that both can occur (possibly leading to a shape like the Guitar's neck). If bubbles form, the simulations might also shed light on details of the morphology, such as the closed appearance at the back of the head and bottom bubbles." }, "0807/0807.2325_arXiv.txt": { "abstract": "In this work, exact solutions of static and spherically symmetric space-times are analyzed in $f(R)$ modified theories of gravity coupled to nonlinear electrodynamics. Firstly, we restrict the metric fields to one degree of freedom, considering the specific case of $g_{tt}\\,g_{rr}=-1$. Using the dual $P$ formalism of nonlinear electrodynamics an exact general solution is deduced in terms of the structural function $H_P$. In particular, specific exact solutions to the gravitational field equations are found, confirming previous results and new pure electric field solutions are found. Secondly, motivated by the existence of regular electric fields at the center, and allowing for the case of $g_{tt}\\,g_{rr}\\neq -1$, new specific solutions are found. Finally, we outline alternative approaches by considering the specific case of constant curvature, followed by the analysis of a specific form for $f(R)$. ", "introduction": "A central theme in cosmology is the perplexing fact that the Universe is undergoing an accelerated expansion \\cite{expansion}. Several candidates, responsible for this expansion, have been proposed in the literature, in particular, dark energy models and modified gravity. Amongst the modified theories of gravity, models generalizing the Einstein-Hilbert action have been proposed, where a nonlinear function of the curvature scalar, $f(R)$, is introduced in the action. These modified theories of gravity seem to provide a natural gravitational alternative to dark energy, and in addition to allow for a unification of the early-time inflation \\cite{Starobinsky:1980te} and late-time cosmic speed-up \\cite{Carroll:2003wy,Fay:2007gg}. These models seem to explain the four cosmological phases \\cite{Nojiri:2006be}. They are also very useful in high energy physics, in explaining the hierarchy problem and the unification of GUTs with gravity \\cite{Cognola:2006eg}. The possibility that the galactic dynamics of massive test particles may be understood without the need for dark matter was also considered in the framework of $f(R)$ gravity models \\cite{darkmatter}. One may also generalize the action by considering an explicit coupling between an arbitrary function of the scalar curvature, $R$, and the Lagrangian density of matter \\cite{Nojiri:2004bi}. Note that these couplings imply the violation of the equivalence principle \\cite{Olmo:2006zu}, which is highly constrained by solar system tests. A fundamental issue extensively addressed in the literature is the viability of the proposed $f(R)$ models \\cite{viablemodels,Hu:2007nk,Sokolowski:2007pk}. In this context, it has been argued that most $f(R)$ models proposed so far in the metric formalism violate weak field solar system constraints \\cite{solartests}, although viable models do exist \\cite{Hu:2007nk,solartests2,Sawicki:2007tf,Amendola:2007nt}. The issue of stability \\cite{Faraoni:2006sy} also plays an important role for the viability of cosmological solutions \\cite{Nojiri:2006ri,Sokolowski:2007pk,Amendola:2007nt,Boehmer:2007tr, deSitter}. In the context of cosmological structure formation observations \\cite{structureform}, it has been argued that the inclusion of inhomogeneities is necessary to distinguish between dark energy models and modified theories of gravity, and therefore, the evolution of density perturbations and the study of perturbation theory in $f(R)$ gravity is of considerable importance \\cite{Boehmer:2007tr,Tsuj,Uddin:2007gj,Bazeia:2007jj}. A great deal of attention has also been paid to the issue of static and spherically symmetric solutions of the gravitational field equations in $f(R)$ gravity \\cite{Clifton:2005aj,SSSsol,Multamaki:2006zb}. Solutions in the presence of a perfect fluid were also analyzed \\cite{Multamaki:2006ym}, where it was shown that the pressure and energy density profiles do not uniquely determine $f(R)$. In addition to this, it was found that matching the exterior Schwarzschild-de Sitter metric to the interior metric leads to additional constraints that severely limit the allowed fluid configurations. An interesting approach in searching for exact spherically symmetric solutions in $f(R)$ theories of gravity was explored in \\cite{Capozziello:2007wc}, via the Noether Symmetry Approach, and a general analytic procedure was developed to deal with the Newtonian limit of $f(R)$ gravity in \\cite{Capozziello:2007ms}. Analytical and numerical solutions of the gravitational field equations for stellar configurations in $f(R)$ gravity theories were also presented \\cite{Kainulainen:2007bt,Henttunen:2007bz,Multamaki:2007jk}, and the generalized Tolman-Oppenheimer-Volkov equations for these theories were derived \\cite{Kainulainen:2007bt}. In the context of $f(R)$ modified theories of gravity, it was recently shown that power-law inflation and late-time cosmic accelerated expansion can be explained by a modified $f(R)$-Maxwell theory \\cite{Bamba:2008ja}, due to breaking the conformal invariance of the electromagnetic field through a non-minimal gravitational coupling. It is interesting to note that such a coupling may generate large-scale magnetic fields. Motivated by these ideas, we consider in this work $f(R)$ gravity coupled to nonlinear electrodynamics, and endeavor to search for exact solutions in a static and spherically symmetric set-up. In contrast to a non-minimal gravitational coupling, here conformal invariance is not broken. In the context of nonlinear electrodynamics, a specific model was proposed by Born and Infeld in 1934 \\cite{BI} founded on a \\emph{principle of finiteness}, namely, that a satisfactory theory should avoid physical quantities to become infinite. The Born-Infeld model was inspired mainly to remedy the fact that the standard picture of a point particle possesses an infinite self-energy, and consisted on placing an upper limit on the electric field strength and considering a finite electron radius. Later, Pleba\\'{n}ski explored and presented other examples of nonlinear electrodynamic Lagrangians \\cite{Pleb}, and showed that the Born-Infeld theory satisfies physically acceptable requirements. Furthermore, nonlinear electrodynamics have recently been revived, mainly because these theories appear as effective theories at different levels of string/M-theory, in particular, in D$p-$branes and supersymmetric extensions, and non-Abelian generalizations (see Ref.~\\cite{Witten} for a review). Much interest in nonlinear electrodynamic theories has also been aroused in applications to cosmological models \\cite{cosmoNLED}, in particular, in explaining the inflationary epoch and the late-time accelerated expansion of the universe \\cite{Novello}. It is interesting to note that the first {\\it exact} regular black hole solution in general relativity was found within nonlinear electrodynamics \\cite{Garcia,Garcia2}, where the source is a nonlinear electrodynamic field satisfying the weak energy condition, and recovering the Maxwell theory in the weak field limit. In fact, general relativistic static and spherically symmetric space-times coupled to nonlinear electrodynamics have been extensively analyzed in the literature: regular magnetic black holes and monopoles \\cite{Bronnikov1}; regular electrically charged structures, possessing a regular de Sitter center \\cite{Dymnikova}; traversable wormholes \\cite{Arellano1} and gravastar solutions \\cite{Arellano3}. Thus, as mentioned above, motivated by recent work on a non-minimal Maxwell-$f(R)$ gravity model \\cite{Bamba:2008ja}, in this paper $f(R)$ modified theories of gravity coupled to nonlinear electrodynamics are explored, in the context of static and spherically symmetric space-times. This paper is outlined in the following manner: In section \\ref{sec2:action}, the action of $f(R)$ gravity coupled to nonlinear electrodynamics is introduced, and the respective gravitational field equations and electromagnetic equations are presented. In section \\ref{sec3:simple}, we restrict the metric fields to one degree of freedom, by considering the specific case of $g_{tt}=-g_{rr}^{-1}$, and using the dual $P$ formalism of nonlinear electrodynamics, we present exact solutions in terms of the structural function $H_P$. Subsequently, in section \\ref{sec4:2dof} we investigate the situation where the two metric fields are related via a power law in $r$, introducing additional parameters, and derive new specific solutions. In section \\ref{sec5:alternative}, we present alternative methods of finding exact solutions, first by considering the specific case of constant curvature, then by choosing a form for the $f(R)$, before we conclude in section \\ref{sec6:conclusion}. ", "conclusions": "\\label{sec6:conclusion} The issue of exact static and spherically symmetric solutions in $f(R)$ modified theories of gravity is an important theme, mainly due to the analysis of weak field solar system constraints, and the generalization of exact general relativistic solutions to $f(R)$ gravity. In this work we have analyzed exact solutions of static and spherically symmetric space-times in $f(R)$ modified theories of gravity coupled to nonlinear electrodynamics. Firstly, the metric fields were restricted to one degree of freedom, by considering the specific case of $g_{tt}=-g_{rr}^{-1}$. Using the dual $P$ formalism of nonlinear electrodynamics, an exact general solution was found in terms of the structural function $H_P$. In particular, exact solutions to the gravitational field equations were found, confirming previous results and new pure electric field solutions were deduced. Secondly, by allowing two degrees of freedom for the metric fields, and motivated by the existence of regular electric fields at the center, new solutions were found. Finally, we have also briefly considered alternative approaches by analyzing the specific case of constant curvature and secondly, by considering a specific form for $f(R)$." }, "0807/0807.0266_arXiv.txt": { "abstract": "We present a time-dependent radiative model for the atmosphere of the transiting planets that take into account the eccentricity of their orbit. We investigate the temporal temperature and flux variations due to the planet-star distance variability. We will also discuss observational aspects with {\\it Spitzer} measurements. ", "introduction": "{\\underline{\\it Why taking into account the eccentricity}}? The major theoretical models of atmospheric structure of exoplanets consider the planet only into two conditions: \\begin{itemize} \\item{} Either the flux received by the planet coming from the star is averaged over an hemisphere of the planet in a static case ; \\item{} or some sort of redistribution day to night of the incoming flux is included. \\end{itemize} In none of this cases the fact that the flux actually varies with time is taken into account. Even in the case of the less eccentric planet, WASP-10 b, with a low eccentricity of 0.057 (\\cite[Christian et al. 2008]{Christian08}) this represents a variation $\\sim$~25\\% in the received flux between periastron and apoastron We focused our study to two planets: HD~17156b, the most eccentric transiting planet, and even if HD~80606b is not transiting, we choose this planet with the largest eccentricity to show the extreme case. Tab.~1 summarize the parameters of these two systems. \\begin{table} \\begin{center} \\caption{Planetary and stellar parameters} \\label{tab1} {\\scriptsize \\begin{tabular}{l|l||c||c} \\hline &Parameters&HD 17156b &HD80606~b\\\\ \\hline Orbit% & $e$ & 0.67 & 0.93 \\\\ & $a$ [AU] & 0.159 & 0.432 \\\\ & $d_{\\rm min}$ [AU] & 0.052 &0.030 \\\\ & $d_{\\rm max}$ [AU] &0.266 & 0.834\\\\ & $P$ [days] &21.22 & 111.78 \\\\ & $\\omega$ [$^\\circ$]& 121 &300\\\\ \\hline Planet % & Radius % [\\rj] & 0.96 & 1.1$^{\\rm a}$\\\\ & Mass % [\\mj] & 3.11 & 5$^{\\rm b}$ \\\\ & $P_{\\rm spin}$ [days] & 3.8 &1.93\\\\ \\hline Star % & Type & G0V & G5\\\\ & Radius [\\rsol]&1.47 &1.05 \\\\ &[Fe/H] & 0.24 & 0.43\\\\ \\hline Refs.& & \\cite{Gillon07}& \\cite{LL08}\\\\ \\hline \\end{tabular}\\label{tab:param} } \\end{center} \\vspace{1mm} \\scriptsize{ {\\it Notes:}\\\\ $^{\\rm a}$ This planet is not transiting. The value of the radius has been predicted by (4) ; \\\\ $^{\\rm b}$We adopted this value of the mass (inclination unknown). } \\end{table} ", "conclusions": "We presented a still under development time-dependent radiative model for the atmosphere of the transiting planets. This model takes into account the eccentricity of their orbit and in its next version will be able to completely map the planet, including the rotation effects on the insolation. The preliminary results shows already that an eccentric planet in under heating conditions is very different that if it were sitting at its semi major axis, even for a low eccentricity even if several steps need to be completed before reaching a fully self-consistent model." }, "0807/0807.2780_arXiv.txt": { "abstract": "We report the discovery that substructures/subhaloes of a galaxy-size halo tend to fall in together in groups in cosmological simulations, something that may explain the oddity of the MW satellite distribution. The original clustering at the time of infall is still discernible in the angular momenta of the subhaloes even for events which took place up to eight Gyrs ago, $z \\sim 1$. This phenomenon appears to be rather common since at least $1/3$ of the present-day subhaloes have fallen in groups in our simulations. Hence, this may well explain the Lynden-Bell \\& Lynden-Bell ghostly streams. We have also found that the probability of building up a flattened distribution similar to the MW satellites is as high as $\\sim 80\\%$ if the MW satellites were from only one group and $\\sim 20\\%$ when five groups are involved. Therefore, we conclude that the `peculiar' distribution of satellites around the MW can be expected with the CDM structure formation theory. This non-random assignment of satellites to subhaloes implies an environmental dependence on whether these low-mass objects are able to form stars, possibly related to the nature of reionization in the early Universe. ", "introduction": "The discrepancy in numbers between the substructures/subhaloes resolved in a galaxy-size cold dark matter (CDM) halo and the satellites around the Milky Way (MW) has been a long standing issue for the `concordance' CDM structure formation theory. It implies a non-trivial mapping between the luminous satellites and the dark matter subhaloes at the (sub) galactic scales \\cite[(Zentner \\etal\\, 2005;]{zentner05} \\cite[Libeskind \\etal\\, 2007)]{libeskind07} through the astrophysics processes with baryons or the theory needs a major modification at a fundamental level (see e.g. \\cite[Kamionkowski \\& Liddle 2000)]{kl00}. In the past ten years new attention has been drawn to the dynamical properties of the MW satellites. Starting with \\cite{lyndenbell95}, the existence of ghostly streams of satellites (dwarf galaxies and globular clusters) was postulated. These objects would share similar energies and angular momenta producing a strong alignment along great circles on the sky. Recently, the anisotropic distribution of satellites around the MW has been argued to be a problem for the CDM theory \\cite[(Kroupa, Theis \\& Boily, 2005;]{kroupa05} \\cite[Metz, Kroupa \\& Jerjen, 2007)]{metz07}. The MW has approximately 20 satellites forming a disk-like structure while the simulated dark matter subhaloes usually distribute almost isotropically. Here we report our findings of subhaloes falling in groups in dark matter simulations and its application on explaining the oddities of the dynamical properties of MW satellites, namely, the Lynden-Bell \\& Lynden-Bell ghostly steams and the great MW satellites disk. Researches in the past showed that clusters of galaxies are built of galaxies coming in groups \\cite[(Knebe \\etal\\, 2004)]{knebe04}, but it was not clear whether a similar picture also applies at the (sub)galactic scale. We refer readers to a more detailed description of our analyses and discussions on the group infall and its link to the environment in \\cite{lihelmi08}. ", "conclusions": "We have revisited the issue of the peculiar distribution and properties of the MW satellites and their link to the dark matter subhaloes. In particular, we have focused on the infall of substructures on to a Milky-Way like dark matter halo in a $\\Lambda$CDM cosmogony utilising a series of high-resolution dark-matter simulations. We have found evidence of group infall on to the MW-like halo, which may explain the ghostly streams proposed by \\cite{lyndenbell95}. We have also explored how this planar configuration may be obtained as a result of the infall of satellites in groups. The observed correlation in the angular momentum orientation of subhaloes naturally gives rise to disk-like configurations. For example, we find that if all subhaloes are accreted from just one group, a disk-like distribution is essentially unavoidable ($\\sim 80\\%$ probability), while for accretion from just two groups, the likelihood of obtaining a distribution as planar as observed is 40\\%. Therefore the disky configuration of satellites is consistent with CDM if most satellites have their origin in a few groups. Note that in our studies, we do not need to invoke the baryon-related physics to account for the dynamical properties of the MW satellites. Thus both the `ghostly streams' and the `planar configuration' are manifestations of the same phenomenon: the hierarchical growth of structure down to the smallest galactic scales. One of the possible implications of the reality of the ghostly streams is that its member galaxies formed and evolved in a similar environment before falling into the MW potential. This would have implications on the (oldest) stellar populations of these objects, such as for example, sharing a common metallicity floor \\cite[(Helmi \\etal\\, 2006)]{helmi06}. On the other hand, this implies that there should be groups that have failed to host any luminous satellites. This would hint at a strong dependence on environment on the ability of a subhalo to retain gas \\cite[(Scannapieco \\etal\\, 2001)]{scannapieco01}, or be shielded from re-ionization by nearby sources \\cite[(Mashchenko, Carignan \\& Bouchard 2004;]{mcb04} \\cite[Weinmann \\etal\\, 2007)]{weinmann07}. Recent proper motion measurements of the Large and Small Magellanic clouds by \\cite{smc-mu}, as well as the simulations by \\cite{bc05} suggest that these systems may have become bound to each other only recently. This would be fairly plausible in the context of our results. The Clouds may well have been part of a recently accreted group and it may not even be necessary for them to ever have been a binary system." }, "0807/0807.0971_arXiv.txt": { "abstract": "{ We find a new analytical solution for the chemical evolution equations, taking into account the delayed contribution of all low and intermediate mass stars (LIMS) as one representative star that enriches the interstellar medium. This solution is built only for star formation rate proportional to the gas mass in a closed box model. We obtain increasing C/O and N/O ratios with increasing O/H, behavior impossible to match with the Instantaneous Recycling Approximation (IRA). Our results, obtained by two analytical equations, are very similar to those found by numerical models that consider the lifetimes of each star. This delayed model reproduces successfully the evolution of C/O$-$O/H and $Y-O$ relations in the solar vicinity. This analytical approximation is a useful tool to study the chemical evolution of elements produced by LIMS when a galactic chemical evolutionary code is not available. } \\resumen{ Encontramos una nueva soluci\\'on anal\\'{\\i}tica para las ecuaciones de evoluci\\'on qu\\'{\\i}mica tomando en cuenta la contribuci\\'on retrasada de todas las estrellas de $m < 8$ \\msun (LIMS) como una estrella representativa que enriquece al medio interestelar. Esta soluci\\'on es construida para tasa de formaci\\'on estelar proporcional a la masa de gas en un modelo de caja cerrada. Obtenemos incrementos en C/O y N/O cuando O/H aumenta, comportamiento imposible de igualar con IRA. Nuestros resultados, obtenidos por dos ecuaciones anal\\'{\\i}ticas, son muy similares a aquellos encontrados por modelos num\\'ericos que consideran el tiempo de vida de cada estrella. Este modelo retrasado reproduce la evoluci\\'on de C/O-O/H y $Y-O$ en la vecindad solar. Esta aproximaci\\'on anal\\'{\\i}tica es una herramienta util para estudiar la evoluci\\'on de elementos producidos por las LIMS cuando no se dispone de un c\\'odigo de evoluci\\'on qu\\'{\\i}mica. } \\addkeyword{galaxies: abundances} \\addkeyword{galaxies: evolution} \\addkeyword{ISM: abundances} \\begin{document} ", "introduction": "\\label{sec:intro} Chemical evolution models are used to describe the temporal variation of the gas mass and the abundances of the different chemical elements that are present in the gas. Their importance relies on the fact that the chemical history of the studied object (i.e. interstellar medium in galaxies and intergalactic medium) can be inferred. Moreover it is possible to get chemical information about stellar population properties and characteristics of the galaxies that we observe in the low redshift Universe. The chemical evolution equations, shown by Tinsley (1974) and corrected later by Maeder (1992), are relatively complex and can be solved through numerical models. There are some analytical approximations that have the advantage of predicting the general behavior of chemical elements in a quick and easy way but some precision may be lost. One of the most well known analytical approximations is the Instantaneous Recycling Approximation (IRA) (Talbot \\& Arnett 1971) where the star lifetimes are negligible compared with the age of the galaxies. This approximation has been widely used because it simplifies the solution to the chemical evolution equations, however, massive star (MS) lifetimes are on the order of $10^{6} - 10^{7}$ years while the lifetimes for the low and intermediate mass stars (LIMS) are on the order of $10^{8} - 10^{10}$ years comparable to the lifetime of a galaxy. Hence, IRA provides only a very rough approximation for elements produced by LIMS. Serrano \\& Peimbert (1983) proposed an analytic approximation related to the delays in chemical enrichment in N/O-O/H relation assuming N and O yields increase with $Z$. They present closed and open models (with and without gas flows, respectively) that takes into account the delay on nitrogen production due to LIMS concluding that nitrogen must be an element mainly secondary. Later on, Pagel (1989) presented another approximation for open models introducing an arbitrary time delay in order to study the chemical evolution of element produced by SNIa and LIMS, such as Fe and Ba (through s-process). This time delay term makes the stars release the processed material to the interstellar medium (ISM) at a single time after the star formation. After this single time delay, the contribution of all type of star is instantaneous, like IRA. The objective of this work is to present an alternative analytical solution to the chemical evolution equations that considers the LIMS lifetimes as only one group with delay times during the whole evolution. Those delay times are different for each chemical element and are computed based on the characteristics of the stellar yields and the stellar population. The delayed contribution to the chemical enrichment of elements produced by this type of star at different times should give results with a precision intermediate between the results obtained by using IRA and the ones obtained by numerical codes. As an application, this work aims to reproduce the C/O \\textit{vs} O/H and $Y(O)$ histories indicated by the stars and HII regions at the solar vicinity. This approximation is a simple tool for theoretical and observational astronomers who need to include chemical aspects in their computations or to interpret observational data when they do not have access to a numerical code of chemical evolution of galaxies. Our equation for the mass abundance of element $i$, $X_i(t)$, would replace the popular equation of $Z(t)$ obtained assuming IRA in the closed box regimen, erroneously applied for elements produced by LIMS (e.g. He, C, N). ", "conclusions": "\\begin{itemize} \\item{We have found analytical equations for chemical evolution in the case of a closed box model and SFR proportional to the gas mass where the delayed enrichment by LIMS is represented by a single type of star. } \\item{The delay of LIMS with respect to the galactic enrichment for He, C and N produces an increase on C/O and N/O with O/H and in Y with O in good agreement with the results obtained by numerical models. With IRA, the C/O and N/O values are constant with O/H in disagreement with observed data and with model results that take into account all star lifetimes.} \\item{For $\\mu = 0.1$ and $Z_{pop}=0.02$, $Y(O)$, C/O, and N/O values show artificial secondary raises due to the O dilution produced by LIMS and not because of the increase of He, C and N produced by the LIMS.} \\item{The delayed approximation was probed successfully in the solar vicinity reproducing the main trends of C/O-O/H relation shown by dwarf stars in agreement with results obtained with a numerical model that considers all star lifetimes. That relation cannot be reproduced at all by the instantaneous recycling approximation.} \\item{ The analytical equation (eq. 5) obtained by our approximation is a useful tool to know the chemical evolution of those elements produced by LIMS when no galactic chemical evolutionary code is available. } \\end{itemize} {\\it Part of this work was submitted in the Physics Undergraduate Program at the Universidad Nacional Aut\\'onoma de M\\'exico.}" }, "0807/0807.2145_arXiv.txt": { "abstract": "We present a source and lens reconstruction for the optical Einstein ring gravitational lens system RXS J1131-1231. We resolve detail in the source, which is the host galaxy of a $z=0.658$ quasar, down to a resolution of 0.045 arc seconds (this is the size of the smallest conclusively resolved structures, rather than the pixel scale), using a Bayesian technique with a realistic model for the prior information. The source reconstruction reveals a substantial amount of complex structure in the host galaxy, which is $\\sim$ 8 kpc in extent and contains several bright compact substructures, with the quasar source residing in one of these bright substructures. Additionally, we recover the mass distribution of the lensing galaxy, assuming a simply-parameterised model, using information from both the quasar images and the extended images. This allows a direct comparison of the amount of information about the lens that is provided by the quasar images in comparison to the extended images. In this system, we find that the extended images provide significantly more information about the lens than the quasar images alone, especially if we do not include prior constraints on the central position of the lens. ", "introduction": "Gravitational lensing can be used as a powerful astrophysical tool for probing the density profiles of galaxies, and is one of the few ways in which dark matter can be detected \\citep[e.g.][]{2005MNRAS.363.1136K}. In addition, it often magnifies source objects by one to two orders of magnitude. This allows us to use the intervening gravitational lens as a kind of natural telescope, magnifying the source so that we can observe more detail than we would have been able to without the lens. This extra magnification provided by lensing has been very beneficial to studies of star formation and galaxy morphology at high redshift. Regions of the galaxy size and luminosity distribution that are inaccessible in unlensed observations are made (more) visible by lensing \\citep[e.g.][]{2000ApJ...528...96P, wayth, 2006ApJ...651....8B, 2007ApJ...671.1196M, 2008arXiv0804.4002D}. The properties of the lens galaxies (typically elliptical galaxies) can also be inferred from their lensing effect \\citep[e.g.][]{2006ApJ...649..599K, 2008arXiv0806.1056T}. Of course, gravitational lensing distorts the image of the source, as well as magnifying it. Thus, techniques have been developed that aim to infer the mass profile of the lens galaxy and the surface brightness profile of the source, given observed images \\citep[e.g.][]{2003ApJ...590..673W, 2006ApJ...637..608B}. The aim of this paper is to carry out this process with the recently discovered gravitationally lensed quasar/host galaxy system RXS J1131-1231 \\citep{2003A&A...406L..43S}. This system consists of a quadruply imaged quasar at redshift $z=0.658$ lensed by a galaxy at $z=0.295$. At the time of its discovery, it was the closest known lensed quasar, with some evidence for an extended optical Einstein ring - the image of the quasar host galaxy. Initial simple modelling suggested that the quasar source was magnified by a factor of $\\sim$ 50. Thus, subsequent observations with the ACS aboard the Hubble Space Telescope \\citep[][hereafter C06]{2006A&A...451..865C} allow the recovery of the morphology of the quasar's host galaxy down to a resolution of about 0.01 arc seconds - at least in principle, for the parts of the source nearest the caustic. Indeed, C06 presented a wide array of results based on HST observations (at 555nm and 814nm with ACS, and 1600nm with NICMOS), including a detailed reconstruction of the extended source. The source reconstruction method used by C06 is based on lensing the image back to a pixellated grid in the source plane, setting the source surface brightnesses to equal the image surface brightness, and using a decision rule (in this case, the median) to decide on the value of a source pixel whenever two or more image pixels disagree about the value of the same source pixel. If the point spread function (PSF) is small or the image has been deconvolved (in C06, the deconvolution step was neglected for the purposes of the extended source reconstruction) and the lens model is correct, this method can expect to be quite accurate. However, in principle, the uncertainty in the lens model parameters and the deconvolution step should always be taken into account. In this paper, we focus our attention on the 555nm ACS image (the drizzled images, as reduced by C06, were provided to us), and the process by which we reconstruct the original, unlensed source from it. Any differences between our reconstruction and the C06 one can be attributed to the following advantages of our approach: PSF deconvolution, averaging over the lens parameter uncertainties, simultaneous fitting of all parameters, and the prior information that Bayesian methods are capable of taking into account: in the case of our model, that is the knowledge that the majority of pixels in an astrophysical sources should be dark \\citep{2006ApJ...637..608B}. The 555nm image is also of particular interest because its rest wavelength (335nm) probes regions of recent star formation in a galaxy with an AGN. In the case of the Einstein Ring 0047-2808 \\citep{2006ApJ...651....8B}, our method was able to resolve structure down to scales of $\\sim$ 0.01 arcsec, a factor of five smaller than that obtainable in an unlensed observation with the Hubble Space Telescope and about double the resolution obtained by \\citet{2005ApJ...623...31D} using adaptive pixellation and least squares {\\it applied to exactly the same data}. This was possible because we used a prior distribution over possible sources that is more realistic as a model of our knowledge of an unknown astrophysical source, that is, we took into account the fact that it should be a positive structure against a dark background, a fact many methods (such as least squares and some popular regularisation formulas) implicitly ignore \\citep{2006ApJ...637..608B}. These differences between methods are likely to be most significant when data are sparse or noisy, and all methods tend to agree as the data quality increases and we approach the regime where the observed image uniquely determines the correct reconstruction. ", "conclusions": "In this paper, we have presented a detailed gravitational lens reconstruction of the optical extended source in the Einstein Ring RXS J1131-1231. The source is a medium sized galaxy ($\\sim$ 8 kpc in visible extent) with several compact bright emission regions. The substructures we found are in general agreement with those found by C06 in terms of their position, but we have shown that they are brighter and more compact than was previously determined. In addition, our reconstruction provides a clearer view of the substructures, including near the central regions of the source. The quasar resides in a bright emission region with an extent of about $\\sim$ 0.15 arcseconds or 1 kpc. It should be noted that the wavelength of the observations in the rest frame is 335 nm, so this reconstruction traces regions of recent star formation in the source galaxy. We have also directly compared point images vs extended images with regard to how well each is able to constrain the lens model. We found that there is a significant gain to be made in taking into account all of the information from the extended images. It has been suggested that this is not true in general \\citep{2007arXiv0710.3159F}, although it really depends on the resolution and number of extended images, which in this case is high. Certainly, in using both, there is nothing to lose but CPU cycles. This system has the potential to become one of the most well-constrained gravitational lenses, with multiple images of the extended ring, quasar image positions and flux ratios in multiple bands, and time delay measurements available \\citep{timedelay, keeton}. Hence, it should be possible to carry out a detailed kiloarametric study of its mass profile to shed some light on the dark matter halo of the lens galaxy. This paper was based on a single image of this system, the 555nm ACS image. Other HST images at different wavelengths (814nm, 1.6$\\mu$m) are available (C06) and can further constrain the lens model. Simultaneous multi-wavelength reconstructions are now becoming routine \\citep[e.g.][]{2007ApJ...671.1196M}. However, all of the structures in these images are in the same locations, and so a multi-wavelength reconstruction would not produce significantly different conclusions to those reached here. C06 note that in the near infrared image, the compact bright images are less pronounced compared to the diffuse background, which is what would be expected if the substructures are regions dense in hot young stars. This study has relied on a number of common assumptions that future research will seek to relax. Extending lens reconstruction techniques to incorporate kiloparametric models of the source and the lens simultaneously is an ambitious task, but some steps are already being taken in that direction \\citep{2008arXiv0804.2827S}. Flexible lens modelling plus information from time delay measurements and other sources would be extremely valuable for studies of galaxy dark matter haloes. Also, explicit modelling of dust absorption by the lens galaxy is proving to be an important ingredient in the inversion of Einstein Rings and would be an essential part of future work on this system." }, "0807/0807.0189_arXiv.txt": { "abstract": "We have observed a region of nebulosity first identified as starlight scattered by interstellar dust by Sandage (1976) using the GALEX ultraviolet imaging telescope. Apart from airglow and zodiacal emission, we have found a diffuse UV background of between 500 and 800 \\phunit\\ in both the \\galex\\ FUV (1350 -- 1750 \\AA) and NUV (1750 -- 2850 \\AA) bands. Of this emission, up to 250 \\phunit\\ is due to \\htwo\\ fluorescent emission in the FUV band. The remainder is consistent with scattering from interstellar dust with forward scattering grains of albedo about 0.4. These are the highest spatial resolution observations of the diffuse UV background to date and show an intrinsic scatter beyond that expected from instrumental noise alone. Further modeling is required to understand the nature of this scatter and its implications for the ISM. ", "introduction": "Ever since the first observations of diffuse ultraviolet radiation by \\citet{Hay69} and \\citet{Lillie}, there has been an effort to understand its distribution and its origin. Unfortunately, because of the difficulty of the observations and the faintness of the background, many of the early observations were conspicuous more by their disagreements than by the light they shed on the topic. The state of the observations and theories before 1990 have been reviewed by \\citet{Bowyer91} and \\citet{RCH91}. There has been significant progress in more recent years, particularly in the far ultraviolet ($\\lambda < $ 1200 \\AA) where \\citet{JM99} and \\citet{JM04} have used spectroscopic data from the \\voyager\\ and \\fuse\\ ({\\it Far Ultraviolet Spectroscopic Explorer}) spacecraft, respectively, to trace the radiation field over many different locations in the sky. There have also been a number of observations at longer wavelengths, most recently by the SPEAR instrument \\citep[][and references therein]{Ryu}, but no systematic study of the UV background. The {\\it Galaxy Evolution Explorer} (\\galex) offers us the opportunity to extend coverage of the diffuse background to a significant fraction of the sky with a sensitivity of better than 100 \\phunit. In this work, we will report on one such observation: that of the nebulosity observed near M82 by \\citet{Sandage76}, as a template for our further work with a much larger data set. This cloud is at a high Galactic latitude (38\\degr) with few nearby stars and was identified by \\citet{Sandage76} as a canonical high latitude dusty cloud illuminated by the Galactic interstellar radiation field (ISRF). Our expectation was that we would be able to differentiate between starlight scattered from the Galactic cloud and extragalactic light which would be shadowed by the foreground cloud. The observations presented here are the first to probe the diffuse UV background at a spatial resolution comparable to other surveys of dust emission, notably the IR. ", "conclusions": "We have obtained the highest spatial resolution images of the diffuse UV background to date with an effective spatial resolution of about 2\\arcmin. These observations are in a region of moderate optical depth with $\\tau\\ > 1$ where \\citet{Sandage76} observed, and correctly identified, starlight scattered from interstellar dust. We have obtained \\galex\\ observations in both the FUV (1350 -- 1750 \\AA) and NUV (1750 -- 2850 \\AA) bands. After subtraction of the foreground airglow and zodiacal light, we were left with about 500 - 800 \\phunit\\ in both the FUV and NUV bands. The FUV/NUV ratio increased with increasing FUV emission suggesting the presence of \\htwo\\ fluorescent emission in the FUV ranging up to an integrated emission of about 250 \\phunit. Our values are consistent with those of \\citet{MHB90} in their scan of this region; however, we observe a spatial variability that was not possible with their observations. When we link our data with the observations of G251.2+73.3 by \\citet{Haikala}, we find that the scattered UV light increases linearly with the IR emission for low optical depths but saturates at optical depths near unity. This is as expected given that the thermal IR emission is from the entire volume of the cloud because of the low optical depth in the IR. We have used the same models as in our earlier studies of the diffuse background and have found that a dust scattered component with $a = 0.4; g = 0.7$ is consistent with the data; ie., the dust is strongly forward scattering with a moderate albedo. However, the data show an intrinsic scatter much greater than can be attributed to photon noise alone which must reflect structure at a spatial scale of at least 2\\arcmin\\, possibly due to variations in the ISRF and in the distribution of the interstellar dust. We are now extending our analysis to a much larger body of \\galex\\ observations, both our own and archival data. Such an investigation will help resolve some of the uncertainties in this work such as the contribution of airglow and the zodiacal light. Previous observations of the scattering by interstellar dust were on much larger spatial scales and indicated a local origin to much of the background; ie., both nearby hot stars and interstellar dust were required. \\galex\\ observations will allow us to probe the diffuse background at much higher spatial resolutions and thus to investigate the small scale structure of the ISM." }, "0807/0807.0376_arXiv.txt": { "abstract": "The GAMA survey aims to deliver 250,000 optical spectra (3--7\\AA~resolution) over 250 sq.\\ degrees to spectroscopic limits of $r_{AB} <19.8$ and $K_{AB}<17.0$ mag. Complementary imaging will be provided by GALEX, VST, UKIRT, VISTA, HERSCHEL and ASKAP to comparable flux levels leading to a definitive multi-wavelength galaxy database. The data will be used to study all aspects of cosmic structures on 1kpc to 1Mpc scales spanning all environments and out to a redshift limit of $z \\approx 0.4$. Key science drivers include the measurement of: the halo mass function via group velocity dispersions; the stellar, HI, and baryonic mass functions; galaxy component mass-size relations; the recent merger and star-formation rates by mass, types and environment. Detailed modeling of the spectra, broad SEDs, and spatial distributions should provide individual star formation histories, ages, bulge-disc decompositions and stellar bulge, stellar disc, dust disc, neutral HI gas and total dynamical masses for a significant subset of the sample ($\\sim 100$k) spanning both the giant and dwarf galaxy populations. The survey commenced March 2008 with 50k spectra obtained in 21 clear nights using the Anglo Australian Observatory's new multi-fibre-fed bench-mounted dual-beam spectroscopic system (AA$\\Omega$). ", "introduction": "Galaxy And Mass Assembly (GAMA) is a major expansion of the Millennium Galaxy Catalogue (MGC) survey (Liske et al 2003; Allen et al 2005; Driver et al 2005) and a natural extension of the extremely productive nearby ``Legacy'' surveys (e.g., SDSS, 2MASS, HIPASS etc). In comparison to the superb SDSS survey GAMA will only sample 250 sq degrees of sky but will extend to significantly fainter spectroscopic limits (12$\\times$ the redshift density of SDSS main, 5$\\times$ stripe 82), to higher spatial ($0.6''$ FWHM) and spectral (3---7\\AA) resolutions, as well as moving to a far broader wavelength coverage (UV to Radio). GAMA has come about by parallel technological developments leading to a suite of new facilities whose survey sensitivities, resolutions, and capabilities are reasonably well matched. Until now the study of galaxies has generally been restricted to either large samples of limited wavelength data or multi-wavelength studies of small (and often biased) samples. However galaxy systems are extremely complex and diverse, exhibiting strong environmental and mass dependencies and containing distinct but interlinked components (AGN, nucleus, bulge, pseudo-bulge, bar, disc etc) and constituents (SMBH, plasma, stars, gas, dust etc). It then follows that a clear understanding of galaxy formation and evolution may only come about via the construction of a comprehensive survey which simultaneously samples all of these facets. The GAMA team aims to provide this data. In addition to the provision of a generic galaxy database, the GAMA project also includes a number of more focussed science goals, in particular: \\begin{description} \\item{1. } Measurement of the Halo Mass Function via virialised group velocity dispersions to directly test the {\\it numerical} prediction from CDM (and WDM) simulations. \\item{2. } Measurement of the dynamic, baryonic, HI and stellar mass functions to LMC masses versus redshift, environment, type, and component (as well as higher order relations, e.g., mass-spin [$M-\\lambda$]). \\item{3. } Measurement of the recent merger rates and star formation rates versus type, mass and environment over a 3---4 Gyr baseline. \\end{description} ", "conclusions": "" }, "0807/0807.0006_arXiv.txt": { "abstract": "The orbital parameters of extra-solar planets have a significant impact on the probability that the planet will transit the host star. This was recently demonstrated by the transit detection of HD 17156b whose favourable eccentricity and argument of periastron dramatically increased its transit likelihood. We present a study which provides a quantitative analysis of how these two orbital parameters effect the geometric transit probability as a function of period. Further, we apply these results to known radial velocity planets and show that there are unexpectedly high transit probabilities for planets at relatively long periods. For a photometric monitoring campaign which aims to determine if the planet indeed transits, we calculate the significance of a null result and the subsequent constraints that may be applied to orbital parameters. ", "introduction": "There have been at least five cases in which planetary transits were detected through photometric follow-up of planets already known via their radial velocity (RV) discoveries. The case of HD 17156b (\\cite{bar07a}) is of particular interest since it is a 21.2 day period planet which happens to have a large eccentricity ($e = 0.67$) and an argument of periastron which places the periapsis of its orbit in the direction toward the observer and close to parallel to the line of sight, resulting in an increased transit probability. Recent work by \\cite{bar07b} and \\cite{bur08} showed that higher eccentricities of planetary orbits will increase their transit probabilities and, consequently, expected yield for transit surveys. We demonstrate the combined effect of the eccentricity, $e$, and argument of periastron, $\\omega$, on transit probability. As shown by \\cite{kan07a}, the place in a planetary orbit where it is possible for a transit to occur (where the planet passes the star-observer plane perpendicular to the planetary orbit) is when $\\omega + f = \\pi / 2$. The probability of such a transit occurring, $P_t$, is given by \\begin{equation} P_t = \\frac{(R_p + R_\\star)(1 + e \\cos (\\pi/2 - \\omega))}{a (1 - e^2)}, \\label{transit_prob} \\end{equation} where $R_p$ and $R_\\star$ are the radii of the planet and star respectively, and $a$ is the semi-major axis. The orbital configuration, especially with regards to the values of $e$ and $\\omega$, plays a major role in determining the likelihood of a planet transiting the parent star. ", "conclusions": "" }, "0807/0807.0230_arXiv.txt": { "abstract": "We analyze the physical conditions of the outflow seen in QSO~2359--1241 (NVSS J235953--124148), based on high resolution spectroscopic VLT observations. This object was previously studied using Keck/HIRES data. The main improvement over the HIRES results is our ability to accurately determine the number density of the outflow. For the major absorption component, level population from five different \\ion{Fe}{2} excited level yields $\\vy{n}{H}=10^{4.4}$ cm$^{-3}$ with less than 20\\% scatter. We find that the \\ion{Fe}{2} absorption arises from a region with roughly constant conditions and temperature greater than 9000~K, before the ionization front where temperature and electron density drop. Further, we model the observed spectra and investigate the effects of varying gas metalicities and the spectral energy distribution of the incident ionizing radiation field. The accurately measured column densities allow us to determine the ionization parameter ($\\log U_H \\approx -2.4$) and total column density of the outflow ($\\log N_H(\\rm{cm}^{-2}) \\approx 20.6$). Combined with the number density finding, these are stepping stones towards determining the mass flux and kinetic luminosity of the outflow, and therefore its importance to AGN feedback processes. ", "introduction": "In recent years, the potential impact of quasar outflows on their environment has become widely recognized (e.g., Silk \\& Rees 1998, King 2003, Cattaneo et al.\\ 2005, Hopkins et al.\\ 2006). Outflows are detected as absorption troughs in quasar spectra that are blueshifted with respect to the systemic redshift of their emission line counterparts. The absorption troughs are mainly associated with UV resonance lines of various ionic species (e.g., \\ion{Mg}{2}~$\\lambda\\lambda$2796.35,2803.53, \\ion{Al}{3}~$\\lambda\\lambda$1854.72,1862.79, \\ion{C}{4}~$\\lambda\\lambda$1548.20,1550.77, \\ion{Si}{4}~$\\lambda\\lambda$1393.75,1402.77, \\ion{N}{5}~$\\lambda\\lambda$1238.82,1242.80). Some quasar outflows show absorption troughs from excited and metastable states. The ratio of the population level between the excited or metastable states and the ground state is sensitive to the number density and temperature of the plasma (Wampler, Chugai \\& Petitjean 1995; de~Kool et~al.\\ 2001). Therefore, accurate measurements of the column densities associated with both excited or metastable states and the ground state of a given ion can yield the gas number density of the outflow. In addition, these measurements and similar ones of troughs from other ions and elements allow us to determine the ionization equilibrium and total column density in the outflow (Arav et~al.\\ 2001; Arav et~al.\\ 2007). Accurate column densities for the outflow's troughs are difficult to determine since the outflow does not cover the emission source homogeneously (Barlow 1997; Telfer et~al.\\ 1998; Arav 1997; Arav et~al.\\ 2003). Over the past several years, we have developed techniques for extracting reliable column densities for such situations (Arav et~al.\\ 1999a; Arav et~al.\\ 1999b; de~Kool et~al.\\ 2001, 2002a,b; Arav et~al.\\ 2002; Scott et al.\\ (2004) Gabel et al.\\ (2005); Arav et~al.\\ 2005). These efforts culminated with the analysis of spectroscopic VLT observations of the outflow seen in QSO~2359--1241 (Arav et~al.\\ 2008; hereafter Paper~I). These data contain absorption troughs from five resonance \\ion{Fe}{2} lines, as well as those from several other metal species and metastable excited state \\ion{He}{1}, with a resolution of $\\sim$7 km~s$^{-1}$ and signal-to-noise ratio per resolution element of order 100. QSO~2359--1241 (NVSS J235953--124148; $E = 15.8$) is an intrinsically reddened ($A_V \\approx 0.5$), luminous ($M_B = -28.7$), radio-moderate, optically polarized ($\\sim 5\\%$), low-ionization broad absorption line quasar, at relatively low redshift $z \\approx 0.868$. See Brotherton et~al.\\ (2001) for further details. Brotherton et~al.\\ (2005) describes its X-ray spectrum. An initial investigation of the physical properties of the outflow in this object using $HST$ FOC and especially Keck HIRES spectra is described in Arav et~al.\\ (2001). The VLT spectral data set of QSO~2359--1241 is described in detail in Paper~I. Its unprecedented high-quality allowed us to test a variety of absorber distribution models needed to derive reliable ionic column densities of the outflow (see Paper~I). In the present paper we report these column densities and use them to determine the physical conditions within the main component of the outflow ({\\bf e}, see Paper~I): the ionization equilibrium, total column density and number density of the absorbing material. To do so we use the photoionization code Cloudy (Ferland et al.\\ 1998) as well as a separate \\ion{Fe}{2} ion model (Bautista \\& Pradhan 1998). The plan of the paper is as follows: In Section~2 we describe the column density measurements. In Section~3 we determine the physical conditions in the main component of the outflow. Finally, in Section~4 we summarize and discuss our results and provide a simple estimate of the outflow's distance from the central continuum source. \\clearpage \\begin{deluxetable}{lrlrrrr} \\tabletypesize{\\scriptsize} \\tablecaption{Absorption lines identified in the VLT spectrum of QSO~2359--1241} \\tablewidth{0pt} \\tablehead{ \\colhead{ $\\lambda$}&\\colhead{$\\log(gf)$\\tablenotemark{a}}& \\colhead{Ion}&\\colhead{E$_{low}$(cm$^{-1}$)}& \\colhead{$g_{low}$}&\\colhead{E$_{up}$(cm$^{-1}$)}& \\colhead{$g_{up}$}} \\startdata 2764.62 &-1.95 &\\ion{He}{1}* &159856 & 3& 196027 & 9\\cr 2829.92 &-1.74 &\\ion{He}{1}* &159856 & 3& 195193 & 9\\cr 2945.98 &-1.58 &\\ion{He}{1}* &159856 & 3& 193801 & 9\\cr 3188.69 &-1.16 &\\ion{He}{1}* &159856 & 3& 191217 & 9\\cr 3889.80 &-0.72 &\\ion{He}{1}* &159856 & 3& 185565 & 9\\cr 2852.97 & 0.270 &\\ion{Mg}{1} & 0 & 1& 35051 & 3\\cr 2796.36 & 0.100 &\\ion{Mg}{2} & 0 & 2& 35761 & 4\\cr 2803.54 &-0.210 &\\ion{Mg}{2} & 0 & 2& 35669 & 2\\cr 1854.72 & 0.060 &\\ion{Al}{3} & 0 & 2& 53917 & 4\\cr 1862.79 &-0.240 &\\ion{Al}{3} & 0 & 2& 53683 & 2\\cr 1808.01 &-2.100 &\\ion{Si}{2} & 0 &2 &55309 &4\\cr 1816.93 &-1.840 &\\ion{Si}{2}m* & 287 &4 &55325 &6\\cr 3934.83 & 0.134 &\\ion{Ca}{2} & 0 &2 &25414 &4\\cr 3969.65 &-0.166 &\\ion{Ca}{2} & 0 &2 &25192 &2\\cr 2576.87 & 0.433 &\\ion{Mn}{2} & 0 &7 &38807 &9\\cr 2594.49 & 0.270 &\\ion{Mn}{2} & 0 &7 &38543 &7\\cr 2606.46 & 0.140 &\\ion{Mn}{2} & 0 &7 &38366 &5\\cr 2344.2139& 0.057& \\ion{Fe}{2} & 0& 10& 42658& 8\\cr 2374.4612& -0.504& \\ion{Fe}{2} & 0& 10& 42115& 10\\cr 2382.7652& 0.505& \\ion{Fe}{2} & 0& 10& 41968& 12\\cr 2586.6500& -0.161& \\ion{Fe}{2} & 0& 10& 38660& 8\\cr 2600.1729& 0.378& \\ion{Fe}{2} & 0& 10& 38459& 10\\cr 2333.5156& -0.206& \\ion{Fe}{2}*& 385& 8& 43239& 6\\cr 2365.5518& -0.402& \\ion{Fe}{2}*& 385& 8& 42658& 8\\cr 2389.3582& -0.180& \\ion{Fe}{2}*& 385& 8& 42237& 8\\cr 2396.3559& 0.362& \\ion{Fe}{2}*& 385& 8& 42115& 10\\cr 2599.1465& -0.063& \\ion{Fe}{2}*& 385& 8& 38859& 6\\cr 2612.6542& 0.004& \\ion{Fe}{2}*& 385& 8& 38660& 8\\cr 2626.4511& -0.452& \\ion{Fe}{2}*& 385& 8& 38459& 10\\cr 2328.1112& -0.684& \\ion{Fe}{2}*& 668& 6& 43621& 4\\cr 2349.0223& -0.269& \\ion{Fe}{2}*& 668& 6& 43239& 6\\cr 2381.4887& -0.693& \\ion{Fe}{2}*& 668& 6& 42658& 8\\cr 2399.9728& -0.148& \\ion{Fe}{2}*& 668& 6& 42335& 6\\cr 2405.6186& 0.152& \\ion{Fe}{2}*& 668& 6& 42237& 8\\cr 2607.8664& -0.150& \\ion{Fe}{2}*& 668& 6& 39013& 4\\cr 2618.3991& -0.519& \\ion{Fe}{2}*& 668& 6& 38859& 6\\cr \\tablebreak 2632.1081& -0.287& \\ion{Fe}{2}*& 668& 6& 38660& 8\\cr 2338.7248& -0.445& \\ion{Fe}{2}*& 863& 4& 43621& 4\\cr 2359.8278& -0.566& \\ion{Fe}{2}*& 863& 4& 43239& 6\\cr 2405.1638& -0.983& \\ion{Fe}{2}*& 863& 4& 42440& 2\\cr 2407.3942& -0.228& \\ion{Fe}{2}*& 863& 4& 42401& 4\\cr 2411.2433& -0.076& \\ion{Fe}{2}*& 863& 4& 42335& 6\\cr 2614.6051& -0.365& \\ion{Fe}{2}*& 863& 4& 39109& 2\\cr 2631.8321& -0.281& \\ion{Fe}{2}*& 863& 4& 38859& 6\\cr 2345.0011& -0.514& \\ion{Fe}{2}*& 977& 2& 43621& 4\\cr 2411.8023& -0.377& \\ion{Fe}{2}*& 977& 2& 42440& 2\\cr 2414.0450& -0.455& \\ion{Fe}{2}*& 977& 2& 42401& 4\\cr 2622.4518& -0.951& \\ion{Fe}{2}*& 977& 2& 39109& 2\\cr 2629.0777& -0.461& \\ion{Fe}{2}*& 977& 2& 39013& 4\\cr 2332.00 & -0.720& \\ion{Fe}{2}*& 1873& 10& 44754& 8\\cr 2348.81 & -0.470& \\ion{Fe}{2}*& 1873& 10& 44447& 8\\cr 2360.70 & -0.700& \\ion{Fe}{2}*& 1873& 10& 44233& 10\\cr 2563.30 & -0.050& \\ion{Fe}{2}*& 7955& 8& 46967& 6\\cr 2715.22 & -0.440& \\ion{Fe}{2}*& 7955& 8& 44785& 6\\cr 2740.36 & 0.240& \\ion{Fe}{2}*& 7955& 8& 44447& 8\\cr 2756.56 & 0.380& \\ion{Fe}{2}*& 7955& 8& 44233& 10\\cr 2166.19 &0.230 &\\ion{Ni}{2}* & 8394 &10 &54557 &10\\cr 2217.14 &0.480 &\\ion{Ni}{2}* & 8394 &10 &53496 &12 \\cr 2223.61 &-0.140 &\\ion{Ni}{2}* & 8394 &10 &53365 &10\\cr 2316.72 &0.268 &\\ion{Ni}{2}* & 8394 &10 &51558 & 8\\cr \\enddata \\tablenotetext{a}{$gf$-values were taken from Kurucz (1995) for all but the \\ion{Fe}{2} lines with wavelengths given more than two decimal figures. $gf$-values for these come from Morton (2003).} \\end{deluxetable} \\clearpage ", "conclusions": "First, we compare the results of this investigation to the analysis of the Keck/HIRES observations of QSO~2359--1214 by Arav et al.\\ (2001; hereafter HIRES paper). Using the same MF87 SED, the HIRES analysis found $\\log N_H = 20.2$ and $\\log U_H = -2.7$, compared to $\\log N_H = 20.6$ and $\\log U_H = -2.4$ for the VLT analysis. These factors of $\\sim2$ differences are mainly attributed to using apparent optical depth methods to extract the \\ion{Fe}{2} and \\ion{He}{1}* column densities. As pointed out in the HIRES paper, the data was not of high enough signal-to-noise ratio to permit more sophisticated analyses. Even so, the HIRES paper already showed that the outflow is not shielded by a hydrogen ionization front, a result confirmed by the VLT analysis. The important leap in diagnostic power for the VLT data came from the ability to accurately measure the population levels of the excited \\ion{Fe}{2} levels, allowing us to pin point the number density of the outflow to $\\log(n_e)=10^{4.4}$ cm$^{-3}$ to better than 20\\% accuracy. This is both qualitatively and quantitatively a great improvement over the lower limit of $\\log(n_e)=10^{5}$ cm$^{-3}$ available from the HIRES data. This result is not accidental. The main reason we invested 6.5 hours of VLT observation on this outflow was precisely to yield a data set from which an accurate $n_e$ could be extracted. This determination of $n_e$ will allow us to determine the distance of the outflow from the central source and thus measure its mass flux and kinetic luminosity. This demonstrates the importance of taking high quality spectra of such outflows. Other important results arising from the measured populations of the \\ion{Fe}{2} levels are the determination of a lower limit to the temperature of the \\ion{Fe}{2} region and the realization that the absorption spectrum forms before the hydrogen ionization front, beyond which the temperature and ionization drop sharply. The temperature determination was crucial in constraining a whole family of SEDs and gas metalicities that would yield very different temperatures in the \\ion{Fe}{2} region, and consequently allowing for a more secure determination of the total gas column density and ionization parameter. That the \\ion{Fe}{2} absorption occurs in a region of nearly constant conditions before the hydrogen ionization front is a key to being able to model the absorption spectrum. Photoionization modeling allowed us to reproduce quite well the observed \\ion{Fe}{2} and \\ion{He}{1}* column densities in the main, {\\bf e}, component of the quasar outflow absorption spectrum. Reiterating, we found $\\log N_H \\approx 20.6$ and $\\log U_H \\approx -2.4$. The dominant error bars to these values come from the uncertainties in the assumed SED and gas metalicities and come to $\\sim$~0.3 dex. Given the above gas density, ionization parameter, an estimate to an {\\em unobscured} incident bolometric luminosity of $\\sim4.7 \\times 10^{47}$ ergs~s$^{-1}$ based on the intrinsic reddening correction in Brotherton et~al.\\ (2001, 2005), and a standard cosmology ($H_o = 70$ km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_\\Lambda = 0.70$, $\\Omega_m = 0.30$), we estimate a distance of component {\\bf e} of the outflow from the central continuum source of $\\sim3$ kpc. In a future paper we will similarly determine the physical conditions in the weaker, lower velocity outflow components {\\bf a--d}, as well as the distances and estimates of the kinetic luminosities for all components in the outflow, important to AGN feedback scenarios of galaxy evolution." }, "0807/0807.1000_arXiv.txt": { "abstract": "Searching for transit timing variations in the known transiting exoplanet systems can reveal the presence of other bodies in the system. Here we report such searches for two transiting exoplanet systems, TrES-1 and WASP-2. Their new transits were observed with the 4.2m William Herschel Telescope located on La Palma, Spain. In a continuing programme, three consecutive transits were observed for TrES-1, and one for WASP-2 during September 2007. We used the Markov Chain Monte Carlo simulations to derive transit times and their uncertainties. The resulting transit times are consistent with the most recent ephemerides and no conclusive proof of additional bodies in either system was found. ", "introduction": "Transiting planets provide a wealth of information about exoplanetary systems. Short-term variations in the mid-eclipse times of the transits may reveal the presence of moons, trojans or other planets (\\cite[Holman \\& Murray 2005]{HolmanMurray2005}, \\cite[Agol \\etal\\ 2005]{Agol_etal05}, \\cite[Ford \\& Holman 2007]{FordHolman07}), whereas long-term variations could result from orbital precession (\\cite[Miralda--Escud\\'{e} 2002]{Miralda02}). This provides further constraints on theories of planetary system formation and evolution, as well on theories of planetary atmospheres and their composition. ", "conclusions": "" }, "0807/0807.3005_arXiv.txt": { "abstract": "Pre-main sequence (PMS) stars are known to produce powerful X-ray flares which resemble magnetic reconnection solar flares scaled by factors up to $10^4$. However, numerous puzzles are present including the structure of X-ray emitting coronae and magnetospheres, effects of protoplanetary disks, and effects of stellar rotation. To investigate these issues in detail, we examine 216 of the brightest flares from 161 PMS stars observed in the Chandra Orion Ultradeep Project (COUP). These constitute the largest homogeneous dataset of PMS, or indeed stellar flares at any stellar age, ever acquired. Our effort is based on a new flare spectral analysis technique that avoids nonlinear parametric modeling. It can be applied to much weaker flares and is more sensitive than standard methods. We provide a catalog with $>30$ derived flare properties and an electronic atlas for this unique collection of stellar X-ray flares. The current study (Paper I) examines the flare morphologies, and provides general comparison of COUP flare characteristics with those of other active X-ray stars and the Sun. Paper II will concentrate on relationships between flare behavior, protoplanetary disks, and other stellar properties. Several results are obtained. First, the COUP flares studied here are among the most powerful, longest, and hottest stellar X-ray flares ever studied. Peak luminosities are in the range $31< \\log L_{X,pk}< 33$ erg~s$^{-1}$; rise (decay) timescales range from 1~hour to 1~day (few hours to 1.5 days); many peak temperatures exceed 100~MK. The scale of their inferred associated coronal structures is $0.5-10$~$R_{\\star}$. Second, no significant statistical differences in peak flare luminosity or temperature distributions are found among different morphological flare classes, suggesting a common underlying mechanism for all flares. Third, comparison with the general solar-scaling laws indicates that COUP flares may not fit adequately proposed power-temperature and duration-temperature solar-stellar fits. Fourth, COUP super-hot flares are found to be brighter but shorter than cooler COUP flares. Fifth, the majority of bright COUP flares can be viewed as enhanced analogs of the rare solar ``long-duration events''. ", "introduction": "} All solar-type stars exhibit their highest levels of magnetic activity during their pre-main sequence (PMS) phase \\citep{Feigelson07}. This includes `superflares' with peak luminosities $\\log L_x \\ga 32$ erg s$^{-1}$ in the $0.5-8$ keV band, $10^4$ more powerful than the strongest flares seen in the contemporary Sun \\citep[e.g.][]{Tsuboi98, Grosso04, Favata05}. PMS stars thus join RS~CVn binary systems \\citep[e.g.][]{Osten07} as laboratories to study the physics of the most powerful magnetic reconnection events. PMS stars are more distant and fainter than the closer RS~CVn systems, but hundreds of flaring PMS stars can be simultaneously studied due to their concentration in rich clusters. The magnetic field structure of PMS stars, and thus the nature of their reconnection and flaring, may (or may not) qualitatively differ from other stars due to the presence of a protoplanetary disk during the early PMS stages. The intense high energy radiation from these PMS reconnection events may affect the physical and chemical properties of the surrounding circumstellar environment and play an important role in the formation of planets \\citep{Glassgold05,Feigelson07}. A consensus has emerged during the past decade that PMS accretion is funneled by magnetic field lines linking the disk inner edge to the stellar surface \\citep[e.g.][]{Hartmann98, Shu00}. However, while early theory assumed a dipolar field morphology, recent studies point to a complex multipolar field structure similar to the Sun's \\citep{Jardine06, Donati07, Long08}. It is also unclear whether the X-ray flares occur primarily in large loops with both footprints anchored on the stellar surface, or in loops linking the stellar photosphere with the inner rim of the circumstellar disk \\citep{Isobe03, Favata05}. The first case may suffer instability due to centrifugal force \\citep{Jardine99} while the second case may load the loop with cool accreting material so that X-rays may not be produced \\citep{Preibisch05}. The 13-day nearly continuous observation of $\\sim 1408$ PMS stars in the Orion Nebula, the Chandra Orion Ultradeep Project \\citep[COUP;][]{Getman05}, enables both studies of individual flare properties and statistical studies of flaring from Orion stars \\citep{Wolk05, Flaccomio05, Stassun06, Caramazza07, Colombo07}. COUP also provided a unique opportunity to study relatively rare superflares and long-duration flares. \\citet{Favata05} have analyzed the strongest 32 flares in the COUP dataset using a long-standing method of time resolved spectroscopy (TRS) modeled as cooling plasma loops. They concluded that at least 1/3 of these are produced by magnetic reconnection in very long coronal $5-20$~R$_{\\star}$ structures. Such structures were predicted in magnetospheric accretion models \\citep[e.g][]{Shu97} but not clearly identified before COUP. \\citet{Favata05} recognized that their sample was too small to quantitatively probe the relationship between long coronal flaring structure and disks or accretion. The aim of the current study is to extend the flare sample of \\citet{Favata05} utilizing a more sensitive technique of flare analysis, the ``method of adaptively smoothed median energy'' (MASME) introduced by \\citet{Getman06}. We combine this method with the astrophysical cooling loop models of \\citet{Reale97} to trace the evolution of the flare plasma in temperature-density diagrams and derive flaring loop sizes. The method allowed us to examine $216$ of the brightest flares from $161$ brightest COUP PMS stars. These constitute the largest homogeneous sample of powerful stellar flares ever acquired in the X-ray band. In \\citet{Getman08} (Paper~II), we use these results to study in detail the relationships between PMS X-ray flares, stellar properties, protoplanetary disks, and accretion. Our flare analysis and the derived flare properties and classifications are presented in \\S \\ref{analysis}. Properties of the stars themselves are also provided. Global properties of our flares are considered in \\S \\ref{results} and compared to published studies of older stars. ", "conclusions": "} We analyze 216 bright X-ray flares from the Chandra Orion Ultradeep Project which provides the longest nearly-continuous observation of a rich PMS stellar cluster in the X-ray band. Our effort is based on a new spectral analysis technique (MASME) that avoids nonlinear parametric modeling and is more sensitive than standard methods. Flare loop parameters are derived from the well-established flare plasma model of \\citet{Reale97}. We thus emerge with the largest dataset of PMS flares, or indeed stellar flares at any stellar age, with peak luminosities in the range $31< \\log L_{X,pk}< 33$ erg~s$^{-1}$, several orders of magnitude more powerful than any solar flare. For each flare we provide a catalog of $>30$ derived flare properties including inferred sizes of associated coronal loops and flare morphological classes. We give an electronic atlas with flare lightcurves, temporal evolution plots of X-ray median energy, plasma temperature, emission measure, and derived temperature density diagram. This collection of empirical and model-dependent information can serve as a valuable testbed for stellar flare models. The powerful COUP flares studied here have rise timescales ranging from 1~hour to 1~day and decay timescales ranging from a few hours to 1.5 days. An important empirical result is that peak plasma temperatures are often 100~MK, in some cases $> 200$~MK. These temperatures are derived from a robust calibration of median energies; traditional time-resolve spectroscopy often does not have the time resolution to detect this brief super-hot phase. No significant differences in peak flare luminosity or temperature distributions are found among the wide range of morphological flare classes: typical fast-rise exponential-decays, step decays, double peaks, and slow-rise flat-top. This suggests that all flare types arise from similar underlying magnetic reconnection mechanisms and geometries. Comparison of the COUP flare properties with the general solar-stellar scaling laws of \\citet{Aschwanden08} presents surprising results. Our flares do not follow the solar-stellar trend between plasma peak emission measure and temperature, $EM_{pk} \\propto T_{obs,pk}^{4.7}$. The trend between flare duration and peak temperature is also absent. Super-hot COUP flares are found to be brighter but shorter in duration than cooler COUP flares. This is further developed in Paper II. Compared to non-PMS systems, the inferred sizes of COUP flaring structures are remarkably large, ranging widely from $L=0.5$ to $10$~R$_{\\star}$. These large flaring structures must be associated with large-scale stellar magnetic fields. Rare long decay solar events associated with the largest known X-ray emitting structures are possible solar analogs to these COUP flares. Our flare sample provides a valuable laboratory for the study of the physics and astronomy of magnetic reconnection events. This study (Paper I) examines flare morphologies, and provides general comparison of COUP flare characteristics with those of other active X-ray stars and the Sun. Paper II concentrates on relationships between flare behavior, protoplanetary disks (both passive and accreting), and other stellar properties including rotational periods and Keplerian corotation radii. Paper II further investigates super-hot COUP flares and magnetic field strength on Orion T-Tauri stars." }, "0807/0807.4139_arXiv.txt": { "abstract": "We present the results of our study of astrometric stability of 200-in Hale (Mt. Palomar) and 10-m Keck II (Mauna Kea) telescopes, both with Adaptive Optics (AO) facilities. A group of nearby visual binaries and multiples was observed in near infrared, relative separations and position angles measured. We have also checked the influence of some systematic effects (e.g. atmospherical refraction, varying plate scale factor) on result and precision of astrometric measurements. We conclude that in visual binaries astrometrical observations it is possible to achieve much better precision than 1 miliarcsecond [$mas$], which in many cases allows detection of the astrometrical signal produced by planetary-mass object. ", "introduction": "Astrometry is thought to be the most promising method of exoplanets detection in the future. On the contrary to radial velocities measurements (RV), astrometry is almost independent on stellar physics, e.g. phenomena like the activity or pulsations, rather than the distance to the object. Todays interferometers are able to achieve precision at the level of mili- or microarcseconds, which is good enough to attempt exoplanet research. Nevertheless, the same level is possible to achieve in small fields by CCD imaging with adaptive optics (AO) systems. To do that, one must subtract and correct some systematical effects in order to obtain a gaussian distribution of the measurements. In such case, the precision improves like $N^{-0.5}$, where $N$ is a number of single measurements. The goal of our studies was to check if a gaussian statistic can really be achieved with two top-class telescopes with AO systems: the 200-in Hale telescope + PHARO camera (Mt. Palomar) and 10-m Keck II telescope + NIRC2 (Mauna Kea), both working in infrared (IR). With these facilities we were imaging 17 objects, majority of which were M-type dwarf binaries, located closer than 20 $pc$ from the Sun. ", "conclusions": "Todays telescopes with AO systems allows us to perform astrometric measurements with precision well bellow 1 $mas$. This refers to a single-epoch observations, as well as to a long-term stability. For many cases such a precision means an ability to detect exoplanets around nearby stars. \\begin{figure} \\includegraphics[width=0.95\\columnwidth]{Helminiak_fig1.eps} \\caption{Example of Allan variance vs. lag for uncorrected (dash-dotted) and corrected (solid line) measurements of separation components $x$ (left) and $y$. A small decrease is normal for finite series of measurements. Dotted line shows a --1-behavior for an ideal, infinitely long white-noise series.} \\end{figure} \\begin{table} \\caption{Smallest $\\sigma_{\\rho}$ and corresponding detection limits for researched stars.} \\begin{tabular}{c|c|c|cc|cc|c} \\hline Star & Lowest & Dist. & Mass A & Limit for A & Mass B & Limit for B& Tel. \\\\ (GJ No.)& $\\sigma\\, [mas]$ & $[pc]$ & $[M_{\\odot}]$ & $[AU \\cdot M_J]$ & $[M_{\\odot}]$ & $[AU \\cdot M_J]$ &\\\\ \\hline 195 & 0.12 & 13.89 & 0.53 & 1.38 & 0.19 & 0.50 & Hale \\\\ 352 & 1.11 & 10.53 & 0.44 & 8.04 & 0.41 & 7.49 & Hale \\\\ 458 & 0.28 & 15.32 & 0.40 & 2.68 & 0.37 & 2.48 & Hale \\\\ 507 & 0.33 & 13.16 & 0.46 & 3.12 & 0.37 & 2.51 & Hale \\\\ 569B & 0.11 & 9.81 & 0.071 & 0.116 & 0.054 & 0.088 & Keck \\\\ 661 & 0.038 & 6.32 & 0.379 & 0.16 & 0.34 & 0.15 & Hale \\\\ 767 & 0.09 & 13.35 & 0.44 & 0.83 & 0.40 & 0.75 & Hale \\\\ 860 & 0.048 & 4.01 & 0.34 & 0.10 & 0.27 & 0.09 & Hale \\\\ 873 & 0.57 & 5.05 & 0.36 & 1.62 & unknown & unknown & Hale \\\\ 9071 & 0.20 & 13.89 & 0.53 & 2.22 & 0.49 & 2.05 & Hale \\\\ \\hline \\end{tabular} \\end{table}" }, "0807/0807.1693_arXiv.txt": { "abstract": "Large volume cosmological simulations succeed in reproducing the large-scale structure of the Universe. However, they lack resolution and may not take into account all relevant physical processes to test if the detail properties of galaxies can be explained by the CDM paradigm. On the other hand, galaxy-scale simulations could resolve this in a robust way but do not usually include a realistic cosmological context. To study galaxy evolution in cosmological context, we use a new method that consists in coupling cosmological simulations and galactic scale simulations. For this, we record merger and gas accretion histories from cosmological simulations and re-simulate at very high resolution the evolution of baryons and dark matter within the virial radius of a target galaxy. This allows us for example to better take into account gas evolution and associated star formation, to finely study the internal evolution of galaxies and their disks in a realistic cosmological context. We aim at obtaining a statistical view on galaxy evolution from z $\\simeq$ 2 to 0, and we present here the first results of the study: we mainly stress the importance of taking into account gas accretion along filaments to understand galaxy evolution. ", "introduction": "The morphology of galaxies in the Local Universe is well constrained by observations, but is still largely unexplained. Indeed, large volume cosmological simulations fail to reproduce realistic galaxies. For instance, the disks formed are often too concentrated : it is the ``angular momentum problem'', well known since the early work of \\cite{Navarro1991}. It is still unclear whether this is an intrinsic problem of the $\\Lambda$CDM paradigm or if something (i.e. resolution, physical processes...) is missing in these simulations. Another puzzle is the question of disk survival till z=0 (\\cite{Koda2007}). For instance, \\cite{Kautsch2006} study a large sample of edge-on spiral galaxies in the SDSS and find that a significant fraction of them (i.e. roughly one third) are bulgeless or ``superthin''. This is still unexplained by cosmological models. Indeed, $\\Lambda$CDM predicts that galaxy interactions are frequent (see e.g. the recent work by \\cite{Stewart2007}). More exactly, major mergers, that are well known to destroy disks to form ellipticals (\\cite{Barnes1991}) are rather rare, but minor mergers are much more common. These minor mergers can thicken disks, and if frequent enough could even form elliptical galaxies (\\cite{Bournaud2007}). The problem is then to find whether $\\Lambda$CDM predicts too many mergers, or if the satellites have properties and orbital parameters such that they have little influence on the galactic disks. Also, gas accretion along filaments could fuel a thin disk and counteract the effect of mergers (\\cite{Dekel2005}, \\cite{Keres2005}, \\cite{Ocvirk2008}). To study the properties of galaxies at low and high redshift, it thus seems necessary to take the full cosmological context into account. Large scale cosmological simulations could of course achieve this goal and give a statistical view on galaxies at each redshift, but for now they mainly lack resolution at the galactic scale. On the contrary, small volume cosmological simulations like the one performed by \\cite{Naab2007} can resolve galactic scales in detail but are so time-consuming that obtaining a statistical sample is for now a challenge. A first method to solve these problem is to use semi-analytical models, i.e. extracting merger trees from cosmological simulations and using different recipes to infer physical properties of galaxies (\\cite{Somerville2001}, \\cite{Hatton2003}, \\cite{Khochfar2005}). The drawback is that approximations are necessary. Another possibility has been explored by \\cite{Kazantzidis2007}, \\cite{Read2007} and \\cite{Villalobos2008} : they extract merger histories from cosmological simulations and re-simulate these histories at higher resolution. Nevertheless, they perform collisionless simulations with no gas component, neither in the main galaxy, nor in satellites, nor in filaments. We here present a new approach where we re-simulate at high resolution a history given by a cosmological simulation, using self consistent realistic galaxies (the main galaxy and the satellites have a gas disk, a stellar disk and a dark matter halo), and we also take into account gas accretion from cosmic filaments. Our goal is to obtain a statistical sample of merger and accretion histories in a $\\Lambda$CDM context to simulate the resulting galaxies and to compare our results to observations at various redshifts. After a description of the technique used, we will present our first results and emphasize the importance of gas accretion along filaments to understand galaxy evolution. ", "conclusions": "In order to study galaxy evolution in cosmological context, we have successfully developed a technique that allows us to perform high resolution simulations taking into account realistic merger and gas accretion histories. The first two simulations shown here do not allow us to draw any general conclusion on galaxy evolution in a $\\Lambda$CDM context. Nevertheless, we can already confirm that even low mass satellites can thicken disks and that ellipticals from both through repeated minor mergers and major mergers. We also emphasize that gas accretion from filaments can allow to rebuild a thin disk in a galaxy, which proves the absolute necessity to take this accretion into account to understand galaxy evolution." }, "0807/0807.0211.txt": { "abstract": "\\noindent We investigate scenarios in which a charged, long-lived scalar particle decouples from the primordial plasma in the Early Universe. We compute the number density at time of freeze-out considering both the cases of abelian and non-abelian interactions and including the effect of Sommerfeld enhancement at low initial velocity. We also discuss as extreme case the maximal cross section that fulfils the unitarity bound. We then compare these number densities to the exotic nuclei searches for stable relics and to the BBN bounds on unstable relics and draw conclusions for the cases of a stau or stop NLSP in supersymmetric models with a gravitino or axino LSP. ", "introduction": "%======================================================================= The early Universe may have been populated by many exotic particles that, especially if charged, should have easily been in thermal equilibrium. No charged relic seems to have survived to the present day. In fact there are very strong upper bounds on the density of electromagnetically % charged and colored and/or colour charged particles with masses below 10--100 TeV from extensive searches for exotic nuclei \\cite{ch-relics}. The standard lore is therefore that only neutral relics may have survived until today. However, it is possible that some unstable but very long-lived charged particle froze-out from thermal equilibrium and decayed much later to a neutral one. A typical example of this kind in supersymmetric models with R-parity conservation is the next-to-lightest supersymmetric particle~(NLSP) if the LSP and Cold Dark Matter is very weakly interacting like the axino~\\cite{axinolsp,crs02,axinolsp2} or the gravitino~\\cite{gravitinolsp, neutralNLSP}. Recently, such candidates have attracted a lot of attention, and indeed the signal of a charged metastable NLSP at colliders would be spectacular~\\cite{stauatcolliders,Fairbairn:2006gg}. In general, strong bounds on the number density of any metastable relic with lifetime of about 1\\,s or longer are provided by Big Bang Nucleosynthesis (BBN)~\\cite{BBNrev}. They come from two classes of processes: on one hand injection of very energetic photons or hadrons from decays during or after BBN adds an additional non-thermal component to the plasma and can modify the abundances of the light elements~\\cite{neutBBN}; on the other hand, if the relic particle is electromagnetically charged, bound states with nuclei may arise that strongly enhance some of the nuclear rates and allow for catalysed production of e.g.\\ $^6Li$~\\cite{CBBN}. The bounds of the first type are very tight for lifetimes of the order of $10^4$\\,s and exclude, for instance, a neutralino NLSP with a gravitino LSP in the CMSSM~\\cite{neutralNLSP}. An electrically charged NLSP like the $\\tilde\\tau$ can instead escape the first class of constraints in part of the parameter space, but it is excluded for long lifetimes by bound state effects~\\cite{stauNLSP}. In the axino LSP case, the NLSP has a shorter lifetime; the BBN bounds are hence much weaker and both, neutralino and stau, NLSP are still allowed \\cite{axinolsp}. In this paper, we investigate the most general case of a scalar charged thermal relic. We compute the number density and compare it to the bounds on exotic nuclei for stable particles and the BBN constraints for unstable ones. Similar studies have been carried out model-independently many years ago~\\cite{wolfram, Nardi:1990ku, unitarity} for stable relics and we will update and improve these computations.\\footnote{Recently the case of general EW charged relics as DM was also considered in full detail~\\cite{Cirelli07}.} % We mostly consider the role of the gauge interaction for two main reasons: i) the annihilation into gauge bosons is often the dominant channel for a charged particle and ii) it depends only on very few parameters, just the mass of the particle and its charge or representation. It is also enhanced by the Sommerfeld effect~\\cite{Sommerfeld}, analogous to heavy quark production at threshold, which has previously been considered for dark matter annihilations in \\cite{gluino-splitsusy, Hisano:2006nn, Cirelli07, Freitas07, DMsommerfeld} and recently also in the context of leptogenesis in \\cite{Strumia:2008cf}. We discuss this Sommerfeld enhancement for the general abelian and non-abelian cases. Moreover, we compare the cross sections with the unitarity bound and update the unitarity limit on the mass of a stable relic. Our main goal is to determine if it is at all possible to evade {\\it completely} either the exotic nuclei bounds or the BBN ones and how strongly the particle has to interact in this case. We then apply our findings to the Minimal Supersymmetric Standard Model and discuss in more detail the cases of the stau and stop NLSP. The paper is organised as follows. In Section~2, we briefly review the computation of the number density from thermal freeze-out. The formulae for the annihilation cross section of a charged particle into gauge bosons are given in Section~3. Here we discuss abelian and non-abelian cases, the Sommerfeld enhancement and the unitarity cross section. Moreover, we compare the thermal averages with the first order in velocity expansion. The resulting relic density is discussed in Section~4. In Section~5, we review the constraints on stable and unstable relics. These are then applied in Section~6 to the concrete examples of relic staus and stops. Section~7 finally contains our conclusions. Details on the computation of the annihilation cross section and the case of massive gauge bosons are given in the Appendices A and B. %======================================================================= ", "conclusions": "%====================================================================== We have studied the number density of a charged relic by computing the annihilation cross section into gauge bosons, including the Sommerfeld enhancement. We have found that the Sommerfeld factor increases the thermally averaged annihilation cross section by 20-50\\% and reduces the final yield even by a factor 2 or 3 for the $SU(3)$ case. Moreover the result is very sensitive on how the higher orders are resummed. Nevertheless the number density surviving the annihilation is still large and BBN constraints are relevant for most relics. They can be avoided completely only for very large $N$ for particles in the fundamental representation of $SU(N)$ ($N> 100 $ for $ m_X \\leq 10 $ TeV) or for cross sections nearly fulfilling the unitarity bound. For the cases of SM gauge groups, the allowed regions only correspond to very light relic masses, where the number density is low enough, or to sufficiently heavy relic masses so that the decay takes place in the first stages of BBN. The latter allowed region depends strongly on the relic decay channel, and, in case of a gravitino LSP with conserved R-parity, also on the gravitino mass. Let us mention here that if R-parity is just marginally broken, the NLSP can decay with shorter lifetime through R-parity violating channels and the BBN constraints can be easily evaded for any NLSP while keeping the gravitino LSP as Dark Matter~\\cite{R-parity}. More specifically, for the stau NLSP the light mass window has nearly completely been excluded by direct searches at LEP, even if the annihilation cross-section is maximal $ \\sim 4 \\sigma(\\tilde\\tau\\tilde\\tau^* \\rightarrow \\gamma\\gamma) $, unless the gravitino is lighter than a few tens of GeV, while the large mass region is unfortunately out of reach at the LHC for gravitino masses $m_{3/2} > 100 \\mbox{GeV} $. The detection of a quasi-stable stau at the LHC would then point to a scenario with relatively light gravitino mass, R-parity breaking or an axino LSP and could probably exclude the gravity mediated supersymmetry breaking scenario. In that case the determination of the stau lifetime and its decays will become crucial in distinguishing the different LSPs \\cite{stauatcolliders,bchrs05}. The stop case is much less constrained thanks to the stronger annihilation cross-section, even if in this case the decay always produces mainly hadrons. We have practically no constraints if the LSP is an axino and even for a gravitino LSP, we can allow for relatively light stops up to approximately 700 GeV ($1\\;$ TeV for lifetimes below $10^{7}\\;$ s), if the annihilation cross section reaches the unitarity one after the QCD phase transition. The window between the present Tevatron bound around 250 GeV and 1 TeV should be surely completely covered by the LHC, the signature being a quasi-stable heavy fermionic meson. The detection of such a state would call for a non-minimal SUSY breaking sector with a coloured NLSP and a very weakly interacting LSP. In this case again only the analysis of the stop decays would allow to distinguish between the lightest states. %======================================================================" }, "0807/0807.4233_arXiv.txt": { "abstract": "The majority of Globular Clusters show chemical inhomogeneities in the composition of their stars, apparently due to a second stellar generation in which the forming gas is enriched by hot-CNO cycled material processed in stars belonging to a first stellar generation. Clearly this evidence prompts questions on the modalities of formation of Globular Clusters. An important preliminary input to any model for the formation of multiple generations is to determine which is today the relative number fraction of ``normal\" and anomalous stars in each cluster. As it is very difficult to gather very large spectroscopic samples of Globular Cluster stars to achieve this result with good statistical significance, we propose to use the horizontal branch. We assume that, whichever the progenitors of the second generation, the anomalies also include enhanced helium abundance. In fact, helium variations have been recently recognized to be able to explain several puzzling peculiarities (gaps, RR Lyr periods and period distribution, ratio of blue to red stars, blue tails) in horizontal branches. We summarize previous results and extend the analysis in order to infer the percentage in number of the first and second generation in as many clusters as possible. We show that, with few exceptions, approximately 50\\% or more of the stars belong to the second generation. In other cases, in which at first sight one would think of a simple stellar population, we give arguments and suggest that the stars might all belong to the second generation. We provide in Appendix a detailed discussion and new fits of the optical and UV data of NGC~2808, the classic example of a multiple helium populations cluster, consistently including a reproduction of the main sequence splittings and an examination of the problem of ``blue hook\" stars. We also show a detailed fit of the totally blue HB of M~13, one among the clusters that are possibly fully made up by second generation stars. We conclude that the formation of the second generation is a crucial event in the life of globular clusters. The problem of the initial mass function required to achieve the observed high fraction of second generation stars can be solved only if the initial cluster was much more massive than the present one and most of the first generation low mass stars have been preferentially lost. As shown by D'Ercole et al. by modelling the formation and dynamical evolution of the second generation, the mass loss due to the explosions of the type II supernovae of the first generation may be the process responsible for triggering the expansion of the cluster, the stripping of its outer layers and the loss of most of the first generation low-mass stars. ", "introduction": "\\label{sec:intro} The observations of Globular Cluster (GC) stars are still to be interpreted in a fully consistent frame. Nevertheless, a general consensus is emerging on the fact that most GCs can not be considered any longer ``simple stellar populations\" (SSP), and that ``self--enrichment\" is a common feature among GCs. This consensus follows from the well known ``chemical anomalies'', already noted in the seventies (such as the variations found in C and N abundances, the Na--O and Mg--Al anticorrelations). Recently observed to be present at the turnoff (TO) and among the subgiants \\citep[e.g.][]{gratton2001,briley2002, briley2004, cohen2005}, they must be attributed to some process of ``self--enrichment\" occurring at the first stages of the cluster life, as the same authors quoted above suggest. There was a first epoch of star formation that gave origin to the ``normal\" (first generation, hereinafter FG) stars, with CNO and other abundances similar to Population II field stars of the same metallicity. Afterwards, there must have been some other epoch of star formation (second generation, hereinafter SG), including material heavily processed through the CNO cycle. This material either comes entirely from the stars belonging to the first stellar generation, or it is a mixture of processed gas and pristine matter of the initial star forming cloud. We can derive this conclusion as a consequence of the fact that there is no appreciable difference in the abundance of elements such as Ca and the heavier ones between ``normal\" and chemically anomalous stars belonging to the same GC. Needless to say, this statement {\\it does not} hold for $\\omega$ Cen, which must indeed be considered a small galaxy and not a typical GC. In the following, we will only examine ``normal clusters\", those which do not show signs of metal enrichment due to supernova ejecta. The homogeneity in the heavy elements is an important fact that tells us, e.g., that it is highly improbable that the chemical anomalies are due to mixing of stars born in two different clouds, as there is no reason why the two clouds should have a unique metallicity. In addition, the clusters showing chemical anomalies have a large variety in metallicities, making the suggestion of mixing of two different clouds even more improbable. The matter must have been processed through the hot CNO cycle, and not, or only marginally, through the helium burning phases, since the sum of CNO elements is the same in the ``normal\" and in the anomalous stars \\citep[e.g.][]{smith1996,ivans1999,cohenmelendez2005}. \\cite{carretta2005} find that actually the CNO is somewhat --but not much-- larger in the SG stars of some GCs. Therefore, the progenitors may be either massive asymptotic giant branch (AGB) stars \\citep[e.g.][]{ventura2001, ventura2002}\\footnote{If the \\cite{carretta2005} CNO data really indicate that a limited number of third dredge up \\citep[e.g.][]{ibenrenzini1983} episodes plays a (small) role in the nuclear processing of the matter giving origin to the SG, the massive AGB progenitors are possibly favoured.} or fast rotating massive stars \\citep{decressin2007}. In both cases, models show that the ejected material must be enriched in helium with respect to the pristine one. The higher helium content has been recognized to have a strong effect on horizontal branch (HB) morphology, possibly helping to explain some features (gaps, hot blue tails, second parameter) which, until now, have defied explanation \\citep{dantona2002}. Along these lines, a variety of problems has been examined: the extreme peculiarity of the HB morphology in the massive cluster NGC~2808, \\citep{dantonacaloi2004}; the second parameter effect in M~13 and M~3 \\citep{caloi-dantona2005}; the peculiar features in the RR Lyr variables and HB of NGC~6441 and NGC~6388 \\citep{caloidantona2007}. The presence of strongly enhanced helium in peculiar HB stars has been confirmed, for NGC 2808 and NGC 6441, by spectroscopic observations \\citep{moeler2004,busso2007}. Beside this spectroscopic evidence, an unexpected feature has recently appeared from photometric data: the splitting of the main sequence in NGC 2808. After first indications from a wider than expected colour distribution \\citep{dantona2005}, recent HST observations by \\cite{piotto2007} leave no doubt that there are at least three different populations in this cluster. This came after the first discovery of a peculiar blue main sequence in $\\omega$~Cen \\citep{bedin2004}, interpreted again in terms of a very high helium content \\citep{norris2004, piotto2005}. The above mentioned cases can be considered as ``extreme'' ones, in the sense that no explanation had been attempted for them before the hypothesis of helium-enriched populations. Less critical situations, such as the HB bimodality in NGC 1851 and 6229, had been tentatively explained in terms of a unimodal mass distribution with a large mass dispersion (0.055 -- 0.10 \\msun, Catelan et al. 1998). But even such a rather artificial assumption could not help in the case of NGC 2808, which, as hinted before, finds a possible solution only in terms of varying helium in multiple stellar generations. Therefore, we consider appropriate to apply to the less peculiar cases the solution found plausible for the most peculiar ones. In fact, notice that split main sequences and strong bimodalities are only the tip of the iceberg of the self--enrichment phenomenon. In most clusters the higher helium abundances remain confined below Y$\\sim 0.30$, and the presence of such stars would not be put in evidence either from main sequence observations \\citep{dantona2002,salaris2006}, or from a naif interpretation of stellar counts on the HB, as we shall discuss in Sect. \\ref{ngc6441}. If we wish to shed light on the entire process of formation of GCs we must have a rough idea of the total number of SG stars. We will then analyze the HB in terms of populations differing in Y, using one or more of the following peculiar features: \\begin{enumerate} \\item bimodal HBs and HB gaps; \\item presence of blue HB stars and very long period RR Lyr's in high Z clusters; \\item peaked number vs. period distribution of RR Lyr's; \\item blue--HB clusters. \\end{enumerate} In this paper we show the results of such an interpretation for several GCs. We start from a reanalyis of the NGC~2808 data, taking into account the results by \\cite{piotto2007} for the main sequence, and the ultraviolet HST data by \\cite{castellani2006}; we summarize the results already published and discuss briefly the other clusters. The table of the derived FG and SG percentages is the basis to discuss the clusters' dynamical evolution required to produce the high fraction of stars presently belonging to the SG. ", "conclusions": "In all GCs examined in this work, a large fraction of the stellar population takes origin from secondary star formation episodes. Notice that we have examined only a fraction of the clusters with HB morphology or RR Lyr period distributions similar to those described here, so that we can suggest that the results of this work probably hold for a larger population of Galactic GCs. While the most massive clusters have extreme helium enhancements, also moderately massive clusters show a considerable degree of helium variation. We reached our goals by examining in detail GCs that have unexplained features in their HBs, and extending the results to clusters with similar features. The HB morphology is one of the important features: clusters having a bimodal or multimodal HB are most easily interpreted by the coexistence of multiple generations with different helium content. Also a unique SG, whose stars had different degrees of mixing with pristine matter and thus ended up with different helium contents, is a possible solution. In any case, the extreme helium rich populations (in \\ocen\\ and NGC~2808 at least) are neatly separated from the other MS stars so that they should have a well defined independent origin. Further, we used the period distribution of RR Lyrs in several clusters in order to reject the hypothesis of a unique Y value with a relatively large spread of mass loss on the RGB, that has been the standard way of interpreting the whole HB colour extension, but is inconsistent with most period distributions. We re--examined in detail the HB distribution in NGC~2808, and obtained a helium distribution consistent with the main sequence recent data by \\cite{piotto2007}. Also the UV data of this cluster find a good interpretation in terms of population with varying helium content, if we make the further hypothesis of deep mixing to understand the location of the blue hook stars. We also show that simulation of M~13 UV data is well explained with populations having different helium. After this analysis, then, we must face the problem that the SG formation is not a peculiarity of a few very massive clusters, but must be the normal way in which a GC is formed in our Galaxy. It is almost obvious, and has often been discussed in the literature, that the ejecta of a unique first stellar generation with a normal initial mass function (IMF) can not produce enough mass to give origin to such a large fraction of second generation stars (see, e.g., the case made by Bekki and Norris (2004), for the blue main sequence of $\\omega$Cen). The only solution to the IMF problem is that {\\it the starting initial mass from which the first generation is born was much larger than today first generation remnant mass} (at least a factor 10 to 20 larger), so that the processed ejecta of the first generation provide enough mass to build up the second one. There are two possible ways of producing this result: the first possibility is that all these GCs formed within a dwarf galaxy environment \\citep{bekki2006,bekki2007}. There, GCs may be formed by mixing of pristine gas with the winds of the very numerous massive AGB stars evolving in the field of the dwarf galaxy, and later on the dwarf galaxy is dynamically destroyed. A second possibility has been recently suggested by \\cite{dercole2008}. They show that the SG stars are preferentially born in the inner core of a FG cluster, where a cooling flow collects the gas lost by the FG stars. The massive stars that explode as SN~II were preferentially concentrated in the cluster core. After the mass loss due to the supernovae type II explosions, the cluster expands, and begins losing the stars ---mainly of FG--- going out of the tidal radius. Thus the cluster may be destroyed, unless the gas lost by the most massive AGB stars begins collecting in the core and forms the SG, that initially does not take part in the cluster expansion. The study of the cluster dynamical evolution, followed by means of N-body simulations, shows that the cluster preferentially loses FG stars; these simulations show that high SG/FG number ratio can be achieved and SG-dominated clusters may survive." }, "0807/0807.1766_arXiv.txt": { "abstract": "We present the first multi-color view of the scattered light disk of the Herbig Ae star HD 163296, based on coronagraphic observations from the \\textit{Hubble Space Telescope} Advanced Camera for Surveys (ACS). Radial profile fits of the surface brightness along the disk's semi-major axis indicates that the disk is not continuously flared, and extends to $\\sim$540 AU. The disk's color (V-I)=1.1 at a radial distance of 3$\\farcs$5 is redder than the observed stellar color (V-I)=0.15. This red disk color might be indicative of either an evolution in the grain size distribution (i.e. grain growth) and/or composition, both of which would be consistent with the observed non-flared geometry of the outer disk. We also identify a single ansa morphological structure in our F435W ACS data, which is absent from earlier epoch F606W and F814W ACS data, but corresponds to one of the two ansa observed in archival HST STIS coronagraphic data. Following transformation to similar band-passes, we find that the scattered light disk of HD 163296 is 1 mag arcsec$^{-2}$ fainter at 3$\\farcs$5 in the STIS data than in the ACS data. Moreover, variations are seen in (i) the visibility of the ansa(e) structures, in (ii) the relative surface brightness of the ansa(e) structures, and in (iii) the (known) intrinsic polarization of the system. These results indicate that the scattered light from the HD 163296 disk is variable. We speculate that the inner disk wall, which Sitko et al. suggests has a variable scale height as diagnosed by near-IR SED variability, induces variable self-shadowing of the outer disk. We further speculate that the observed surface brightness variability of the ansa(e) structures may indicate that the inner disk wall is azimuthally asymmetric. ", "introduction": "\\label{intro} Investigations of the nature and evolution of dust grains in proto-planetary systems are motivated in part by our desire to understand the birth and evolution of planetary bodies, which originate from these systems. It is well established that Herbig Ae/Be stars \\citep{her60} are intermediate mass pre-main-sequence stars, analogous to the more familiar low mass T Tauri stars, which contain copious amounts of circumstellar gas and dust \\citep{wat98}. A wealth of observational evidence \\citep{man97,oud99,man00, vin02,eis04} suggests that the geometry of this gas and dust takes the form of a circumstellar disk. By contrast, our understanding of fundamental properties such as grain composition and size distribution, as well as the time evolution of these parameters, are much less understood. Analysis of ISO spectra led \\citet{mee01} to develop a 2-part evolutionary classification of Herbig systems; Group I sources were characterized as being slightly younger systems with flared outer disks, while Group II sources were characterized as slightly more evolved systems which have flatter disks owing to grain growth and settling. While \\citet{mee01} invoked dust settling to explain the differences in flaring between Group I and II sources, it has also been suggested that the outer regions of some Herbig disks may experience self-shadowing, owing to a ``puffed-up'' inner circumstellar disk rim \\citep{nat01,dul01,dul04,ise05}. Recent models which include inflated inner disk rims have proven to be successful in explaining the near-IR and interferometric properties of several Herbig Ae stars \\citep{ise06}. Conclusive confirmation of the geometry of the outer disk regions of Herbig Ae/Be stars inferred by SED-based studies has generally not occurred. HD 163296 is a young (4 Myr; \\citealt{van00}), nearby (122 pc; \\citealt{van98}) Herbig Ae star. While the star is not deeply embedded in a natal star formation dust cloud \\citep{the85}, it still displays clear signs of active accretion via the presence of jets and Herbig Haro knots \\citep{gra00,dev00,was06}. Hence, HD 163296 appears to be in a transition phase between optically thick, extremely young pre-main-sequence stars and the much more optically thin, near zero-age-main-sequence debris disk type stars \\citep{ard04,kal05,kal06,gol06}. Indirect techniques \\citep{bjo95} and resolved imaging \\citep{man97,gra00,ise07} have confirmed the presence of a circumstellar disk associated with HD 163296. HST STIS white-light coronagraphic observations spatially resolved the outer regions of this disk, and detected evidence of disk structure including an inner annulus of reduced scattering and a bright outer ring or ansae of material \\citep{gra00}. The inner region of HD 163296's circumstellar disk, as diagnosed by near-IR (0.8 - 5 $\\mu$m) SED monitoring, exhibits evidence of variability, possibly owing to changes in the inner disk wall \\citep{sit06}. Attempts to link processes which dominate the inner regions of Herbig Ae disks to morphological features observed in the outer disk regions, i.e. determining the relative roles of self-shadowing, dust settling, and disk flaring \\citep{dul04}, requires a wealth of observational data which diagnose the inner and outer disk regions. In this paper, we present a multi-epoch, multi-color view of the HD 163296 scattered light disk, providing a significant improvement in the documented behavior of the outer regions of HD 163296's circumstellar disk. We describe the observational details and data reduction procedures applied to our ACS data in Section \\ref{obs-red}. In Section \\ref{acs-charact}, we document the basic features of the HD 163296 scattered light disk in our ACS data, and compare these results to earlier epoch scattered light imaging in Section \\ref{stis}. The possible mechanisms which might explain the observed scattered light variability are discussed in Section \\ref{origin}. We provide a summary of our main results in Section \\ref{summary}. ", "conclusions": "\\label{summary} We have presented the first multi-color, multi-epoch analysis of resolved optical scattered light imaging of the Herbig Ae star HD 163296. To summarize the observational properties of these data: \\begin{enumerate} \\item We spatially resolved the HD 163296 scattered light disk over radial distances of 2$\\farcs$9-4$\\farcs$4 (350-540 AU) in the HST/ACS F435W, F606W, and F814W filters. \\item Radial profiles of the surface brightness along the semi-major axis of the disk follow a $\\sim$r$^{-4}$ power law behavior, which is indicative of a non-continuously flared disk. \\item The (V-I) color of the disk at 3$\\farcs$5, $\\sim$1.1, is significantly redder than the stellar (V-I) color, 0.15. This red disk color appears to be spatially uniform at the SNR of our data, and is consistent with that expected for a disk which has experienced grain growth and is at least partially self-shadowed. \\item A single ansa structure is present in the SE disk quadrant of the 2004 epoch ACS data (F435W filter), while no structure was observed in the 2003 epoch ACS data (F606W and F814W filters), despite the fact that the limiting detection magnitude of these latter data was sufficient to probe structure at the surface brightness observed in 2004. This morphological feature is spatially coincident with the fainter ansa reported in 1998 STIS observations \\citep{gra00} of HD 163296; the brighter ansa seen in the NW disk quadrant of the STIS data are absent from our ACS epoch data. \\item The scattered light disk was observed to be significantly brighter in each of three filters of ACS observations from 2003-2004, as compared to white-light STIS observations in 1998. Combining the multi-color ACS data to crudely approximate the STIS bandpass suggests that the ACS epoch data are $\\sim$1 mag arcsec$^{-2}$ brighter than the STIS data. Along with the variability of the visibility and surface brightness of ansa structure(s) in the disk, these results \\textbf{\\textit{suggest the scattered light disk of HD 163296 is variable}}. \\end{enumerate} We expect that the characterization of the basic behavior of HD 163296's scattered light disk that we have provided in this study will serve as an important foundation to future efforts to model the multi-wavelength behavior of the system. Although speculative, we suggest that one plausible explanation for the origin of the observed variability is: \\begin{enumerate} \\item The scale height of the inner disk wall is believed to inflate and deflate on time-scales of less than a few years, based on IR SED monitoring \\citep{sit06}. The variable scale height of the inner disk wall could induce variable self-shadowing of the outer disk, hence produce the observed overall variability of the scattered light disk. \\item We suggest that the ansa structure(s), which appear to be discernible only during periods of enhanced self-shadowing, represent a localized region of scatterers which are at a scale height (at least marginally) above the projected shadow. As the features are not discernible during periods of more complete illumination of the outer disk, we believe that a localized enhancement in the scale height of the disk is more likely to produce the observed phenomenon than a localized density enhancement of the scatterers (i.e. a clump). \\item The relative surface brightness variability of the NW versus SE ansa structures during epochs in which they are visible suggests that the inner disk wall might be azimuthally asymmetric, hence produce an azimuthally asymmetric shadow on the outer disk. Based on the short ($\\leq$1 week, Sitko 2007 personal communication) orbital time-scale of material located at the inner wall, we suggest that a series of resolved scattered light images of HD 163296, obtained during an epoch of enhanced self-shadowing, would provide a test of this particular suggested phenomenon. \\end{enumerate}" }, "0807/0807.3555_arXiv.txt": { "abstract": "Near-infrared photometric observations of the Type IIn SN 2005ip in NGC 2906 reveal large fluxes ($>$1.3 mJy) in the $K_s$-band over more than 900 days. While warm dust can explain the late-time $K_s$-band emission of SN 2005ip, the nature of the dust heating source is ambiguous. Shock heating of pre-existing dust by post-shocked gas is unlikely because the forward shock is moving too slowly to have traversed the expected dust-free cavity by the time observations first reveal the $K_s$ emission. While an infrared light echo model correctly predicts a near-infrared luminosity plateau, heating dust to the observed temperatures of $\\sim$1400-1600 K at a relatively large distance from the supernova ($\\ga 10^{18}$~cm) requires an extraordinarily high early supernova luminosity ($\\sim1\\times10^{11}$~\\lsolar). The evidence instead favors condensing dust in the cool, dense shell between the forward and reverse shocks. Both the initial dust temperature and the evolutionary trend towards lower temperatures are consistent with this scenario. We infer that radiation from the circumstellar interaction heats the dust. While this paper includes no spectroscopic confirmation, the photometry is comparable to other SNe that do show spectroscopic evidence for dust formation. Observations of dust formation in SNe are sparse, so these results provide a rare opportunity to consider SNe Type IIn as dust sources. ", "introduction": "\\label{sec_intro} For nearly 40 years, core-collapse SN events have been considered as possible sources of dust in the universe \\citep{cernuschi67,hoyle70}. More recent studies have proposed that the core-collapse supernovae may be the primary sources of dust in the early universe \\citep{todini01, nozawa03, dwek07}. Several models \\citep{todini01,nozawa03,nozawa08} succeed in producing the large amounts of dust observed at high redshifts, and recent data present the first evidence for a supernova origin for dust in an object at $z>6$ \\citep{maiolino04}. Nonetheless, direct observational evidence for dust formation in supernovae remains sparse, even in the local universe \\citep[][and references therein]{meikle07}. Any newly formed dust would produce a late-time near-infrared excess in comparison to the blackbody optical spectrum. \\citet{merrill80} was the first to detect such an infrared excess from a supernova (SN 1979C). Since then, only a handful of near-infrared excesses associated with core-collapse events have been observed \\citep[e.g.][and references therein]{dwek92,gerardy02,pozzo04,meikle06,smith08a}. Late-time near-infrared emission, however, is not unique to dust formation. Several possible mechanisms may give rise to such an excess due to the complex nature of the surrounding environment of most core-collapse supernovae (see Figure \\ref{f1}). (1) As the ejecta expand and cool, dust condenses and is radiatively heated. (2) The hot gas behind the forward shock collisionally heats pre-existing circumstellar gas and dust. (3) The SN peak luminosity heats pre-existing dust and produces an ``IR echo.'' (4) Dust condenses in either the forward or reverse shocks and is heated by radiative shock emission. The first scenario requires dust to condense from the ejecta as it expands and cools. The physical conditions in an expanding SN support this possibility. In particular, large abundances of the necessary elements, cooling of the ejecta via expansion, and dynamical instabilities can result in dust formation \\citep{lucy91,dwek92a,meikle93,roche93}. In this scenario the primary energy source is the radioactive power released in the expanding gas, which sets the luminosity of the dust. It is possible, however, for other energy sources to exist, such as circumstellar interactions. The second scenario describes pre-existing circumstellar dust heated directly by shock interactions. The scenario is more likely to occur at later times. Assuming a spherically symmetric circumstellar distribution of dust, \\citet{dwek83b} shows that a SN peak-luminosity of $\\sim10^{10}~$\\lsolar~creates a dust free cavity, via photo-evaporation, with a radius of $\\sim3\\times10^{17}$~cm (0.1 light years). For typical expansion velocities ($\\sim5000$~\\kms), this cavity delays the onset of emission by a couple of years. In addition, dust sputtering may limit the grain radiation efficiency so that only a small portion of the total shock power is converted into radiation \\citep{draine81}. \\begin{figure*}[t] \\plotone{f1.eps} \\caption{An illustration of the environment surrounding core-collapse supernovae similar to SN 2005ip (adapted from \\citet{smith08a}). Dust may form in a number of locations. Following the initial supernova explosion, the SN ejecta expand and cool sufficiently for dust formation to occur. A shocked region forms in between a forward shock in the circumstellar gas and a reverse shock in the supernova ejecta. If the gas is able to cool efficiently, dust condensation is likely to occur in the cool, dense shell behind the shock. Radiative cooling is more likely at the reverse shock because of the higher density, lower shock velocity, and the possibility of heavy element enrichment. Cooling at the forward shock is less likely, but the presence of clumps in the circumstellar medium, as indicated by the relatively narrow lines observed in SNe IIn, may allow radiative shocks.} \\label{f1} \\end{figure*} In the third scenario, the energy from the SN peak-luminosity once again creates a dust free cavity. The remaining shell of dust warms to high temperatures but does not vaporize. Due to light travel time effects, the thermal radiation from the dust grains reaches the observer over an extended period of time \\citep{dwek83b}. The infrared luminosity plateau occurs on year long time scales, which are set by the light travel time across the inner edge of the cavity. An IR echo's spectrum and temporal evolution are useful for estimating the mass-loss history of the progenitor star, and to investigate the geometry and composition of the circumstellar structure \\citep{dwek83b,dwek85,emmering88}. In the final scenario, the ejecta collide with the pre-existing circumstellar medium, which creates a forward shock. The circumstellar interaction decelerates the blast wave, thereby simultaneously creating a reverse shock. The deceleration converts some of the kinetic energy of the shock into X-rays and visible light. Dust grains are likely to condense in the cool, dense shell that forms behind the radiative shocks as they undergo a thermal instability \\citep{pozzo04}. Typically, the post-shock conditions surrounding reverse shocks are more conducive to dust condensation as the chemically rich, dense ejecta pass through the shock front and cool. The decelerated blast wave associated with the reverse shock is more likely to cool in sufficiently dense circumstellar environments. Type IIn events typically have the densest environments of any SNe, as they are defined by the ``narrow'' H lines that originate from clumps in the pre-existing dense H-rich shell that is moving at a relatively slow velocity \\citep{schlegel90}. Therefore, one might suppose dust formation in the cool, dense shell behind the reverse shock to be associated with Type IIn events. Unfortunately, Type IIn events are rare, consisting of only $\\sim$2-3\\% of all core-collapse SNe \\citep{galyam07}, and few well-studied events exist. This paper presents an analysis of the observed late-time near-infrared emission in the Type IIn SN 2005ip from days \\firstday-\\lastday~post-discovery. In \\S \\ref{sec_observations}, we present the observations, data reduction techniques, and photometry. In \\S \\ref{sec_analysis}, we discuss the different possible dust heating mechanisms and compare the SN to SNe with similar characteristics. The observations suggest that either an IR echo or dust formation in the post-shocked gas explain the long lived IR excess, but that dust formation is more likely. \\S \\ref{sec_conclusion} summarizes these results. ", "conclusions": "\\label{sec_conclusion} This paper presented near-infrared observations of SN 2005ip for the first \\lastday~days following detection. A large $K_s$-band luminosity persists even as the supernova's $J$-band luminosity falls. Among a variety of potential mechanisms, dust condensation in the cool, dense shell downstream from the reverse shock is the likely source of the observed infrared emission. We are able to rule out other mechanisms for late-time near-infrared emission. While dust condensation in the supernova ejecta is a likely possibility, the duration of the observed near-infrared light curve is inconsistent with the quickly declining radioactive heating source. Shock/mechanical heating of pre-existing dust grains from prior mass loss is not possible because the near infrared excess appears quickly ($<$100 days), providing insufficient time for the expanding ejecta to cross the dust-free cavity formed by the supernova. An IR echo successfully explains the fairly uniform luminosity over the period of observation. The duration of the infrared excess, however, implies a cavity size that requires a peak supernova luminosity that is much larger than observed for SN 2005ip and is comparable to the most luminous supernova ever observed. On the other hand, there is mounting evidence to support dust condensation in the shock interaction region. SN 2005ip shares many properties with SNe 1998S and 2006jc, for which spectra reveal direct evidence of dust condensation in the reverse shock \\citep{pozzo04,smith08b}. While we currently have no such spectra for SN 2005ip, the observed infrared luminosities and temperatures are consistent with dust forming in shocked, cooled gas. The observed X-rays indicate the infrared emission is most likely associated with the shock interaction region. Furthermore, the initial dust temperature of $\\sim1400-1600$~K is consistent with the dust condensation temperature of carbon-rich grains, which we expect to find in the ejecta encountering the reverse shock. The temperature evolution is also somewhat consistent with the decline predicted by an expanding cool, dense shell and diluted radiation field. Unlike the IR echo, this scenario is physically plausible and a large initial luminosity is not required. Of course, near-infrared photometry is only a small part of a much larger picture. Infrared spectra can better constrain the dust temperature, infrared flux, and emission/absorption efficiency than can be obtained with near-infrared photometry alone. Mid-infrared spectra also hold the prospects of revealing the mineralogy of the heated dust, distinguishing between carbon-rich (graphite) and oxygen-rich (silicate) grains. Based on the dust composition and mass loss history, the nature of the progenitor system may be deduced. Optical spectra are also of interest in that they can potentially reveal more evidence for dust formation and heating via circumstellar interaction \\citep[e.g.][]{fransson02} The overall similarity between SN 1995N and SN 2005ip suggests that SN 1995N be considered in the context of the dust condensation model, as opposed to the echo model favored by \\citet{gerardy02}. Indeed, \\citet{pastorello05} show that the late-time optical and infrared luminosities of SN 1995N are comparable, which suggests a correlation between the ejecta/circumstellar medium interaction and the warm dust. The IR echo model also predicts a optical light echo from the scattered light, as well as an infrared echo produced by the absorption of the supernova peak luminosity \\citep{chevalier86}. Yet detailed spectroscopic observations of SN 1995N suggest the late visible emission is due to circumstellar interaction and not an echo \\citep{fransson02}. These spectra reveal evidence for absorption in the diminution of the red sides of emission lines compared to the blue sides, as would be expected with dust formation. Furthermore, the authors present evidence for an increased helium abundance, CNO-burning products, and Ly$\\alpha$-pumped fluorescence of Fe II lines, all of which suggest a complex interaction between the ejecta and a pre-existing, dense circumstellar medium. The dust temperature of SN 1995N on day 730, $700-800$~K, is slightly smaller than for SN 2005ip. While this difference may be not be statistically significant, it is possible that SN 1995N had either a significantly higher expansion velocity or a different grain composition. Our discussion of SN 1995N and the data for SN 2005ip indicate that Type IIn SNe may be sources of dust in our universe. This association would explain the consistent late-time infrared emission observed in several Type IIn events \\citep[e.g.][]{pastorello02,gerardy02,pozzo04}. The rarity of Type IIn events, which represent only $\\sim$2-3\\% of all core-collapse SNe \\citep{galyam07}, makes collecting complete light curves difficult. Nonetheless, an interesting study would include follow-up observations of all Type IIn events over the past several years to determine which events are still bright in the infrared. We also reiterate the call by \\citet{gerardy02} for simultaneous UV and X-Ray observations to better constrain $\\tau$ and the amount of shock radiation being reprocessed by the dust grains." }, "0807/0807.4143_arXiv.txt": { "abstract": "\\noindent We investigate the possibility that a (light) hidden sector extra photon receives its mass via spontaneous symmetry breaking of a hidden sector Higgs boson, the so-called \\emph{hidden-Higgs}. The hidden-photon can mix with the ordinary photon via a gauge kinetic mixing term. The hidden-Higgs can couple to the Standard Model Higgs via a renormalizable quartic term -- sometimes called the \\emph{Higgs Portal}. We discuss the implications of this light hidden-Higgs in the context of laser polarization and light-shining-through-the-wall experiments as well as cosmological, astrophysical, and non-Newtonian force measurements. For hidden-photons receiving their mass from a hidden-Higgs we find in the small mass regime significantly stronger bounds than the bounds on massive hidden sector photons alone. ", "introduction": "Many extensions of the Standard Model contain so-called hidden-sectors which interact only very weakly with the {known particles from the visible sector}. Due to their feeble interactions particles in these hidden sectors can be easily missed in conventional collider experiments. Therefore, the bounds on their masses are often very weak and even masses in the sub-eV regime are possible. Such small masses, however, open the possibility that these particles may be detectable in low energy high precision experiments. Moreover, they could leave observable footprints in astrophysics and cosmology. This could therefore open a new window into particle physics which could give us crucial complementary information about the underlying laws of nature. One interesting class of hidden sector particles is additional U(1) gauge bosons, {\\it i.e.}~hidden-photons. For example many models arising from string compactifications contain extra U(1) gauge particles under which the Standard Model particles are uncharged. Accordingly the only renormalizable interaction of the hidden-photon with the Standard Model is via mixing of the hidden-photon with the ordinary electromagnetic photon~\\cite{Okun:1982xi,Holdom:1985ag,Foot:1991kb}. Current constraints on this mixing are shown in Fig.~\\ref{bounds1}. As can be seen from Fig.~\\ref{bounds1} the bounds depend crucially on the mass of the extra photon. In particular for very small masses the bounds become very weak. In this note we want to investigate if knowledge of the mechanism which generates a mass for the hidden-photon can improve the bounds. In principle a mass for the hidden-photon can be generated either via a Higgs mechanism or via a St\\\"uckelberg mechanism \\cite{Stueckelberg:1938}. Here, we focus mainly on the case of the Higgs mechanism. \\begin{figure}[tb] \\begin{center} \\includegraphics[width=0.7\\linewidth]{HP.eps} \\end{center} \\vspace{-2ex} \\caption[...]{\\small Current bounds on hidden-sector photons from analyzing the magnetic fields of Jupiter and Earth~\\cite{Goldhaber:1971mr}, Coulomb law tests \\cite{Williams:1971ms,Bartlett:1988yy} (gold), electroweak precision data \\cite{Feldman:2007wj} (lightgray), searches of solar hidden-photons with the CAST experiment (purple) \\cite{Popov:1991,Popov:1999,Andriamonje:2007ew,Redondo:2008aa} and light-shining-through-walls (LSW) experiments \\cite{Cameron:1993mr,Robilliard:2007bq,Chou:2007zz,Ahlers:2007rd,Ahlers:2007qf} {(grey)} as well as CMB measurements of the effective number of neutrinos and the blackbody nature of the spectrum (black) \\cite{Mangano:2006ur,Ichikawa:2006vm,Komatsu:2008hk,Jaeckel:2008fi}. Improvements of the solar bounds can be achieved using the SuperKamiokande detector or upgrading the CAST experiment \\cite{Gninenko:2008pz}. The region $m_{\\gamma^{\\prime}}\\lesssim {\\rm few}\\,\\,{\\rm meV}$ could be tested by an experiment using microwave cavities \\cite{Jaeckel:2007ch,Penny} or experiments searching for magnetic fields leaking through a superconducting shielding~\\cite{Jaeckel:2008sz}. \\label{bounds1}} \\end{figure} The crucial difference between the Higgs and the St\\\"uckelberg mechanism is that the gauge boson acquires a mass from the expectation value of a physical boson. As we will see in the following this additional boson will open new avenues of detection. Moreover, the additional physical boson also allows for a new possible renormalizable coupling to the Standard Model. The hidden sector Higgs can mix with the Standard Model Higgs via a quartic term~\\cite{Foot:1991bp}, sometimes called the Higgs-Portal~\\cite{portalref}. The paper is organized as follows. In the next Sect.~\\ref{gauge} we will present the essentials of the hidden-photon hidden-Higgs system including a gauge kinetic mixing term with the photon. This will already lead us to our first main conclusion. In processes where the momentum transfer is greater than the mass of the hidden-photon the hidden-Higgs behaves essentially like a minicharged particle and corresponding (strong) astrophysical and cosmological bounds apply. Then in Sect.~\\ref{magnetic} we discuss the effects of a strong magnetic field as relevant for laser polarization and light-shining-through-walls experiments. Again we find that the bounds improve significantly. Moreover we suggest possible ways to experimentally distinguish between the Higgs and St\\\"uckelberg mechanisms. In Sect.~\\ref{Portal} we then include effects of electroweak symmetry breaking and a possible mixing of the hidden-Higgs with the Standard Model Higgs via a Higgs-Portal term. Bounds from fifth-force experiments provide interesting constraints on the Higgs-Portal term which are independent of the size of the kinetic mixing. Finally, in Sect.~\\ref{conclusions} we summarize and conclude. ", "conclusions": "\\begin{figure}[tb] \\begin{center} \\includegraphics[width=0.7\\linewidth]{HP_MCP.eps} \\end{center} \\vspace{-2ex} \\caption[...]{\\small Bounds on the kinetic mixing parameter for massive hidden sector photons (cf. the caption of Fig.~\\ref{bounds1}). Regions labeled in italic are the the bounds that apply if the mass arises from a Higgs mechanism, and the Higgs boson appears as a minicharged particle. We have bounds from a SLAC beamdump experiment, invisible orthopositronium decays, light-shining-through-walls experiments (LSW), big bang nucleosynthesis (BBN), and energy loss considerations in supernovae~(SN1986a), white dwarfs and red giants (see~\\cite{Davidson:1993sj,Badertscher:2006fm,Davidson:2000hf,Ahlers:2007qf}). Notice that we have assumed $g_Xq_X=e$ (so $|\\chi|=|q_\\theta|$) and $m_{\\gamma'}\\simeq m_h$. For $g_X q_X \\ll e$, typically $m_h\\gg m_{\\gamma'}$ and the new bounds move upwards and to the left.\\label{newbounds}} \\end{figure} Extra `hidden' U(1) gauge bosons appear in many extensions of the Standard Model. The bounds on these hidden sector photons depend crucially on their mass. This mass can arise either via a St\\\"uckelberg mechanism or from a Higgs mechanism. In this paper we have investigated if one can use knowledge about the mechanism that generates the mass to improve the bounds. In particular, we have focused on the case of the Higgs mechanism. The crucial point in the case of the Higgs mechanism is that it provides an extra degree of freedom which leads to additional experimental and observational constraints. Indeed, at large momentum transfer as, {\\it e.g.}, in the interior of stars, a light hidden-Higgs behaves as a minicharged particle. A similar behavior is found inside strong magnetic fields. This can be used to translate bounds on minicharged particles into bounds on massive hidden sector photons. As can be seen from Fig.~\\ref{newbounds} this leads to a dramatic strengthening of the (astrophysical as well as laboratory) bounds for small masses. The hidden-Higgs field $\\theta$ also provides new potential couplings to the Standard Model. In particular, it allows for a renormalizable interaction with the Standard Model Higgs $\\phi$ via a so-called Higgs-Portal term $\\kappa|\\theta|^2|\\phi|^2$. This coupling leads to fifth-force type couplings which can be used to obtain bounds on this coupling which are independent of the size of the kinetic mixing between photon and hidden-photon (cf.~Fig.~\\ref{nonnewton}). This opens the `Higgs-Portal' to the physics of light hidden sectors." }, "0807/0807.3588_arXiv.txt": { "abstract": "We show that a consistent fit to observed secondary eclipse data for several strongly irradiated transiting planets demands a temperature inversion (stratosphere) at altitude. Such a thermal inversion significantly influences the planet/star contrast ratios at the secondary eclipse, their wavelength dependences, and, importantly, the day-night flux contrast during a planetary orbit. The presence of the thermal inversion/stratosphere seems to roughly correlate with the stellar flux at the planet. Such temperature inversions might caused by an upper-atmosphere absorber whose exact nature is still uncertain. ", "introduction": "Theoretical modeling of atmospheres of extrasolar giant planets (EGP) is a young field, about one decade old, and as youngsters usually are, it is very active, restless, and sometimes quite unpredictable. It had recently undergone a transition from a purely care-free stage (no observations were available) to a more difficult stage where there already are some observations to be fit by the theory. In this paper, we briefly describe our recent efforts in this area. We will try to convince the reader that despite its young age and many shortcomings, the theory is actually doing quite well. ", "conclusions": "We have found (\\cite{BHBKC08}; \\cite{BBH08}) that a consistent fit to all data at secondary eclipse for several strongly irradiated transiting planets (HD 209458b, HD 149026b, and possibly HD 189733b), and very likely a non-transiting planet $\\upsilon$ And b, requires that their atmospheres have temperature inversions -- stratospheres -- at altitude. Such a thermal inversion affects: (i) planet/star contrast ratios at the secondary eclipse; (ii) their wavelength dependences; and (iii) day-night flux contrast during a planetary orbit. Moreover, the presence of the thermal inversion/stratosphere seems to roughly correlate with the total irradiated flux. Temperature inversion is caused either by TiO/VO, as first suggested by \\cite{HBS03}, or by another, as yet unidentified, opacity sources. These may be tholins, polyacetylenes, or various non-equilibrium compounds. We invoke such extra absorbers, because a cold-trap effect can operate to deplete the upper atmosphere of TiO/VO. However, one may speculate that with ongoing mass loss and/or rotational shear instabilities the atmosphere may be partially replenished in TiO/VO. Therefore, while TiO/VO might be responsible for the formation of thermal inversions in the strongly irradiated planets, the exact nature of the absorber must be viewed as very uncertain." }, "0807/0807.4682_arXiv.txt": { "abstract": "We consider the Friedmann-Robertson-Walker cosmologies of theories of gravity that generalise the Einstein-Hilbert action by replacing the Ricci scalar, $R$, with some function, $f(R)$. The general asymptotic behaviour of these cosmologies is found, at both early and late times, and the effects of adding higher and lower powers of $R$ to the Einstein-Hilbert action is investigated. The assumption that the highest powers of $R$ should dominate the Universe's early history, and that the lowest powers should dominate its future is found to be inaccurate. The behaviour of the general solution is complicated, and while it can be the case that single powers of $R$ dominate the dynamics at late times, it can be either the higher or lower powers that do so. It is also shown that it is often the lowest powers of $R$ that dominate at early times, when approach to a bounce or a Tolman solution are generic possibilities. Various examples are considered, and both vacuum and perfect fluid solutions investigated. ", "introduction": "We study here the dynamics of Friedmann-Robertson-Walker (FRW) universes in $f(R)$ theories of gravity. These theories are derived from generalisations of the usual Einstein-Hilbert Lagrangian of General Relativity (GR), such that \\begin{equation} \\label{gravL} \\mathcal{L} = f(R), \\end{equation} and have been considered extensively in the literature (see e.g. \\cite{early1,early2,early3,early4}). Specification of the function $f(R)$ defines the theory, and GR can be seen to be the special case $f =R$. Such theories have drawn considerable interest as they found success in early attempts to create a perturbatively re-normalisable quantum field theory of gravity \\cite{qgrav}, as well as turning up more recently in the effective actions of string theory \\cite{string1,string2}. In cosmology these theories have been used extensively in attempts to explain the late-time accelerating expansion of the Universe \\cite{accel1,accel2}, cosmological inflation \\cite{infl1,infl2,infl3} and the nature of the initial singularity \\cite{singul1,singul2,singul3}. For a recent review see \\cite{review}. In considering generalised $f(R)$ theories of gravity it is often implicitly assumed that at late times in the evolution of the Universe it should be the lowest powers of $R$ that dominate the gravitational Lagrangian. That is, at late times we should have $R \\rightarrow 0$, and the Universe should behave as if it were governed by a gravitational Lagrangian of the form \\begin{equation} \\label{low} \\mathcal{L}_0 = \\lim_{R \\to 0} f(R), \\end{equation} which is often presumed to correspond to the Einstein-Hilbert Lagrangian, although other limits have been considered in attempts to address the apparent late-time acceleration of the Universe \\cite{accel1,accel2}. Conversely, the introduction of higher powers of $R$ into the gravitational Lagrangian has often been assumed to mean that at early times the Universe should behave as if governed by the vacuum dynamics of a Lagrangian \\begin{equation} \\label{high} \\mathcal{L}_{\\infty} = \\lim_{R \\to \\infty} f(R). \\end{equation} The picture is then one of a universe that starts off at high $R$, dominated by a Lagrangian of the form (\\ref{high}), and that subsequently expands until $R$ becomes small and the gravitational dynamics are dominated by a Lagrangian of the form (\\ref{low}). It is the purpose of this paper to determine the veracity of such assumptions. This is achieved by studying the dynamical evolution of FRW universes, governed by theories with general $f(R)$. The asymptotic behaviour of the general solution to the Friedmann equations is then investigated, and used to evaluate the extent to which the afore mentioned behaviour may be considered generic. The general solutions of FRW cosmologies governed by $f(R)$ theories of gravity have been studied previously by a number of authors, in a number of different contexts. Much of this work has made use of the dynamical systems approach, which has been used to study specific classes of $f(R)$ in isotropic cosmologies in \\cite{Power, FRW1,FRW2}, and anisotropic cosmologies in \\cite{Bianchi1,Bianchi2,Bianchi3}. Exact analytic expressions have been found for the general FRW solutions of some $f(R)$ theories in \\cite{exact}, and the dynamical systems approach applied to general $f(R)$ has been considered in \\cite{gen}\\footnote{See \\cite{crit} for a criticism of this work. The present study is free from the defects pointed out in \\cite{crit}.}. For studies of spherically symmetric and weak field solutions see \\cite{review,Power,weak,ppn}, and references therein. The approach used here is a generalisation of the analysis performed in \\cite{Power}, where theories of the form $f \\propto R^n$ were considered. We find here that the late-time attractors of FRW cosmologies for general $f(R)$ have various different forms, and that the asymptote toward which the general solution is attracted depends upon the initial conditions. Some of these solutions correspond to the lowest powers of $f(R)$ dominating at late-times, and others to the highest powers. Expanding universes with powers of $R$ lower than $R^2$ dominating their dynamics generically appear to exhibit the former behaviour, while universes with powers of $R$ greater than $R^2$ dominating appear to generically exhibit the latter. Expanding universes with higher powers of $R$ dominating their early evolution therefore appear unlikely to evolve to a state where the Einstein-Hilbert term dominates. We also find that there usually exist multiple early-time attractors for the general solution. These can take on different forms, but generically it appears that they either evolve as $a \\sim t^{\\frac{1}{2}}$, toward a big bang singularity in the past, or that they approach a point of inflexion, where the scale factor is constant. This is in good agreement with the analytic general solutions for $f \\propto R^n$ found in \\cite{exact}. These results do not mean that a period of inflation cannot occur (as indeed it appears to if $f \\sim R^2$), but it does mean that the general solution does not generally start off inflating (even if $f \\sim R^2$). It also means that the picture of the highest powers of $R$ dominating the earliest stages of the Universe's evolution may not be an accurate one. We begin in section 2 by giving the FRW field equations for $f(R)$ theories, together with some simple power-law particular solutions that will later appear as asymptotes of the general solution. In section 3 we use a dynamical systems approach to determine the form of the general solution for vacuum cosmologies. The phase space of the general solution is two dimensional, and the location and stability of critical points in this space is determined. In section 4 we perform a similar analysis for the case of perfect fluid cosmologies. The phase space of the general solution is now three dimensional, and the location and stability of critical points is again determined. In section 5 we consider the effect of adding higher and lower powers of $R$ to the Einstein-Hilbert action. Section 6 provides a discussion of the results, and the appendix gives some special cases that are of particular interest. ", "conclusions": "We have considered in this paper the evolution of spatially flat FRW universes governed by $f(R)$ theories of gravity. The Friedmann equations (\\ref{Friedmann1})-(\\ref{conservation}) were transformed into an autonomous system of first-order differential equations, and a dynamical systems analysis was performed. The location and stability of all critical points in the phase space were found, for both vacuum and perfect fluid cosmologies, and for general $f(R)$. It was shown that the simple power-law solutions, given by equations (\\ref{vacuum})-(\\ref{radiation}), often act as the early and late-time asymptotes of the general solution. The general behaviour of $f(R)$ FRW cosmologies is complicated. The phase space of solutions is often divided into sub-spaces by invariant manifolds, through which the trajectories describing the general solution cannot pass. The asymptotic past of general solutions can contain points of inflexion, or big-bang singularities that can be approached in different ways. Similarly, future behaviour can be seen to be able to asymptote toward matter dominated expansion, (\\ref{matter}), vacuum domination, (\\ref{vacuum}), or various other forms. Whatsmore, the three dimensional phase space of solutions, in the presence of a perfect fluid, allows for the possible existence of strange attractors, and chaotic behaviour. Nevertheless, despite the complicated behaviour exhibited by the general solutions, it is still possible to make statements about the effects of modifying the Einstein-Hilbert action to more general functions of $R$. It can be said that theories that contain lower powers of $R$ often have a stable asymptote that corresponds to the expanding vacuum dominated solution, (\\ref{vacuum}), of that lowest power, and that therefore behave as if governed by a gravitational Lagrangian of the form (\\ref{low}) at late times. However, if a theory contains any higher powers of $R$ then there is also often a stable asymptote that corresponds to the expanding vacuum dominated solution, (\\ref{vacuum}), of that highest power, and that therefore behaves as if governed by a Lagrangian of the form (\\ref{high}) at late times. Theories containing both higher and lower powers of $R$ can then asymptote, at late times, to regimes in which either the lowest {\\it or} highest powers or $R$ dominate. In either case the consequent evolution is that of a gravitational Lagrangian dominated by a single power of $R$. Which solutions asymptote to high $R$ domination, and which to low $R$ domination, depends on the initial conditions, and the form of $f(R)$. Using illustrative examples we have shown that if a power of $R$ higher than $R^2$ dominates at some point then the generic behaviour of expanding solutions is to higher $R$. If a power of $R$ lower than $R^2$ dominates, then the trend is to lower $R$. A term $R^2$ in the Lagrangian is then a special case; if it dominates then the expanding attractor corresponds to exponential growth, and acts as a separatrix between the higher or lower powers of $R$ dominating the future dynamics of the Universe. Those solutions with low $R$ dominating at late-times expand eternally, while those with high $R$ dominating approach either a big-rip singularity, or exponential expansion. Big-rips often occur if there exists a single power of $R$ that dominates at late times, and exponential expansion occurs if $B \\rightarrow 1$, as is the case for some infinite power series, such as $f \\sim \\exp \\{ R \\}$. This late-time acceleration does not appear to be a good candidate for the apparent acceleration we observe, however, as it cannot follow from a period of Einstein-Hilbert domination in which $R \\rightarrow 0$. The asymptotic past of the general solutions is similarly complicated. A variety of behaviours seems possible, including big-bang singularities and bounces, where the scale factor reaches a non-zero minimum. Big-bang singularities are often approached with the scale-factor behaving as in the Tolman solution, (\\ref{radiation}). It is interesting to note that even trajectories which undergo an early period of inflation (such as those solutions in which a power of $R^2$ dominates at some point) do not generically follow such expansion indefinitely into the past, but rather have a big bang or bounce at some point in their past, prior to the onset of inflation. While the behaviour found here is quite complicated, with numerous different asymptotes possible, it is certainly not the most general case that one may consider. We have limited ourselves here to spatially flat FRW universes. Relinquishing the criterion of spatial flatness would lead to more complicated behaviours still, and one may also consider inhomogeneous and/or anisotropic cosmologies, or cosmologies with multiple fluids. It is not clear whether or not the behaviour identified above would hold in these more general cases or not. What does seem clear, however, is that the simple picture of the highest powers of $R$ dominating at early times, and lower powers dominating at late times, is unlikely to be accurate. \\appendix" }, "0807/0807.4820_arXiv.txt": { "abstract": "Data points are placed in bins when a histogram is created, but there is always a decision to be made about the number or width of the bins. This decision is often made arbitrarily or subjectively, but it need not be. A jackknife or leave-one-out cross-validation likelihood is defined and employed as a scalar objective function for optimization of the locations and widths of the bins. The objective is justified as being related to the histogram's usefulness for predicting future data. The method works for data or histograms of any dimensionality. ", "introduction": "There are many situations in experimental science in which one is presented with a collection of discrete measurements $\\xx_j$ and one must bin those points into a set of finite-sized bins $i$, with centers $\\XX_i$ and full-widths $\\DD_i$, to create a histogram of numbers of points $N_i$, or the equivalent when the points have non-uniform weights $w_j$. The problem of binning comes up, for example, when one needs to plot a data histogram, when one needs to perform least-square fitting of a probability distribution function, and when one wants to compute entropies or other measurements on the inferred data probability distribution function. The choice of bin centers and widths often seems arbitrary. However, there is a non-arbitrary choice, derived below, which emerges when the histogram is thought of as an estimate of the probability distribution function of whatever process generated the data. If the binning is too coarse, the histogram does not give much information about the shape of the probability distribution function. If the binning is too fine, bins become empty and the histogram becomes noisy, so it in some sense ``overfits'' the data. The best binning lies in between these extremes and can be found simply and quickly by a ``jackknife'' or cross-validation method, that is, by excluding data subsamples and using the non-excluded data to predict the excluded data. This is not the only data-based binning-choice approach\\footnote{\\raggedright see, for example, Knuth,~K.~H., ``Optimal data-based binning for histograms,'' arXiv:physics/0605197, and references cited therein.}, but it is simple and sensible. In what follows, we are going to consider a data histogram, which we imagine as a set of bins $i$, with centers $\\XX_i$ and widths (or multi-dimensional volumes) $\\DD_i$. Equivalently (and perhaps more usefully), the parameterization of the bins can be described by a set of edges $\\XX_{(i-1/2)}$ so the centers become $\\XX_i=\\left(\\XX_{(i-1/2)}+\\XX_{(i+1/2)}\\right)/2$ and the widths become $\\DD_i=\\left|\\XX_{(i+1/2)}-\\XX_{(i-1/2)}\\right|$. These bins will get filled by a set of (possibly multi-dimensional) data points $\\xx_j$, leading to each bin $i$ containing a number of data points $N_i$. We will also make reference to the binning function $i(\\xx)$ which, for a given data value $\\xx$, returns the bin $i$. ", "conclusions": "I have shown that when a histogram of data needs to be made, there \\emph{is} a non-arbitrary choice of binning. Some qualitative observations follow. \\begin{itemize} \\item The optimal bin widths get smaller as the number of data points goes up or as the features in the (true) probability distribution function get narrower. \\item The results are more sensitive to the smoothing parameter $\\alpha$ when the number of empty or near-empty bins becomes significant. \\item The jackknife likelihood makes discontinuous jumps as the bin edges cross individual data points. For this reason, the likelihood does not have well-defined derivatives. Some care must be taken that the optimization technique does not depend on having a differentiable likelihood function. \\item There is nothing special about one-dimensional or two-dimensional distribution functions; this is easily generalized to $n$-dimensional distributions. However, it takes a lot of data points to measure a distribution function in $n$ dimensions when $n$ is large; I understand that the required number of data points scales worse than $e^n$ [need ref]. \\item There is nothing special about equal-width binning; I simply chose this to make the optimization problem easily tractable and the results easily presentable. \\item This method makes no reference to the \\emph{errors} or \\emph{uncertainties} on the measurements $\\xx_j$. Effectively, I have assumed that the errors are small relative to any real features in the probability distribution function. In practice, it is rarely useful to have more than a few bins per the width of your error distribution, if all the points have similar uncertainties. \\item There is often an additional choice about what minimum and maximum data values to allow for histogramming. This choice also ought to be made in a non-arbitrary fashion if there are data points that will be excluded by the choice. \\item Finally, there is nothing special about the ``tophat'' binning model used in the above examples. Everything can be generalized to smoothly overlapping bins, in which points are assigned fractionally to multiple bins. In general, smoother binning models make for more well-behaved derivatives of the jackknife likelihood and therefore more straightforward optimization. This can also all be generalized to kernel-smoothing techniques for density estimation, which ought to be made the subject of a separate note. \\end{itemize} \\paragraph" }, "0807/0807.1634_arXiv.txt": { "abstract": "We study direct and indirect detection possibilities of neutralino dark matter produced non-thermally by e.g. the decay of long-lived particles, as is easily implemented in the case of anomaly or mirage mediation models. In this scenario, large self-annihilation cross sections are required to account for the present dark matter abundance, and it leads to significant enhancement of the gamma-ray signature from the Galactic Center and the positron flux from the dark matter annihilation. It is found that GLAST and PAMELA will find the signal or give tight constraints on such nonthermal production scenarios of neutralino dark matter. ", "introduction": "While there are lots of cosmological evidence of the dark matter in the universe \\cite{Jungman:1995df,Bertone:2004pz}, its detailed properties remain largely undetermined. Requirements for the dark matter particle are the following. (1) It reproduces the present matter density of the universe. In terms of the density parameter, $\\Omega_m h^2 \\sim 0.11$ must be satisfied where $h(\\sim 0.70)$ is the Hubble parameter in units of 100~km/s/Mpc \\cite{Komatsu:2008hk}. (2) It is electrically neutral. (3) It is cold, which means that its free-streaming length ($\\lambda_{\\rm FS}$) is not so long as to seed the structure formation satisfactory, and this requires $\\lambda_{\\rm FS} \\lesssim 1$~Mpc. In fact many candidates of dark matter are proposed in the framework of physics beyond the standard model. In particular, supersymmetry (SUSY) provides interesting candidates. If $R$-parity is conserved, the lightest SUSY particle (LSP) becomes stable and contributes present matter density of the universe. Among SUSY particles, the gravitino and (lightest) neutralino are possible candidates of the LSP dark matter. From the viewpoint of detection possibility, the gravitino dark matter is undesirable because its interaction strength with ordinary matter is Planck-suppressed.\\footnote{ Recently it is pointed out that the detection of inflationary gravitational wave background can help the situation \\cite{Nakayama:2008ip}. } In the following our focus is the neutralino dark matter, which may have distinct signatures of direct and/or indirect detection. Usually neutralinos are assumed to be produced thermally as in the following scenario \\cite{Jungman:1995df,Kolb:1990}. In the early universe with temperature $T\\gtrsim 1$~TeV, SUSY particles including neutralinos are thermalized and their number density is given by $\\sim T^3$. As the temperature decreases, their thermal abundance receives Boltzmann suppression factor and eventually they decouple from thermal bath at the freeze-out temperature $T_f\\sim m_{\\rm L}/20$ where $m_{\\rm L}$ denotes the LSP mass. After that the number density of the LSP per comoving volume remains constant until now and hence contributes as dark matter of the universe. The resultant abundance of the LSP is estimated as \\begin{equation} Y_{\\rm L}\\equiv \\frac{n_{\\rm L}}{s} \\sim \\frac{1}{T_f M_P \\langle \\sigma v \\rangle}, \\end{equation} where $\\langle \\sigma v \\rangle$ denotes the annihilation cross section of the LSP and $M_P$ is the reduced Planck scale $(=2.4\\times 10^{18}$~GeV). However, such a thermal relic scenario does not always hold in realistic SUSY models. For example, there often exists Polonyi or moduli field in order to break SUSY and gives rise to correct order of gaugino masses. Those singlet scalar fields generally have long lifetime and decay after freeze-out of the LSP, yielding substantial amount of LSPs. Actually Polonyi/moduli dominate the universe before they decay, and hence reheat the universe again with very low reheating temperature of $O(1)$~MeV-$O(1)$~GeV depending on their masses \\cite{Moroi:1994rs,Kawasaki:1995cy}. In this case a large amount of LSPs are produced non-thermally by the Polonyi/modulus decay, and hence large annihilation cross section is needed to account for the present dark matter abundance. Therefore, taking into account nonthermal production mechanism may significantly change the properties of the LSP and its direct/indirect detection signatures. Thus in this paper we study direct/indirect detection signatures of non-thermally produced neutralino dark matter. As for direct detection, there are some ongoing and planned projects devoted to detect scattering signals of the LSP with nucleons such as CDMS \\cite{CDMS} and XENON \\cite{XENON}. As for indirect detection, many possible ways are proposed. First, neutralinos accumulated in the Galactic Center annihilate each other and produce line and continuum gamma rays. Such gamma-ray signals can be searched by satellite experiments (GLAST \\cite{GLAST}) or ground-based Cerenkov telescope (HESS \\cite{HESS}, MAGIC \\cite{MAGIC}, CTA \\cite{CTA}). Second, anti-matter such as positrons or anti-protons are produced by the annihilation of the neutralinos. Since these particles are diffused by galactic magnetic fields during their propagation to the Earth, we need to solve its propagation in a diffusion model to discuss their flux on the Earth \\cite{Longair}. Fortunately, the positron flux is less sensitive to the precise diffusion model since magnetic fields easily dissipates their energy through the propagation and positrons come only from near the Earth. These anti-matter signals can be detected PAMELA \\cite{PAMELA} and AMS-02 \\cite{AMS02}. Third, neutralinos trapped in the Sun annihilate and produce high-energy neutrinos. Super Kamiokande \\cite{Desai:2004pq}, AMANDA \\cite{Andres:1999hm}, IceCube \\cite{Ahrens:2003ix} and planned KM3NeT \\cite{KM3NeT} experiments search high energy muon signals, which arise from high-energy neutrino interaction with Earth matter. We investigate characteristic signals of nonthermal neutralino dark matter on these experiments. For the sake of concreteness, we stick to two SUSY breaking models : minimal anomaly-mediated SUSY breaking model \\cite{Randall:1998uk} and mirage-mediation model \\cite{Choi:2004sx}. The former model predicts the wino-like neutralino LSP in broad parameter regions. Since the wino-like neutralino with a mass of ${\\cal O}(100)$ GeV has too large annihilation cross sections, its thermal abundance becomes too small to account for the present dark matter abundance. Hence we need to consider some non-thermal production processes of the neutralino dark matter. The latter model contains a heavy modulus field and non-thermal production of the neutralino dark matter is expected naturally. As we will see later, the large annihilation cross section of the neutralino dark matter is a general feature of non-thermal production scenario, and hence our results are less sensitive to the model construction. A similar subject was studied in Ref.~\\cite{Profumo:2004ty}, where it was pointed out that the large annihilation cross section of the neutralino yields enhancement of the anti-matter signals. We emphasize that such an enhancement is rather generic feature when considering the nonthermal production scenario of the dark matter, and its detection may be directly related to the early Universe cosmology, in particular, the existence of late-decaying particles and their decay temperature. Also we have performed more detailed parameter analyses both in the anomaly-mediated SUSY breaking and mirage-mediation models, including the gamma-ray signature as well as anti-matter searches. This paper is organized as follows. In Sec.~\\ref{sec:NT} we review some non-thermal production mechanisms of the neutralino dark matter. The minimal anomaly-mediated SUSY breaking model and the mirage-mediation mode are taken as examples. In Sec.~\\ref{sec:detection} detection possibilities of nonthermal dark matter are discussed. These include direct detection using recoil of nuclei by the neutralino, gamma-ray flux from the neutralino annihilation at the Galactic Center, positron flux from the annihilation near the Earth, and high energy neutrino flux from the annihilation in the Sun. Sec.~\\ref{sec:conclusion} is devoted to our conclusions. For calculating these direct/indirect detection rates, we have utilized DarkSUSY code \\cite{Gondolo:2004sc}. ", "conclusions": "\\label{sec:conclusion} We have discussed direct and indirect detection signatures of neutralino dark matters produced non-thermally with a very low reheating temperature $T_d\\sim {\\cal O}(1) \\rm{MeV} -{\\cal O}(1)\\rm{GeV}$. In this scenario, the self-annihilation cross section of dark matter should be large enough to account for the present relic abundance, within the rage consistent with bounds provided in \\cite{Mack:2008wu}. In SUSY models, such a large annihilation cross section is naturally realized for the neutralino dark matter with significant wino or Higgsino components. In the case of bino-like neutralino, such a large annihilation cross section can be obtained by s-channel Higgs resonance. In both cases, the large annihilation cross section leads to the enhancement of gamma-ray signals and the positron flux from the dark matter annihilation, and it becomes promising to detect the non-thermally produced dark matter with $T_\\chi\\lsim 1{\\rm GeV}$ by these indirect detection experiments. In other words, the indirect detection experiments may give us clues to explore the history of the universe with the temperature up to $1~{\\rm GeV}$. Obviously, the consideration with other observations is important and essential to make definitive conclusion about the non-thermal production scenario. For example, it is known that the large annihilation cross section of the LSP affects the Big-Bang-Nucleosynthesis (BBN), and non-thermally produced dark matter with $T_\\chi\\lsim {\\cal O}(100) {\\rm MeV}$ may be severely constrained by the observation of $^6$Li abundance \\cite{Jedamzik:2004ip}. And also the combination with collider experiments may be most important. The upcoming Large Hadron Collider (LHC) experiments are expected to discover new particles relevant to the EWSB in the standard model, and there may appear dark matter candidates. Once such a dark matter candidate is discovered, we may have insight on its production mechanism in the universe by comparing the theoretical calculation of the cross section with cosmological and astrophysical observations, such as the dark matter abundance and its direct and indirect detection signatures. As explained in this paper, large indirect detection signals are characteristic features for the non-thermally produced dark matter, and combined with LHC experiments we may probe the nature of dark matter by these experiments. {\\it Note added:} While finalizing this manuscript, Ref.~\\cite{Grajek:2008jb} was submitted to the preprint server, which studied similar subject to our present work. While they focus on non-thermally produced wino and higgsino like dark matter, we have also studied bino-like one in the s-channel resonance region, and performed detailed parameter analyses in MAMSB and mMSB models. The main conclusion seems to be consistent with ours." }, "0807/0807.3966_arXiv.txt": { "abstract": "Using the Spitzer Space Telescope, we have obtained 3.6--24\\,$\\mu$m photometry of 38 radio galaxies and 24 quasars from the 3CR catalog at redshift $11$). In order to assess galaxy and AGN evolution in the universe, we therefore need to understand this AGN/starburst degeneracy for a population of luminous high-redshift sources. A crucial step towards this is to study the orientation dependence of the NIR and MIR emission of high-redshift AGN. Orientation-dependent effects can only be tested and quantified with AGN samples having type~1 (unobscured) and type~2 (obscured) subsamples matched in isotropic emission. The clean AGN tracers --- optical, [\\ion{O}{3}]~$\\lambda$5007\\,\\AA, NIR, and X-ray ($\\la$10~keV) --- all fail to fulfill this requirement. The [\\ion{O}{2}]~$\\lambda$3727\\AA ~emission, while isotropic (Hes et al.\\ 1993), is probably dominated by extended starbursts and shocks (Best et al.\\ 2000) rather than by the AGN. Therefore, the only feasible way is low-frequency (meter-wavelength) radio selection because the integrated emission from the radio lobes is optically thin and essentially isotropic. This makes radio-loud AGN particularly attractive for studying orientation-dependent properties at other wavelengths and, after sorting out the influence of radio jets/lobes on the emission, for generalizing conclusions about orientation-dependent effects to the much larger population of radio-quiet AGN. The brightest low-frequency-selected AGN sample is the 3CR compilation (Spinrad et al.\\ 1985). The powerful double-lobed radio galaxies (henceforth simply called radio galaxies) are supposed to be misaligned quasars (Barthel 1989). Based on {\\it IRAS} coadded scans and a few individual detections, Heckman et al.\\ (1992, 1994) already noted an average MIR/FIR difference between 3CR quasars and radio galaxies. More comprehensive MIR and FIR spectrophotometry from {\\it ISO} is in hand (as compiled by Siebenmorgen et al.\\ 2004 and by Haas et al.\\ 2004) as well as from {\\it Spitzer} (e.g., Shi et al.\\ 2005, Haas et al.\\ 2005, Ogle et al.\\ 2006, Cleary et al.\\ 2007), providing a basis to study the $z < 1$ 3CR objects. These sources are, however, a factor of five less radio-luminous on average than the most powerful radio sources seen at higher redshift, and the lower indicated accretion power may reflect different source physics. The higher-luminosity population can be sampled by the 3CR sources at $11$ prior to normalization. 'x' symbols denote quasars; circles and squares denote radio galaxies. Superposed crosses indicate radio galaxies that show evidence of silicate absorption (\\S\\ref{sec_rg}). The vertical long-dashed lines mark the range of our luminosity-matched quasar and radio galaxy subsamples. The dotted lines indicate $L_{8\\micron}/L_{\\rm 178\\,MHz}$ ratios of 1, 10, and 100. The radio galaxies are grouped into several SED classes in Fig.\\,\\ref{fig_nir_mir_cc} and \\S\\ref{sec_rg}. The color-coding and symbols are: green circle (A), red circle (B), red square (C), blue square (D), blue circle (E). The two low-excitation radio galaxies 3C\\,68.1 and 3C\\,469.1 are labeled with their 3C numbers, as are sources outside the luminosity range of our analysis. } \\label{fig_ir_radio_lum} \\end{center} \\end{figure*} ", "conclusions": "The 3CR sample at $11$) and those without ($\\gamma<1$), where $\\gamma^4=\\kapa/\\kapp$. Phase measurements coupled to inversion determinations will uniquely place you in a given quadrant.} \\label{fig:diagram} \\end{figure}" }, "0807/0807.4294_arXiv.txt": { "abstract": "TeV $\\gamma$ rays from distant astrophysical sources are attenuated due to electron-positron pair creation by interacting with ultraviolet/optical to infrared photons which fill the universe and are collectively known as the extra-galactic background light (EBL). We model the $\\sim$0.1--10~eV starlight component of the EBL derived from expressions for the stellar initial mass function, star formation history of the universe, and wavelength-dependent absorption of a large sample of galaxies in the local universe. These models are simultaneously fitted to the EBL data as well as to the data on the stellar luminosity density in our local universe. We find that the models with modified Salpeter A initial mass function together with Cole et al. (2001) or Hopkins \\& Beacom (2006) star formation history best represent available data. Since no dust emission is included, our calculated EBL models can be interpreted as the lower limits in the $\\sim$0.1--1~eV range. We present simple analytic fits to the best-fit EBL model evolving with redshift. We then proceed to calculate $\\gamma$-ray opacities, and absorption of $\\sim$10--300~GeV $\\gamma$-rays coming from different redshifts. We discuss implications of our results for the Fermi Gamma Ray Space Telescope and ground-based Air Cherenkov Telescopes. ", "introduction": "Stars are the dominant sources of electromagnetic radiation in the universe after the cosmic microwave background (see, e.g., Fukugita \\& Peebles 2004). They emit radiation longward from ultraviolet to infrared wavelengths. However, photons with wavelength $\\lesssim 2~\\mu$m are highly absorbed by the dust in the host galaxies and only a fraction of the radiation emitted by the stars escape to the inter-galactic medium and form a diffuse background or EBL (see, e.g., Baldry \\& Glazebrook 2003; Driver et al. 2008). The dust in the host galaxies, heated by the starlight, also radiate in the infrared wavelengths and contribute to the EBL density at $\\sim 10^{12}$ Hz. It is the direct starlight component, $\\lesssim 2~\\mu$m or $\\gtrsim 0.1$~eV, that affects the propagation of $\\lesssim 5$~TeV $\\gamma$-rays from distant sources. Indeed, the very soft spectral energy distribution ($dN/dE \\propto E^{-\\Gamma}$) with $\\Gamma \\gtrsim 3$ observed from several TeV blazars at high redshift ($z\\gtrsim 0.1$) such as PKS 2155-304 (Aharonian et al. 2005), H 2356-309 (Aharonian et al. 2006a); 1ES 1218+304 (Albert et al. 2006); 1ES 1101-232 (Aharonian et al. 2006b); 0347-121 (Aharonian et al. 2007), 1ES 1011+496 (Albert et al. 2007) and 3C 279 (Albert et al. 2008), and their cutoff at $\\gtrsim 1$~TeV are hints that high energy $\\gamma$ rays from these sources are absorbed by the EBL UV/optical photons (Persic \\& de Angelis 2008). Lower energy ($<$TeV) $\\gamma$-rays from high redshift sources such as gamma-ray bursts (GRBs) and blazars can also probe the EBL starlight component. Calculation of the opacity of the universe to $\\gamma$-rays by $\\gamma\\gamma \\to e^+e^-$ process dates back to Nishikov (1961), followed by Gould \\& Shr\\'eder (1966) and Fazio \\& Stecker (1970). More recently Malkan \\& Stecker (1998, 2001); Primack et al. (1999); Kneiske, Mannheim \\& Hartmann (2002); Kneiske et al. (2004); Primack, Bullock and Somerville (2005); Stecker, Malkan \\& Scully (2006) calculated EBL models adopting either a phenomenological approach or Monte Carlo galaxy formation code. These models trace the general trend of the data, which may be fitted with a combination of two or more modified blackbody spectra for its two distinct peaks at the infrared and optical wavebands (Dermer 2007). Significant uncertainty in data and large dispersion among models led to an indirect method to constrain the EBL, namely by estimating change in spectral slope from distant TeV blazars due to $\\gamma\\gamma$ absorption (Stecker \\& de Jager 1993; Stanev \\& Franceschini 1998; Mazin \\& Raue 2007). However, such a method generally does not include possible absorption at the source (see, e.g., Reimer 2007) and presumes a source spectrum. In this paper, we build models of the EBL starlight component ($\\sim$0.1--10~eV) directly from the stellar thermal surface radiation. Emission from an individual star during its main-sequence lifetime is well approximated as a blackbody with a mass-dependent temperature. The post-main-sequence lifetime of a star is very short compared to its main-sequence lifetime and their contribution at the UV-optical wave bands is not significant. They can, however, contribute significantly to longer wavelengths due to their increased luminosity in the post-main-sequence phase (Finke et al., in preparation). Only a small fraction of the stars with mass $\\gtrsim 8M_\\odot$ produce supernovae and even a smaller fraction produce GRBs. Emission from these sources dominate the diffuse MeV background (Watanabe et al. 1999; Ruiz-Lapuente, Cass\\'e \\& Vangioni-Flam 2001; see, however, Strigari et al. 2005; Inoue, Totani \\& Ueda 2008). Emission from quasars and AGNs, on the other hand, dominate the diffuse X-ray background (Mushotzky et al. 2000). Contributions from all these sources add only a small fraction to the total cosmic electromagnetic energy density in the $\\sim$ 0.1--10~eV range, and we also ignore that. The estimated lifetimes of individual stars depend on their masses and the assumed cosmology, which is the standard $\\Lambda$CDM with ($h, ~\\Omega_m, ~\\Omega_\\Lambda$) = (0.7,~0.3,~0.7), and the Hubble constant $H_0 = 70 h_{0.7}$~km~s$^{-1}$~Mpc$^{-1}$. Summing over contributions from stars of all masses formed in the history of the universe then gives us the diffuse emission or EBL. A sum over contributions from individual stars radiating at a given redshift corresponds to the luminosity density or the stellar energy emissivity of the universe at that redshift. The initial mass function (IMF), which is the distribution of stars by mass, and the star formation rate (SFR), which is the mass that forms stars per unit comoving volume per unit time, are two uncertain but related parameters in our calculation. We form classes of models by choosing different combinations of these parameters and compare, in the UV-optical band, the luminosity density data of the local universe found from the surveys of nearby galaxies. The same models are then compared with EBL data. Note that there are no adjustable free parameters in our calculation once we choose a particular model. Finally we use one of our best-fit models to calculate the $e^\\pm$ pair production opacities in the $\\sim$10--300~GeV energy range at different redshifts. These results are applicable to high-energy emission from distant sources such as GRBs and blazars detected by the currently operating Fermi Gamma Ray Space Telescope and Air Cherenkov Telescopes such as HESS, MAGIC and VERITAS. In Sec. \\ref{sec:formalism} we outline the formalism of our method and introduce different models in Sec. \\ref{sec:SFR_models} upon which we base our EBL calculation. We report our results in Sec. \\ref{sec:results}, compare with results from previous authors as well as calculate EBL evolution with redshift. We discuss implications of our results for $\\gamma$-ray astronomy in Sec. \\ref{sec:implications}, calculating $\\gamma\\gamma \\to e^+e^-$ absorption opacity ($\\tau_{\\gamma\\gamma}$) for high energy $\\gamma$ rays from sources at different redshifts. Conclusions of the work for EBL and $\\gamma$ ray absorption are given in Sec. \\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} We have derived a class of well-defined models for the spectral energy density of the EBL. The models with modified Salpeter A initial mass function, a single power-law mass-luminosity relation and Cole et al. (2001) or Hopkins \\& Beacom (2006) star formation history reasonably fit the EBL UV-optical data and the luminosity density in our local universe. Our models are based on the underlying assumption that the bulk of the EBL radiation between $\\sim 0.1$ -- 10 eV is due to stellar radiations absorbed by dust, which can be determined from recent analyses of galaxies by Driver et al.\\ (2008). This approach differs from models by Stecker, Malkan \\& Scully (2006) based on luminosity evolution of galactic spectral energy distributions, which is limited to the accuracy of the available observational data used in the survey. Without need of a population synthesis code (e.g., models by Primack, Bullock \\& Somerville 2005) or fits to the results of such a code (e.g., models by Kneiske, Mannheim \\& Hartmann 2002), our model is based on well-studied results from stellar astronomy. The sources of uncertainties in our models are the (i) main-sequence age of the star and stellar luminosity, discussed earlier, (ii) star formation history and initial mass function, and (iii) dust absorption. We have already discussed point (ii) in some details using five SFR+IMF models. Note that in all those models, the IMF was assumed to be independent of redshift. In principle the normalization of the IMF or even its shape may depend on $z$. Nevertheless, a universal IMF fits SFR data reasonably well. The evolution of the dust absorption model with redshift, which we have not taken into account, is of potentially greater concern. At high redshift it is more reasonable to assume that the dust absorption (escape) fraction would be higher (lower), so that the stellar contribution to the EBL at high $z$ would be less than if using a constant absorption fraction. A more detailed examination of these issues are under further study (Finke et al., in perperation). We have provided an analytic fit to our best-fit EBL model and its evolution with redshift. This result can be used to calculate, as we have done in this work, opacity of the universe to $\\sim$10--300~GeV $\\gamma$-rays relevant for the high energy data from the {\\em Fermi Gamma Ray Space Telescope} and Air Cherenkov Telescopes, and estimating unknown quantities such as the spectrum and energy at production of the distant GRBs and blazars such as 3C 279." }, "0807/0807.3820_arXiv.txt": { "abstract": "A detailed analysis of new and existing photometric, spectroscopic and spatial distribution data of the eccentric binary V731~Cep was performed. Spectroscopic orbital elements of the system were obtained by means of cross-correlation technique. According to the solution of radial velocities with {\\em UBVR$_c$} and {\\em I$_c$} light curves, V731~Cep consists of two main-sequence stars with masses M$_{1}$=2.577 (0.098) M$_{\\odot}$, M$_{2}$=2.017 (0.084) M$_{\\odot}$, radii R$_{1}$=1.823 (0.030) R$_{\\odot}$, R$_{2}$=1.717 (0.025) R$_{\\odot}$, and temperatures T$_{eff1}$=10700 (200) K, T$_{eff2}$=9265 (220) K separated from each other by \\textit{a}=23.27 (0.29) R$_{\\odot}$ in an orbit with inclination of 88$^{\\circ}$.70 (0.03). Analysis of the O--C residuals yielded a rather long apsidal motion period of $U$$=$10000(2500) yr compared to the observational history of the system. The relativistic contribution to the observed rates of apsidal motion for V731~Cep is significant (76 per cent). The combination of the absolute dimensions and the apsidal motion properties of the system yielded consistent observed internal structure parameter (log$\\bar{k}_{2,obs}$ = $-$2.36) compared to the theory (log$\\bar{k}_{2,theo}$ = $-$2.32). Evolutionary investigation of the binary by two methods (Bayesian and evolutionary tracks) shows that the system is $t=$133(26) Myr old and has a metallicity of $[M/H]=-0.04(0.02)$ dex. The similarities in the spatial distribution and evolutionary properties of V731~Cep with the nearby ($\\rho\\sim$3$^{\\circ}$.9) open cluster NGC 7762 suggests that V731~Cep could have been evaporated from NGC~7762. ", "introduction": "Study of eclipsing binaries is still the most effective way of determining the absolute parameters of stars, especially from the spectroscopic and photometric analysis of detached double-lined eclipsing binaries, masses and radii of the components can be obtained with a precision of $\\sim$1 per cent (e.g. Southworth et al. 2005; Bak\\i\\c{s} et al. 2008), the limit precision for stellar evolutionary tests (Andersen 1991). Among the stars, those in the upper main sequence band are particularly useful in empirical tests of the convection formulae used in various evolution codes, while systems containing unevolved stars are useful in testing opacity and metallicity effects in near-ZAMS (Zero Age Main Sequence) models. V731~Cep, the binary discussed in the present paper, is of the latter type. The variability of V731~Cep (GSC~4288~0168; brightness at maximum V$\\sim$10.5 mag; orbital period $P$$\\sim$6.06 d) was discovered by Bak\\i\\c{s} et al. (2003) (hereafter B03). A brief history of V731~Cep was given by Bak\\i\\c{s} et al. (2007) (hereafter B07) who presented photometric light curves in {\\em BVR$_c$} and {\\em I$_c$} bands and limited spectroscopic observations. Within the scope of a project to study close eclipsing binaries of SB2 type, we were able to obtain new times of minima as well as new spectroscopic data for V731~Cep. The system has an eccentric orbit ($e=0.0165$), which makes it an important astrophysical tool for the investigation of internal structures of its components, if the rotation period of the apsides is precisely determined. The apsidal motion period together with system geometry allow the computation of the observed internal structure constant (ISC) to be compared with theoretical internal structure computations. The primary motivation for the present paper is to obtain the parameters of the close binary system V731~Cep, to discuss the evolutionary status of the system using the system parameters with the latest evolutionary models, and to investigate the apsidal motion of the orbit for estimation of the internal structures of the component stars. ", "conclusions": "\\subsection{Absolute Dimensions and Distance of the System} Combination of spectroscopic orbital elements (Table~6) with light curve elements (Table~7) yields the absolute dimensions of the system, which are presented in Table~8. The adopted temperature T$_{eff1}$$=$10700 K and mass $M_{1}$$=$2.577$M_{\\odot}$ of the primary component correspond to the spectral type of a normal B8.5-type main sequence star, while the adopted temperature T$_{eff2}$$=$9265 K and mass $M_{2}$$=$2.017$M_{\\odot}$ of the secondary component are in good agreement with the spectral type of a normal A1.5-type main sequence star (i.e. Strai\\v{z}ys \\& Kuriliene 1981). However, the radii of the components, $R_{1}$$=$1.823$R_{\\odot}$ and $R_{2}$$=$1.717$R_{\\odot}$, are in better agreement with the same spectral type stars but at closer locations to ZAMS, suggesting a young age of the system, as determined in \\S6.2. The synchronization time-scale for the component stars of the V731~Cep system is, following Zahn (1977), in the order of 15 Myr, which is smaller than the age of 120 Myr estimated from the isochrones (see \\S6.2). To compare the observed rotational velocities with the synchronization velocities listed in Table~8, we have modelled Si II doublets (6347.091 \\AA, 6371.359 \\AA) with model atmosphere grids using ATLAS9 and SYNTHE codes under Linux (Kurucz 1993). The modeling yielded equatorial rotational velocities of V$_{rot1}$$=$19(3) km s$^{-1}$ and V$_{rot2}$$=$18(3) km s$^{-1}$ for the primary and secondary components, respectively. Although errors in the observed rotational velocities are in the order of 15 per cent, the synchronization velocities seem to be slightly below the observational measurements. The asynchronization of the components with the orbit should be confirmed by analyzing new spectra with a higher S/N ratio and more absorption lines in a larger spectral range. \\begin{table*} \\small \\caption{Close binary stellar parameters of V731~Cep. Errors of parameters are given in parenthesis.} \\label{table8} \\begin{tabular}{lccc}\\hline\\hline Parameter & Symbol & Primary & Secondary \\\\ \\hline Mass (M$_\\odot$) & \\emph{M} & 2.577(0.098) & 2.017(0.084) \\\\ Radius (R$_\\odot$) & \\emph{R} & 1.823(0.030) & 1.717(0.025) \\\\ Separation (R$_\\odot$) & \\emph{a} & \\multicolumn{2}{c}{23.27(0.29)} \\\\ Surface gravity (cgs) & log $g$ &4.304(0.011)& 4.273(0.011) \\\\ Integrated visual magnitude (mag) & \\emph{V} &\\multicolumn{2}{c}{10.54(0.01)}\\\\ Integrated colour index (mag) & $B-V$ &\\multicolumn{2}{c}{0.09(0.01)}\\\\ Colour excess (mag) &$E(B-V)$&\\multicolumn{2}{c}{0.13(0.03)}\\\\ Visual absorption (mag) & $A_{V}$ &\\multicolumn{2}{c}{0.40}\\\\ Intrinsic colour index (mag) & $(B-V)_{0}$&\\multicolumn{2}{c}{-0.04(0.02)}\\\\ Component intrinsic colour index (mag) & $(B-V)$ & -0.073(0.020) & 0.016(0.020) \\\\ Temperature (K) & $T_{eff}$ & 10700(200) & 9265(220) \\\\ Spectral type & Sp & B8.5 V & A1.5 V \\\\ Luminosity (L$_\\odot$) & log \\emph{L}& 1.618(0.035) & 1.292(0.043)\\\\ Computed synchronization velocities (km s$^{-1}$)& V$_{synch}$ & 15.6(0.2) & 14.3(0.2) \\\\ Observed rotational velocities (km s$^{-1}$) & V$_{rot}$ & 19(3) & 18(3) \\\\ Bolometric magnitude (mag) &$M_{bol}$& 0.705(0.088) & 1.519(0.108) \\\\ Velocity amplitudes (km s$^{-1}$) &$K_{1,2}$& 85.18(1.72) & 108.84(1.73) \\\\ Absolute visual magnitude (mag) &$M_{v}$ & 1.104(0.054) & 1.631(0.070) \\\\ Bolometric correction (mag) &\\emph{BC}& -0.399 & -0.112 \\\\ Distance (pc) &\\emph{d} & \\multicolumn{2}{c}{809(30)} \\\\ Systemic velocity (km s$^{-1}$) &$V_{\\gamma}$ & \\multicolumn{2}{c}{0.62(0.94)} \\\\ Parallax (mas) &$\\pi$ & \\multicolumn{2}{c}{1.236(0.044)*} \\\\ Proper motion (mas yr$^{-1}$) &$\\mu_\\alpha cos\\delta$, $\\mu_\\delta$ & \\multicolumn{2}{c}{-1.5 (2.6), -3.1 (2.5)**} \\\\ Space velocities (km s$^{-1}$) & $U, V, W$ & \\multicolumn{2}{c}{7.59(9.03), 4.71(4.31), -9.75(9.62)}\\\\ \\hline {\\em * In this study.} \\\\ {\\em ** NOMAD Catalog (Zacharias 2005).} \\end{tabular} \\end{table*} Using the brightness of the system listed in Table~1 together with the light contributions of the components listed in Table~7, the intrinsic magnitudes of the components were calculated and are presented in Table~9. During the derivation of the intrinsic magnitudes, the interstellar extinction in {\\em B} and {\\em V} bands were adopted from the $Q$-method while the following relations of Fiorucci \\& Munari (2003) were used for the determination of extinction in $R_c$ and $I_c$. \\begin{eqnarray} \\nonumber (R_{c})_0 = R_c - 2.494 \\times E(B-V), \\\\ (I_{c})_0 = I_c - 1.753\\times E(B-V). \\end{eqnarray} \\noindent where ($R{_c})_0$ and ($I{_c})_0$ are the de-reddened magnitudes. \\begin{table} \\begin{center} \\caption{De-reddened magnitudes of stars in V731~Cep system.} \\label{table9} \\begin{tabular}{lccccc}\\hline\\hline & {\\em B} & {\\em V} & {\\em R$_c$} & {\\em I$_c$} & Err.\\\\ \\hline {\\em Primary} & 10.59 & 10.66 & 10.67 & 10.74 & 0.02 \\\\ {\\em Secondary} & 11.21 & 11.19 & 11.17 & 11.20 & 0.02 \\\\ \\hline \\end{tabular} \\end{center} \\end{table} The de-reddened visual apparent magnitude and optical absolute magnitude presented in Table~8 allowed us to derive a distance of 809(30) pc to the system. To compare the distance of V731~Cep system using a different method, a luminosity-colour relation (Bilir et al. 2008) which has been formed for detached binary systems with main-sequence components was used in this study. The near-infrared magnitudes of the system were taken from the {\\em 2MASS} Point Sources Catalogue of Cutri et al. (2003) and are shown in Table~1. For de-reddenig near-infrared magnitude and colours of the system, the following formulae (Bilir, G\\\"uver, Aslan 2006; Ak et al. 2007; Bilir et al. 2008) were used: \\begin{eqnarray} \\nonumber J_{o}=J-0.884\\times E(B-V), \\\\ (J-H)_{o}=(J-H)-0.322\\times E(B-V), \\\\ \\nonumber (H-K_{s})_{o}=(H-K_{s})-0.187\\times E(B-V). \\end{eqnarray} All the colours and magnitude with subscript ``0'' show the de-reddened ones. The colour excess $E(B-V)=0.13$ was estimated in a direction to V731~Cep by using {\\em Q}-method (see \\S5.1). The near-infrared absolute magnitude of V731~Cep system was estimated by the luminosity-colour relation, $M_{J}=5.228(J-H)_{o}+6.185(H-K_{s})_{0}+0.608$, of Bilir et al. (2008) and the distance of the system calculated as $733(50)$ pc by using the photometric parallaxes method. The photometric distance of $809(30)$ pc given in Table~8 is consistent with the $733(50)$ pc distance estimated by luminosity-colour relation obtained for detached binary systems. \\subsection{Internal Structure} Binary systems with apsidal motion allow us to determine the ISC, which is an important parameter of stellar evolution models. The observed apsidal motion period of $\\textit{U}$ = 10000(2500) yr, corresponding to a total rate of $\\dot{\\omega}$ = 0.00060(0.00015) $^\\circ$ cycle$^{-1}$ was obtained in \\S6. The relativistic contribution to the apsidal motion in case of V731~Cep is substantial $\\dot{\\omega}_{rel}$ = 0.00045 $^\\circ$ cycle$^{-1}$, or about 75 per cent of the total observed rate (Gim\\'enez 1985). After correcting for this effect, an average ISC was derived to be log $\\bar{k}_{2,obs}$ = $-$2.36 under the assumption that the component stars rotate pseudosynchronously. This value is in very good agreement with the theoretical prediction of log $\\bar{k}_{2,theo}$ = $-$2.34 according to new evolutionary models of Claret (2004) with the standard chemical composition of (X,Z) = (0.70, 0.02). It should, however, be noted that the present apsidal motion solution is still tentative due to relatively short observational history of V731 Cep compared to the apsidal motion period. Therefore, accurate eclipse timings are strongly needed in a decade or more in order to say more definite on the apsidal motion parameters and the related ISC. \\subsection{Evolutionary Stage and Age of the System} We investigated the evolutionary status of the system by means of the Bayesian method and constructing an H-R diagram for the component masses in log T$_{eff}$-log L plane. We used a slightly modified version of the Bayesian estimation method idealized by J{\\o}rgensen \\& Lindegren (2005), which is designed to avoid statistical biases and to take error estimates of all observed quantities into consideration. Estimation of age and metal abundance of the components was made by using the web interface\\footnote{http://stev.oapd.inaf.it/$\\sim$lgirardi/cgi-bin/param} based on the Bayesian method of da Silva et al. (2006). Including their errors, the effective temperatures, visual brightness, metal abundance of components and distance to the system were the parameters used in the web interface to obtain the best matching model parameters (i.e. surface gravity (log $g$), radii ($R_{1,2}$), masses ($M_{1,2}$) and age of components ($t_{1,2}$)) to Padova isochrones by the Bayesian method. Since the metal abundance of the components was initially not known, a range of metal abundance (i.e. $-$0.10 $<$ $[M/H]$ $<$ +0.10 dex) was selected and the values in this range were used with 0.02 dex steps. The output model parameters for each component were compared by means of $\\chi^2$ test with the absolute dimensions of the components listed in Table~8. Consequently, the minimum $\\chi^2$ yielded simultaneously the metal abundance and age of the components as $[M/H]_1=-0.06(0.02)$ dex and $t_{1}=116(15)$ Myr for the primary and $[M/H]_2=-0.02(0.02)$ dex and $t_{2}=150(15)$ Myr for the secondary component. From these values, we adopted the mean metal abundance and mean age of the system to be $[M/H]=-0.04(0.02)$ dex and $t=133(26)$ Myr, respectively. Interpretation of the evolutionary status of V731~Cep requires also the construction of H-R diagrams using the latest theoretical evolutionary models. In Fig.~8, the components of V731~Cep are shown in the log T$_{eff}$-log $L$ plane together with Yonsei-Yale (Y2) evolutionary tracks (Yi, Demarque, Kim et al. 2001; hereafter YDK) for different masses. Evolutionary tracks for the exact masses of the components were also computed using the code provided by YDK. Among the tracks computed for the exact masses with their errors, those with $[M/H]$$=$-0.024 dex ($Z$$=$0.0172) metallicity models match the locations of the components within the error limits in the log T$_{eff}$-log $L$ plane. We also computed a set of isochrones using a metallicity of $[M/H]$$=$ $-$0.024 dex in Y2 models. In Fig.~8, two isochrones ($t=$100 Myr and $t=$120 Myr) are plotted for comparison. It was found that 120 Myr age is the best fitting isochrone to the locations of both components. Although, the most reliable method of metal abundance determination is the atmosphere modeling of spectral lines, in the present study, it seems fair to conclude that the two methods (Bayesian methods and evolutionary tracks) used for estimation of the metallicity and age of the system are in excellent agreement within the error limits. The metallicity we found for the components should be confirmed with the atmosphere modeling of metallic absorption lines in spectra taken in wider wavelength range. \\begin{figure} \\centering \\resizebox{90mm}{!} {\\includegraphics[]{fig8.eps}} \\caption{Evolutionary tracks for individual component masses and isochrone curves best matching the location of the components in log T$_{eff}$ - log $L$ plane. The primary and secondary stars are shown with filled and empty circles, respectively.} \\label{fig8} \\end{figure} \\subsection{Possible Origin of V731~Cep} One of the formation regions of early-type stars is open clusters. To find a possible formation region for V731~Cep, nearby young open clusters were investigated. Among others, NGC~7762 was found to be the closest open cluster and the most similar in chemical composition to V731~Cep. Chincarini (1966) estimated the distance of NGC~7762 to be 750 pc and the age 266 Myr. In a more recent study relating on the cluster, Patat \\& Carraro (1995) proposed a similar distance of 800 pc to NGC~7762. Nevertheless, Patat \\& Carraro (1995) estimated an older age for the cluster at 1.8 Gyr and less metallicity compared to the Sun. The age (133 Myr) we adopted for V731~Cep system seems to agree more with the age (266 Myr) that Chincarini (1966) calculated for NGC~7762. In addition to this, the distance and metal abundance estimation of NGC~7762 by Patat \\& Carraro (1995) are in agreement with the distance and metal abundance of V731~Cep found in this study. The most reliable evidence that V731~Cep is evaporated from NGC~7762 can be obtained from the spatial distribution and the metallicity of V731~Cep. In calculation of the kinematical properties of V731~Cep, the systemic velocity, distance and proper motion components listed in Table~8 were used in the algorithm given by Johnson \\& Soderbloom (1987). The computed space velocity components with their errors are given in Table~8. The total space velocity of 14 km s$^{-1}$ for V731~Cep is in agreement with the space velocities of young stars. Although the space velocity of the cluster could not be computed due to its unavailable RV data, the distance of 55 pc between V731~Cep and NGC~7762, and similar age and metallicity distribution of V731~Cep with NGC~7762, suggest that V731~Cep could be evaporated from NGC~7762. However, it is necessary to carry out precise photometric and spectroscopic observations of the member stars of NGC~7762 in order to form more definite conclusions on the history of V731~Cep in relation with NGC~7762. \\\\ \\\\ \\textbf{Acknowledgements} \\\\ The authors would like to thank to Dr. Antonio Cabrera-Lavers for his help in extracting extinction data from RC stars and to the referee, Prof. E. F. Guinan, for his useful comments that improved the readability of this paper. This study is part of a project funded by \\c{C}OMU Scientific Research Foundation under project code BAP2008/37. The research of MW was supported by the Research Program MSM0021620860 of the Ministry of Education of Czech Republic. Participation of MZ in this study was endorsed by the grant GA CR 205/06/0217 of the Czech Science Foundation. \\\\" }, "0807/0807.4631_arXiv.txt": { "abstract": "We present the magnetic landscape of the polar region of the Sun that is unprecedented in terms of high spatial resolution, large field of view, and polarimetric precision. These observations were carried out with the Solar Optical Telescope aboard \\emph{Hinode}. Using a Milne-Eddington inversion, we found many vertically-oriented magnetic flux tubes with field strength as strong as 1 kG that are scattered in latitude between $70\\,^{\\circ}$ $\\sim$ $90\\,^{\\circ}$. They all have the same polarity, consistent with the global polarity of the polar region. The field vectors were observed to diverge from the center of the flux elements, consistent with a view of magnetic fields that expand and fan out with height. The polar region is also covered with ubiquitous horizontal fields. The polar regions are the source of the fast solar wind channelled along unipolar coronal magnetic fields whose photospheric source is evidently rooted in the strong field, vertical patches of flux. We conjecture that vertical flux tubes with large expansion around the photosphere-corona boundary serve as efficient chimneys for Alfv\\'en waves that accelerate the solar wind. ", "introduction": "The Sun's polar magnetic fields are thought to be the direct manifestation of the global poloidal fields in the interior, which serve as seed fields for the global dynamo that produces the toroidal fields responsible in active regions and sunspots. The polar regions are also the source of fast solar winds. Although the polar regions are of crucial importance to the dynamo process and acceleration of the fast solar winds, its magnetic properties are poorly known. Magnetic field measurements in the solar polar regions have long been a challenge: variable seeing combined with the strong intensity gradient and the foreshortening effect at the solar limb greatly increases the systematic noise in ground-based magnetographs. Nevertheless, pioneering observations have been carried out for the polar regions \\citep{tw90, lvz94, o04a, of04, hkm97, br07}. These observations typically have provided only the measurements of the line-of-sight magnetic component. Full Stokes polarimetry has also been carried, but as with most of the ground-based observations described above, the spatial resolution of those measurements was limited by seeing \\citep{bw96}. Another limitation of the past polar observations is that they have been restricted to individual polar faculae within a small field of view, and have not provided us with a global magnetic landscape of the polar region except for \\emph{GONG/SOLIS} \\citep{jh07}. We investigated the properties of photospheric magnetic field in polar regions using the Solar Optical Telescope, SOT \\citep{st07, ys07, ki07, ts07}. SOT is a diffraction-limited (0\".2\\,-\\,0\".3) Gregorian telescope with filter-graph and spectro-polarimeter aboard the satellite \\emph{Hinode} \\citep{tk07}. These observations are unprecedented in terms of their very high spatial resolution, wide field of view, and high polarimetric sensitivity and accuracy in measurements of vector magnetic fields. ", "conclusions": "We discovered that the poloidal field near the pole has a form of unipolar flux tubes scattered in the polar region rather than a weak extended field. If the polar field with the same total magnetic flux $\\Phi \\sim BfS$ is uniformly distributed ($S$ is the total magnetic area), the estimated effective field strength would be about 10G as described above. The total magnetic energy is then proportional to $B^{2}fS = B\\Phi$. Thus, the surface poloidal magnetic energy is approximately 90 times larger than the case for the uniform magnetic field, if we take $B \\sim 900$G, corresponding to the peak of the energy PDF in Figure 5 (b). The equi-partition field strength $B_{e}$ is the field strength where magnetic energy is equal to kinetic energy of surface granular motion: $B_{e}=\\sqrt{4 \\pi \\rho v^{2}}$. The typical equipartition field strength $B_{e}$ is about 400 Gauss for granules with a velocity of $v = 2 \\times10^{5}$ cm s$^{-1}$ and $\\rho$ given in section 4.2. The magnetic field strength for the majority of patches is larger than the equi-partition field strength. The observed unipolar strong flux tubes scattered in the polar region are considered to represent seed poloidal fields for toroidal fields \\citep{wns89a, wns89b}. Magnetic flux is transported to the polar regions with meridional flows and supergranular diffusion in the flux-transport dynamo model \\citep{dc99}. Since magnetic field has a form of such isolated flux tubes with super equi-partition strength instead of diffuse weak mean-field assumed in the flux transport dynamo \\citep{dc99}, flux transport on the sun would be done via an aerodynamic (drag) force against magnetic tension force, and may be more difficult than the case for the mean field case assumed in the models. If the flux tubes seen on the surface of the Sun are maintained inside the Sun, this would affect a known difficulty in $\\Omega$-mechanism \\citep{wme56} to generate intense toroidal field: smaller amplification factor is needed to generate the same toroidal field from the poloidal field with intrinsic field strength of 1kG than from the averaged 10G field, and thus may be achievable within a solar cycle. We, however, recognize that there would remain a serious energetic problem, if the toroidal field strength indeed reaches 100kG \\citep{sc96, mr06}. Total flux of vertical magnetic field at the polar region estimated here is at most $7.2 \\times 10^{21}$ Mx at the solar minimum, while various measurements on the total magnetic flux of single active region indicate $\\sim 10^{22}$Mx \\citep{lc07, jc07, mt08}. Thus, the measured total polar flux barely corresponds to that of single active region. The total toroidal flux would increase with time during the winding-up process by differential rotation, and the concept of the $\\Omega$-mechanism would be viable with the observations presented here. The transient horizontal magnetic field discovered in the polar region appears to have properties similar to those found in quiet Sun and in active regions \\citep{bwl07, rc07, os07, ri07, ri08}. In particular, PDFs of magnetic field strength for the polar region (Figure 5 (a)), quiet Sun, and active regions \\citep{ri08} are remarkably similar, suggesting a common local dynamo process \\citep{fc99} taking place all over the Sun. The X-ray telescope and EUV imaging spectrometer aboard \\emph{Hinode} observed remarkable activity in the polar regions in a form of micro flares and jets \\citep{sa07, ci07}. The lateral spreading of the vertical flux tubes to large area may be located well above the formation height of the two Fe lines, since there is no clear positional correlation between the horizontal fields and the vertical fields as seen in Figure 4. These X-ray jets could be due to magnetic reconnection at the lateral magnetic contacts with the horizontal fields and/or transient emergence of separate bipolar field lines \\citep{ssh98, sh92}. In conclusion, the magnetic landscape of the polar region is characterized by vertical kG patches with super equi-partition field strength, a coherency in polarity, lifetime with time scale of $5\\,-\\,15$ hours, and the ubiquitous weaker transient horizontal fields. The life time of the magnetic concentrations in the quiet Sun is estimated to be 2500 s for $2.5 \\times 10^{18}$Mx, and 40 ks for $10 \\times 10^{18}$Mx with \\emph{MDI} \\citep{hage99}. It is important to clarify similarities and differences between the polar region and the quiet Sun with \\emph{Hinode}. This will be discussed in our subsequent paper." }, "0807/0807.0224_arXiv.txt": { "abstract": "{}{Theory of random processes provides an attractive mathematical tool to describe the fluctuating signal from accreting sources, such as active galactic nuclei and Galactic black holes observed in X-rays. These objects exhibit featureless variability on different timescales, probably originating from an accretion disc.}{We study the basic features of the power spectra in terms of a general framework, which permits semi-analytical determination of the power spectral density (PSD) of the resulting light curve. We consider the expected signal generated by an ensemble of spots randomly created on the accretion disc surface. Spot generation is governed by Poisson or by Hawkes processes. The latter one represents an avalanche mechanism and seems to be suggested by the observed form of the power spectrum. We include general relativity effects shaping the signal on its propagation to a distant observer.}{We analyse the PSD of a spotted disc light curve and show the accuracy of our semi-analytical approach by comparing the obtained PSD with the results of Monte Carlo simulations. The asymptotic slopes of PSD are $0$ at low frequencies and they drop to $-2$ at high frequencies, usually with a single frequency break. More complex two-peak solutions also occur. The amplitude of the peaks and their frequency difference depend on the inherent timescales of the model, i.e., the intrinsic lifetime of the spots and the typical duration of avalanches.}{At intermediate frequencies, the intrinsic PSD is influenced by the individual light curve profile as well as by the type of the underlying process. However, even in cases when two Lorentzians seem to dominate the PSD, it does not necessarily imply that two single oscillation mechanisms operate simultaneously. Instead, it may well be the manifestation of the avalanche mechanism. The main advantage of our approach is an insight in the model functioning and the fast evaluation of the PSD.} ", "introduction": "It is widely accepted that massive black holes with accretion discs reside in cores of active galactic nuclei, where most activity originates and X-rays are produced (e.g., \\citeauthor{1990agn..book.....B} \\citeyear{1990agn..book.....B}). The observed light curves, $f\\,\\equiv\\,f(t)$, show irregular, featureless fluctuations with a very complex behaviour, practically at every studied frequency \\citep{2006ASPC..360.....G}. Variability has been traditionally analysed by the Fourier method \\citep{1992scma.book.....F}. Remarkably, a number of similarities appear between the properties of massive black holes in galactic nuclei and those in X-ray binaries, suggesting that some kind of universal rescaling operates according to central masses of these systems \\citep{1998Natur.392..673M}. This concerns also the X-ray power spectra (e.g., \\citeauthor{2003ApJ...593...96M} \\citeyear{2003ApJ...593...96M}; \\citeauthor{2006Natur.444..730M} \\citeyear{2006Natur.444..730M}). Light curves can be characterised by an appropriate estimator of the source variability which, in the mathematical sense, is a functional: $f\\rightarrow\\mathcal{S}\\left[f\\right]$. We accept the idea that the signal resulting from a spotted accretion disc is intrinsically stochastic, likely originating from turbulence. Hence, $\\mathcal{S}\\left[f\\right]$ is a random value. It can be a number (for example, the `rms' characteristic), function of a single variable (for example, the power-spectral density -- PSD) or of many variables (e.g., poly-spectra, rms--flux relation, etc). The appropriate choice depends on the type of information we seek and the quality of data available. The PSD is a traditional and widely utilised method to examine variable signals, and the AGN light curves are no exception. A typical signal can be represented by a broad band PSD with the tendency towards flattening at low frequency \\citep{1987Natur.325..694L,1987Natur.325..696M,1993ApJ...414L..85L,1993ARA&A..31..717M,2002MNRAS.332..231U}. There is an ongoing debate about the overall shape of the PSD and the occurrence of the break frequency or, possibly, two break frequencies at which the slope of the PSD can change \\citep{1999ApJ...510..874N,2003ApJ...593...96M}. In the case of a widely-studied Seyfert galaxy, MCG--6-30-15, \\cite{2005MNRAS.359.1469M} have closely examined the slope of PSD, namely its bending, with RXTE and XMM-Newton data. It is worth noticing that the accurate fits to the X-ray sources seem to exhibit a multi-Lorentzian structure rather than a simple power law. The same is true for the best studied example, the Seyfert~1 galaxy Akn~564 \\citep{2007MNRAS.382..985M}. It was proposed \\citep{1991A&A...245..454A,1991A&A...246...21Z,1992AIPC..254..251W} that hot spots contribute to the variability of the AGN variability, or that they could even be the dominant process shaping the variability pattern. These spots should occur on the disc surface following its intermittent irradiation by localised coronal flares \\citep{1979ApJ...229..318G,2001MNRAS.328..958M,2004A&A...420....1C}. Here, the ``spots'' represent a somewhat generic designation for non-axisymmetric features evolving on the disc surface in connection with flares. They share the bulk orbital motion with the underlying disc. The observed signal is thus modulated by relativistic effects as photons propagate towards a distant observer. Various schemes have been discussed in which the fluctuations of the disc emissivity at different points of space and time are mutually interconnected in some way. In particular, the avalanche model \\citep{1999MNRAS.306L..31P,2002MNRAS.333..800Z,2005MNRAS.359..308Z} seems to be physically substantiated within the framework of magnetically-triggered flares and spots. It is also a promising model capable to reproduce, for example, the broken power-law PSD profiles. Notice, however, that other promising ideas were proposed (e.g., \\citeauthor{1994PASJ...46...97M} \\citeyear{1994PASJ...46...97M}; \\citeauthor{1997MNRAS.292..679L} \\citeyear{1997MNRAS.292..679L}), provoking the question of whether a common mathematical basis could reflect the entire range of models and provide us with general constraints, independent of (largely unknown) model details. We add to this model by applying the method of random point processes \\citep{1965cox}. Interestingly enough, a rather formal approach can provide useful analytical formulae defining the basic form of the expected power spectrum. Apart from this practical aspect, we suggest that the concept discussed here offers much better insight into various influences that shape the expected form the power spectrum. These are very attractive features especially with respect to avalanche models, which may have different flavours, typically with a vast parameter space, thus proving very demanding to examine in a systematic manner. Even more important is that the adopted formalism provides a very general tool and allows for a broader perspective on different mechanisms of variability \\citep{2007arXiv0711.4772P}. We develop the idea in a systematic way and give the explicit correspondence between our approach and some of the above-mentioned and widely-known scenarios \\citep{1991A&A...245..454A,1999MNRAS.306L..31P}. This description provides semi-analytical solutions, convenient to search through a broad parameter space. Our results can help to identify how the intrinsic properties of individual flares and the relativistic effects influence the overall PSD. In particular, we can identify those situations in which a doubly-broken power law occurs. We consider stochastic processes (e.g., \\citeauthor{1971aitp.book.....F} \\citeyear{1971aitp.book.....F}; \\citeauthor{1994hsmp.book.....G} \\citeyear{1994hsmp.book.....G}) in the framework appropriate for modelling the accretion disc variability. In particular, in Sect.~\\ref{models} we consider a simple model of a spotted accretion disc constrained by the following three assumptions about the creation and evolution of spots: (i)~each spot is described by its time and place of birth ($t_j$, $r_j$ and $\\phi_j$) in the plain of the disc; (ii)~every new occurrence starts instantaneously; afterwards, the emissivity decays gradually to zero (the total radiated energy is of course finite); and (iii)~the intrinsic emissivity is fully determined by a finite set of parameters which form a vector, $\\mbox{\\boldmath $\\xi$}_j$, defining the light curve profile. Later on, we will consider the modulation of the intrinsic emission by the orbital motion and relativistic lensing. The disc itself has a passive role in our considerations; we will treat it as a geometrically thin, optically thick layer lying in the equatorial plane. Because of the apparently random nature of the variability, we adopt a stochastic model in which the creation of spots is governed by a random process. The assumption that spots are mutually statistically independent seems to be a reasonable (first) approximation, however, we find that we do need to introduce some kind of relationship between them. This connection is discussed in Sect.~\\ref{relationship}. The statistical dependence among spots can be introduced in several ways. In Sect.~\\ref{results}, we explore in detail the specific models of interrelated spots using the formalism of Hawkes-type processes. Conclusions are summarised in Sect.~\\ref{conclusions}. Finally, in the Appendix we provide some mathematical prerequisites, which the reader may find useful to understand the general background of the paper, and we also summarise the mathematical notation. ", "conclusions": "\\label{conclusions} We adopted the viewpoint that the variability pattern is determined by the interplay among the bulk orbital motion, relativistic effects, and the intrinsic changes of the inner accretion disc. We concentrated our attention solely on the PSD characteristics. The spots have a certain kind of memory in our model. We gave several examples in which the PSD changes the slope and certain break frequencies. The frequency of the break depends on the interplay of model properties, i.e., the intrinsic form of the spot light curves, which determine the individual contributions to the total signal together with the avalanche mechanism. {\\em The location of spots on the disc and the inclination of the source define the importance of relativistic effects.} In some cases, a double break occurs and the overall PSD profile is then approximated by a broken power law. This is a promising feature in view of applications to real sources with accreting black holes. The broken power-law profile either resembles a combination of the Lorentzians or, in some cases, an intermediate power law develops and connects the two peaks across the middle frequencies. The change of the PSD slope is clearly visible and well-defined in some cases, though under most circumstances it appears rather blurry. The low-frequency limit of the PSD slope is a constant; the-high frequency behaviour depends mainly on the shape of the spot emission profile, including the general relativity effects. In our calculations the emissivity was decaying exponentially and the slope of the PSD was equal to $-2$ at high frequencies. {\\em In between those two limits the intrinsic PSD is influenced by both the individual light curve profile and the underlying process.} It is interesting to notice that the doubly-broken power law occurs only for certain assumptions about the intrinsic light curves of the individual spots or avalanches -- their onset and the decay; in other cases the break frequencies are not well defined, or the broken power-law PSD is not preferred at all. {\\em We stress that if two Lorentzian seem to dominate the PSD (i.e., two peaks show up), it still does not necessarily mean that two single oscillation mechanisms operate simultaneously. Instead, it may well be the manifestation of the avalanche mechanism.} We employed a general statistical approach to the variability of a black hole accretion disc with orbiting spots that continuously arise and decay. The origin and evolution of spots were described by Poissonian and Hawkes' processes, the latter one representing a category of avalanche models. We derived analytical formulae for the PSD, Eqs.\\ (\\ref{PoissPSD}) and (\\ref{HawkPSD}), and we discussed their limitations and accuracy. {\\it The main advantage of the analytical form is the insight into the properties and the fast evaluation that captures the main trends of the PSD shape.} It is worth noting that the PSD does not maintain all information about the light curves that can be studied by Fourier methods \\citep{1992ApJ...391..518V}. Extensions have been discussed and compared with real data \\citep{1993ApJ...402..432K,1997MNRAS.288...12K,1999ApJ...510..874N,2008arXiv0802.0391V}, but this would go beyond the scope of our present work. Our approach allows us to investigate the influence of the assumed mechanism, which describes the creation of parent spots and of subsequent cascades of daughter spots. In particular, we can discuss the PSD slope at different frequency ranges and locate the break frequencies depending on the model parameters. The relationship between the mathematical nature of the process and the PSD of the resulting signal is an interesting consequence of this investigation, as it provides a way to grasp and rigorously constrain the physical models of the source. Therefore we believe that the method that we described is very helpful for identifying the underlying mechanisms that shape the PSD in black hole accreting sources." }, "0807/0807.2017_arXiv.txt": { "abstract": "We study second-order cosmological perturbations in scalar-tensor models of dark energy that satisfy local gravity constraints, including $f(R)$ gravity. We derive equations for matter fluctuations under a sub-horizon approximation and clarify conditions under which first-order perturbations in the scalar field can be neglected relative to second-order matter and velocity perturbations. We also compute the skewness of the matter density distribution and find that the difference from the $\\Lambda$CDM model is only less than a few percent even if the growth rate of first-order perturbations is significantly different from that in the $\\Lambda$CDM model. This shows that the skewness provides a model-independent test for the picture of gravitational instability from Gaussian initial perturbations including scalar-tensor modified gravity models. ", "introduction": "The constantly accumulating observational data \\cite{Perl} continue to confirm that the Universe has entered the phase of an accelerated expansion after the matter-dominated epoch. To reveal the origin of dark energy (DE) responsible for this late-time acceleration is one of the most serious stumbling block in modern cosmology \\cite{review,CST}. The first step toward understanding the nature of DE is to find a signature whether it originates from some modification of gravity or it comes from some exotic matter with negative pressure. If gravity is modified from Einstein's General Relativity, this leaves a number of interesting experimental and observational signatures that can be tested. Especially local gravity experiments generally place tight bounds for the parameter space of modified gravity models. So far many modified gravity DE models have been proposed-- ranging from $f(R)$ gravity \\cite{fR} ($R$ is a Ricci scalar), scalar-tensor theory \\cite{stensor,Boi} to braneworld scenarios \\cite{brane}. The $f(R)$ gravity is presumably the simplest generalization to the $\\Lambda$-Cold Dark Matter ($\\Lambda$CDM) model ($f(R)=R-\\Lambda$). Nevertheless it is generally not easy to construct viable $f(R)$ models that satisfy all stability, experimental and observational constraints while at the same time showing appreciable deviations from the $\\Lambda$CDM model. In order to avoid that a scalar degree of freedom (scalaron) as well as a graviton becomes ghosts or tachyons we require the conditions $f_{,RR}>0$ and $f_{,R}>0$ \\cite{Star07}. These conditions are also needed for the stability of density perturbations \\cite{staper}. For the existence of a matter-dominated epoch followed by a late-time acceleration, the models need to be close to the $\\Lambda$CDM model ($m \\equiv Rf_{,RR}/f_{,R} \\approx +0$) in the region $R \\gg R_0$ ($R_0$ is the present cosmological Ricci scalar) \\cite{AGPT}. Moreover the mass of the scalaron field in the region $R \\gg R_0$ is sufficiently heavy for the compatibility with local gravity experiments \\cite{Nava,CT,TUT,Brax}. Finally, for the presence of a stable de Sitter fixed point at $r \\equiv -Rf_{,R}/f=-2$, we require that $0 \\le m(r=-2) \\le 1$ \\cite{Faraoni,AGPT}. The models proposed by Hu and Sawicki \\cite{Hu07} and Starobinsky \\cite{Star07} satisfy all these requirements. They take the asymptotic form, $f(R) \\simeq R-\\mu R_c [1-(R/R_c)^{-2n}]$ ($\\mu>0, R_c>0, n>0$), in the region $R \\gg R_c$ ($R_c$ is roughly the same order as $R_0$). See Refs.~\\cite{Li,AT,Appleby,Tsuji08,NO07} for other viable $f(R)$ models. The main reason why viable $f(R)$ models are so restrictive is that the strength of a coupling $Q$ between dark energy and non-relativistic matter (such as dark matter) is large in the Einstein frame ($Q=-1/\\sqrt{6}$) \\cite{APT}. In the region of high-density where local gravity experiments are carried out, the scalaron field $\\phi$ needs to be almost frozen \\cite{Nava,CT} with a large mass through a chameleon mechanism \\cite{KW} to avoid that the field mediates a long ranged fifth force. Cosmologically this means that the field does not approach a kinematically driven $\\phi$ matter-dominated era ($``\\phi$MDE'' \\cite{coupled}) in which the evolution of scale factor is non-standard ($a \\propto t^{1/2}$ \\cite{APT}). The deviation from the $\\Lambda$CDM model becomes important as the mass of the scalaron gets smaller so that the field begins to evolve slowly along its potential. In other words the effect of modified gravity manifests itself from the late-time matter era to the accelerated epoch \\cite{Star07,Hu07}. This leaves a number of interesting observational signatures for the equation of state of DE \\cite{AT,Tsuji08}, matter power spectra \\cite{staper,Star07,Tsuji08} and convergence spectra in weak lensing \\cite{TT,Schmidt}. One can generalize the analysis in $f(R)$ gravity to the theories that have arbitrary constant couplings $Q$ \\cite{TUMTY}. In fact this is equivalent to Brans-Dicke theory \\cite{BD} with a scalar-field potential $V(\\phi)$. By designing the potential so that the field mass is sufficiently heavy in the region of high density, it is possible to satisfy both local gravity and cosmological constraints even when $|Q|$ is of the order of unity \\cite{TUMTY}. The representative potential of this type is given by $V(\\phi)=V_0[1-C(1-e^{-2Q\\phi})^p]$ $(V_0>0, C>0, 040000$ \\cite{Hoyle} from solar-system experiments. This shows that the evolution of the scale factor in the matter era is very close to the standard one: $a(t) \\propto t^{2/3}$. We note that an effective gravitational ``constant'' that appears as a coefficient of matter density perturbations is also subject to change in Brans-Dicke theory. However it was found that the skewness in such a case is given by $S_3=(34\\omega_{\\rm BD}+56)/ (7\\omega_{\\rm BD}+12)$ \\cite{BD2} during the matter era, which is very close to the standard one ($S_3=34/7$) under the condition $\\omega_{\\rm BD}>40000$. The $f(R)$ gravity corresponds to theory with the Brans-Dicke parameter $\\omega_{\\rm BD}=0$ \\cite{Chiba}. Even in this situation, if the scalaron field has a potential whose mass is sufficiently large in the region of high density, the $f(R)$ models can pass local gravity constraints as in the models proposed in Refs.~\\cite{Hu07,Star07}. In such cases, compared to Brans-Dicke theory with a massless field, it is expected that the skewness may show significant deviations from that in General Relativity. Since the evolution of scale factor and matter perturbations is different from that in the massless case, we can not employ the result of the skewness presented above. In this paper we study second-order perturbations and the skewness for Brans-Dicke theory in the presence of a potential $V(\\phi)$. This is equivalent to the scalar-field action given in Eq.~(\\ref{action}) by identifying the coupling $Q$ with the Brans-Dicke parameter $\\omega_{\\rm BD}$ via the relation $1/(2Q^2)=3+2\\omega_{\\rm BD}$. In the massless case the solar-system constraint, $\\omega_{\\rm BD}>40000$, gives the bound $|Q| \\lesssim 10^{-3}$, but it is difficult to find some deviations from General Relativity in such a situation. Our interest is the case in which the coupling $Q$ is of the order of $0.1 \\lesssim |Q| \\lesssim 1$ with the field potential that has a sufficiently large mass in the high-density region. This analysis includes viable $f(R)$ models \\cite{Hu07,Star07} recently proposed in the literature. We would like to investigate how much extent the skewness differs from that in the $\\Lambda$CDM model. We also derive conditions under which the contribution coming from first-order field perturbations can be neglected relative to second-order matter and velocity perturbations by starting from fully relativistic second-order perturbation equations. ", "conclusions": "\\label{conclude} In this paper we have studied the evolution of second-order matter density perturbations in a class of modified gravity models that satisfy local gravity constraints. We have considered the scalar-tensor action (\\ref{action}), which is equivalent to Brans-Dicke action (\\ref{action0}) with the correspondence $1/(2Q^2)=3+2\\omega_{\\rm BD}$. In the presence of a field potential it is possible to satisfy local gravity constraints (LGC) even when $|Q|$ is of the order of unity. In fact the potential (\\ref{potential}) is designed to have a large mass in the region of high density for the consistency with LGC. This covers the models proposed by Hu and Sawicki \\cite{Hu07} and Starobinsky \\cite{Star07} in the context of $f(R)$ gravity ($Q=-1/\\sqrt{6}$). Starting from second-order relativistic equations of cosmological perturbations, we have derived the equation (\\ref{basic3}) of matter density fluctuations approximately. In so doing we employed the approximation that first-order perturbations in the scalar field $\\phi$ is neglected relative to second-order matter and velocity perturbations. This is valid under the conditions (\\ref{QH}) and (\\ref{delcon}), both of which can be naturally satisfied for the values of $Q$ we are interested in ($0.1 \\lesssim |Q| \\lesssim 1$). Compared to the $\\Lambda$CDM model, the effective gravitational constant $G_{\\rm eff}$ is subject to change at the late epoch of the matter era. This leads to the larger growth rate of first-order matter perturbations ($\\delta_k^{(1)} \\propto t^{(\\sqrt{25+48Q^2}-1)/6}$) compared to the standard case ($\\delta_k^{(1)} \\propto t^{2/3}$). The skewness of matter distributions is determined by the second-order growth factor $E_a$ relative to the squared of the first-order growth factor $D$. In the ``scalar-tensor regime'' where the effective gravitational constant is given by $G_{\\rm eff} \\simeq (1+2Q^2)/8\\pi F$, we have derived the analytic expression (\\ref{anaes}) of the skewness in the matter-dominated epoch. In the ``General Relativistic regime'' where $G_{\\rm eff} \\simeq 1/8\\pi F \\simeq G$, we have reproduced the standard value $S_3=34/7$ in the Einstein-de Sitter Universe. In modified gravity models with $|Q| \\lesssim 1$, the analytic value (\\ref{anaes}) of the skewness in the asymptotic regime of the matter era is different from the value $34/7$ only less than a few percent. Even if we take into account the evolution of perturbations during the accelerated phase, the difference of the skewness relative to the $\\Lambda$CDM model remains to be small. The above result comes from the fact that the ratio of the second-order growth rate relative to the first-order one has a weak dependence on the coupling $Q$. When $|Q|={\\cal O} (1)$ the growth rate of first-order matter perturbations is significantly different from that in the $\\Lambda$CDM model. This gives rise to large modifications to the matter power spectrum as well as to the convergence spectrum in weak lensing, while the skewness is hardly distinguishable from that in $\\Lambda$CDM model. This fact can be useful to discriminate large coupling scalar-tensor models among many other dark energy models from future high-precision observations." }, "0807/0807.0012_arXiv.txt": { "abstract": "We investigate the GeV emission from gamma-ray bursts (GRBs), using the results from the Energetic Gamma Ray Experimental Telescope (EGRET), and in view of the {\\it Gamma-ray Large Area Space Telescope (GLAST)}. Assuming that the conventional prompt and afterglow photons originate from synchrotron radiation, we compare an accompanying inverse-Compton component with EGRET measurements and upper limits on GeV fluence, taking Klein-Nishina feedback into account. We find that EGRET constraints are consistent with the theoretical framework of the synchrotron self-Compton model for both prompt and afterglow phases, and discuss constraints on microphysical parameters in both phases. Based on the inverse-Compton model and using EGRET results, we predict that {\\it GLAST} would detect GRBs with GeV photons at a rate $\\gtrsim 20$ yr$^{-1}$ from each of the prompt and afterglow phases. This rate applies to the high-energy tail of the prompt synchrotron emission and to the inverse-Compton component of the afterglow. Theory predicts that in a large fraction of the cases where synchrotron GeV prompt emission would be detected by {\\it GLAST}, inverse-Compton photons should be detected as well at high energies ($\\gtrsim 10$ GeV). Therefore {\\it GLAST} will enable a more precise test of the high-energy emission mechanism. Finally, we show that the contribution of GRBs to the flux of the extragalactic gamma-ray background measured with EGRET is at least 0.01\\% and likely around 0.1\\%. ", "introduction": "\\label{sec:Introduction} Cosmological gamma-ray bursts (GRBs) have released a tremendous amount of energy in the past and present Universe. Their emission covers very wide range of frequencies: a highly variable prompt phase radiates $\\sim$100 keV gamma rays, while a subsequent afterglow radiates radio to X-ray photons. It is likely that the bulk of these photons are emitted by gyration of relativistic electrons in magnetic fields---e.g., synchrotron radiation. The relativistic electrons are accelerated in either internal dissipation (for prompt emission) or external shocks (for afterglows). For reviews, see, \\citet{Piran2005,Meszaros2006,Nakar2007}. GeV photons were detected as well from several GRBs by the Energetic Gamma Ray Experimental Telescope (EGRET) on board the {\\it Compton Gamma Ray Observatory (CGRO)} \\citep{Schneid1992,Sommer1994,Hurley1994,Schneid1995,Gonzalez2003}. The data are still not sufficient for us to firmly infer emission mechanisms of these GeV gamma rays, but the most promising mechanism is synchrotron self-Compton (SSC) scattering \\citep*[e.g.,][]{Meszaros1994,Waxman1997,Wei1998,Chiang1999, Panaitescu2000,Zhang2001,Sari2001,Guetta2003}. This is because the relevant emission parameters such as the energy fraction of the GRB jets going to electrons ($\\epsilon_e$) and magnetic fields ($\\epsilon_B$) are relatively well measured from the afterglow spectra as well as light curves; the typical values are $\\epsilon_e = 0.1$ and $\\epsilon_B = 0.01$ \\citep[e.g.,][]{Panaitescu2001,Yost2003}. In the prompt emission, $\\epsilon_e$ is similar or even higher, as evident from the high efficiency of this phase, while $\\epsilon_B$ is not well constrained. Thus, there should be a significant inverse-Compton (IC) component accompanying the synchrotron radiation in both the afterglow and prompt emission. The luminosities of the synchrotron and IC are expected to be comparable as IC-to-synchrotron luminosity ratio is roughly given by $(\\epsilon_e / \\epsilon_B)^{1/2}$, according to theory \\citep[e.g.,][]{Sari2001}. In this paper, we explore the GeV gamma-ray emission of GRBs in the context of SSC mechanism.\\footnote{Our analysis and conclusions are applicable also if the MeV and/or radio-X-ray afterglow emission mechanism is not synchrotron but another type of emission from relativistic electrons that gyrate in a magnetic field, such as jitter radiation \\citep{Medvedev00}.} Besides the several GRBs detected by EGRET, there are many others for which upper bounds on the fluence were obtained \\citep{Gonzalez2005}. These $\\sim$100 GRBs should also be compared with the predictions of SSC model, because the fluence upper limits in the EGRET energy band are comparable to the fluence of prompt emission collected by Burst And Transient Source Experiment (BATSE) instrument onboard {\\it CGRO}. As the experimental bound is already strong, while theoretical models of SSC process predict a large fluence for the EGRET energy range, we derive meaningful constraints from EGRET data analysis on the physics of the high-energy emission mechanisms of GRBs. This approach is different from (and therefore complementary with) that in previous studies \\citep*[e.g.,][and references therein]{Dermer00, Asano2007, Ioka2007, Gou2007, Fan2008, Murase2008, Panaitescu2008}, where the prediction of gamma-ray flux relies only on theoretical models and sub-GeV observations. We instead use EGRET data in order to infer the GeV emission and constrain the theoretical models. We use our results to predict the expected number of GRBs that would be detected by the {\\it Gamma-ray Large Area Space Telescope (GLAST)}. The {\\it GLAST} satellite is equipped with the Large Area Telescope (LAT), which is an upgraded version of EGRET. Since revealing the high-energy emission mechanisms of GRBs are one of the important objectives of {\\it GLAST}, our prediction should give a useful guideline. Finally, we apply our results to estimate the contribution of GRBs to the diffuse extragalactic gamma-ray background (EGB), which was also measured by EGRET \\citep[][see, however, \\citealt*{Keshet2004b} for a subtle issue of Galactic foreground subtraction]{Sreekumar1998,Strong2004}. This paper is organized as follows. In \\S~\\ref{sec:IC}, we summarize the predictions of SSC model for the prompt (\\S~\\ref{sub:prompt}) and afterglow (\\S~\\ref{sub:afterglow}) phases. Section~\\ref{sec:Constraint on high-energy emission with EGRET} is devoted for analysis of the GRB fluence data by EGRET, from which distributions of fluence in the GeV band are derived. We then use these distributions to argue prospects for GRB detection with {\\it GLAST} in \\S~\\ref{sec:GLAST}, and implications for EGB from GRB emissions in \\S~\\ref{sec:EGB}. In \\S~\\ref{sec:conclusions}, we give a summary of the present paper. ", "conclusions": "\\label{sec:conclusions} The {\\it GLAST} satellite would enable us to test high-energy emission mechanisms of GRBs. If this emission will be found to be consistent with SSC then its observations would constrain physical parameters such as $\\epsilon_{e}/\\epsilon_B$ ratio and the bulk Lorentz factor of the jet, $\\Gamma_{b}$. The EGRET instrument on board {\\it CGRO}, while less sensitive than the {\\it GLAST}-LAT detector, identified several BATSE GRBs with GeV photons. In addition, stringent upper limits for $\\sim$100 GRBs were put on fluences in the GeV band by analyzing the EGRET data \\citep{Gonzalez2005}. In this paper, we further extended this EGRET result, comparing with the SSC emission model. Following theoretical models of SSC, we assumed that there is a linear correlation between fluences in BATSE and EGRET energy bands, and that the proportionality coefficient $\\eta$ follows a log-normal distribution. We found that the predictions from the SSC model using canonical parameter values is fully consistent with EGRET fluence measurements and upper limits for both the prompt and afterglow phases. During the course of showing this result, we properly took the Klein-Nishina feedback effect into account in the theoretical calculation. The best-fit value of the coefficient was $\\log \\eta \\simeq -1.5$ for both the prompt and afterglow emissions, and it is already stringent enough to test the SSC scenario. The limits for the prompt emission phase are for the synchrotron radiation, and thus if we consider the IC component as well, the value of $\\eta$ could be larger by up to one order of magnitude. The obtained $\\eta$ distribution, together with the BATSE fluence distribution, gives the expected fluence distribution in the GeV band, which is shown in Figure~\\ref{fig:dndf_egret}. As the {\\it GLAST}-LAT detector covers EGRET energy band, we can predict the detectable number of GRBs with {\\it GLAST} from the distribution of $F_{\\rm EGRET}$, given the {\\it GLAST}-LAT sensitivity. Our conservative estimate using the five-photon criterion is that about $\\sim$20 GRBs among those detected with GBM would be detected with {\\it GLAST}-LAT each year. This number could be even larger if we use fewer-photon criteria. The fluence distribution can also be used to estimate the GRB contribution to the EGB intensity. We found that the contribution would be at least $\\sim$0.01\\% but is likely to be as large as $\\sim$0.1\\%." }, "0807/0807.4340_arXiv.txt": { "abstract": "We estimate the temporal change of magnetic flux perpendicular to the solar surface in a decaying active region by using a time series of the spatial distribution of vector magnetic fields in the photosphere. The vector magnetic fields are derived from full spectropolarimetric measurements with the Solar Optical Telescope aboard \\textit{Hinode}. We compare a magnetic flux loss rate to a flux transport rate in a decaying sunspot and its surrounding moat region. The amount of magnetic flux that decreases in the sunspot and moat region is very similar to magnetic flux transported to the outer boundary of the moat region. The flux loss rates [$(dF/dt)_{loss}$] of magnetic elements with positive and negative polarities are balanced each other around the outer boundary of the moat region. These results suggest that most of the magnetic flux in the sunspot is transported to the outer boundary of the moat region as moving magnetic features, and then removed from the photosphere by flux cancellation around the outer boundary of the moat region. ", "introduction": "How and where is magnetic flux in sunspots removed from the photosphere? Sunspot umbrae are sometimes split by formation of a light bridge, which is a bright lane crossing the umbra \\citep{Bray1964}. Small magnetic features with a typical size less than 2$\\arcsec$ called moving magnetic features \\citep[MMFs;][]{Harvey1973} are generally observed in the moat region that surrounds a decaying, mature sunspot. It has been reported that MMFs appear not only in the decaying phase of sunspots but also in the growing phase \\citep[e.g.][]{Wang1991}. The MMFs mostly appear around the outer boundary of the sunspot, moving almost radially outward during their lifetime ranging from a few minutes to 10 hr \\citep{Harvey1973, Zhang2003, Hagenarr2005}. The formation of the light bridges and MMFs is closely related to the fragmentation and disintegration of the sunspot magnetic flux. Indeed, it has been observed that the net flux carried away from the sunspot by MMFs is larger than the flux decrease in the sunspot \\citep{Martinez2002, Kubo2007a}. This indicates that MMFs can be responsible for the flux loss of the sunspot. The mutual apparent loss of magnetic flux is often observed in the line-of-sight magnetograms when one magnetic polarity element collides with another polarity magnetic element in the photosphere. This apparent flux loss is called ``magnetic flux cancellation'' as a descriptive term. It is observed that moving magnetic features often collide with apparently static opposite polarity magnetic features around the outer boundary of the moat region \\citep{Martin1985, Yurchyshyn2001, Chae2004, Bellot2005} and widely believed that understanding the flux cancellation process around the moat boundary is the key to understand the dissipation of the sunspot flux from the photosphere. Three models have been proposed by \\citet{Zwaan1987} to describe the flux cancellation: (1) retraction of magnetic fields that connect an emerged bipole, (2) submergence of $\\Omega$-loop formed by magnetic reconnection between the canceling two bipoles above the photosphere, and (3) emergence of U-loop due to reconnection below the photosphere. As expected in these models, horizontal magnetic fields are formed between canceling magnetic elements \\citep{Wang1993, Chae2004, Kubo2007b}. However, whether upward or downward motions are observed in the cancellation sites depends on the events and the positions in the cancellation sites \\citep{Harvey1999, Kubo2007b}. Therefore, the nature of the physical process driving magnetic flux cancellation is still an open issue. This study attempts to address a basic question how much magnetic flux is carried away from the sunspot to the outer boundary of the moat region and is subsequently removed from the photosphere. Because it has been difficult to measure the magnetic field vector under stable seeing conditions for the period longer than a typical lifetime of MMFs, the flux loss rate of the sunspot, flux transport rate due to MMFs, and flux cancellation rate have been independently estimated by using different data sets. A time series of spectropolarimetric measurements with the Solar Optical Telescope \\citep[SOT;][]{Tsuneta2008} aboard the \\textit{Hinode} satellite \\citep{Kosugi2007} allows us, for the first time, to estimate an accurate flux change without any effects of atmospheric seeing. Moreover, the high spatial resolution observations with the SOT decrease the likelihood of spurious magnetic cancellation events, i.e., those for which magnetic elements with opposite polarities are located entirely within a resolution element and will dramatically increase the reliability of the results presented. ", "conclusions": "The averaged flux change and averaged flux transport in Figure~\\ref{flux_summary} are described in units of Mx s$^{-1}$, so that we can compare these values directly. The observed flux change [$(dF/dt)_{Obs}$] in each region and the observed flux transport [$(F_v)_{Obs}$] at its boundaries would have a relationship: \\begin{equation} (\\frac{dF}{dt})_{Obs} = (\\frac{dF}{dt})_{Emerge} - (\\frac{dF}{dt})_{Loss} \\pm (F_v)_{Obs}.\\label{eq_flux_relation} \\end{equation} The increase of magnetic flux due to flux emergence [$(dF/dt)_{Emerge}$] can be neglected in this case, because we select the period without any significant flux emergence. Thus, we can estimate an actual flux loss rate [$(dF/dt)_{Loss}$] in each region from the observations. The total of flux decrease rates in the sunspot and unipolar regions ($dF/dt$ = -3.2 - 4.8 = -8.0$\\times 10^{15}$ Mx s$^{-1}$) is almost equal to the flux transport rate at the outer boundary of the unipolar region for the positive polarity ($F_v$ = 7.4$\\times 10^{15}$ Mx s$^{-1}$). This means that most of magnetic flux that disappeared in the sunspot and unipolar regions is carried away to the mixed polarity region. Note that we do not trace each magnetic element, and all of the MMFs that separated from the sunspot may not reach at the outer boundary of the unipolar region. However, we make Figure~\\ref{flux_summary} from the observations for about 12 hr, which is twice as long as the period that magnetic elements with the average horizontal speed of 0.5 km s$^{-1}$ need to move through the unipolar region of a width of 14$\\arcsec$. The increase of the positive flux in the mixed polarity region supports the migration of positive flux into the mixed polarity region. Both the increase of the positive flux in the mixed polarity region and the flux transport for the positive polarity elements at the outer boundary of the mixed polarity region are smaller than the positive flux transported from the unipolar region. Therefore, the magnetic flux that is carried away from the sunspot (and moat region) mostly disappears in the mixed polarity region, especially near the outer boundary of the moat region. One issue is that the positive flux carried away from the sunspot region ($F_v$ = 7.8$\\times 10^{15}$ Mx s$^{-1}$) is bigger than decrease of the positive flux in the sunspot region ($dF/dt$ = -3.2$\\times 10^{15}$ Mx s$^{-1}$). This tendency was also reported in the previous work with a lower (about $1\\arcsec$) spatial resolution \\citep{Kubo2007a}. As a result of no flux emergence in the sunspot region, the flux transport rate should be less than the flux decrease rate in the sunspot region. That is to say that the flux transport rate is overestimated at the outer boundary of the sunspot region. In the calculation of horizontal velocities, we use the apodization window with 1$\\arcsec$, which is lower than the spatial resolution of the magnetic field maps. Such a lower spatial resolution of the horizontal velocity maps probably causes the overestimation of the flux transport rate at the outer boundary of the sunspot region. Fuzzy, small magnetic elements with a short lifetime have been observed around the outer boundary of decaying sunspots \\citep{Zhang2007, Kubo2008}. These fuzzy magnetic elements have higher outward motion and smaller magnetic flux than those of usual MMFs. In the estimation of flux transport at the outer boundary of the sunspot, magnetic flux is mostly represented by MMFs with small horizontal velocity and large magnetic flux, but its horizontal velocity is represented by the fuzzy magnetic elements. We believe that these fuzzy magnetic elements correspond to a fluctuation of field strength or a fluctuation of inclination of penumbral magnetic fields, and thus do not contribute to the flux loss of the sunspot. Further investigation using spectropolarimetric measurements with a higher cadence will be necessary to know the nature of such fuzzy magnetic elements and their impact on the presented calculations. The magnetic flux of negative polarity decreases in the mixed polarity region, although the negative flux converges from the inner and outer boundaries of the mixed polarity region. Considering that the negative flux moves into the mixed polarity region with the average rate of 1.6$\\times 10^{15}$ Mx s$^{-1}$, the actual flux loss rate [$(dF/dt)_{Loss}$] in the mixed polarity region may be as large as 3.9$\\times 10^{15}$ Mx s$^{-1}$ from Equation~(\\ref{eq_flux_relation}). This flux loss rate of negative polarity is balanced by the actual flux loss rate of the positive polarity (3.9$\\times 10^{15}$ Mx s$^{-1}$), which is a difference between the flux transport rate ($F_v$ = 7.4 - 0.8 = 6.6$\\times 10^{15}$ Mx s$^{-1}$) into the mixed polarity region and the flux increase rate ($dF/dt$ = 2.7$\\times 10^{15}$ Mx s$^{-1}$) there. The flux loss rates with both polarities in the mixed polarity region are consistent with the cancellation rates in active regions \\citep{Chae2000,Chae2004,Kubo2007b}. Furthermore, most of the magnetic elements with negative polarity are located in contact with the positive elements. These results suggest that magnetic flux cancellation at the outer boundary of the moat region is essential for the removal of the sunspot magnetic flux from the photosphere." }, "0807/0807.1565_arXiv.txt": { "abstract": "Using three newly identified galaxy clusters at $z\\sim1$ (photometric redshift) we measure the evolution of the galaxies within clusters from high redshift to the present day by studying the growth of the red cluster sequence. The clusters are located in the Spitzer Infrared Array Camera (IRAC) Dark Field, an extremely deep mid-infrared survey near the north ecliptic pole with photometry in 18 total bands from X-ray through far-IR. Two of the candidate clusters are additionally detected as extended emission in matching Chandra data in the survey area allowing us to measure their masses to be $M_{500}= 6.2 \\pm 1.0 \\times 10^{13}$ and $3.6 \\pm 1.1 \\times 10^{13}$ \\msun. For all three clusters we create a composite color magnitude diagram in rest-frame B-K using our deep HST and Spitzer imaging. By comparing the fraction of low luminosity member galaxies on the composite red sequence with the corresponding population in local clusters at $z=0.1$ taken from the COSMOS survey, we examine the effect of a galaxy's mass on its evolution. We find a deficit of faint galaxies on the red sequence in our $z\\sim1$ clusters which implies that more massive galaxies have evolved in clusters faster than less massive galaxies, and that the less massive galaxies are still forming stars in clusters such that they have not yet settled onto the red sequence. ", "introduction": "The redshift range from $z=1$ to the present day is a particularly dynamic epoch in the history of groups and clusters as evidenced by the evolution of the morphology-density relation and increasing fraction of blue galaxies with increasing redshift \\citep{capak2007, butcher1984}. Interestingly, cluster ellipticals at $z\\sim1$ already have a narrow distribution of red colors \\citep[the red cluster sequence (RCS);][]{blakeslee2003,vandokkum2001}. There is some debate about the mechanism by which these cluster galaxies arrive onto the red sequence. It is difficult to distinguish whether these red ellipticals all formed their stars and did their merging at $z>3$, then stopped forming stars when they entered the cluster environment \\citep{ford2004}; or if they are the product of the merging of gas-poor systems which do not produce star formation \\citep{vandokkum2005}. We investigate whether the red population is still in the process of forming at $z=1$ or if indeed assembly has already finished at higher redshift by studying the presence of the faint end of the RCS at $z=1$ and comparing it to the present epoch. We measure the ratio of faint to bright RCS galaxies in a sample of three $z\\sim1$ clusters from the IRAC Dark Field (described below). These clusters have the benefit of extremely deep $3.6\\,\\micron$ data which allows us to study the faint end of the luminosity function at rest frame near-IR, which traces the peak of the spectral energy distribution in galaxies. A confirmed deficit of faint galaxies on the RCS would imply that more massive galaxies have evolved in clusters faster than less massive galaxies. A constant fraction of faint red galaxies between $z=1$ and the present would require a formation mechanism where galaxies of all masses have already joined the red sequence at redshifts higher than one. There is evidence that the faint end of the RCS is not completely in place by z=0.8 \\citep[][and references therein]{delucia2007,koyama2007}, although at least some clusters at these redshifts appear to have complete RCSs to M*+3.5 \\citep{andreon2006}. These questions are ideally addressed with deep infrared surveys of clusters at high redshift. In the last four years deep and wide area surveys in the mid-infrared using the {\\it Spitzer Space Telescope} have substantially opened a new window on galaxy and star formation at $0$ 0.02) SNe Ia, $M_B = 0.98 \\times s_{(B)}^{-1}- 2.28 \\times (B - V)_{max} - 19.95$ with an r.m.s. of 0.27 mag. The r.m.s. becomes 0.12 mag when we select only SNe Ia hosted by E or S0 galaxies. In \\S7, we statistically investigate the properties of dust in host galaxes, and find that the ratio of total to selective extinction $R$ can be consistent with that of dust in the Milky Way if we assume that SNe Ia have some diversity in their intrinsic colours. Based on these results, in \\S8, we discuss how to select subsets of SNe Ia which may be good distance indicators for cosmology. We find two possibilities: one is to use only ``BV bluest'' SNe Ia with broad light curves, and the other is to use only SNe Ia inside E or S0 galaxies. To use these samples for precision cosmology, we need to study the properties of SNe Ia from the low-z to the high-z universe." }, "0807/0807.2566_arXiv.txt": { "abstract": "{ Very long baseline interferometric observations of the radio galaxy 3C~120 show a systematic presence of gradients in Faraday rotation and degree of polarization across and along the jet. These are revealed by the passage of multiple superluminal components throughout the jet as they move out from the core in a sequence of 12 monthly polarimetric observations taken with the VLBA at 15, 22, and 43 GHz. The degree of polarization has an asymmetric profile in which the northern side of the jet is more highly polarized. The Faraday rotation measure is also stratified across the jet width, with larger values for the southern side. Superposed on this structure we find a localized region of high Faraday rotation measure ($\\sim$ 6000 rad m$^{-2}$) between approximately 3 and 4 mas from the core. This region of enhanced Faraday rotation may result from the interaction of the jet with the ambient medium, which may also explain the stratification in degree of polarization. The data are also consistent with a helical magnetic field in a two-fluid jet model, consisting of an inner emitting jet and a sheath of nonrelativistic electrons. ", "introduction": "Helical magnetic fields may play an important role in the dynamics and emission of relativistic jets in active galactic nuclei (AGN), specially in the formation and collimation processes \\citep{2002Sci...295.1688K,2008Natur.452..966M}. If jets are surrounded by a sheath of nonrelativistic electrons, it is possible to search for these helical magnetic fields by looking for Faraday rotation measure (RM) gradients across the jet \\citep*[e.g.,][]{Blandford:1993fk}. Such gradients have been observed across the jet in 3C~273 \\citep{2002PASJ...54L..39A,2008ApJ...675...79A,2005ApJ...626L..73Z,2005ApJ...633L..85A} and other sources \\citep[e.g.,][]{2004MNRAS.351L..89G}, however they do not seem to be a universal feature \\citep{2003ApJ...589..126Z}. \\begin{figure*}[t!] \\center{\\includegraphics[scale=0.85]{Gomez_fig1.eps}} \\caption{\\footnotesize Map of the mean value of the rotation measure across epochs. Data with standard deviation larger than 1000 rad m$^{-2}$ were discarded. Bars indicate the mean value of the RM-corrected EVPAs, with all displayed pixels having a standard deviation smaller than 30$^{\\circ}$ (96\\% under 20$^{\\circ}$). Contours show the 22 GHz total intensity at epoch 2001.00 for reference. The thick lines indicate the direction of the slices shown in Fig.~\\ref{slices}.} \\label{rm_mean} \\end{figure*} Thanks to its proximity ($z=0.033$), Very Long Baseline Array (VLBA) observations of the radio galaxy 3C~120 are capable of resolving the jet across its width, revealing a very rich structure in total and polarized flux \\citep{1998ApJ...499..221G,1999ApJ...521L..29G,2000Sci...289.2317G,2001ApJ...561L.161G,2008ApJ...681L..69G,2001ApJ...556..756W,2005AJ....130.1418J,2002Natur.417..625M,2007ApJ...665..232M}. In addition, evidence for the presence of a helical magnetic field has been found in 3C~120 by analyzing the motion and polarization of superluminal components \\citep{2001ApJ...561L.161G,2005ApJ...620..646H}. ", "conclusions": "" }, "0807/0807.2799_arXiv.txt": { "abstract": "Observations of the intergalactic medium (IGM) suggest that quasars reionize HeII in the IGM at $z \\approx 3$. We have run a set of $190$ and $430$ comoving Mpc simulations of HeII being reionized by quasars to develop an understanding of the nature of HeII reionization and its potential impact on observables. We find that HeII reionization heats regions in the IGM by as much as $25,000 \\, \\Kelvin$ above the temperature that is expected otherwise, with the volume-averaged temperature increasing by $\\sim 12,000 \\, \\Kelvin$ and with large temperature fluctuations on $\\sim 50$ Mpc scales. Much of this heating occurs far from quasars by photons with long mean free paths. We find a temperature-density equation of state of $\\gamma -1 \\approx 0.3$ during HeII reionization, but with a wide dispersion in this relation having $\\sigma_{T} \\sim 10^4$ K. HeII reionization by the observed population of quasars cannot produce an inverted relation ($\\gamma - 1 < 0$). Our simulations are consistent with the observed evolution in the mean transmission of the HeII Ly$\\alpha$ forest. We argue that the heat input from HeII reionization is unable to cause the observed depression at $z \\approx 3.2$ in the HI Ly$\\alpha$ forest opacity as has been suggested. We investigate how uncertainties in the properties of QSOs and of HeII Lyman-limit systems influence our predictions. ", "introduction": "\\label{firstpage} In the standard picture for the reionization history of the Universe, radiation from Population II stars ionized the intergalactic HI at $z>6$ as well as the HeI, converting the vast majority of the intergalactic helium to HeII. However, these stars cannot ionize HeII, and at $z \\approx 3$ quasars, with their harder UV spectrum, doubly ionize the intergalactic helium. To test this model, many observations are targeting high redshifts to probe hydrogen reionization (e.g., \\citealt{fan02, taniguchi05, bouwens07, kashikawa06, stark07, totani06}). In this picture, HeII reionization occurs at redshifts for which there is substantially more data on the state of the intergalactic medium (IGM). In fact, a number of observations suggest that HeII reionization happened at $z \\sim 3$. Two measurements of the mean transmission in the HI Ly$\\alpha$ forest have noted an upward bump at $z \\approx 3.2$ \\citep{bernardi03, faucher07}, which \\citet{theuns02} interpreted as arising from a temperature increase of the IGM during HeII reionization (but see \\citet{faucher07} for alternative explanations). An increase in the average temperature of the IGM would also decrease the small-scale fluctuations in the HI Ly$\\alpha$ forest. \\citet{ricotti00} and \\citet{schaye00} measured the temperature from the widths of the narrowest lines in the HI Ly$\\alpha$ forest and claimed to have detected a sudden increase in the temperature of $\\Delta T \\sim 10^4$ K between $z = 3.5$ and $3$. Photo-heating during HeII reionization is the only known process that could be responsible for such an effect (e.g., \\citealt{miralda94, abel99}). However, a subsequent study by \\citet{mcdonald01b} using a similar method and \\citet{zaldarriaga01} using the HI Ly$\\alpha$ forest power spectrum did not confirm this sudden increase in temperature, but rather found a constant temperature at mean density of $T_0 \\approx 17,000$ K for $2 < z < 4$. Temperatures of $17,000\\; K$ are difficult to explain without HeII reionization occurring at $z \\sim 3$ \\citep{hui03}, and it is unclear whether as sudden an increase in temperature as \\citet{ricotti00} and \\citet{schaye00} find is even expected theoretically. If a substantial fraction of the helium is in HeII ($\\gtrsim 1\\%$), this would produce a Gunn-Peterson absorption trough in the spectra of high-redshift quasars at wavelengths blueward of HeII Ly$\\alpha$. Observations of HeII Ly$\\alpha$ forest absorption at $2.8 < z < 3.3$ find $10$s of comoving Mpc regions with no detected transmission \\citep{jakobsen94, davidsen96, hogan97, reimers97, heap00}. These troughs may signify the presence of diffuse intergalactic HeII. However, current data, which consist of only a few quasar sight-lines, do not rule out the intergalactic HeII being primarily ionized and in photo-ionization equilibrium with a weak background \\citep{giroux97, fardal98,heap00}. The Cosmic Origins Spectrograph, which NASA plans to install on the Hubble Space Telescope in 2009, will increase the quantity and quality of HeII Ly$\\alpha$ forest sight-lines. As the intergalactic HeII becomes progressively more ionized, the extragalactic UV background will harden around the ionization energy of HeII at $54.4 \\; \\eV$. This hardening will affect the ionization state of intergalactic metals. \\citet{songaila98} observed a sharp evolution at $z \\approx 3$ in the column density ratios in SIV (Ionization Potential $= 45.1$ eV) to CIV ($64.5$ eV) absorbers. \\citet{boksenberg03} found evidence for a more gradual hardening of the background between $2 < z <4$ from the column density ratios of NV ($98 \\, {\\rm eV}$) to CIV. Finally, by simultaneously fitting to multiple metal lines that originate from the same absorption systems, \\citet{agafonova05} and \\citet{agafonova07} inferred a background spectrum that is hardening at $z \\approx 3$ near $4$ Ry. Measurements from the $z \\sim 3$ HI Ly$\\alpha$ forest ignore the effects of a patchy HeII reionization process. For example, estimates of the photo-ionizing background and the IGM temperature from the forest assume a power-law temperature-density relation. Cosmological parameter studies from the HI Ly$\\alpha$ forest power spectrum make a similar assumption. Different regions can have vastly different HeII reionization histories, resulting in a more complicated distribution of temperatures and pressure-smoothing scales than is commonly adopted. Realistic simulations of HeII reionization will help quantify the level at which this process biases these measurements.\\footnote{Of note, \\citet{lai06} used simple models for HeII reionization to show that it has a surprisingly small effect on the Ly$\\alpha$ forest power spectrum on large scales, modifying it at a $ \\lesssim 5\\%$ level for wavevectors $k \\lesssim 5 \\, {\\rm comoving} \\; \\Mpc^{-1}$.} The aim of this paper is to run realistic simulations of HeII reionization to understand the morphology of this process as well as its effect on observables. We concentrate primarily on its impact on quantities that are sensitive to the IGM temperature, but we also study the effect of HeII reionization on the transmission in the HeII Ly$\\alpha$ forest. \\citet{sokasian02} and \\citet{paschos07} have performed the most realistic simulations of patchy HeII reionization to date. There are several differences between our work and these earlier investigations. Both of these studies employed volumes $\\leq 100^3$ comoving $\\Mpc^3$. Here, we examine HeII reionization in $186^3$ and $429^3$ comoving Mpc$^3$ volumes, providing a more representative cosmic sample. However, both \\citet{sokasian02} and \\citet{paschos07} simulated HeII reionization as a post-processing step on top of cosmological simulations that included gas dynamics. Our study instead uses N-body simulations, which result in a less realistic model for the gas distribution, but afford a larger dynamic range. Furthermore, \\citet{sokasian02} assumed sharp ionizing fronts and ignored the detailed temperature evolution. \\citet{paschos07} did not calculate the gas temperature self-consistently. Our calculations capture the width of the ionization fronts and the temperature in a consistent manner. Finally, in contrast to previous studies, our work presents a large set of radiative transfer simulations in order to survey the parameter space. In Section \\ref{code}, we describe the details of our code. The models for the quasar sources are described in Section \\ref{sources}. Section \\ref{sims} presents the simulations. Finally, Section \\ref{observations} addresses the implications HeII reionization has on observations of the HI and HeII Ly$\\alpha$ forests. Throughout, we use a $\\Lambda$CDM cosmology with $n_s = 1$, $\\sigma_8 = 0.8$, $\\Omega_m = 0.27$, $\\Omega_{\\Lambda} = 0.73$, $\\Omega_b = 0.046$, and $h = 0.7$, consistent with the most recent WMAP results \\citep{komatsu08}. All distances are in comoving units unless specified otherwise. An overbar over a variable signifies a volume average, and $x_Y$ is the fraction of helium/hydrogen that is in ionization state $Y$. ", "conclusions": "We have run a set of simulations of HeII reionization to understand the structure of HeII reionization and its effect on several observables. We find that for a late reionization of HeII by quasars: \\begin{itemize} \\item The popular assumptions that the HeIII ionization fronts are sharp and that there is uniform heating within the front (and no heating outside of it) do not yield a realistic model for HeII reionization. While the ionization fronts are still fairly localized, hard photons stream far from their sources and are absorbed ahead of the front. These photons inject at large distances a significant fraction of the energy radiated above the HeII Lyman-limit. \\item The average temperature at the mean density is increased by $\\approx 12,000 \\; {\\rm K}$ over the average temperature of the gas in the absence of HeII reionization for $\\bar{\\alpha}_{\\rm UV} = 1.6$. This temperature increase is consistent with estimates that assume that the IGM absorbs all photons with energies less than a few hundred eV during HeII reionization. Regions that are ionized last are ionized by the hardest radiation, reaching $T > 30,000 \\; {\\rm K}$ in our fiducial model. If the spectral index of the QSOs is different from our fiducial value of $\\bar{\\alpha}_{\\rm UV} = 1.6$, the average temperature can be be significantly altered. \\item Poisson fluctuations in the number of QSOs rather than their spatial clustering shape the structure of the ionization and temperature fluctuations on $\\lesssim 50$ Mpc scales because rare $L \\sim L_*$ quasars ionize the HeII. HeII reionization produces large temperature fluctuations on $50$ Mpc scales, and it results in the ionization and HeII photo-heating fluctuations being essentially uncorrelated with the density fluctuations on the Jeans scale, the scale most relevant for HI and HeII Ly$\\alpha$ forest statistics. \\item Measurements of the $z \\sim 3$ forest suggest $T_0 \\approx 20,000\\, {\\rm K}$. Without invoking an exotic heating mechanism, a late HeII reionization epoch is required to produce this temperature. The amount of additional heat provided by HeII reionization in our simulations is enough to produce the inferred temperatures. \\item To the extent that there is a $T$-$\\Delta_b$ relation, HeII reionization by quasars leads to a temperature-density equation of state of $\\gamma -1 \\approx 0.3$ for $0.1 \\lesssim \\bar{x}_{\\rm HeIII} \\lesssim 0.9$. HeII reionization by QSOs cannot result in an inverted relation ($\\gamma - 1 < 0$) as has been claimed. \\item Our simulations of a $z\\sim 3$ HeII reionization process produce a similar evolution in the HeII Ly$\\alpha$ mean transmission to what has been estimated. Better observations of the mean transmission and its scatter will be able to rule out models presented here. \\item In our simulations, the heating from HeII reionization by quasars is unable to produce any semblance of the observed depression in the $z \\approx 3.2$ HI Ly$\\alpha$ forest opacity. The HeII reionization process is too extended in redshift ($\\Delta z > 1$) to be responsible for this narrow feature ($\\Delta z \\lesssim 0.4$). \\end{itemize} The morphology of HeII reionization is considerably different than that of hydrogen reionization (e.g., \\citealt{mcquinn07}). During hydrogen reionization, current models find that the growth of ionized bubbles is more collective; hundreds or even thousands of galaxies within a bubble contribute to its growth, leading to the bubble structure tracing the distribution of galaxies and to tens of Mpc HII regions (e.g., \\citealt{zahn07}). During HeII reionization, the growth is more stochastic. Regions that happen to host a ``nearby'' quasar (within $\\sim 30$ Mpc) are ionized by that quasar. Because the QSO bubbles are so extended, the ionized structures are larger than during HI reionization. Another significant difference stems from the m.f.p. of the ionizing photons to be absorbed in diffuse gas. The spectrum from quasars is harder than that of stars -- our best guess for what ionizes the hydrogen -- the number density of helium is $70$ times smaller at $z= 3$ than hydrogen at $z= 6$, and the cross section is $4$ times smaller. These three factors conspire to make the typical m.f.p. for a HeII ionizing photon megaparsecs rather than kiloparsecs, as it is for HI ionizing photons during hydrogen reionization. We have seen that some HeII ionizing photons free stream far from their sources, partially ionizing and heating up these regions. If stars reionize the hydrogen, ionization and heating occurs within HII bubbles. A definitive identification of the redshifts of HeII reionization will place constraints on the sources that produce $> 4$ Ry photons and will aid studies of the HI Ly$\\alpha$ forest. The data from previous observations, if analyzed properly, may be able to confirm whether quasars reionize HeII at $z \\approx 3$. In addition, future observations of the HI and HeII Ly$\\alpha$ forest will soon be available with Sloan Digital Sky Survey III\\footnote{www.sdss3.org} and the Cosmic Origins Spectrograph on the Hubble Space Telescope. These telescopes will significantly increase the number of sight-lines in the HI and HeII Ly$\\alpha$ forests. It is timely to make predictions for the effect of HeII reionization on these observations." }, "0807/0807.1942_arXiv.txt": { "abstract": "The outer disks of galaxies present a unique laboratory for studying the process of disk formation. A considerable fraction of observed disks exhibit a break in their surface brightness profiles. The ubiquity of these features points to a crucial aspect of disk formation which must be explained. Recent theoretical work suggests that such breaks are related to significant amounts of radial migration. We discuss the current observational evidence which supports this picture. ", "introduction": "The majority of the thin disk forms out of gas quiescently cooling and collapsing inside the host dark matter halo following the last major merger \\citep{Brook:2004}. While the inner parts of the stellar thin disk are entangled with the pre-merger material, the outer parts evolve in relative solitude, modulo interactions with substructure components present within the host halo. The outer parts of galactic disks therefore provide us with a direct view of disk assembly in progress. Since \\citet{van-der-Kruit:1979,van-der-Kruit:1987} it has been known that the surface brightness profiles of disk galaxies may not always follow a simple single-exponential law. In a recent work using data from the Sloan Digital Sky Survey (SDSS), \\citet{Pohlen:2006} showed that $\\sim60\\%$ of nearby disk galaxies have downward-bending surface brightness breaks, traditionally termed ``truncations''. Disk breaks have also been observed in the distant universe out to a redshift of z$\\sim$1 \\citep{Perez:2004,Trujillo:2005,Azzollini:2008a}, further implying that they are a generic feature of disk formation. Several theories for the formation of downward-bending breaks have been suggested. The most common interpretations include angular momentum-limited collapse \\citep{van-der-Kruit:1987, van-den-Bosch:2001}, star formation threshold either due to a critical gas density \\citep{Kennicutt:1989} or a lack of a cool equilibrium ISM phase \\citep{elmegreen:1994,Schaye:2004}. Alternatively, breaks have also been attributed to angular momentum redistribution \\citep{Debattista:2006, Foyle:2008}. ", "conclusions": "We have argued that outer disk breaks are a phenomenon that is not only common in observed systems, but that its mechanism of formation may provide important insights into disk evolution as a whole. Substantial observational evidence taken in the context of recent theoretical models suggests that at least a fraction of outer parts of late-type galactic disks form due to substantial radial migration of fully-formed stellar material. Such evolution affects these extreme outer regions of galaxies, and profoundly impacts the properties of stellar populations in the entire disk." }, "0807/0807.1490_arXiv.txt": { "abstract": "According to previous investigations the effect of diffusion in the stellar atmospheres and envelopes of subdwarf B (sdB) stars with luminosities $10 \\la L / L_{\\odot} \\la 100$ strongly depends on the presence of weak winds with mass loss rates $\\dot M \\la 10^{-12} M_{\\odot}/ \\rm yr$. These calculations with the mass loss rate as a free parameter have shown that it is hardly possible to reproduce the measured abundances of helium and metals simultaneously. A possible reason is the decoupling of metals, which preferably absorb the photon momentum, from hydrogen and helium in the wind region. In the present paper it will be investigated if ``chemically homogeneous\" winds, as assumed in previous investigations, with mass loss rates $\\dot M \\leq 10^{-12} M_{\\odot} / \\rm yr$ can exist. From an observational point of view the existence of weak winds in sdB stars is unclear. Only in the most luminous ones possible wind signatures have been detected. Therefore it will be investigated if according to the theory of radiatively driven winds the existence of weak winds is plausible. A stellar mass $M_{*} = 0.5 M_{\\odot}$ is assumed. The results for effective temperatures $T_{\\rm eff} = 35000$, $30000$ and $25000 \\rm K$, metallicities $0.1 \\leq Z/Z_{\\odot} \\leq 1$ predict decreasing mass loss rates with increasing surface gravity. Dependent on the luminosity and metallicity the mass loss rates are between about $10^{-11} M_{\\odot} / \\rm yr$ and zero. If at all, chemically homogeneous winds can exist for the most luminous sdB stars only. For the other ones selective winds are expected which should lead to additional changes of the surface composition. In sdB stars, hot white dwarfs and HgMn stars (which are chemically peculiar main sequence stars) the measured metal abundances are tendencially lower than the ones predicted from diffusion calculations which assume an equilibrium between gravitational settling and radiative levitation. Only for helium in almost all cases the measured abundances are larger than the predicted ones, but usually lower, below the solar value. This may be an indication that the abundance anomalies of metals are preferably due to the selective winds, whereas the helium deficiencies are due to gravitational settling, which for still unknown reasons is less effective than expected in an undisturbed stellar atmosphere. ", "introduction": "The abundance anomalies in subdwarf B (sdB) stars, white dwarfs and chemically peculiar main sequence stars are believed to be at least partially due to the effect of diffusion in the stellar atmosphere and envelope. Several attempts have been made to explain the abundances with the effect of gravitational settling which may be counteracted by radiative levitation. As the radiative force on an element decreases with increasing abundance due to saturation effects, the surface composition can be predicted from the equilibrium condition between the inward gravitational force and the outward radiative force. However, an agreement between predicted and measured abundances requires the absence of disturbing processes like mass loss or convective mixing. This may be the reason why in many cases the agreement is not satisfactory. A comparison between predicted and measured metal abundances for sdB stars \\citep[e.g.][]{t12_ber88,t12_ohl00,t12_chay06,t12_nat08} and hot white dwarfs \\citep[e.g.][]{t12_son02,t12_son05,t12_good05} shows that the measured abundances are tendencially lower than the predicted ones, although a few exceptions exist (e.g.\\ silicon in hot white dwarfs). According to recent spectral analyses of sdB stars \\citep[e.g.][]{t12_gei08,t12_gei08b, t12_otol06,t12_edel06,t12_blan06} especially the hotter ones with $T_{\\rm eff} > 30000 \\rm K$ in many cases show strong deficiencies of light metals like Al, Mg, O and Si by more than a factor of $100$ in comparison to the solar abundances, whereas enrichments of elements heavier than iron by a factor of $100$ are not unusual. Qualitatively, these abundance patterns show some similarity to those found in HgMn stars, which are a subgroup of the chemically peculiar main sequence stars reviewed by \\citet{t12_smith96}. The HgMn stars with $10000 \\rm K \\la T_{\\rm eff} \\leq 16000 \\rm K$, $\\log g \\approx 4.0$ are characterized by low rotational velocities and weak or non-detectable magnetic fields. Some light metals (e.g.\\ Al, N) tend to be deficient, whereas heavy metals (e.g.\\ Hg, Mn, Pt, Sr, Ga) may be enriched up to several orders of magnitude. Some of the most recent spectral analyses are from \\citet{t12_zav07} and \\citet{t12_adel06}. The measured abundances of iron group elements \\citep{t12_sea96,t12_jom99} as well as nitrogen \\citep{t12_roby99} tend to be lower than predicted from equilibrium diffusion calculations. Only for mercury abundances larger than predicted have been detected \\citep{t12_prof99}. So in hot white dwarfs, sdB and HgMn stars a common tendency seems to be present, according to which the metal abundances are lower than expected from the equilibrium condition between gravitational settling and radiative levitation. In addition there is a large scatter of abundances from star to star. Even for stars with similar stellar parameters the abundances may be different. This points to some time-dependent process and not to an equilibrium state. In main sequence stars the chemically peculiar phenomenon is restricted to stars with low rotational velocities. In addition the presence of magnetic fields may be of importance, because magnetic fields may change the radiative acceleration or suppress convection \\citep{t12_tur03}. White dwarfs and sdB stars, however, are always more or less chemically peculiar. Up to now no correlation between abundance anomalies and magnetic field strengths has been found \\citep{t12_otol05}. In hot DA white dwarfs no outer convection zones should exist, because hydrogen is preferably ionized and helium is strongly deficient. In sdB stars a thin superficial convection zone with a mass depth of the order $10^{-12} M_{*}$ may be present only for helium abundances $\\rm He / \\rm H \\geq 0.01$ by number \\citep{t12_gro85}. So it seems to be unlikely that magnetic fields are of decisive importance for the explanation of the surface compositions. In hot DAO white dwarfs helium is detectable, but in many cases it is deficient in comparison to the solar value \\citep[e.g.][]{t12_napi99}. The same is true for the majority of sdB's \\citep[e.g.][]{t12_edel03,t12_lisk05} and the helium deficient main sequence stars (e.g.\\ in HgMn stars helium usually is deficient). In contrast to the metal abundances, however, the abundances of helium are always larger than predicted from equilibrium diffusion calculations \\citep{t12_ven88,t12_mic89,t12_mic79}. These results may be explained with the effect of gravitational settling which, however, somehow must be disturbed. Possible reasons for this disturbance of the equilibrium may be the presence of turbulence \\citep{t12_vauc78} or mass loss. \\citet{t12_fon97} and \\citet{t12_ub98} predicted helium abundances as a function of time in the presence of weak winds. The results have shown that for mass loss rates of the order $10^{-13} M_{\\odot} / \\rm yr$ helium sinks much more slowly than in the case of an undisturbed stellar atmosphere. Within the lifetimes of sdB stars near the extended horizontal branch ($\\approx 10^{8} \\rm yr$) the helium abundance would gradually decrease from the solar value to $\\rm He / \\rm H \\approx 10^{-4}$ by number. This could explain why the helium abundances usually are in this range. If this scenario with weak winds were the correct explanation for both the helium and the metal abundance anomalies, then it should be possible to find a mass loss rate for which all abundances can be explained simultaneously. The calculations of \\citet{t12_ub01} for the elements H, He, C, N and O have shown that this is hardly possible. According to these calculations for solar initial composition helium should always be more deficient than the metals. No mass loss rate exists which leads to deficiencies of C and O by more than a factor of $100$, whereas helium is deficient by a factor of ten only. This, however, is not an unusual composition in sdB stars \\citep[e.g.][]{t12_heb00}. Moreover in the presence of winds with mass loss rates of the order $10^{-13} M_{\\odot} / \\rm yr$ which are required to explain the helium abundances, the proposed pulsation mechanism of some sdB stars \\citep{t12_char97,t12_fon03} would become questionable. As mass loss tends to level out concentration gradients, the reservoir of iron (or other iron group elements) in the stellar envelope needed to explain the pulsations should be destroyed in time scales which are much shorter than the lifetimes of sdB stars. According to \\citet{t12_chay04} and \\citet{t12_fon06} for a mass loss rate of $\\dot M = 6\\times 10^{-15} M_{\\odot} / \\rm yr$ the reservoir would be destroyed on about $10^{7} \\rm yr$. For $\\dot M = 10^{-13} M_{\\odot} / \\rm yr$ the matter in mass depths $\\la 10^{-7} M_{*}$, where the reservoir is expected, would be blown away within one million years only. Probably the most crucial assumption in these diffusion calculations with mass loss has been that the winds are ``chemically homogeneous\". If $\\dot M_{l}$ is the mass loss rate of an element $l$ and $\\zeta_{l}$ its mass fraction in the photosphere, then this assumption states that $\\dot M_{l} = \\zeta_{l} \\dot M$, where $\\dot M$ is the total mass loss rate. Such a chemically homogeneous wind prevents (if $\\dot M \\ga 10^{-11} M_{\\odot} / \\rm yr$) or retards (if $\\dot M < 10^{-11} M_{\\odot} / \\rm yr$) gravitational settling. However, it does not directly change the surface composition. The opposite case would be a ``selective\" wind in which the mass loss rates of the individual elements are essentially independent of each other. A selective wind should lead to additional changes of the surface composition, which have not yet been taken into account in the calculations. In radiatively driven winds of hot stars the photon momentum is absorbed preferably by the metals \\citep[see e.g.][]{t12_abb82,t12_vink01}, whereas the contribution of hydrogen and helium is small. Thus the metals are accelerated and move outwards. If the flow of metals is sufficiently large, then due to Coulomb collisions with the metals hydrogen and helium are accelerated as well. For this purpose, in dense winds a small velocity difference between the metals and hydrogen and helium is sufficient. Then it should be a good approximation that all elements have the same velocity and that the wind is chemically homogeneous. If, however, the flow of metals is small, then it may happen that the coupling of the various constituents due to collisions is not sufficiently effective and that hydrogen and helium are left behind. This scenario may lead to pure metallic winds such as investigated by \\citet{t12_bab95} for main sequence A stars. Mass loss has been detected in subdwarf O stars (sdO) which are more luminous than sdB's \\citep{t12_ham81,t12_rau93}. Up to now there is no observational proof for the existence of winds in sdB stars. From a quantitative analysis of $\\rm H \\alpha$ line profiles of 40 sdB stars \\citep{t12_max01}, a comparison of synthetic NLTE $\\rm H \\alpha$ line profiles from static model atmospheres with the observations revealed perfect matches for almost all stars. Only in the four most luminous sdB's anomalous $\\rm H \\alpha$ lines with a small emission at the line center have been detected, which possibly are signatures of weak winds \\citep{t12_heb03}. For the case $T_{\\rm eff} = 36000 \\rm K$, $\\log g = 5.5$ and $\\log L/L_{\\odot} = 1.51$, from a spectral synthesis of $\\rm H \\alpha$ with his wind code \\citet{t12_vink04} found a similar behaviour of $\\rm H \\alpha$ if the existence of a weak wind with $\\dot M \\approx 10^{-11} \\rm M_{\\odot} / \\rm yr$ is assumed. In Sect.~2 it will be investigated for a stellar mass $M_{*} = 0.5 M_{\\odot}$ if winds with mass loss rates $\\dot M \\la 10^{-12} M_{\\odot} / \\rm yr$ can be chemically homogeneous. The arguments are similar as in the investigations for more luminous stars e.g.\\ from \\citet{t12_ow2002}, \\citet{t12_spr92} and \\citet{t12_krt03}. All metals are lumped together into one mean metal which is accelerated due to the absorption of photon momentum. This is some simplification, in \\citet{t12_krt06} the individual elements are considered separately. Hydrogen, helium and the free electrons are denoted as ``passive plasma\" which can only be accelerated due to collisions with the outflowing metals. From the calculations e.g.\\ of \\citet{t12_abb82} and \\citet{t12_vink01} as well as from own calculations as described in Sect.~3 there is no indication that the radiative force on hydrogen and helium is not negligible. The mean radiative acceleration on the metals exceeds the one on hydrogen and helium by at least a factor of $100$. In Sect.~3 the results of mass loss calculations for sdB stars according to the original theory of radiatively driven winds of \\citet{t12_ca1975} are presented and are compared with the predictions from the mass loss recipe of \\citet{t12_vink02}. For these wind models, which are obtained from a one component description of the wind, it is again checked if the metals may be coupled to hydrogen and helium. In Sect.~4 the consequences of the results for the surface composition of sdB stars are discussed. ", "conclusions": "In the $T_{\\rm eff} - \\log g$ diagram of Fig.~2 for $25000 \\mathrm{K} \\leq T_{\\rm eff} \\leq 40000 \\rm K$ and for various metallicities the lines are shown above which according to the mass loss calculations as described in Sect.~3 chemically homogeneous winds may exist. \\begin{figure} \\centering \\includegraphics[width=11cm,bb=132 75 625 425,clip=true]{klaus2.eps} \\caption{Lines in the $T_{\\rm eff} - \\log g$ diagram above which chemically homogeneous winds may exist for $Z/Z_{\\odot} = 0.1$, $1/3$ and $1$, respectively. Squares and circles represent the sdB stars analyzed by \\citet{t12_max01} and \\citet{t12_lisk05}, respectively. Filled symbols represent the sdB's with peculiar $\\rm H \\alpha$ line profiles, which may indicate the presence of a weak wind.} \\end{figure} It can be seen that for the majority of sdB stars this is not possible if the metallicity is solar or subsolar. They are below the line for $Z / Z_{\\odot} = 1$. The sdB stars introduced in the diagram have been analyzed by \\citet{t12_max01} and \\citet{t12_lisk05}. According to the assumptions of the present paper in those ones with peculiar $H \\alpha$ line profiles (represented by filled symbols) chemically homogeneous winds may indeed exist if the metal abundances are not too far below the solar value. The predicted mass loss rates agree with the ones from the mass loss recipe of \\citet{t12_vink02} and are of the order $10^{-10}$ to $10^{-11} M_{\\odot} / \\rm yr$. However, as explained in Sect.~3, this agreement does not exclude that the mass loss rates are overestimated. So this result is still questionable. The existence of winds may depend on the abundance of one element only, which preferably contributes to the radiative acceleration. From the results it is clear that chemically homogeneous winds with mass loss rates $\\dot M \\la 10^{-12} M_{\\odot} / \\rm yr$ cannot exists. From arguments similar as in the paper of \\citet{t12_ow2002}, this result can be obtained without the calculation of a wind model. The mass loss rate of the metals alone must be at least of the order $10^{-14} M_{\\odot} / \\rm yr$. Otherwise hydrogen and helium cannot be accelerated throughout the wind. This result could be questioned only if a wind solution could be found for which the terminal velocity of the metals is clearly lower than the surface escape velocity. \\citet{t12_krt00} suggested that the terminal velocity could indeed be lower than $v_{\\rm esc}$ (by a factor of the order two only, however), if the wind switches to a ``shallow\" solution with an abrupt lowering of the velocity gradient. However, \\citet{t12_ow2002} and \\citet{t12_krt02} argued that these solutions are unstable. If hydrogen and helium cannot be expelled from the star, then pure metallic winds may exist. As the outflowing metals in the stellar atmosphere not only have to overcome the gravitational force, but in addition the frictional force due to collisions with protons and helium particles, the metal abundances should be lower than predicted from equilibrium diffusion calculations (if concentration gradients are negligible). For these metals, for which the mass loss rate is sufficiently small, measured and predicted abundances should be in agreement. For metals with non-zero mass loss rate this scenario should lead to abundances varying with time as has been discussed by \\citet{t12_sea96, t12_sea99} for iron group elements in the envelopes of HgMn stars. If both hydrogen and helium are in hydrostatic equilibrium, then measured helium abundances should be approximately in agreement with the ones predicted from equilibrium diffusion calculations. The fact that in the various types of helium deficient chemically peculiar stars the measured ones are larger, seems to indicate that gravitational settling in general is less effective than expected in an undisturbed stellar atmosphere. Moreover the existence of two distinct sequences of sdB stars which are characterized by an offset in the helium abundance \\citep{t12_edel03,t12_lisk05} can hardly be explained with one atmospheric effect alone. It may point to a dependence on the star's history. Several scenarios of single star and binary evolution of the sdB and the hotter sdO stars are under discussion \\citep{t12_str07}." }, "0807/0807.1729_arXiv.txt": { "abstract": " ", "introduction": "Recent times have seen cosmology enter a precision era. Measurements of the cosmic microwave background (CMB) have lead to estimates of the standard 6 cosmological parameters (see, for example, the 5 year results from WMAP~\\cite{Dunkley:2008ie} and references therein). Of particular interest here are the estimates of the amplitude, $A_{\\rm s}$, and spectral index, $n_{\\rm s}$, of the scalar density fluctuations which are measured to be $n_{\\rm s}=0.963^{+0.014}_{-0.015}$ and $A_{\\rm s}=(2.41\\pm0.01)\\times 10^{-9}$. It is presumed that the density fluctuations are created during an inflationary era through quantum effects. Constraining the nature of this epoch on the basis of the observations is now the focus of much effort~\\cite{Peiris:2006sj,Kinney:2006qm,Martin:2006rs} with the ultimate aim of making contact with particle physics models around the Grand Unified Theory (GUT) scale~\\cite{Lyth:1998xn}. In addition to these fluctuations of quantum origin, there may be others created by topological defects formed naturally as a consequence of a phase transition at the end of hybrid inflation~\\cite{Linde:1993cn}. Such fluctuations cannot be the primary source of CMB anisotropies, but they could provide a sub-dominant component~\\cite{Wyman:2005tu}, which can have some interesting effects. In particular, it has been shown that if cosmic strings are formed, then $n_{\\rm s}=1$ can be made compatible with the data~\\cite{Battye:2006pk,Bevis:2007gh,Battye:2007si}. This is of interest since many of the models of inflation motivated by fundamental theories predict larger values for the spectral index than the WMAP best fit value. For example, SUSY $F$- and $D$-term hybrid inflationary models~\\cite{Copeland:1994vg,Dvali:1994ms,DTerminf} predict $n_{\\rm s}>0.98$~\\cite{Panagiotakopoulos:1997qd}, while even larger values may be required if one considers gravitino constraints on the reheat temperature or the bounds on the cosmic string tension in $D$-term models. Brane inflation~\\cite{Dvali:1998pa} in its simplest form predicts $n_{\\rm s}>0.97$, but again, more sophisticated implementations may lead to a scale-invariant spectral index~\\cite{Haack:2008yb}. Most work on inflation plus defect scenarios has focused on strings. However, the recent suggestion~\\cite{Cruz:2007pe,Cruz:2008sb} that the previously identified cold spot in the CMB~\\cite{Cruz:2006fy} could be due to cosmic textures~\\cite{Turok:1989ai,Turok:1990gw} motivates our consideration of hybrid inflation scenarios involving global defects. Such dynamically unstable defects produce a spectrum of anisotropies qualitatively similar to strings~\\cite{Bevis:2004wk} and are likely to have similar properties in respect of the degeneracy with $n_{\\rm s}$. We note, in addition, the other cold spots detected on smaller scale in the CMB which do not appear to be associated with the Sunyaev-Zel'dovich effect~\\cite{GenovaSantos:2008hf}. Textures are most easily implemented in non-supersymmetric models, where they can emerge from a scalar potential of the form~\\cite{VilenkinShellard} \\begin{equation} \\label{V:texture} V(\\Phi)=\\frac14 \\lambda (\\Phi_i\\Phi_i-\\phi_0^2)^2\\,, \\end{equation} where $\\lambda$ is a dimensionless coupling constant and $\\phi_0$ is the symmetry breaking scale. The fields $\\Phi_i$ are a fundamental representation of a global ${\\rm O}(4)$-symmetry. This potential induces spontaneous symmetry breaking ${\\rm O}(4)\\to{\\rm O}(3)$, for which the vacuum manifold is~$S^3$, where $S^3$ denotes the 3-dimensional sphere. Since $\\pi_3(S^3)=\\mathbbm{Z}$, there is the possibility of non-trivial topological configurations called textures. However, they cannot be stable by Derrick's theorem and hence, once formed at the phase transition, they unwind. This unwinding process takes place in a self-similar fashion such that the density of textures scales relative to the background. The latter has been shown to be a more general feature of global field models with a broken $O(N)$ symmetry~\\cite{Turok:1991qq}. The process of unwinding creates density fluctuations and CMB anisotropies~\\cite{Turok:1990gw,Durrer:1990mk,Notzold:1990jt}. The CMB anisotropies result in a distribution of hot and cold spots, which by the central limit theorem, appear as approximately Gaussian on small scales. On larger scales, they may be revealed by more detailed analysis. By taking into account various observational biases, the analysis of Cruz {\\it et al.}~\\cite{Cruz:2007pe} has suggested that the large cold spot identified in the CMB would result from such a model with \\begin{equation} \\label{eta:obs} \\phi_0=(8.7^{+2.1}_{-3.0})\\times 10^{15}~{\\rm GeV}\\,. \\end{equation} In this paper, we study supersymmetric (SUSY) scenarios of $F$-term hybrid inflation that realise the global symmetry ${\\rm SU}(2)\\times {\\rm U}(1)_X$. We also consider the possibility that the ${\\rm U}(1)_X$-symmetry is gauged, giving rise to a semi-local model. Even though in the context of SUSY hybrid inflation, models with local symmetries are more commonly discussed, we note that the present paper is not the first one in which the use of global symmetries is considered. In one of the seminal papers on SUSY $F$-term inflation, the use of a spontaneously broken global ${\\rm U}(1)_X$-symmetry is suggested~\\cite{Copeland:1994vg}, and the possibilities of larger symmetries and textures are mentioned. To the best of our knowledge, however, neither an explicit analysis of the particle spectrum has been performed so far, nor have the consequences of the higher symmetry been investigated in detail within the framework of $F$-term hybrid inflation. An important aspect of the model-building is that unless additional ${\\rm SU}(2)$-invariant lifting terms are included, the vacuum manifold is larger than ${\\rm SU}(2)\\sim S^3$. This is because the holomorphic nature of the superpotential is consistently maintained by doubling the number of degrees of freedom in the symmetry-breaking fields, such as the field $\\Phi$ in~(\\ref{V:texture}). We describe here in detail how these additional degrees of freedom in the vacuum manifold can be lifted such that it is indeed reduced to $S^3$. Furthermore, we include the effect of radiative corrections on the predictions for the initial power spectrum created during the inflationary phase as first performed in Ref.~\\cite{Dvali:1994ms}. The breaking of a local instead of a global symmetry was first proposed in this work and following this reference most articles on SUSY-hybrid inflation consider only local symmetries. This may also be due to the fact that quantum gravitational effects can violate global symmetries. Since no complete model of quantum gravity exists yet, we do not consider this possibility here and take ${\\rm SU}(2)$ to be an exact symmetry. Such an assumption is not entirely unrealistic, since ${\\rm SU}(2)$ is the only group that is self-non-anomalous, independently of the representation of the chiral fermions in the theory. The organisation of the paper is as follows: in Section~2 we describe in detail the SUSY hybrid texture model based on the global symmetry ${\\rm SU(2)}\\times {\\rm U(1)}_X$. We also calculate its scalar mass spectrum and discuss mechanisms for giving masses to the massless moduli fields present in the model. Formal aspects of the model-building have been relegated to the appendices. In Section~3 we analyse possible scenarios for reheating after inflation, including constraints from possible overproduction of gravitinos and Big Bang Nucleosynthesis (BBN). In addition, we estimate the effect that the texture Goldstone bosons may have on stellar cooling. Section~4 discusses a semi-local realisation of the SUSY hybrid model, where the aforementioned global ${\\rm U(1)}_X$ symmetry is promoted to a local one. In Section~5 we study the inflationary dynamics of the SUSY hybrid model and analyse the possibility whether the texture scale $\\phi_0$ given in~(\\ref{eta:obs}) can naturally be implemented within this model, without being in conflict with other cosmological and astrophysical constraints discussed in Section~3. Finally, Section~6 summarises the main conclusions of our study. ", "conclusions": "" }, "0807/0807.3206_arXiv.txt": { "abstract": "In this paper we use coherently integrated visibilities (see separate paper in these proceedings\\cite{jorgensen:2008}) to measure the properties of binary stars. We use only the phase of the complex visibility and not the amplitude. The reason for this is that amplitudes suffer from the calibration effect (the same for coherent and incoherent averages) and thus effectively provide lower accuracy measurements. We demonstrate that the baseline phase alone can be used to measure the separation, orientation and brightness ratio of a binary star, as a function of wavelength. ", "introduction": "\\noindent Binary stars are important for calibrating evolutionary stellar models. Because stellar models are sensitive to the parameters of the stars, it is important to obtain the highest possible accuracy of the stellar parameters. In this paper we will demonstrate how to extract brightness ratio and separation vector from measurements of the complex visibility phase only. This is significant because the visibility phase does not suffer from the calibration effects of visibility amplitudes, and therefore much higher precision can be obtained. Further, there are more complex visibility baseline phases than closure phases such that using baseline phases instead of closure phases yields more information. Finally, complex visibility phases generally have better SNR than closure phases. ", "conclusions": "\\noindent We have demonstrated that coherently integrated visibility phases can be used to obtain high-quality measurements of the fundamental parameters of binary stars. The measurements are accurate to the point where careful calibration of wavelengths and bandpasses is necessary. This calibration was not available for the data sets considered in this paper, but is routinely performed on more recent NPOI data sets." }, "0807/0807.3801_arXiv.txt": { "abstract": " ", "introduction": "By the end of twentieth century the overall picture of star formation process in dense interstellar clouds became fully formed. Contraction (collapse) of protostellar clumps is initiated by gravitation and external pressure; thermal pressure, rotation of the clumps and magnetic field counteract contraction. Thus, general scenario of star formation is defined by a complex interaction of these factors. However, many fundamental problems of star formation remain unsolved. For instance, the mechanisms that stimulate the collapse of low-mass prestellar cores remain unclear. Even more enigmatic is formation of massive stars. Apparently, more complex processes, like competitive accretion and merger of protostellar fragments are involved in the latter process in addition to spherical and disc accretion. \\begin{table*}[p!] \\caption{Atomic and molecular lines in the 10~mkm to 2~cm range which are used for the study of star formation regions.} \\begin{tabular}{|l|c|l|} \\hline Molecule & Lines & Radiation mechanism \\& Specific problem \\\\ \\hline H$_{2}$O & 22~GHz & Maser emission, Studies of protostellar discs, astrometry \\\\ CH$_{3}$OH & 25~GHz & Maser and thermal emission, Astrometry, determination of \\\\ & & physical conditions in protoplanetary \\\\ & &discs and regions of massive star formation \\\\ NH$_{3}$ & 24~GHz & Thermal emission, Tracer of physical conditions \\\\ & &in the dense gas, especially in the late stages \\\\ & &of the evolution of prestellar cores \\\\ CS & 49~GHz & A tracer of physical conditions in the dense gas\\\\ & 98~GHz & \\\\ & 147~GHz & \\\\ & 244~GHz & \\\\ HCN & 89~GHz & A tracer of physical conditions in the dense gas \\\\ & 266~GHz & \\\\ HCO$^{+}$ & 89~GHz & Tracer of physical conditions \\\\ & & in the dense gas, including ionization degree \\\\ & 268~GHz & \\\\ HNC & 91~GHz & A tracer of physical conditions in the dense gas \\\\ & 272~GHz & \\\\ N$_{2}$H$^{+}$ & 93~GHz & Tracer of physical conditions \\\\ & & in the dense gas, especially in the late stages \\\\ & & of the evolution of prestellar cores and \\\\ & & in the regions of massive star formation \\\\ CO & 115~GHz & The main tracer of the presence of diffuse molecular gas\\\\ & 230~GHz &\\\\ H$_{2}$CO & 140~GHz & Tracer of physical conditions in the dense gas\\\\ NO & 150~GHz & Tracer of physical conditions in the dense gas \\\\ & 250~GHz &\\\\ H$_{2}$D$^{+}$ & 372~GHz & Tracer of physical conditions \\\\ & & in the dense gas, especially, of the \\\\ & & kinematics of the central regions of prestellar cores \\\\ C & 492~GHz & Tracer of physical conditions in diffuse gas, \\\\ & & PDR-regions, ultracompact HII regions \\\\ & 809~GHz & \\\\ C$^{+}$ & 1.9 THz & A tracer of physical conditions in diffuse gas, \\\\ & &PDR-regions, ultracompact HII regions\\\\ Si$^{+}$ & 8.6 THz & A tracer of physical conditions in the protoplanetary discs\\\\ H$_{2}$ & 10.7 THz & A tracer of physical conditions in the protoplanetary discs\\\\ Fe$^{+}$ & 11.5 THz & A tracer of physical conditions in the protoplanetary discs\\\\ S & 12.0 THz & A tracer of physical conditions in the protoplanetary discs\\\\ Fe & 12.5 THz & A tracer of physical conditions in the protoplanetary discs\\\\[1mm] \\hline \\end{tabular} \\end{table*} \\begin{table*}[t!] \\caption{Requirements to the angular resolution in observations of different stages of star formation} \\begin{tabular}{l|c|c|c} \\hline Stage & Typical & Typical & Angular resolution\\\\ & distance & scale & (10 diagrams per object) \\\\ \\hline Prestellar cores & 140 pc & 0.1 pc\\phantom{0} & $15^{\\prime\\prime}$\\phantom{0,} \\\\ Hot cores, UCHII$^*$ & 2--4 kpc and more & 0.1 pc\\phantom{0} & ${<}1^{\\prime\\prime}$\\phantom{00} \\\\ protostellar objects & 400 pc & 0.01 pc & \\phantom{0}$0.5^{\\prime\\prime}$ \\\\ Outer regions of the discs & 100 pc & 1000 a.u. & $1^{\\prime\\prime}$\\phantom{,} \\\\ Inner regions of the discs, & 100 pc & \\phantom{0}100 a.u. & \\phantom{0}$0.1^{\\prime\\prime}$ \\\\ brown dwarfs discs &&&\\\\ \\hline \\end{tabular} \\end{table*} \\vspace*{5pt} For the later evolutionary stages, when the planetary systems form, the number of unsolved problems is none the less. There are still no unique solutions for the problems of the nature of angular momentum transfer in the protoplanetary discs, their physical and chemical structure, the role of mixing in the formation of chemical and mineralogical composition of protoplanetary matter (including protosolar nebula). These uncertainties are related, first, to the low energetics of the transformation of gas into stars, especially in its initial stages, which makes impossible studies of this process in the visual waverange. For instance, in the prestellar cores the temperature does not exceed 10~K and it is only slightly higher at the periphery of protostellar objects, protoplanetary and more evolved debris discs. Therefore, a significant fraction of their radiation is emitted in submillimeter and millimeter waveranges. However, these waveranges are very informative \"--- the spectral range from 100~mkm to 20~mm contains thousands of lines of many dozens of interstellar molecules (table~1) which, in the absence of observed emission of molecular hydrogen, are the only source of information on composition, temperature, and kinematics of molecular clouds and star formation regions. Second, prestellar and, especially, protostellar objects are rather compact, but at the same time their structure is complex and its study demands high angular resolution (table~2). Just for this reason there is world-wide growing interest to construction of sensitive detectors for millimeter and submillimeter wavebands, including interferometers. Of course, the main expectations are related to the submillimeter interferometric system ALMA (located in the Atacama desert in Chile) which will consist of several tens of 12-m antennas with the maximum distance between them of 12~km~[1]. A somewhat lower scale project works successfully already. It is submillimeter interferometer SMA~[2] in Hawaii (USA) which consists of eight 6-m antennas. However, capabilities of the best of existing and planned ground-based instruments for the studies in millimeter and submillimeter ranges are limited by disturbances created by the Earth atmosphere. In the submillimeter range there are only several transparent windows in the Earth atmosphere, with transmittance factor that does not exceed \\vspace*{5pt} 60$\\%$ even in the Earth regions with the best astroclimate \\footnote{http://www.eso.org/projects/alma/specifications/ FreqBands.html.}. To some extent this problem will be solved with the launch of the space-born infrared telescope ``Herschel'', but its spatial resolution will be limited by its relatively small 3.5-m dish. More, in the long wavelength band ``Herschel'' will be sensitive up to 670!mkm only (ALMA will be sensitive up to 1~cm). Therefore, this telescope Will be unable to observe both the coldest clouds and many astrophysically interesting molecular lines, including, for instance, the lines associated with low-level rotational transition of the CO molecule. A prerequisite to substantial increase in our knowledge of star formation would be construction of an extra-atmospheric submillimeter telescope which would conjoin sensitivity and spectral resolution with high angular resolution (which would mean possibility of its usage in the regime of an interferometer). It is by no means unimportant that location of the telescope in the space would allow homogeneous observations both in northern and southern celestial hemispheres. Space telescope ``Millimetron'' which is a part of the Russian Federal Space Programme may become such an instrument. In Russia, problems of star formation are studied in most of astronomical research centers (see, for instance, the collection of works edited by Wiebe and Kirsanova~[3]). For this reason, high demand for observations by submillimeter telescope is expected from Russian scientists working in this field of astrophysics. In the present paper we describe some of the problems for which space-born submillimeter observatory will give crucially important results and justify some of requirements to its parameters. ", "conclusions": "Currently, several projects of millimeter and submillimeter range telescopes are already implemented or developed. The problems related to different stages of star formation take an important place in their programmes. However, the sensitivity of ground-based instruments is naturally limited by atmospheric absorption. A submillimeter telescope in the space, even operating in the one-mirror would allow to carry out extremely important investigations both of close and distant star formation regions (thanks to high sensitivity) and will allow to solve a lot of enigmas of star formation not only in the vicinity of the Sun but also in a more extended region of the Galaxy. To some extent this goal will be achieved by ``Herschel'' project, but the angular resolution of `Herschel'' will be not high enough to solve many fundamental problems of star formation theory. In this respect, capabilities of the ``Earth--Space'' interferometer will be unbeaten and will allow to obtain highly valuable information on the structure of the star formation regions and protostellar (young stellar) objects that is unavailable by other means." }, "0807/0807.0175_arXiv.txt": { "abstract": "{The mechanism of formation of the \\ion{He}{i} 10830~\\AA\\ triplet in cool stars has been subject of debate for the last 30 years. A relation between the X-ray luminosity and the \\ion{He}{i} 10830~\\AA\\ flux was found in cool stars, but the dominant mechanism of formation in these stars (photoionization by coronal radiation followed by recombination and cascade, or collisional excitation in the chromosphere), has not yet been established.} {We use modern instrumentation (NOT/SOFIN) and a direct measurement of the EUV flux, which photoionizes \\ion{He}{i}, to investigate the formation mechanism of the line for the most active stars which are frequently excluded from analysis.} {We have observed with an unprecedented resolution ($R\\sim 170,000$) the \\ion{He}{i} 10830~\\AA\\ triplet in a set of 15 stars that were also observed with the Extreme Ultraviolet Explorer (EUVE) in order to compare the line strengths with their EUV and X-ray fluxes.} {Active dwarf and subgiant stars do not exhibit a relation between the EUV flux and the equivalent width of the \\ion{He}{i} 10830~\\AA\\ line. Giant stars however, show a positive correlation between the strength of the \\ion{He}{i} 10830~\\AA\\ absorption and the EUV and X-ray fluxes. The strength of the \\ion{C}{iv} 1550~\\AA\\ emission does not correlate with coronal fluxes in this sample of 15 stars.} {Active dwarf stars may have high chromospheric densities thus allowing collisional excitation to dominate photoionization/recombination processes in forming the \\ion{He}{i} 10830~\\AA\\ line. Active giant stars possess lower gravities, and lower chromospheric densities than dwarfs, allowing for photoexcitation processes to become important. Moreover, their extended chromospheres allow for scattering of infrared continuum radiation, producing strong absorption in \\ion{He}{i} and tracing wind dynamics.} ", "introduction": "Observations of the solar corona and chromosphere reveal that regions with copious X-ray emission (emitted by bright points or active regions in the corona) also have enhanced \\ion{He}{i} 10830~\\AA\\ absorption that arises in the chromosphere (Fig.~\\ref{fig:solar}). Conversely, chromospheric regions located below solar coronal holes, where X-rays are diminished, show weakened \\ion{He}{i} absorption \\citep{zir75,she81,dup96}. The \\ion{He}{i} 10830~\\AA\\ line, actually a triplet (10829.081~\\AA, 10830.250~\\AA, and 10830.341~\\AA), is a transition between the lower, metastable level (2$^3$S) in the triplet series of \\ion{He}{i} and the 2$^3$P level (Fig.~\\ref{fig:levels}). The metastable 2$^3$S level can be populated only through collisional excitation from the ground level or through recombination and deexcitation from upper levels. Two mechanisms have been proposed for the population of the lower 2$^3$S level. In the photoionization-recombination (PR) mechanism \\citep{gol39}, X-rays and EUV radiation ($\\lambda <$504~\\AA) from the corona photoionize the neutral helium from the ground state; photoionization is followed by recombination, and the electrons cascade to populate the lower levels of \\ion{He}{i}, especially the metastable 2$^3$S level. Scattering of the local infrared continuum produces an absorption line. Models suggest \\citep{and97} the PR process is important in the Sun at temperatures $<$10,000~K. However, the opacity of the chromosphere can limit the efficiency of the photoionization mechanism. In an alternative mechanism, electron collisions from the ground and the metastable 2$^3$S level of \\ion{He}{i} dominate photoionization. In this case, a temperature $\\sim$ 20,000-30,000~K is required, and high densities in the chromosphere and transition region enhance the process. A combination of the two processes can exist as well. Arguments for each of the mechanisms can be found in \\citet{zir82,sim82,smi83,wol84,obr86, zar86,lan95,and95,and97,piet04}, and references therein. If the PR mechanism dominates the formation of the 10830~\\AA\\ transition, a correlation is expected between the strength of the helium line and the radiation field at $\\lambda <$504~\\AA\\ located in the EUV and X-ray bands. Several studies have compared the equivalent widths (EW) of the \\ion{He}{i}~10830~\\AA\\ line and the X-ray flux in late-type stars in order to establish a relation between those parameters. Theses studies show that stronger X-ray emission yields stronger 10830~\\AA\\ absorption in both dwarfs (of spectral type F7 or later) and giants \\citep{zir82,obr86,zar86}. However the RS~CVn active binary systems were generally not included in these analyses. Many of the previous observations made use of photographic plates to measure the 10830~\\AA\\ line, and correlated the line strengths with the X-ray fluxes observed by the Einstein satellite (IPC and HRI instruments) which spanned the energy range 0.1~--~4 keV. In this paper we present measurements of the 10830~\\AA\\ line in very active cool stars and binaries taken with modern instrumentation and high resolution. The use of high spectral resolution is essential in order to discern blends with telluric lines. Additionally, EUV fluxes are obtained as measured directly with the Extreme Ultraviolet Explorer (EUVE). Since the photoionization edge of \\ion{He}{i} is located at 504~\\AA\\ (0.02 keV), the EUV fluxes are expected to relate closely to the photoionization rate of helium. The chromospheric line of \\ion{He}{i} at 10830~\\AA\\ can indicate bulk mass motions in the atmospheres of cool luminous stars. The lower level of the transition is metastable, and a large population can build up in this level. If the gas itself is moving outward in a wind, the helium in the metastable level in the outflow scatters photospheric infrared radiation, and the line profile can reveal the dynamics of the atmosphere. If there is a significant contribution by the photoionization-recombination process, then the level population is independent of the local thermodynamic conditions in the chromosphere. In all cases, detailed modeling demonstrates \\citep[see][]{dup92} that in luminous stars, the Helium atom is formed further out in the atmosphere than other optical diagnostics of mass flow such as H$\\alpha$ and \\ion{Ca}{ii} H \\& K lines, making it an extremely sensitive probe of the acceleration region of a stellar wind. Even in a dwarf star, such as the Sun, models demonstrate that the 10830~\\AA\\ line is formed above the \\ion{Ca}{ii} K core and H$\\alpha$ \\citep{avr98} so that this Helium transition is more favorable for detecting regions where acceleration is likely to occur. Most importantly, observations of the 10830~\\AA\\ line in both Sun and stars demonstrate that high outflow velocities are observed in the line profiles of cool stars \\citep[see e.g.][]{obr86, dup92, dup96, edw03, dup05}. Given the high spectral resolution of our observations, the presence of winds in the stars of the sample can be explored. The cool giants are especially good targets. In Sect. 2 we describe the observations. Sect. 3 compares the \\ion{He}{i} line strengths to other parameters. Results are discussed in Sect. 4; conclusions can be found in Sect. 5. \\begin{figure} \\centering \\hspace{10mm} \\includegraphics[width=0.45\\textwidth]{figure1.eps} \\caption{Spectra of different solar regions: coronal hole, quiet Sun, and an active region \\citep[from][with spectra from an active region courtesy of M. Penn]{dup96}.} \\label{fig:solar} \\end{figure} ", "conclusions": "The \\ion{He}{i} $\\lambda$10830 line responds differently to the EUV radiation field between the dwarf and giant stars in our sample. Active dwarf stars reach a 'saturated' equivalent width in $\\lambda$10830 in the presence of a strong radiation field. This behavior appears consistent with model calculations in which high chromospheric densities allow collisional excitation to dominate photoionization/recombination processes in forming the line. Giant stars, with lower chromospheric densities than dwarfs, show increased \\ion{He}{i} absorption related to an increased EUV radiation field and the absorption is strengthened by an extended expanding atmosphere scattering the line. The \\ion{He}{i} line in giant stars has a photoionization-recombination component that appears to dominate the line-forming process. Detailed radiative transfer calculations would be helpful to assess the contribution of collisions to line formation in the giant stars." }, "0807/0807.4929_arXiv.txt": { "abstract": "By observing the transits of exoplanets, one may determine many fundamental system parameters. I review current techniques and results for the parameters that can be measured with the greatest precision, specifically, the transit times, the planetary mass and radius, and the projected spin-orbit angle. ", "introduction": "Henry Norris Russell~(1948) once delivered a lecture here in Cambridge entitled ``The royal road of eclipses,'' about the determination of accurate parameters for eclipsing binary stars, and the promise that such systems held for progress in stellar astrophysics. Given the rapid progress on display at this meeting, it is clear that exoplanetary science too has its royal road: the royal road of transits. Figure~1 illustrates the happy situation in which the planet's orbit is viewed nearly edge-on, and the planet undergoes transits and occultations. \\begin{figure}[b] \\begin{center} \\includegraphics[width=29pc]{Pedagogical.eps} \\caption{Illustration of transits and occultations. During a transit, the planet blocks a fraction of the starlight. Afterwards, the planet's brighter dayside comes into view and the total flux rises. The total flux drops again when the planet is occulted by the star.} \\label{fig1} \\end{center} \\end{figure} \\begin{table} \\caption{Properties that have been measured, or that might be measured in the future, through precise observations of transiting planets.} \\label{tab1} \\begin{tabular}{lclc}\\hline {\\bf Property} & {\\bf Refs.} & ~~~~~~~{\\bf Property} & {\\bf Refs.} \\\\ \\hline Orbital period & 1,2 & ~~~~~~~Planet-planet interactions (short-term) & 19,20 \\\\ Orbital inclination & 1,2 & ~~~~~~~Planet-planet interactions (long-term) & 21,22 \\\\ Planetary mass & 1,2 & ~~~~~~~Mutual orbital inclinations & 20,23 \\\\ Planetary radius & 1,2 & ~~~~~~~Planetary rings & 24,25 \\\\ Stellar obliquity & 3,4 & ~~~~~~~Satellites & 9,24 \\\\ Orbital eccentricity & 5,6 & ~~~~~~~Relativistic precession & 26,27 \\\\ Stellar limb darkening & 7 & ~~~~~~~Parallax effects & 28, 29 \\\\ Star spots & 8,9 & ~~~~~~~Apsidal motion constant & 30 \\\\ Thermal emission & 5,10 & ~~~~~~~Stellar differential rotation & 31 \\\\ Absorption spectrum & 11,12 & ~~~~~~~Oblateness and obliquity & 32,33 \\\\ Albedo & 13,14 & ~~~~~~~Variations in stellar radius & 34 \\\\ Phase function & 15 & ~~~~~~~Yarkovsky effect & 35 \\\\ Effective radiative time constant & 16 & ~~~~~~~Planetary wind speed & 36 \\\\ Trojan companions & 17,18 & ~~~~~~~Artificial planet-sized objects & 37 \\\\ \\hline \\end{tabular} {\\it Non-exhaustive list of references:} (1) Charbonneau et al.~(2000). (2) Henry et al.~(2000). (3) Queloz et al.~(2000). (4) Winn et al.~(2005). (5) Charbonneau et al.~(2005). (6) Bakos et al.~(2007). (7) Knutson et al.~(2007a). (8) Silva~(2003). (9) Pont et al.~(2007). (10) Deming et al.~(2005). (11) Charbonneau et al.~(2002). (12) Vidal-Madjar et al.~(2003). (13) Rowe et al.~(2006). (14) Winn et al.~(2008a). (15) Knutson et al.~(2007b). (16) Langton \\& Laughlin~(2008). (17) Ford \\& Gaudi~(2006), (18) Madhusudhan \\& Winn (2008). (19) Holman \\& Murray~(2005). (20) Agol et al.~(2005). (21) Miralda-Escud\\'e (2002). (22) Heyl \\& Gladman~(2007). (23) Fabrycky, D., these proceedings. (24) Brown et al.~(2001). (25) Barnes \\& Fortney~(2004). (26) P\\'al \\& Kocsis~(2008). (27) Jordan \\& Bakos (2008). (28) Scharf~(2007). (29) Rafikov~(2008). (30) Ragozzine \\& Wolf~(2008). (31) Gaudi \\& Winn~(2007). (32) Seager \\& Hui~(2002), (33) Barnes \\& Fortney~(2003). (34) Loeb~(2008). (35) Fabrycky~(2008). (36) Spiegel et al.~(2007). (37) Arnold~(2005). \\end{table} Table~1 summarizes the information that has been obtained---or that is obtainable in principle---through precise observations of transits and occultations. This table is surely incomplete. Every few months, a new and creative application of transit observations is proposed. I was asked to discuss some of the measurements that can be made with the highest signal-to-noise ratio. In the best cases, we can measure orbital periods with 8 significant digits; transit times to within a fraction of a minute; the planetary mass and radius to within a few per cent; and the stellar obliquity (or at least its sky projection) to within a few degrees. ", "conclusions": "" }, "0807/0807.1021_arXiv.txt": { "abstract": " ", "introduction": "This is the third paper of a series that studies a relativistic cosmology modelling the relativistic generalization of the single fractal Newtonian model advanced by Pietronero (1987; see also Coleman \\& Pietronero 1992), in which the galactic clustering problem is studied by assuming that the large-scale structure of the universe can be described as being a self-similar fractal system.\\footnote{ \\ The models investigated in this series of papers are in the realm of classical cosmology. No hypotheses concerning inflationary cosmology have been considered.} In Ribeiro (1992a, hereafter paper I) I argued that the recent all sky redshift surveys (de Lapparent, Geller \\& Huchra 1986; Saunders et al. 1991) present observations consistent both with the old Charlier hypothesis of hierarchical clustering and with fractals, where the latter is in essence a more precise conceptualization of the scaling idea implicit in the hierarchical clustering hypothesis. In paper I Pietronero's (1987) basic hypotheses were assumed in order to propose similar ones in a relativistic context, and I obtained observational relations compatible with fractals in Tolman's spacetime and devised a numerical strategy for finding particular Tolman solutions representing a fractal behaviour along the backward null cone. In Ribeiro (1992b, hereafter paper II) I studied analytically the Einstein-de Sitter model in the context of the theory developed in paper I. By treating the Einstein-de Sitter model as a special case of Tolman's spacetime, I found that it does not appear to remain homogeneous along the past null geodesic, has a volume (average) density which vanishes asymptotically and that it also shows no single fractal features along the backward null cone. The apparent inhomogeneity of the Einstein-de Sitter model is explained by the fact that densities measured along the geodesic go through different hypersurfaces of constant $t$, where each one has a different value for the proper density. This paper continues the study of these cosmologies and presents relativistic fractal solutions obtained by following the numerical simulation strategy already devised in paper I. By {\\it fractal solutions} I mean solutions where the fractal system is smoothed-out and its average density follows the de Vaucouleurs density power-law. These solutions represent fractal behaviour along the backward null cone and they were obtained for all three types of Tolman dust models, namely, elliptic, parabolic and hyperbolic. By analysing these solutions we conclude that the only ones with features that may represent real astronomical observations are of hyperbolic type. As we are studying the Tolman spacetime as a region (or regions) possibly surrounded by a Friedmann universe (see paper I), if we adopt the fitting condition approach for interpreting the match between the two spacetimes (Ellis \\& Stoeger 1987; Ellis \\& Jaklitsch 1989) we then conclude that the Friedmann background required is also hyperbolic and we would be living in an open, ever expanding, universe. This paper is organized as follows. In \\S 2 is shown a summary of the observational relations developed in papers I and II plus some minor extensions which will be necessary here, and in \\S 3 I discuss the initial conditions and the algorithm used to find Tolman numerical solutions along the past light cone. In \\S 4 I present the numerical solutions for the spatially homogeneous special cases and \\S 5 shows fractal solutions for all types of Tolman models. \\S 6 discusses these fractal solutions in terms of comparison with the spatially homogeneous cases, fitting condition, evolution of the most realistic fractal model in terms of real observations and relations with homothetic self-similarity. In this section I also express my criticism of criticisms of fractal cosmology. I finish in \\S 7 with the conclusions on this and the preceding papers. ", "conclusions": "\\subsection{A Metric for a Smoothed-out and Averaged Fractal} In the previous section it was shown that Tolman fractal solutions do exist, and that the only ones compatible with observations are the hyperbolic type solutions obtained by means of the specializations given by equations (\\ref{t29}) and (\\ref{t30}).\\footnote{ \\ Notice however that fractal solutions of elliptic and parabolic types may, in principle, be obtained from other more complex specializations of the arbitrary functions than the ones considered in this paper, and these solutions could be compatible with observations.} As in equations (\\ref{t29}) the function $\\beta(r)$ is not constant, this means that the model has no simultaneous big bang. In other words, in a model of this sort the big bang singularity hypersurface occurred at different proper times in different locations, and the age of the universe is different for different observers at different radial coordinates. More specifically, as $\\beta(r)$ in equations (\\ref{t29}) is an increasing function, regions at smaller $r$ are younger than at bigger $r$, and the youngest region of the model is ``here'', at $r=0$. An universe model where some regions are older than others is not as odd in terms of accepted ideas of galaxy formation as it might seem at first. Inasmuch as the observed universe is lumpy, in a spatially homogeneous Friedmann universe where $\\rho = \\rho (t)$ and the big bang singularity is simultaneous, there must be density fluctuations $\\delta \\rho / \\rho$ of some kind in order to form galaxies, and it is necessary to have some sort of metric perturbations for that to happen. So, at the era of galaxy formation, which may be defined as a hypersurface of constant time in order to agree with our intuition in an unperturbed metric, in the perturbed metric the overdensities $\\rho_{\\rm o}$ occur at time $t_{\\rm o}$ and the underdensities $\\rho_{\\rm u}$ occur at $t_{\\rm u}$, and $t_{\\rm o}$ must be different from $t_{\\rm u}$, otherwise there would not be any fluctuation at all. In other words, due to the density fluctuations $\\delta \\rho / \\rho$, in the perturbed metric a hypersurface of constant density no longer coincides with a hypersurface of constant time. Therefore, a deviation of the Friedmann metric from spatial homogeneity, even if it is small, is essential for lumpiness and, hence, some regions will inevitably have different local times than others. In the perturbed metric we may even define the era of galaxy formation as being a hypersurface of constant density. In other words, a non-simultaneous big bang seems inevitable in order to form galaxies in the standard scenario, even if those differences in local times are small. Note that this discussion assumes that a hypersurface of simultaneity is defined by a specific value of the proper time, which is a logical thing to do in an unperturbed metric. In a perturbed metric, however, one could define the hypersurfaces of simultaneity in a different way, which would mean a different choice of gauge by which the perturbed and the non-perturbed spacetimes are related. Nevertheless, considering it is desirable that fractal models be as close as possible to their Friedmann counterparts, and also considering mathematical simplicity, I shall take the specializations given by equations (\\ref{t30}) as the best modelling of a relativistic hierarchical (fractal) cosmology by Tolman's spacetime. Let us now write this model explicitly. Its metric is expressed as \\begin{equation} dS^2=dt^2-\\left( \\frac{{R'}^2}{\\cosh^2 r} \\right) dr^2 - R^2 (d \\theta^2+\\sin^2 \\theta d \\phi^2); \\ \\ \\ r \\ge 0, \\ R[r,t(r)] \\ge 0. \\label{t31} \\end{equation} The Einstein field equation for this metric may be written as an energy equation (Bondi 1947; paper I) \\begin{equation} \\frac{\\dot{R}^2}{2} - U(R) = E(r), \\label{t32} \\end{equation} where \\begin{equation} U(R) = \\frac{\\alpha r^p}{4R} \\label{t33} \\end{equation} is the effective potential energy, and \\begin{equation} E(r) = \\frac{1}{2} \\sinh^2 r \\label{t34} \\end{equation} is the total energy within $r$. The solution of equation (\\ref{t32}) is \\begin{equation} R = \\frac{ \\alpha r^p}{8 E(r)} \\left( \\cosh 2 \\Theta - 1 \\right), \\label{t35} \\end{equation} where $\\Theta$ is given by \\begin{equation} 4 \\left[ t(r) + \\beta_0 \\right] { \\left[ 2 E(r) \\right] }^{3/2} = \\alpha r^p \\left( \\sinh 2 \\Theta - 2 \\Theta \\right), \\label{t36} \\end{equation} $t(r)$ is the solution of the past radial null geodesic \\begin{equation} \\frac{d t}{d r} = - \\frac{R'}{ \\cosh r}, \\label{tt3366} \\end{equation} and the local density is expressed as \\begin{equation} \\rho = \\frac{4 p { \\left[ E(r) \\right] }^2}{ \\pi \\alpha r^{p+1} R' { \\left( \\cosh 2 \\Theta - 1 \\right) }^2 }. \\label{tt3377} \\end{equation} The essentially new physical feature of the model above is its single difference from the open Friedmann one: the form of the function for the gravitational mass. That is given by $F = \\alpha r^p$, while in Friedmann this function must be $F = b_2 \\sinh^3 r$. Remembering that $\\alpha = 10^{-4} - 10^{-5}$ and $p=1 - 2.5$, the fractal metric (\\ref{t31}) appears to have a more rarefied dust than its Friedmann equivalent. As fractal models are characterized by a power law nature of their average densities, with fractional exponents smaller than 3, it is hardly surprising that such models have a more rarefied distribution of mass. \\subsection{Evolution of the Fractal Model and Comparison with the Spatially Homogeneous Case} The constant $\\beta_0$ in equation (\\ref{t36}) gives the universal big bang time if we define ``now'' as $t=0$, and this fact allows us to study the evolution of the fractal model (\\ref{t31}) through the easily computable manner of simply varying $\\beta_0$. This investigation permits us to answer the question of whether or not the fractal features of the metric (\\ref{t31}) are present at different epochs. Bearing this point in mind, I carried out simulations for different values of $\\beta_0$ in the interval $0 \\le r \\le 0.07$, but keeping $\\alpha = 10^{-4}$ and $p=1.4$ in all of them. I found out that the model under consideration remains fractal in integrations where $\\beta_0 =$ 1.5, 2, 2.5, 3.6, 4.5 and 6, with the resulting fractal dimensions being $D=$ 1.5, 1.5, 1.5, 1.4, 1.4 and 1.4, respectively. These results show that the metric (\\ref{t31}) effectively models a fractal distribution of dust at different epochs, with a remarkable constancy in $D$. All those integrations had $r = 0.07$, the final value of the integrating parameter, corresponding to $z \\cong 0.07$, although the luminosity distance varied from $d_l \\cong 110$ Mpc when $\\beta_0 = 1.5$ to $d_l \\cong 450$ Mpc when $\\beta_0 = 6$. This variation may be physically explained as due to changes in the Hubble constant itself, whose values get bigger at earlier epochs. Finally, those results together with some simulations on different values of $p$ suggest it is possible to propose a simple, but very restricted, relationship between $p$ and $D$. For $0 \\le r \\le 0.07$, $1.5 \\le \\beta_0 \\le 6$ and $ 1.4 \\le p \\le 2.5$ we can say approximately that $D = p \\pm 0.1$. An interesting question about the model (\\ref{t31}) is to see how it would compare with the spatially homogeneous open Friedmann one. Looking at the results of the latter in figure 6 and the former in figure 20 we can qualitatively see that the absolute value of the difference in $\\log \\rho_v$ between the two models starts as zero, but increases rapidly. In the analytical study of the Einstein-de Sitter model presented in paper II it was shown that at the big bang singularity hypersurface the volume density $\\rho_v$ vanishes and the luminosity distance $d_l$ goes to infinity, and this effect is a consequence of the definition (\\ref{t10}) of the volume density: at the big bang the volume (\\ref{t9}) is infinity, but the total mass is finite. As already said at the end of \\S 4, we can therefore expect a similar asymptotic effect in both the open Friedmann case and the fractal model (\\ref{t31}), and this means that the difference in $\\rho_v$ between these two models should start decreasing after reaching a maximum in its increase. Actually, figure 6 already shows a sharp decrease of $\\rho_v$ in the open Friedmann case, but as the integrations of the model (\\ref{t31}) in figure 20 did not go far enough in the past, the results cannot show where this maximum might be. Nevertheless, based on this reasoning we can deduce that the observational relations of the fractal model (\\ref{t31}) appear to be asymptotically Friedmann when calculated along the past light cone. \\subsection{Relations to Homothetic Self-Similarity} Another topic which deserves some investigation is the relation, if any, between self-similarity due to fractals and self-similarity due to homothetic Killing vectors (Cahill \\& Taub 1971). There is some interest in this point because some attempts have been made to explain large-scale voids and clusters by self-similar perturbations of a Friedmann universe (Carr \\& Yahil 1990), and also because the first attempt to propose a workable relativistic hierarchical cosmology was done assuming a homothetic self-similar metric (Wesson 1978, 1979). In general relativity a spacetime is called self-similar if all metric components can be put in a form in which they are functions of a single independent dimensionless variable which is a combination of the spacetime coordinates. Mathematically, this corresponds to the existence of homothetic Killing vectors, meaning that spherically symmetric similarity solutions are unchanged by coordinate transformations of the form $t \\rightarrow bt$, $r \\rightarrow br$, for any constant $b$ obeying the conformal transformation $g_{\\mu \\nu}(r,t) \\rightarrow \\frac{1}{b^2} g_{\\mu \\nu}(r,t)$ (Cahill \\& Taub 1971). Physically, spherically symmetric similarity spacetimes with $\\Lambda = 0$ contain no fundamental scales or dimensional constraints (Henriksen \\& Wesson 1978), and it would seem that these features make homothetic self-similarity a possible mathematical version of the scaling idea behind the empirical hierarchical clustering concept. However, homotheties, as defined in general relativitity, are basically geometrical features which will not necessarily translate themselves in observable quantities. In other words, the geometrical scaling features of the model do not necessarily mean that its observational relations are also scaling. For this reason it seems that homothetic self-similarity provides an unsatisfactory manner of modelling hierarchy. It is beyond the aims of this work to make a general discussion about the possible relations between these two types of self-similarity. However it is of interest to investigate whether or not the fractal solutions presented in \\S 5 are homothetic. Recently Lemos \\& Lynden-Bell (1989) and Ponce de Leon (1991) studied homotheties in Tolman models and found that specific criteria are necessary to maintain the assumed similarity symmetry in the solutions. For models where $\\Lambda = 0$ (our case here), they showed that the first criterion is for the function $f(r)$ to be constant, that is, each comoving shell must have the same total energy. That immediately tells us that all fractal solutions of elliptic and hyperbolic type studied here, including the metric (\\ref{t31}), are not homothetic, leaving only the parabolic solutions still to investigate. The next criterion says that homothetic solutions with $f(r) = 1$ restrict the mass distribution to have the form $F(r) = \\mbox{constant} \\times r^{2 \\wp +3}$, where $\\wp$ is a constant, and this is the case in the solutions (\\ref{t24}) and (\\ref{t25}). The final requirement for homothetic symmetry to hold in our Tolman solutions demands that for $f=1$ and $\\wp \\not= 0$ the big bang hypersurface must be of the form $\\beta(r)= a + b r^{-\\wp}$, where $a$ and $b$ are constants. The solution (\\ref{t24}) satisfies these three requirements only if $p=3-2q$. The solution (\\ref{t25}) can have $\\beta(r)$ reduced to its form in equations (\\ref{t24}), but then the value $p=1$ must hold to satisfy the third requirement. Therefore, very restricted cases of the fractal solutions (\\ref{t24}) and (\\ref{t25}) have also homothetic self-similarity. In conclusion, from what we have seen it does appear valid to say that fractal self-similarity is a much weaker requirement on the solutions than homotheties, and although we have reached this conclusion looking in more detail only at the Tolman spacetime, based on the self-similar requirements for both cases it seems reasonable to suppose that this conclusion may well be valid in general. \\subsection{Fitting Condition} In the previous section of this paper we dealt with the problem of finding a specific Tolman model which best represents the observed inhomogeneous distribution of galaxies, and in that respect it was concluded that the metric (\\ref{t31}) is the simplest one to achieve this aim. In other words, what was being sought was the optimal way of fitting the Tolman metric to the real lumpy large-scale structure of the universe. It was discussed in paper I that it may be desirable for us to have a Friedmann metric as background spacetime to the inhomogeneous Tolman region (or regions) used here to describe a fractal distribution of galaxies, and having found the specific forms for this inhomogeneous region an important question arises at once: what are the implications that the fractal metric (\\ref{t31}) and the hyperbolic solution (\\ref{t29}) bring to a possible Friedmann spacetime background? Answering this question is equivalent to finding a response to the ``fitting problem'' in cosmology, in the specific context of this work. Ellis \\& Stoeger (1987) have outlined the fitting problem as being the search for an ideal Friedmann model which best fits another cosmological model that gives a realistic representation of the universe, including all inhomogeneities down to some specified length scale. In Ellis \\& Stoeger's words, ``the approach resembles that used in geodesy, where a perfect sphere is fitted to the pear-shaped earth; deviations of the real earth from the idealised model can then be measured and characterized''. Various ways in which this approach might be tried are discussed in detail by them, however here we shall restrict ourselves to the specific one outlined by Ellis \\& Jaklitsch (1989) where the matching between the Tolman and Friedmann spacetimes is interpreted as a fitting condition. It was shown in paper I that the Darmois junction conditions between the two spacetimes under consideration require that $f=g'$ on the joining surface. Here $g= \\sin r$, $r$, $\\sinh r$ is the function which determines the curvature of the Friedmann model. As both the solution (\\ref{t29}) and the metric (\\ref{t31}) are of hyperbolic type with $f= \\cosh r$, there is no way of satisfying this condition when $r > 0$ unless we have $g = \\sinh r$. Therefore, the first response to the fitting problem in this context says that our Tolman fractal solutions imply an open Friedmann background model. It was also shown in paper I that the matching between these two spacetimes severely restricts the gravitational mass inside the Tolman cavity. If $m(r) = F(r)/4$ is the gravitational mass of the Tolman region within a comoving radius $r$, and $\\overline{m}(x) = 4 \\pi \\mu a^3(t) g^3(x)/3$ is its Friedmann equivalent for a radius $x$ and dust density $\\mu$, the junction conditions demand \\begin{equation} m ( \\Sigma_0 ) = \\overline{m} ( \\Sigma_0 ), \\label{ultimaIII} \\end{equation} where $\\Sigma_0$ is the constant that defines the joining surface $r=x=\\Sigma_0$ between the two spacetimes. Therefore, as discussed by Ellis \\& Jaklitsch (1989), equation (\\ref{ultimaIII}) allows us to choose the Friedmann background model whose density is appropriate to our lumpy Tolman model. Let us see in more detail how this background spacetime can be specified. In the open Friedmann model the local density is given by \\[ \\mu = \\frac{3 b_2}{16 \\pi a^3(t)}, \\] and as $F = \\alpha r^p$ in our fractal models, we can then write the equation (\\ref{ultimaIII}) as \\begin{equation} \\alpha {\\Sigma_0}^p = b_2 \\sinh ^3 \\Sigma_0. \\label{ultimaIV} \\end{equation} The value of the parameter $b_2$ is what we are seeking in order to determine precisely the Friedmann background and give a more accurate answer to the fitting problem in this context. Equation (\\ref{ultimaIV}) shows that $b_2$ is dependent on the other three parameters $\\alpha$, $p$, $\\Sigma_0$, and hence, there is a certain degree of flexibility in choosing the mass of the background spacetime. Thus, even when the interior region is determined by known values of $\\alpha$ and $p$, different values of $b_2$ are obtained according to exactly where the joining surface is located. We can work out how the Friedmann background is in the case of the numerical integrations of the model (\\ref{t30}) shown in figure 20. We have $\\alpha = 10^{-4}$, $p=1.4$, and if we take $\\Sigma_0 = 0.07$ (the value where the numerical evaluation ends) we get $b_2 = 0.007$, $\\Omega_0 \\cong 0.002$ and $H_0 = 83$ km/s/Mpc. This low value for $\\Omega_0$ is in the lower limits of the interval where it has been reportedly measured. However, it is important to notice that in the approach used in this work no kind of dark matter was considered, but only the luminous matter associated with galaxies. Galactic luminous matter gives a value for $\\Omega_0$ of the same order of magnitude as the one found for the background model above (see White 1990, p. 38). As a final remark, we have so far considered the interior Tolman region joining directly to the exterior Friedmann metric. That does not need to be always the case and we can envisage an interior region surrounded by one or more intermediary regions with higher or lower densities, in a scheme designed to model specific observational features. For example, a structure like the ``Great Wall'' (Geller \\& Huchra 1989; Ramella, Geller \\& Huchra 1992) could be modelled by an intermediary overdensity region before the background spacetime is reached, and with an underdensity interior fractal Tolman region (see Bonnor \\& Chamorro 1991 on how to join an underdensity Tolman region to an overdensity intermediary section). Such a modelling will obviously increase the value of $\\Omega_0$ for the fitted Friedmann background. This method, however, demands more detailed work in order to achieve a model where this sort of structure is precisely characterized. I shall not pursue further this study here. \\subsection{Criticism of Criticisms of Fractal Cosmology} As the final issue of this section, I shall discuss some of the objections raised by some authors against a fractal cosmology. The first type of criticism is contrary to a possible unlimited fractal pattern for the large-scale clustering of galaxies, and although some of the critical voices do accept fractals at small scales, their objections are usually based either on reasoning from the 2-point angular correlation function (Peebles 1989), or on supposed strong theoretical limitations of the standard Friedmannian cosmology. Therefore, in one way or another those authors see the strong need for a crossover to homogeneity on the fractal structure, at a scale yet to be agreed upon. Criticisms based on the angular correlation function have been addressed by Coleman \\& Pietronero (1992) and I shall not discuss them here, although this point was briefly mentioned in paper II. The theoretical criticisms, on the other hand, must be addressed in this paper, and for this purpose I shall reproduce here two quotations from Mart\\'{\\i}nez (1991) which well represent this point of view, although he is by no means the only one to raise such kind of objections. Mart\\'{\\i}nez states that ``...~it should be noted that in the standard cosmology, the distribution of mass must tend to a non-zero finite density when averaged over large volumes'' \\footnote{ \\ The expression ``large volumes'' used in this quotation is imprecise, and can be interpreted as meaning either big local volumes or volumes which are big enough to be no longer considered as local. In a cosmological context the latter is more appropriate and from now on I shall assume the expression ``large volumes'' to mean non-local ones.}. From a relativistic point of view, the problem with this statement is its failure to specify where this average is supposed to be carried out. It is correct to say that if in a Friedmannian cosmology we make averages of density at spacelike hypersurfaces of constant $t$, those averages cannot be zero as this cosmology is spatially homogeneous. However, at large volumes such averages would be observationally irrelevant as astronomy in the electromagnetic spectrum is actually made along the backward null cone and {\\it not} at such spacelike hypersurfaces. Therefore, the statement above is only true in an unobservable situation, and considering that voids and clusters of galaxies were, and still are being, identified in the so-called redshift space, which lies on the past null cone, this is where the average of density must be carried out. Hence, the quotation above is inappropriate as an objection to a fractal structure for the distribution of galaxies. These two types of averages will only coincide locally, and it will depend on the model and the value assumed for the Hubble constant in order to establish what scales are local, although, in any case they will certainly differ at large volumes. In effect, in this imprecise formulation this statement may actually reinforce, or be taken by, a common misinterpretation of the standard model, which, due to inappropriate Newtonian analogies, confuses the model's geometry with its observable quantities . Mart\\'{\\i}nez (1991) goes on and states that ``... a fractal universe without a crossover to homogeneity (...) implies a vanishing density for very large volumes and this idea cannot be accepted without creating important additional problems''. It was shown in paper II (and rediscussed in \\S 4) that if averages on density are made along the backward null cone, at the big bang singularity hypersurface the luminosity distance goes infinite and this average density is zero. That happens in the Einstein-de Sitter model, the most popular of the cosmological models, and no additional hypothesis or change in the metric was done to achieve this result. The point is, once more, where the average is made and which definition of density is adopted. Thus, again the statement is in fact an untenable objection since the standard cosmology does have a vanishing average density without any important additional problem. Having or not having zero average density in the model is just a question of interpretation. It should be clearly understood that the above criticisms of Mart\\'{\\i}nez's (1991) statements are made solely on the grounds of the standard Friedmannian cosmology, and there is no need whatsoever to mention any fractal hypothesis in order to show the impreciseness and inappropriateness of such statements. The important point being that even accepting the spatially homogeneous standard Friedmannian cosmology, this model tells us we would only be able to see, through our telescopes, its homogeneity locally. Closely related to this point of local homogeneity is the issue, one could argue, of how we would understand in this context the reported uniform distribution of some deep samples like radio sources. In the first place it must be made very clear that the discussion made so far is aimed at showing that from a theoretical point of view there is no constraint to an unlimited fractal distribution, even from within a Friedmannian framework, but that does not imply the fractal system is indeed limitless, as an upper cutoff to homogeneity is not yet ruled out. \\footnote{ \\ The proposal for an upper cutoff to homogeneity in the fractal system appears to have been initially advanced by Ruffini, Song \\& Taraglio (1988), although Pietronero (1987) had already made a discussion about this issue.} Nonetheless, a simple calculation in the Einstein-de Sitter model, as presented in paper II, will show that a 30\\% decrease in $\\rho_v$ (from the value at present time $t=0$) occurs at $z~\\approx~0.1$ ($d_l~\\approx~500$ Mpc), and this means that even considering such high error in the determination of the volume density, this is, roughly speaking, the maximum approximate range where the homogeneity of this model could be observed. Beyond this range the homogeneity of the Einstein-de Sitter model would no longer be observed in the past light cone. Thus the first issue raised by this result is a problem of methodology: curvature effects occur in Friedmann models at much closer ranges than usually assumed, and this means that those surveys must consider in their data analysis expressions along the past light cone. Currently, calculations of observational relations where the backward null cone is taken into account is a very much neglected problem in cosmology. Secondly, if it is confirmed that in those deep surveys the distribution is really uniform, that would put the Friedmann model in even greater difficulties as it would appear to predict inhomogeneity in deeper ranges where this would not be observed. Thirdly, the sceptical viewpoint on this issue would be to argue that usually deeper observations are less precise than shallower ones, and previous claims of the so-called homogeneous ``fair-sample'' finally being observed did not stand once more refined and complete observations were made. Historically, the range at where the homogeneity is, or would be, finally reached has being pushed further and further away as more complete data become available and observational techniques improve, and so, the sceptic may say, we may not have necessarily seen the end of this story. In addition to the points discussed above, a second kind of criticism to the fractal cosmology has been voiced by Peebles, Schramm, Turner \\& Kron (1991) in the following form. ``If the galaxy distribution had been observed to follow a pure scale-invariant fractal, (...) the closely thermal spectrum and isotropy of the cosmic background radiation in this highly inhomogeneous Universe would have been a deep puzzle''. First of all, it should be said that in an inhomogeneous model with a Friedmann background, the apparent discrepancy between the inhomogeneity of the model and the isotropy of the microwave background is not really an issue as the junction conditions already require that an overdensity must be compensated by an underdensity before the uniform region is reached, in order that the average densities will be the same (see paper~I, \\S 4). Nevertheless, the most important point is the result already obtained in paper II and extended in this paper for the other Friedmann models: the standard Friedmannian cosmology may be taken to be inhomogeneous depending on how we look at it. That means that the apparent contradiction between inhomogeneity and the isotropy of the microwave background may not be a contradiction at all. This is an essential point in order to put Peebles, Schramm, Turner \\& Kron's statement above into perspective, as we have already seen in this paper and in paper~II that at relatively modest luminosity distances and redshifts, even the standard spatially homogeneous Friedmannian cosmology becomes highly inhomogeneous on the past light cone because $\\rho_v$ departs considerably from its constant value in our constant time hypersurface the further into the past we look, and $\\rho$ is dependent on $r$. Considering that so far even the deepest all-sky redshift surveys have failed to reach the so-called ``fair sample'' where the homogeneity is supposed to be, waiting for us to discover it, perhaps it is about time to ask if the cosmic background radiation could be accommodated in a different cosmology, or wonder whether it is really a deep puzzle. After all this discussion, we are left with an important point to consider. If the supposed theoretical need for a crossover to homogeneity in the fractal system is much weaker than previously thought, we have the question: is it really necessary? Since we have seen here and in paper II that even the Friedmann models do not seem to remain homogeneous along the past null geodesic (see figures 1, 2, 6 and 7), if the homogeneity of the standard model does not survive, where is the strong need for a crossover? In paper I it was assumed that the Tolman metric would eventually join a Friedmann background and, among other things, I argued the need for that was to make the model compatible with a different interpretation of the Copernican principle. However, in the light of inhomogeneity even in the Friedmann metric, it could be argued that from an observational point of view, that is, in calculations along our past null cone, if the relativistic fractal cosmology developed in this series of papers has or has not a Friedmann background might well be irrelevant." }, "0807/0807.1956_arXiv.txt": { "abstract": "Recent observational and theoretical studies on the three-dimensional (3D) space motions of the Large and the Small Magellanic Clouds (LMC and SMC, respectively) have strongly suggested that the latest proper motion measurements of the Magellanic Clouds (MCs) are consistent with their orbital evolution models in which the MCs have arrived in the Galaxy quite recently for the first time. The suggested orbital models appear to be seriously inconsistent with the tidal interaction models in which the Magellanic Stream (MS) can be formed as a result of the mutual tidal interaction between the MCs and the Galaxy for the last $\\sim 2$ Gyr. Based on orbital models of the MCs, we propose that if the MCs have a common diffuse dark halo with the mass larger than $\\sim 2 \\times 10^{10} M_{\\odot}$, the MCs can not only have the present 3D velocities consistent with the latest proper motion measurements but also interact strongly with each other and with the Galaxy for the last 2 Gyr. These results imply that if the observed proper motions of the MCs are true ones of the centers of mass for the MCs, the common halo of the MCs would need to be considered in constructing self-consistent MS formation models. We discuss whether the origin of the possible common halo can be closely associated either with the past binary formation or with the MCs having been in a small group. ", "introduction": "Recent proper motion measurements of the MCs by the Advanced Camera for Surveys (ACS) on the {\\it Hubble Space Telescope (HST)} have reported that the LMC and the SMC have significantly high Galactic tangential velocities ($367 \\pm 18$ km s$^{-1}$ and $301 \\pm 52$ km s$^{-1}$, respectively) and thus suggested that the MCs could be unbound from each other (Kallivayalil et al. 2006; K06). Piatek et al. (2008) have independently analyzed the same data sets as those used by K06 and confirmed the proper motions of the MCs derived from K06. Besla et al. (2007, B07) extensively investigated the long-term orbital evolution of the MCs by using the results of K06 and thereby suggested that the MCs have recently arrived in the Galaxy for the first time (``the first passage scenario''). These observational and theoretical results on the 3D space motions of the MCs appears to be seriously inconsistent with the tidal interaction models in which the MS can be formed as a result of the strong tidal interaction between the MCs and the Galaxy (e.g., Gardiner \\& Noguchi 1996, GN): the MCs are required not only to keep their binary status but also to interact strongly with the Galaxy at least for the last $\\sim 2$ Gyr in the models. Recently Bekki \\& Chiba (2008) have shown that the tidal interaction models consistent with the results by K06 can not reproduce well the observed location of the MS on the sky, though they did not investigate all possible orbits consistent with K06. Given that the tidal interaction models of the MS formation can explain not only the fundamental properties of the MS but also the presence of the leading arm features (GN), it is worth while to discuss whether the tidal interaction models consistent with K06 can be constructed by considering some new physical processes that have not been so far included in the previous models of the MC evolution. The purpose of this Letter is to propose that if the MCs have a common diffuse dark halo, the halo can play an important role in the long-term orbital evolution of the MCs. Based on orbital models of the MCs with and without common halos, we investigate whether the MCs can keep their binary status at least for the last 2 Gyr for their present 3D velocities consistent with K06. We demonstrate that some models with common halos can keep the binary status of the MCs and thus show strong tidal interaction between the LMC and the SMC and between MCs and the Galaxy for the last 2 Gyr. This work is inspired by recent numerical simulations which have clearly shown the concurrent accretion of multiple satellite systems onto galaxy-scale halos in a hierarchical galaxy formation scenario (e.g., Sales et al. 2007; Li \\& Helmi 2008; Ludlow et al. 2008). ", "conclusions": "The present study has first shown that if the MCs have a common halo, they can not only have the present 3D velocities consistent with K06 but also keep their binary status within the last more than 2 Gyr in some models. It should be however stressed that the common diffuse halo needs to have (i) the present velocity ($|v|$) significantly smaller than that of the LMC and (ii) the mass larger than $\\sim 2 \\times 10^{10}$ in order for the MCs to keep their binary status for the last $\\sim 2$ Gyr. These two requirements would be very hard to be confirmed directly by observations: differences in velocities between the center of the common halo and those of the MCs and the mass of the halo can be inferred from numerical simulations on the binary formation of the MCs. Then, how could they have formed a common halo in the histories of the MCs ? We here suggest the following two scenarios for the common halo formation. The first is that the MCs might have dynamically coupled recently ($<4$ Gyr) to form a common halo: dynamical relaxation processes of the two pre-existing halos of the MCs during binary galaxy formation can be responsible for the common halo formation. The orbital evolution models including dynamical friction between the MCs by BC05 showed that the MCs could become dynamically coupled for the first time about $3-4$ Gyr ago. Recent cosmological N-body simulations of the pair galaxy formation based on a $\\Lambda$ cold dark matter ($\\Lambda$CDM) cosmology have shown that the pair formation like the MCs can occur at $z<0.33$ corresponding to less than 3.7 Gyr ago for a canonical set of cosmological parameters (Ishiyama et al. 2008). These results imply that the common halo formation of the MCs might have happened recently ($<4$ Gyr ago). We suggest that the required larger mass of the common halo (i.e., $M_{\\rm ch} \\ge 2\\times 10^{10} M_{\\odot}$) for the binary statue of the MCs in the last $\\sim 2$ Gyr can come from the stripped dark matter halos of the MCs at the epoch of their binary formation: their original masses are significantly larger than the present ones. The second scenario is that a small group of galaxies including the MCs in its central region fell onto the outer region of the Galaxy and then lost most of the halo and the group member galaxies via tidal stripping by the Galaxy: the MC system with the common halo is the remnant of a destroyed group. Li \\& Helmi (2008) have investigated merging histories of subhalos in a Milky Way-like halo using high-resolution simulations based on a $\\Lambda$CDM model and thereby demonstrated that about one-third of the subhalos have been accreted in groups (see also Sales et al. 2007 and Ludlow et al. 2008). The demonstrated higher incidence of the group infall appears to suggest that the MCs can originate from a group thus that the second scenario is also viable. It is currently unclear which of the two scenarios are more consistent with other observations. If the MCs really have a common halo, then the common halo would have the following possible dynamical effects on the Galaxy and the LMC. Firstly, the MC system embedded in the common halo, which is more massive than the LMC, can more strongly influence the outer part of the Galaxy than the LMC alone is demonstrated to be able to do (e.g., Tsuchiya 2002) so that the observed HI warp of the Galaxy (e.g., Diplas \\& Savage 1991) can be more naturally explained in terms of the common halo scenario. Secondly, the common halo can weakly influence the disk of the LMC so that the combined tidal effect of the Galaxy, the SMC, and the common halo could form a off-center bar that is more pronounced than the simulated one in the last LMC-SMC interaction about 0.2 Gyr ago (Bekki \\& Chiba 2007). Thirdly, the common halo enables the LMC and the SMC to have their stellar halos extended much beyond their optical radii. The present study suggests that it would be difficult to observationally determine the 3D velocities of the MC system solely from proper motion measurements of the MCs in the common halo scenario. Furthermore, it would be even more difficult for theoretical and numerical works to predict precisely the long-term orbital evolution of the MCs with a common halo owing to additional two (or more) parameters for physical properties of the halo. If the observed proper motions of the MCs (K06) are really true ones of {\\it the centers of mass for the LMC and the SMC}, some new physical processes need to be incorporated into the tidal models of the MS formation for self-consistency. The hypothesized common halo of the MCs in the present study would be just one of possible physical ingredients that need to be considered in constructing a more self-consistent model of the MS formation." }, "0807/0807.0741_arXiv.txt": { "abstract": "{Shell galaxies are considered the debris of recent accretion/merging episodes. Their high frequency in low density environments suggests that such episodes could drive the secular evolution for at least some fraction of the early-type galaxy population.} {We present XMM-Newton X-ray observations of two shell galaxies, NGC~7070A and ESO~2400100, and far UV observations obtained with the Optical Monitor for these and for an additional shell galaxy, NGC 474, for which we also have near and far UV data from GALEX. We aim at gaining insight on the overall evolution traced by their star formation history and by their hot gas content.} {The X-ray and the far UV data are used to derive their X-ray spatial and spectral characteristics and their UV luminosity profiles. We use models developed ad hoc to investigate the age of the last episode of star formation from the (UV - optical) colors and line strength indices. } {The X-ray spatial and spectral analysis show significant differences in the two objects. A low luminosity nuclear source is the dominant component in NGC~7070A (log L$_X$=41.7 erg~s$^{-1}$ in the 2-10 keV band). In ESO~2400100, the X-ray emission is due to a low temperature plasma with a contribution from the collective emission of individual sources. In the Optical Monitor image ESO~2400100 shows a double nucleus, one bluer than the other. This probably results from a very recent star formation event in the northern nuclear region. The extension of the UV emission is consistent with the optical extent for all galaxies, at different degrees of significance in different filters. The presence of the double nucleus, corroborated by the (UV - optical) colors and line strength indices analysis, suggests that ESO~2400100 is accreting a faint companion. We explore the evolution of the X-ray luminosity during accretion processes with time. We discuss the link between the presence of gas and age, since gas is detected either before coalescence or several Gyr ($>3$) after. } {} ", "introduction": "In $\\Lambda$CDM cosmology, galaxies are assembled hierarchically over an extended period by mergers of smaller systems. In this framework, early-type galaxies showing {\\it fine structure} occupy a special position since they are a testimony of the effects of past merging/accretion events, and as such they fill the gap between on-going mergers and the relaxed elliptical galaxy population. Among examples of {\\it fine structure}, shells are faint, sharp-edged stellar features \\citep{Malin83} that characterize a significant fraction ($\\approx$ 16.5\\%) of the field early--type galaxy population \\citep{Malin83, Schweizer92,Reduzzi96,Colbert01}. Different scenarios for the origin of shells emerge from the rich harvest of simulations performed since their discovery in the early 80's, mostly involving galaxy-galaxy interactions. These range from merging/accretion between galaxies of different morphological type or masses \\citep[mass ratios typically 1/10 - 1/100, ][]{Quinn, Dupraz86, Hernquist87a,Hernquist87b} to significantly weaker interaction events \\citep{Thomson90,Thomson91}. A few models that invoke gas ejection due to the central AGN or the power of supernovae \\citep{Fabian80, Williams85} do not associate the shell formation with environment. Most models can reproduce qualitatively basic characteristics such as spatial distribution, frequency and shape of observed shell systems \\citep[see e.g.][and references therein]{Wilkinson00,Pierfederici04,Sikkema07}. In the accretion models, the more credited scenario for their formation, shells are density waves formed by infall of stars from a companion. A major merger may also produce shells \\citep{Barnes92,Hernquist92,Hernquist95}. The fact that shells are frequently found in the field, rather then in the cluster environment, finds a direct explanation in the week interaction hypothesis, since the galaxy-galaxy ``harassment'' within the cluster tends to destroy shells, while poorer environments are much better suited, because the group velocity dispersion is of the order of the internal velocity of the member galaxies \\citep[see e.g.][]{Aarseth, Barnes85, Merritt}. The class of the so-called ``internal models'' for shell formation suggests that star formation within a giant shell is the result of shocked interstellar gas. In such a case shells are expected to be bluer than the parent galaxy, up to $\\approx$0.5 mag in U-B and B-V \\citep[see e.g.][]{Fabian80}. In a few cases this has been observed, \\citep[see e.g.][and reference therein]{Sikkema07} but attributed to multiple accretion events of galaxies of different intrinsic color. In the framework of a hierarchical cosmology, all galaxies, including early-type galaxies, are expected to contain multi-epoch stellar populations. Shell galaxies, among the early-type class of galaxies, are the ideal candidate to contain also a young stellar population since some simulations indicate a shell dynamical age of 0.5--2 Gyr \\citep{Hernquist87a,Nulsen89}. The star formation history of shell early-type galaxies has been analyzed by \\citet{Longhetti00} using line--strength indices. They show that shell-galaxies encompass the whole range of ages inferred from the H$\\beta$ vs. MgFe plane, indicating that among them recent and old interaction/acquisition events are equally probable. If shells are formed at the same time at which the ``rejuvenating'' event took place, shells ought to be long--lasting phenomena. Recently, \\citet{Rampazzo07} combining {\\it GALEX} far UV data and line--strength indices show that the peculiar position of some shell galaxies in the (FUV-NUV) vs. H$\\beta$ plane could be explained in terms of a recent (1-2 Gyr old) rejuvenation episode. A rejuvenation of stellar population requires the presence of fresh gas during the accretion event. Therefore, a study of the cold, warm and hot gas phases is important in order to consider many of the elements relevant for the evaluation of shell galaxy evolution. \\citet{Rampazzo03} found that the warm ionized (H$\\alpha$) gas and stars appear often decoupled suggesting an external acquisition of the gas, as predicted by merging models \\citep{Weil93}. At the same time, a set of observations showing a clear association between cold (\\hi/CO) gas and stars challenge present merging models which do not predict it unless cold gas behaves differently from the ionized gas \\citep{Schiminovich94,Schiminovich95,Charmandaris00,Balcells01}. The bulk of the interstellar medium (ISM) in early-type galaxies emits in X-rays, and only comparatively small quantities are detected in the warm and cold phases of the ISM \\citep{Bregman92}. The hot ISM is believed to build up primarily through stellar mass loss ($\\approx$ 1 M$_\\odot$/yr in an old passively evolving elliptical galaxy with M$_B=-22$~mag) and Type Ia supernova ejecta \\citep[$\\approx$0.03 M$_\\odot$/yr in a galaxy with M$_B=-22$~mag; see details in][]{Greggio05}. The two processes produce an almost identical mass in metals although the SNIa ejecta are dominated by Fe. \\citet{HB06}, using proper spectral fits to excellent Chandra data, were able to better determine the metal abundances of the X-ray emitting gas for a sample of early-type galaxies. By estimating stellar abundances from optical line strength indices, adopting simple stellar population models, they showed that the hot ISM and the stars have similar abundances. The link between the stellar and ISM metallicities could be masked/contaminated by the accretion of a gas-rich system. At the same time, the content of hot X-ray emitting gas could be correlated with the ``age'' of the rejuvenation episode. Early-type galaxies with {\\it fine structure} (e.g. shells), which are considered {\\it bona fide} signatures of the ``dynamical youth'' of the galaxy, tend to be less X-ray luminous than more relaxed, ``mature'' ellipticals with little evidence of {\\it fine structure}. These are also characterized by extended, and generally stronger, X-ray emission \\citep{Sansom00}. Simulations suggest that mergers \\citep[see e.g.][]{Cox06} or interactions \\citep[see e.g.][]{dercole00} among galaxies produce indeed a wide range of X-ray luminosities. One extreme example in this picture is NGC 474, which shows a well developed shell system and has an X-ray luminosity consistent with the low end of the expected emission from discrete sources \\citep{Rampazzo06}. The X-ray domain may further disclose an otherwise hidden AGN activity, that could also be a result of the merging episodes. Again an extreme example of nuclear activity in a merger remnant can be found in the newly discovered spectacular shell system in the elliptical host of the QSO MC2 1635+119 by \\citet{Canalizo07}. So, whereas accretion/merging events are widely believed to be at the origin of shell galaxies, all the details such as: the age of the event and duration of the shell structure, the global secular evolution of the stellar and gas components of the host galaxy as well as the timing for triggering the AGN, its duty cycle and feedback, are far from being firmly established. In light of the high fraction of shell galaxies in the field, interaction/accretion/merging events seem to have played a significant role in the evolution of the early-type class as a whole. At the same time, there is still the general open question of whether a link exists between shell galaxies and the early phases of merging processes (ULIRGs, AGN, E+A galaxies etc.) on one side and the general class of ``normal'' early-type galaxies on the other. In the above framework, we discuss the X-ray (XMM-{\\it Newton}) observations of two shell galaxies, NGC~7070A and ESO~2400100, taken from the \\citet{Malin83} compilation. We further present far UV XMM-{\\it Newton} Optical Monitor (OM) observations of these objects, and we add {\\it GALEX} far UV observations of NGC~474, for which the results from XMM-Newton X-ray observations have been already presented in \\citet{Rampazzo06}. Through far UV photometry we aim at inferring whether these galaxies have ongoing/recent star formation activity and its distribution across the galaxy. We finally aim to correlate the above information with those extracted from their hot gas content and properties in light of our current understanding of these components. The plan of the paper is the following. Section~2 describes the relevant properties of our sample gathered from the literature. Section~3 presents the X-ray and Far UV observations and data reduction. Results are presented in Section~4 and discussed in Section~5. H$_0$=75 km~s$^{-1}$ is used throughout the paper. \\begin{table*} \\caption{Relevant photometric, structural and kinematic properties} \\begin{tabular}{llllc} \\hline\\hline & NGC~474 & NGC~7070A &ESO 2400100 & Ref. \\\\ \\hline Morphol. Type & (R')SA(s)0 &(R')SB(l)0/a & SAB0: pec & [1] \\\\ Mean Hel. Sys. Vel. [km~s$^{-1}$] &2366$\\pm$16 &2391$\\pm$18 &3184$\\pm$14 & [2] \\\\ Adopted distance [Mpc] & 32.5 & 31.9 & 42.4 & [3]\\\\ $\\rho_{(x,y,z)}$ [gal Mpc$^{-3}$] & 0.19 & & & [3] \\\\ Environment & Cetus-Aries & & & [3] \\\\ cloud & 52-12 & & & [3]\\\\ & & & &\\\\ {\\bf Apparent magnitude,} & & & &\\\\ {\\bf colours, indices}: & & & &\\\\ B$_T$ & 12.36$\\pm$0.16& 13.35$\\pm$0.15&12.63$\\pm$0.24 & [2] \\\\ K$_T$ (2MASS) & 8.555$\\pm$0.039 &9.130$\\pm$0.026 & 8.698$\\pm$0.028 & [1] \\\\ 6 cm [mJy] & & 1.3 & & [5] \\\\ S$_{60\\mu m}$ (IRAS) [Jy] & & 0.26$\\pm$0.040 & & [1]\\\\ S$_{100\\mu m}$ (IRAS) [Jy] & &0.75$\\pm$0.077 & & [1]\\\\ Mg2 & & & 0.28/0.21 & [6]\\\\ H$\\beta$ & & & 1.54/2.79 & [6]\\\\ $\\Delta4000$ & & & 2.31/2.01 & [6]\\\\ H+K(CaII) & & & 1.18/1.24 & [6]\\\\ H$\\delta$/FeI & & & 0.97/0.88 & [6]\\\\ & & & & \\\\ {\\bf Galaxy structure}: & & & & \\\\ Effective Surf. Bright. $\\mu_e$(B) &21.99$\\pm$0.31&22.48$\\pm$0.33 &21.84$\\pm$0.33 &[2]\\\\ Diam. Eff. Apert., A$_e$ [\\arcsec] & 64.6 & & &[2] \\\\ Average ellipticity & 0.21 &0.28 & 0.46 & [2]\\\\ P.A. [deg] & 75 &6.5 & 132.5& [2]\\\\ Fine structure ($\\Sigma$) & 5.26 & & & [4]\\\\ & & & & \\\\ {\\bf Kinematic parameters} & & & & \\\\ Vel.disp. $\\sigma_0$ stars [km~s$^{-1}$] &163.9$\\pm$5.1& 101.0$\\pm$20.2 & 225/223&[1,6] \\\\ App.Max. rotation V$_{max}$ star [km~s$^{-1}$] & 30$\\pm$6 &0.0$\\pm$0.0 & & [2] \\\\ App.Max. rotation V$_{max}$ gas [km~s$^{-1}$] &158.4$\\pm$9.3 & & & [2] \\\\ & & & & \\\\ \\hline \\end{tabular} \\label{table1} \\medskip References: [1] {\\tt NED http://nedwww.ipac.caltech.edu/}; [2] {\\tt HYPERLEDA http://leda.univ-lyon1.fr/}; [3] \\citet{Tully88} (H$_0$=75 km~s$^{-1}$~Mpc$^{-1}$); [4] \\citet{Sansom00}; [5] \\citet{Sadler89}; [6] \\citep{Longhetti99,Longhetti00} data refer respectively to the $a/b$ nuclei embedded in the envelope of ESO~2400100. \\end{table*} ", "conclusions": "We have studied the characteristics of the X-ray emission of two shell systems, NGC~7070A and ESO~2400100 using XMM-Newton. We also analyzed their far UV emission using the XMM-Newton Optical Monitor. We include another shell galaxy NGC~474, for which we have already presented the X-ray characteristics in \\citet{Rampazzo06}, exploiting both original XMM-Newton OM images and {\\it GALEX} archival data. The XMM-Newton spatial and spectral analysis suggest that a nuclear source is the dominant component in NGC 7070A. In ESO~2400100, the emission is due to a low temperature plasma with a contribution from the collective emission of individual sources in the galaxy. {\\it GALEX} data of NGC~474 show that the extension of NUV emission is comparable with that of the optical image, while the FUV emission shows up only in the central regions of the galaxy. Also in the UVW1 and UVM2 filters NGC~474, ESO~2400100 and NGC~7070A have extensions similar to that of the optical image. XMM-OM UV images of ESO~2400100 show the presence of a double nucleus. The shape of the luminosity profiles suggest that a disk component is present, confirming the morphological classification provided by {\\tt NED} and the suggestion, from the kinematical study of \\citet{Sharples83}, that NGC 7070A is an S0 seen face-on. This study further suggests that the prominent dust lane is not yet in equilibrium, indicating a recent accretion episode. We model line--strength indices and the far UV - optical colors of the galaxies to infer the time elapsed from the last significant episode of star formation. From the comparison between our galaxies and shell galaxies in the \\citet{Rampazzo07} sample we suggest that the (UV - optical) colors of NGC~474 and NGC~7070A are consistent with a recent burst of star formation. We argue that the double nucleus in ESO~2400100 is indicative of an ongoing accretion event. The combined UV and line-strength indices analysis suggests indeed a very recent star formation episode in the northern nucleus \\citep[see also simulations by ][]{Kojima97}. Using the above estimates of the time elapsed from the last significant episode of star formation we investigate whether the evolutionary scheme for gas-rich systems that would lead to mature ellipticals proposed by \\citet{Brassington07} could be applied to the same early-type galaxies undergoing merging episodes. The time range is very poorly sampled, with only 4 objects, but we span almost the full range, since \\eso\\ should be at $Time \\sim -200$ Myr, i.e. before coalescence. We notice that shell galaxies are systematically underluminous relative to gas-rich systems at similar evolutionary stages, and that the $L_x/L_B$ or $L_x/L_K$ values vary only a factor of a few, compared to factors of 100. The only indication of a possible difference is in the gas content, which is present either before or several Gyr after coalescence. This is consistent with a picture in which several gigayears are required to refurbish a galaxy of a hot gaseous halos after a merging responsible for its depletion \\citep{Sansom00, Osul01}. To fully understand whether shell galaxies are the precursors of relaxed ellipticals, their more mature counterparts according to hierarchically evolutionary scenarios, we need to better define the dynamical and photometric time-scales of the accretion/merging event and understand whether there is an evolution of their X-ray properties linked to their ``age\" based on a larger and better studied sample of objects. The study of the UV emission in early-type galaxies in connection with optical line-strength indices \\citep[see e.g.][]{Rampazzo07} could provide useful insight for timing their photometric evolution." }, "0807/0807.0431_arXiv.txt": { "abstract": "A comprehensive statistical analysis of the broadband properties of EGRET blazars is presented. This analysis includes sources identified as blazars in the Sowards-Emmerd publications. Using this sample of 122 sources, we find that there is a relationship $L_\\gamma \\propto {L_r}^{0.77 \\pm 0.03} $ as well as a correlation between $\\alpha_{og}$ and $\\alpha_{ro}$, and a correlation between radio luminosity and $\\alpha_{og}$. Through the use of Monte Carlo simulations, we can replicate the observed luminosity relationship if a synchrotron self-Compton model is assumed. However, this relationship can not be replicated if an external Compton scattering model is assumed. These differences are primarily due to beaming effects. In addition it has been determined that the intrinsic radio luminosity of the parent sample falls in the range $10^{21} < L < 10^{30}\\, {\\rm Watts\\,Hz^{-1}}$ and that the bulk Lorentz factors of the source are in the range $ 1 < \\Gamma < 30 $, in a agreement with VLBI observations. Finally, we discuss implications for GLAST, successfully launched in June 2008. ", "introduction": "During the lifetime of the Energetic Gamma-Ray Experiment Telescope (EGRET) instrument on board {\\it Compton Gamma-Ray Observatory} (CGRO) 271 sources were detected with 66 being confidently identified as blazars in the Third EGRET Catalog \\citep{har99}. Many differing statistical analyses of these gamma ray detected blazars have been conducted, among them \\citet{fos98};\\citet{muc97}. Some of these analyses concentrated on the direct statistical relationship between luminosities in the gamma-ray band and radio bands \\citep{sal96,ste93,fan98}. Of these, most have used single dish radio data, but some have used VLBI fluxes \\citep{zho97,mat97}, all at various radio frequencies $>$ 1 GHz. In addition, \\citet{muc97} and \\citet{imp96} use Monte Carlo simulations to aid in interpreting these relationships. Though significant correlations are reported in all of these works, \\citet{muc97} show that in some cases these will result from a combination of variability and selection effects. To investigate the variability effects further, \\citet{zha01} have compared their statistical results using time averaged data for the entire sample to similar results for a restricted sample for which data were available during high and low states. They report a similar significant correlation in each case and show that using time averaged data tends to under estimate the underlying linear regression slope (as applied to the logarithmic data). Various models have been invoked to explain the origin of the gamma-ray emission of blazars and specifically, the radio gamma-ray correlation. Among these are the synchrotron self Compton (SSC) model \\citep{blo96,ghi85} and various models in which the source of seed photons for scattering is from a source external to the jet (henceforth called ECS models for ``external Compton scattering''). \\citet{der93} use the accretion disk as the source of soft photons whereas others \\citep{ghi96} use broad-line region clouds as the source of soft photons. Since the end of the EGRET mission there have also been several reanalyses of the significance of identifications of EGRET sources, particularly those of \\citet{mat97,mat01} and \\citet{sow03,sow04}.The \\citet{sow03,sow04} survey excludes sources with $|b|<10^{\\circ}$ (the Mattox papers exclude sources with $|b|<3^{\\circ}$), and thus may exclude additional sources that are thought to be blazars \\citep{sgu04}. Likewise, some work continues on identifying other individual sources, such as 3EG J0416+3650, possibly identified with 3C 111 \\citep{sgu06}. In the eight years since the end of the CGRO mission, there has been substantial modification to the identifications given in the last catalog published by the EGRET team \\citep{har99}, so we present a comprehensive statistical analysis including all potential identifications, using homogeneous criteria for inclusion in the sample ({\\S 2}). We have also included various statistical techniques of survival analysis for the inclusion of upper and lower limits in the data ({\\S 3}). We later discuss whether or not the results of the analyses are dependent on the precise source list. The approach used in understanding the physical implications of our results is to use a Monte Carlo technique \\citep{lis97} to generate multiple simulated samples under differing assumptions of overall theoretical or phenomenological model ({\\S 4,5}), distribution of physical parameters such as bulk Lorentz factor (e.g., Gaussian, power-law, etc.) and selection effects. Additional physical implications of observed correlations and implications for the Gamma Ray Large Area Space Telescope (GLAST), successfully launched in June 2008, are discussed in {\\S 6,7}. ", "conclusions": "After taking into account statistical tests and Monte Carlo analysis, we find the following: \\begin{enumerate} \\item For this sample of 122 gamma-ray blazars there is a strong correlation between radio and gamma ray luminosity which persists even after the effects of redshift and limits are taken into account. The correlation is of the form $L_{\\gamma} \\propto {L_r}^{0.77}$. This correlation remains with similar regression coefficients even when only the strongest 76 candidates are included in the sample. \\item There is a correlation between $\\alpha_{og}$ and $\\alpha_{ro}$ as well as a correlation between $L_r$ and $\\alpha_{og}$. Each correlation is consistent with the increasing dominance of the SED by gamma ray luminosity as the maximum relativistic electron energy decreases and the increasing importance of the ECS process at high radio luminosities and low maximum electron energy. \\item A detailed simulation of source statistics using Monte Carlo techniques shows that the relationship $L_{\\gamma} \\propto {L_r}^{0.77}$ can only be reproduced assuming the SSC model. Though the assumed intrinsic physical model in the source frame is of the form $L_{\\gamma} \\propto L_r$, selection effects lead to a simulated sample with the relationship described above. This effect is much less significant when ECS is assumed as the underlying model. However, upon also comparing the observed and simulated distributions of the luminosities, flux densities and redshifts, both the SSC and ECS models are only weakly consistent with the data, assuming linear dependence of intrinsic gamma ray luminosity on intrinsic radio luminosity. Taken together with the previous results, this would suggest that if either SSC or ECS is indeed responsible for the gamma ray emission, a more complex model is likely needed. In addition, effects of evolution and assumed cosmologies can be explored in more detail. \\end{enumerate} In the GLAST era, our findings can be clarified in the following ways. Acquiring more gamma-ray data, including, particularly, detections of new dim sources near the GLAST detection limit will clarify whether the correlations we find are caused by truncation effects due to the flux limits of our instruments, or whether this is due, at least in part, to physical causes, at least down to the new limit established by GLAST. In order to predict what we might see with GLAST, we can extend our observed radio/gamma ray luminosity correlation down to lower luminosities than what are covered in Figure 1. A radio luminosity of $1.25 \\times 10^{24} {\\rm Watts\\,Hz^{-1}}$ would lead to a gamma ray luminosity of approximately $4.98 \\times 10^{13} {\\rm Watts \\, Hz^{-1}}$. We can then covert these luminosities to flux densities for a range of potential redshifts. For z=0.1-1.1 the radio flux density would be in the range of 0.8-60 mJy and the gamma-ray flux density would be in the range, $3.3 \\times 10^{-14}$-$2.4 \\times 10^{-12}$ Jy. Greater values of redshift would lead to even lower flux density values. Most of the range of radio flux densities ($>$ few mJy) would be detectable by the Green Bank Telescope (GBT), with the lower limit detectable with the Very Large Array (VLA)\\citep{con08}. Assuming that the sensitivity of GLAST is approximately $2.8 \\times 10^{-13}$ Jy at 100 MeV, these sources would only be be detected if $z< 0.2$. A radio luminosity of $1.58 \\times 10^{23} {\\rm Watts \\, Hz^{-1}}$ corresponds to a gamma-ray luminosity of $1 x10^{13} {\\rm Watts \\, Hz^{-1}}$ . The radio flux density would fall in the range of 0.1-8 mJy for z=0.1-1.1. These sources could potentially be radio detectable with the VLA under optimal conditions. At the higher flux density limit it would also be possible to detect with GBT, especially at higher frequencies (i.e., 40-50 GHz)\\citep{min08}. However, if the regression slope were actually closer to 1, implying a d! irect proportionality between radio and gamma ray luminosities, then radio sources in the same luminosity, redshift and flux ranges discussed above would not be detected in gamma rays at all. In short, for the observed luminosity relationship for EGRET blazars to be appreciably extended down to lower luminosities with GLAST detections, then a large percentage of GLAST sources near the detection limit would have radio flux densities about 10 mJy or greater (and $z< 1$). Several authors have stated that the predicted number of blazars to be detected by GLAST rougly matches with the sky density of flat spectrum radio sources down to 50 mJy at 5 GHz \\citep{pad07} and 65 mJy at 8.4 GHz \\citep{hea07}. However, our results show that a large number of GLAST sources would be detected at even lower radio flux densities if the luminosity relationship we observe for EGRET sources also holds up for GLAST sources. It is very likely that many new radio observations would be required in a! ny of the cases mentioned above, but especially for cases in w! hich the putative radio source may not even be in any previous catalog. >From a theoretical perspective, confirmation of the previously determined correlation between gamma-ray and radio luminosities will lead to confirmation of SSC, though the precise agreement would have to be re-analyzed with these new data." }, "0807/0807.2434_arXiv.txt": { "abstract": "We present limits on transit timing variations and secondary eclipse depth variations at 8 microns with the Spitzer Space Telescope IRAC camera. Due to the weak limb darkening in the infrared and uninterrupted observing, Spitzer provides the highest accuracy transit times for this bright system, in principle providing sensitivity to secondary planets of Mars mass in resonant orbits. Finally, the transit data provides tighter constraints on the wavelength-dependent atmospheric absorption by the planet. ", "introduction": "The extremely precise 33 hour phase function measurement of HD~189733b \\cite[Knutson et al.\\ (2007)]{Knutson2007} observed with the 8 micron IRAC camera on the Spitzer Space Telescope yielded the most precise times of transit and secondary eclipse for any extrasolar planet, 6 and 24 seconds, respectively. This led us to propose to observe an additional six transits and eclipses of this system over time with the goal of measuring: \\begin{itemize} \\item precise transit-timing (\\cite[Agol et al.\\ 2005]{Agol2005}, \\cite[Holman \\& Murray 2005]{Holman2005}) to search for the presence of resonant (or near-resonant) terrestrial-mass planets captured by migration (\\cite[e.g.\\ Mandell, Raymond \\& Sigurdsson 2007]{Mandell2007}) or on longer timescales precession of an eccentric orbit (\\cite[Miralda-Escud{\\'e} 2002]{MiraldaEscude2002}, \\cite[Heyl \\& Gladman 2007]{Heyl2007}; also Fabrycky and Wolfe, these proceedings); \\item variations in the depth of the secondary eclipses with time which might be caused by large-scale variable atmospheric features (\\cite[Rauscher et al. 2007]{Rauscher2007}; also Showmand and Dobbs-Dixon, these proceedings); \\item precise transit depth for comparison with atmospheric-absorption models to contrain the molecular composition (\\cite[e.g. Tinetti et al. 2007]{Tinetti2007}; also Tinetti et al., Fortney et al., and Hubeny et al., these proceedings); \\item improved system parameters for better characterization of the planet, host-star, and orbit properties (\\cite[Winn et al.\\ (2007)]{Winn2007}; also Winn, these proceedings). \\end{itemize} For librating planets in a low-order mean motion resonance, the times of transit vary with an amplitude: \\begin{equation} \\delta t_{2:1} \\approx P_{trans} \\left({M_{pert} \\over M_{trans}}\\right) \\approx 3 {\\rm \\ min} \\left({P_{trans} \\over 3 {\\rm \\ day}}\\right) \\left({M_{pert} \\over M_{\\oplus}}\\right) \\left({M_{J} \\over M_{trans}}\\right), \\end{equation} and libration period of \\begin{equation} P_{2:1} \\approx P_{trans}\\left({M_* \\over M_{trans}}\\right)^{2/3} \\approx 150 {\\rm \\ days}\\left({P_{trans} \\over 3 {\\rm \\ day}}\\right) \\left({M_{*} \\over M_{\\odot}}\\right)^{2/3} \\left({M_{J} \\over M_{trans}}\\right)^{2/3}, \\end{equation} where $M_\\oplus, M_J, M_*$ are masses of the Earth, Jupiter, and host star, $P_{trans}$ is the period of the transiting planet, and the numbers have been estimated for the libration amplitude for planets starting on circular orbits with exact commensurability \\cite[(Agol et al.\\ 2005)]{Agol2005}; the actual value depends on the libration amplitude. This timescale requires observations separated by months with $\\approx$ seconds precision in timing, and in principle could be sensitive to sensitive to Mars mass planets. ", "conclusions": "The five transits we have observed are consistent with no transit timing variations greater than 5 seconds. The five secondary eclipses are consistent with no eclipse depth variations at greater than the 10\\% level. From just an analysis of the lightcurves we can place constraints on the eccentricity and longitude of periastron: $\\vert e \\cos{\\omega} \\vert = 0.0002 \\pm 0.0001$ and $\\vert e \\sin{\\omega}\\vert = 0.015 \\pm 0.012$; similar to the results of \\cite[Winn et al.\\ (2007)]{Winn2007}. We are collecting two more transits and two more secondary eclipses in Summer 2008; thus the results presented here are preliminary and require further more careful analysis." }, "0807/0807.4481_arXiv.txt": { "abstract": " ", "introduction": "A major unsolved problem in theoretical physics is to reconcile the classical theory of general relativity with quantum mechanics. Of the numerous attempts, some have postulated new and so far unobserved ingredients, while others have proposed radically new principles governing physics at the as yet untested Planckian energy scale. Here we report on a much more mundane approach using only standard quantum field theory. In a sum-over-histories approach we will attempt to define a nonperturbative quantum field theory which has as its infrared limit ordinary classical general relativity and at the same time has a nontrivial ultra\\-vio\\-let limit. From this point of view it is in the spirit of the renormalization group approach, first advocated long ago by Weinberg \\cite{weinberg}, and more recently substantiated by several groups of researchers \\cite{reuteretc}. However, it has some advantages compared to the renormalization group approach in that it allows us to study (numerically) certain geometric observables which are difficult to handle analytically. We define the path integral of quantum gravity nonperturbatively using the lattice approach known as {\\it causal dynamical triangulations} (CDT) as a regularization. In Sec.\\ \\ref{cdt} we give a short description of the formalism, providing the definitions which are needed later to describe the measurements. CDT establishes a nonperturbative way of performing the sum over four-geometries (for more extensive definitions, see \\cite{ajl4d,blp}). It sums over the class of piecewise linear four-geometries which can be assembled from four-dimensional simplicial building blocks of link length $a$, such that only {\\it causal} spacetime histories are included. The continuum limit of such a lattice theory should ideally be obtained as for QCD defined on an ordinary fixed lattice, where for an observable $\\cO(x_n)$, $x_n$ denoting a lattice point, one can measure the correlation length $\\xi(g_0)$ from \\beq\\label{1.1} -\\log (\\la \\cO(x_n) \\cO(y_m) \\ra) \\sim |n-m|/ \\xi(g_0) + o(|n-m|). \\eeq A continuum limit of the lattice theory may then exist if it is possible to fine-tune the bare coupling constant $g_0$ of the theory to a critical value $g_0^c$ such that the correlation length goes to infinity, $\\xi(g_0) \\to \\infty$. Knowing how $\\xi(g_0)$ diverges for $g_0 \\to g_0^c$ determines how the lattice spacing $a$ should be taken to zero as a function of the coupling constants, namely \\beq\\label{1.2} \\xi(g_0) = \\frac{1}{|g_0-g_0^c|^\\n},~~~~~a(g_0) = |g_0-g_0^c|^\\n. \\eeq The challenge when searching for a {\\it field theory} of quantum gravity is to find a theory which behaves in this way. The challenge is three-fold: (i) to find a suitable nonperturbative formulation of such a theory which satisfies a minimum of reasonable requirements, (ii) to find observables which can be used to test relations like \\rf{1.1}, and (iii) to show that one can adjust the coupling constants of the theory such that \\rf{1.2} is satisfied. Although we will focus on (i) in what follows, let us immediately mention that (ii) is notoriously difficult in a theory of quantum gravity, where one is faced with a number of questions originating in the dynamical nature of geometry. What is the meaning of distance when integrating over all geometries? How do we attach a meaning to local spacetime points like $x_n$ and $y_n$? How can we define at all local, diffeomorphism-invariant quantities in the continuum which can then be translated to the regularized (lattice) theory? -- What we want to point out here is that although (i)-(iii) are standard requirements when relating critical phenomena and (Euclidean) quantum field theory, gravity {\\it is} special and may require a reformulation of (part of) the standard scenario sketched above. We will return to this issue when we discuss our results in Sec.\\ \\ref{discussion}. Our proposed nonperturbative formulation of four-dimensional quantum gravity has a number of nice features. Firstly, it sums over a class of piecewise linear geometries, which -- as usual -- are described without the use of coordinate systems. In this way we perform the sum over geometries directly, avoiding the cumbersome procedure of first introducing a coordinate system and then getting rid of the ensuing gauge redundancy, as one has to do in a continuum calculation. Our underlying assumptions are that 1) the class of piecewise linear geometries is in a suitable sense dense in the set of all geometries relevant for the path integral (probably a fairly mild assumption), and 2) that we are using a correct measure on the set of geometries. This is a more questionable assumption since we do not even know whether such a measure exists. Here one has to take a pragmatic attitude in order to make progress. We will simply examine the outcome of our construction and try to judge whether it is promising. Secondly, our scheme is background-independent. No distinguished geometry, accompanied by quantum fluctuations, is put in by hand. If the CDT-regularized theory is to be taken seriously as a potential theory of quantum gravity, there has to be a region in the space spanned by the bare coupling constants where the geometry of spacetime bears some resemblance with the kind of universe we observe around us. That is, the theory should create dynamically an effective background geometry around which there are (small) quantum fluctuations. This is a very nontrivial property of the theory and one we are going to investigate in detail in the present piece of work. New computer simulations presented here confirm in a much more direct way the indirect evidence for such a scenario which we provided earlier in \\cite{emerge,semi}. They establish the de Sitter nature of the background spacetime, quantify the fluctuations around it, and set a physical scale for the universes we are dealing with. The main results of our investigation, without the numerical details, were announced in \\cite{agjl} (see also \\cite{desitter}). The rest of the article is organized as follows. In Sec.\\ \\ref{cdt} we describe briefly the regularization method of quantum gravity named CDT and the set-up of the computer simulations. In Sec.\\ \\ref{S4} we present the evidence for an effective background geometry corresponding to the four-dimensional sphere $S^4$, i.e.\\ Euclidean de Sitter spacetime. Sec.\\ \\ref{effective} deals with the reconstruction of an effective action for the scale factor of the universe from the computer data, and in Sec.\\ \\ref{fluctuations} we analyze the quantum fluctuations around the ``classical'' $S^4$-solution. Sec.\\ \\ref{S3} contains an analysis of the geometry of the spatial slices of our computer-generated universe. In Sec.\\ \\ref{size} we determine the physical sizes of our universes expressed in Planck lengths and try to follow the flow of the gravitational coupling constant of the effective action under a change of the bare coupling constants of the bare, ``classical'' action used in the path integral. Finally we discuss the results, their interpretation and future perspectives of the CDT-quantum gravity theory in Sec.\\ \\ref{discussion}. ", "conclusions": "} The CDT model of quantum gravity is extremely simple. It is the path integral over the class of causal geometries with a global time foliation. In order to perform the summation explicitly, we introduce a grid of piecewise linear geometries, much in the same way as when defining the path integral in quantum mechanics. Next, we rotate each of these geometries to Euclidean signature and use as bare action the Einstein-Hilbert action\\footnote{Of course, the full, effective action, including measure contributions, will contain all higher-derivative terms.} in Regge form. That is all. The resulting superposition exhibits a nontrivial scaling behaviour as function of the four-volume, and we observe the appearance of a well-defined average geometry, that of de Sitter space, the maximally symmetric solution to the classical Einstein equations in the presence of a positive cosmological constant. We are definitely in a quantum regime, since the fluctuations of the three-volume around de Sitter space are sizeable, as can be seen in Fig.\\ \\ref{fig1}. Both the average geometry and the quantum fluctuations are well described in terms of the mini\\-superspace action \\rf{n5}. A key feature to appreciate is that, unlike in standard (quantum-)cosmological treatments, this description is the {\\it outcome} of a nonperturbative evaluation of the {\\it full} path integral, with everything but the scale factor (equivalently, $V_3(t)$) summed over. Measuring the correlations of the quantum fluctuations in the computer simulations for a particular choice of bare coupling constants enabled us to determine the continuum gravitational coupling constant $G$ as $G\\approx 0.42 a^2$, thereby introducing an absolute physical length scale into the dimensionless lattice setting. Within measuring accuracy, our de Sitter universes (with volumes lying in the range of 6.000-47.000 $\\ell_{Pl}^4$) are seen to behave perfectly semiclassically with regard to their large-scale properties. We have also indicated how we may be able to penetrate into the sub-Planckian regime by suitably changing the bare coupling constants. By ``sub-Planckian regime\" we mean that the lattice spacing $a$ is (much) smaller than the Planck length. While we have not yet analyzed this region in detail, we expect to eventually observe a breakdown of the semiclassical approximation. This will hopefully allow us to make contact with attempts to use renormalization group techniques in the continuum and the concept of asymptotic safety to study scaling violations in quantum gravity \\cite{reuteretc}. On the basis of the results presented here, two major issues suggest themselves for further research. First, we need to establish the relation of our effective gravitational coupling constant $G$ with a more conventional gravitational coupling constant, defined directly in terms of coupling matter to gravity. In the present work, we have defined $G$ as the coupling constant in front of the effective action, but it would be desirable to verify directly that a gravitational coupling defined via the coupling to matter agrees with our $G$. In principle it is easy to couple matter to our model, but it is less straightforward to define in a simple way a set-up for extracting the semiclassical effect of gravity on the matter sector. Attempts in this direction were already undertaken in the ``old'' Euclidean approach \\cite{js,newton}, and it is possible that similar ideas can be used in CDT quantum gravity. Work on this is in progress. The second issue concerns the precise nature of the ``continuum limit''. Recall our discussion in the Introduction about this in a conventional lattice-theoretic setting. The continuum limit is usually linked to a divergent correlation length at a critical point. It is unclear whether such a scenario is realized in our case. In general, it is rather unclear how one could define at all the concept of a divergent length related to correlators in quantum gravity, since one is integrating over all geometries, and it is the geometries which dynamically give rise to the notion of ``length\". This has been studied in detail in two-dimensional (Euclidean) quantum gravity coupled to matter with central charge $c \\le 1$ \\cite{corr2d}. It led to the conclusion that one could associate the critical behaviour of the matter fields (i.e.\\ approaching the critical point of the Ising model) with a divergent correlation length, although the matter correlators themselves had to be defined as non-local objects due to the requirement of diffeomorphism invariance. On the other hand, the two-dimensional studies do not give us a clue of how to treat the gravitational sector itself, since they do not possess gravitational field-theoretic degrees of freedom. What happens in the two-dimensional lattice models which can be solved analytically is that the only fine-tuning needed to approach the continuum limit is an additive renormalization of the cosmological constant (for fixed matter couplings). Thus, fixing the two-dimensional spacetime volume $N_2$ (the number of triangles), such that the cosmological constant plays no role, there are no further coupling constants to adjust and the continuum limit is automatically obtained by the assignment $V_2 = N_2 a^2$ and taking $N_2 \\to \\infty$. This situation can also occur in special circumstances in ordinary lattice field theory. A term like \\beq\\label{k5} \\sum_i c_1 (\\phi_{i+1}-\\phi_i)^2 + c_2(\\phi_{i+1}+\\phi_{i-1}-2\\phi_i)^2 \\eeq (or a higher-dimensional generalization) will also go to the continuum free field theory simply by increasing the lattice size and using the identification $V_d = L^d a^d$ ($L$ denoting the linear size of the lattice in lattice units), the higher-derivative term being sub-dominant in the limit. It is not obvious that in quantum gravity one can obtain a continuum quantum field theory without fine-tuning in a similar way, because the action in this case is multiplied by a dimensionful coupling constant. Nevertheless, it is certainly remarkable that the infrared limit of our effective action apparently reproduces -- within the cosmological setting -- the Einstein-Hilbert action, which is the unique diffeomorphism-invariant generalization of the ordinary kinetic term, containing at most second derivatives of the metric. A major question is whether and how far our theory can be pushed towards an ultraviolet limit. We have indicated how to obtain such a limit by varying the bare coupling constants of the theory, but the investigation of the limit $a \\to 0$ with fixed $G$ has only just begun. \\subsection*" }, "0807/0807.3742_arXiv.txt": { "abstract": "This paper presents a photometric and spectroscopic study of the bright blue eclipsing binary LMC-SC1-105, selected from the OGLE catalog as a candidate host of very massive stars ($\\geq 30\\,\\msun$). The system is found to be a double-lined spectroscopic binary, which indeed contains massive stars. The masses and radii of the components are $\\rm M_{1}=30.9\\pm1.0\\;\\msun$, $\\rm M_{2}=13.0\\pm0.7\\;\\msun$, and $\\rm R_{1}=15.1\\pm0.2\\;\\rsun$, $\\rm R_{2}=11.9\\pm0.2\\;\\rsun$, respectively. The less massive star is found to be filling its Roche lobe, indicating the system has undergone mass-transfer. The spectra of LMC-SC1-105 display the Struve-Sahade effect, with the \\ion{He}{1} lines of the secondary appearing stronger when it is receding and causing the spectral types to change with phase (O8$+$O8 to O7$+$O8.5). This effect could be related to the mass-transfer in this system. To date, accurate ($\\leq 10\\%$) fundamental parameters have only been measured for 15 stars with masses greater than 30~\\msun, with the reported measurements contributing valuable data on the fundamental parameters of very massive stars at low metallicity. The results of this work demonstrate that the strategy of targeting the brightest blue stars in eclipsing binaries is an effective way of studying very massive stars. ", "introduction": "\\label{section:intro} \\vspace{-0.4cm} The fundamental parameters of very massive stars ($\\geq 30\\,\\msun$) remain uncertain, despite the large impact massive stars have in astrophysics, both individually and collectively \\citep[see review by][]{Massey03}. The equations of stellar structure allow for stars with arbitrarily large masses, however the mechanisms to form massive stars \\citep[accretion and mergers; e.g.][]{Bally05} and the associated instabilities \\citep[see][and references therein]{Elmegreen00, Zinnecker07} are not well understood, hindering theoretical predictions on the existence of an upper limit on the stellar mass. The ``mass discrepancy'' problem, i.e. the disagreement between masses derived from parameters determined by fitting stellar atmosphere models to spectra and from evolutionary tracks \\citep[see e.g.][for a comparison]{Repolust04, Massey05}, still affects studies of single massive stars, even though significant progress has been made in both stellar atmosphere \\citep[see review by][]{Herrero07} and stellar evolution models \\citep[e.g.][]{Meynet03}. The parameters of single stars also suffer from suspected multiplicity, which in many cases cannot be determined. Pismis 24-1 demonstrates this problem: its inferred evolutionary mass $>200\\,\\msun$ \\citep{Walborn02} contradicted the upper stellar mass limit of $\\sim 150\\,\\msun$ suggested by statistical arguments based on observations \\citep{Figer05,Oey05}. \\citet{Maiz-Apellaniz07} resolved it into a visual binary with the {\\it Hubble Space Telescope}, thereby removing the discrepancy. One of its components is also a spectroscopic binary, illustrating the systematic effects often accompanying ``single'' stars. \\vspace{-0.2cm} The only model-independent way to obtain accurate fundamental parameters of distant massive stars and to resolve the ``mass discrepancy'' problem is to use eclipsing binaries \\citep[see review by][]{Andersen91}. In particular, double-lined spectroscopic binary systems exhibiting eclipses in their light curves are extremely powerful tools for measuring masses and radii of stars. Specifically, the light curve provides the orbital period, inclination, eccentricity, the fractional radii and flux ratio of the two stars. The radial velocity semi-amplitudes determine the mass ratio; the individual masses can be solved for by using Kepler's third law. Furthermore, by fitting synthetic spectra to the observed ones, one can infer the effective temperatures of the stars, solve for their luminosities and derive the distance \\citep[e.g.\\@][]{Bonanos06}. The most massive stars measured in eclipsing binaries are galactic Wolf-Rayet stars of WN6ha spectral type: NGC3603-A1 \\citep[M$_1=116 \\pm 31\\; \\msun$, M$_2=89 \\pm 16\\; \\msun$;][]{Schnurr08}, and WR~20a \\citep[M$_1=83.0 \\pm 5.0\\; \\msun$ and M$_2=82.0 \\pm 5.0\\; \\msun$][]{Rauw04, Bonanos04} in Westerlund~2, presenting a challenge for both stellar evolution and massive star formation models \\citep{Yungelson08, Zinnecker07} and raising the issue of the frequency and origin of ``binary twins'' \\citep{Pinsonneault06, Lucy06, Krumholz07}. Such systems are of particular interest, since massive binaries might be progenitors of gamma-ray bursts \\citep[e.g.\\@][]{Fryer07}, especially in the case of Population III, metal-free stars \\citep[see][]{Bromm06}. Analogs of these heavyweight champions, if not more massive binaries, are bound to exist in the young massive clusters at the center of the Galaxy (Center, Arches, Quintuplet), in nearby super star clusters (e.g.\\@ Westerlund\\,1, R136), in Local Group galaxies (e.g.\\@ LMC, SMC, M31, M33) and beyond (e.g.\\@ M81, M83, NGC 2403). A systematic wide-ranging survey of these clusters and galaxies is currently underway. The goal is to provide data with which to test star formation theories, stellar atmosphere and stellar evolution models for both single and binary stars as a function of metallicity, and the theoretical predictions on the upper limit of the stellar mass. The adopted strategy involves two steps: a variability survey to discover eclipsing binaries in these massive clusters and nearby galaxies, which is followed by spectroscopy to derive parameters of the brightest -- thus most luminous and massive -- blue systems. However, characterizing massive stars requires the availability of 8-m class telescopes and high resolution near-infrared spectrographs (since massive stars in the Galaxy are extincted and extragalactic ones are faint) and has only become feasible in the past few years. \\citet{Bonanos07} demonstrated that this method efficiently finds massive candidates, by performing the first variability survey of the Westerlund~1 super star cluster and discovering 4 massive eclipsing binary systems. Figure~\\ref{massradius} illustrates the extent of our knowledge of precise fundamental parameters of massive stars. It presents published mass-radius measurements from eclipsing binaries, accurate to better than $10\\%$ for the more massive component. The zero-age main sequences (ZAMS) at both Z=0.02 \\citep{Schaller92} and Z=0.008 \\citep{Schaerer93} are overplotted as a reference. The Galactic data are mainly taken from the compilations of \\citet{Andersen91} and \\citet{Gies03} with additions and updates from \\citet{Vitrichenko07}, but also \\citet{Gonzalez05} for the Large Magellanic Cloud (LMC), \\citet{Harries03} and \\citet{Hilditch05} for the SMC, and \\citet{Ribas05} for M31. A literature search was done to include all accurate measurements of stars in eclipsing binaries with masses $\\geq 30\\,\\msun$, which are presented in Table~\\ref{massradiusdata}. This Table is, to my knowledge, complete at present and consists of only 14 very massive stars with better than $10\\%$ mass-radius measurements, located in 3 galaxies. Of these, WR~20a and M33 X-7 \\citep{Orosz07} are the most massive and noteworthy. M33 X-7 contains a very massive $70.0 \\pm 6.9\\; \\msun$ O-type giant and a record-breaking $15.65\\;\\msun$ black hole, challenging current evolutionary models, which fail to explain such a large black hole mass. Without accurate measurements for a large sample of massive stars, theoretical models will remain unconstrained. A survey to determine accurate parameters for several massive eclipsing binaries in the low metallicity (Z$=0.008$) LMC was undertaken, with the purpose of increasing the sample and improving our understanding of these rare systems. Several candidates were selected from the OGLE-II catalog of eclipsing binaries in the LMC \\citep{Wyrzykowski03} as the brightest systems with $B-V<0$. LMC-SC1-105, or OGLE053448.26-694236.4, is one of the brightest eclipsing binaries with $I_{max}=13.04$~mag, $V_{max}=12.97$~mag, $B_{max}=12.81$~mag and a preliminary semi-detached classification. This work presents the analysis of the follow-up observations obtained for LMC-SC1-105. The paper is organized as follows: \\S2 describes the spectroscopy and data reduction, \\S3 the spectral classification, \\S4 the radial velocity curve, \\S5 the light curve analysis, \\S6 the evolutionary status, and \\S7 the conclusion. ", "conclusions": "This paper presents accurate (to better than $5\\%$) fundamental parameters of LMC-SC1-105, one of the brightest blue eclipsing binary stars in the LMC found by the OGLE survey. The aim of this work is twofold: 1) to demonstrate that targeted surveys of the brightest blue eclipsing binaries in nearby galaxies do indeed select very massive stars and 2) to measure accurate parameters for one of these rare systems. The parameters of LMC-SC1-105 were determined from the light curves available from the OGLE and MACHO surveys and newly acquired high resolution spectroscopy that targeted quadrature phases, in part applying the strategy proposed by \\citet{Gonzalez05} to constrain the radial velocity curve with a small number of spectra. The system was found to contain a very massive main sequence primary ($30.9\\pm1.0\\,\\msun$) and a possibly evolved Roche lobe-filling secondary. The spectra display the Struve-Sahade effect, which is present in all the \\ion{He}{1} lines, causing the spectral classification to change with phase, and could be related to the mass transfer occurring in the system. LMC-SC1-105 could further be used as a distance indicator to the LMC. However, in addition to accurate radii, accurate flux (i.e. effective temperatures) and extinction estimates are necessary for accurate distances. Eclipsing binaries have been used to derive accurate and independent distances to the LMC \\citep[e.g.][]{Guinan98, Fitzpatrick03}, the Small Magellanic Cloud \\citep{Harries03,Hilditch05}, M31 \\citep{Ribas05} and most recently to M33 \\citep{Bonanos06}. The accurate parameters determined herein for LMC-SC1-105 contribute valuable data on very massive stars, increasing the current sample of 14 very massive stars with accurate parameters to 15, which despite their importance remain poorly studied. Such data serve as an external check to resolve the ``mass discrepancy'' problem, as \\citet{Burkholder97} have shown, and to constrain stellar atmosphere, evolution and formation models. Further systematic studies of massive binaries in nearby galaxies are needed to extend the sample of 50 SMC eclipsing binaries \\citep{Harries03, Hilditch05} to higher masses and metallicities and populate the sparsely sampled parameter space (mass, metallicity, evolutionary state) with accurate measurements of their masses and radii. The method of targeting very massive stars in bright blue eclipsing binaries can therefore be employed towards this goal." }, "0807/0807.3748_arXiv.txt": { "abstract": "Radiative cooling is central to a wide range of astrophysical problems. Despite its importance, cooling rates are generally computed using very restrictive assumptions, such as collisional ionization equilibrium and solar relative abundances. We simultaneously relax both assumptions and investigate the effects of photo-ionization of heavy elements by the meta-galactic UV/X-ray background and of variations in relative abundances on the cooling rates of optically thin gas in ionization equilibrium. We find that photo-ionization by the meta-galactic background radiation reduces the net cooling rates by up to an order of magnitude for gas densities and temperatures typical of the shock-heated intergalactic medium and proto-galaxies ($10^4\\,\\K \\la T \\la 10^6\\,\\K$, $\\rho / \\left<\\rho\\right> \\la 100$). In addition, photo-ionization changes the relative contributions of different elements to the cooling rates. We conclude that photo-ionization by the ionizing background and heavy elements both need to be taken into account in order for the cooling rates to be correct to order of magnitude. Moreover, if the rates need to be known to better than a factor of a few, then departures of the relative abundances from solar need to be taken into account. We propose a method to compute cooling rates on an element-by-element basis by interpolating pre-computed tables that take photo-ionization into account. We provide such tables for a popular model of the evolving UV/X-ray background radiation, computed using the photo-ionization package \\textsc{cloudy}. ", "introduction": "\\label{sec:intro} Dissipation of energy via radiative cooling plays a central role in many different astrophysical contexts. In general the cooling rate depends on a large number of parameters, such as the gas density, temperature, chemical composition, ionization balance, and the radiation field. In the absence of radiation, however, the equilibrium ionization balance depends only on the temperature. In that case the cooling rate in the low density regime, which is dominated by collisional processes, is simply proportional to the gas density squared, for a given composition. Thus, the cooling rates for a plasma in collisional ionization equilibrium (CIE) can be conveniently tabulated as a function of the temperature and composition (metallicity) of the gas \\cite[e.g.,][]{Cox1969,Raymond1976,Shull1982,Gaetz1983,Boehringer1989,Sutherland1993,Landi1999,Benjamin2001,Gnat2007,Smith2008}, and such tables are widely used for a large variety of problems. Although it is convenient to ignore radiation when calculating cooling rates, radiation is generally important for the thermal and ionization state of astrophysical plasmas. For example, \\cite{Efstathiou1992} investigated the effect of the extragalactic UV background on the cooling rates for gas of primordial composition (in practice this means a pure H/He plasma) and found that including photo-ionization can suppress the cooling rates of gas in the temperature range $T\\sim 10^4 - 10^5\\,\\K$ by a large factor. Although the effects of radiation are often taken into account for gas of primordial composition, photo-ionization of heavy elements is usually ignored in the calculation of cooling rates (but see \\citealt{Leonard1998,Benson2002}). In this paper we will investigate the dependence of cooling rates of gas enriched with metals on the presence of ionizing radiation, focusing on the temperature range $T\\sim 10^4 - 10^8\\,\\K$ and optically thin plasmas. We will show that, as is the case for gas of primordial composition \\cite[][]{Efstathiou1992}, photo-ionization can suppress the metallic cooling rates by a large factor. Moreover, the suppression of the cooling rate is significant up to much higher temperatures than for the primordial case. We will also investigate the relative contributions of various elements to the cooling rates. If the relative abundances are similar to solar, then oxygen, neon, and iron dominate the cooling in the temperature range $T \\sim 10^4 - 10^7\\,\\K$. However, we will show that the relative contributions of different elements to the cooling rate are sensitive to the presence of ionizing radiation. Although we will illustrate the results using densities and radiation fields that are relevant for studies of galaxy formation and the intergalactic medium (IGM), the conclusion that photo-ionization significantly reduces the cooling rates of enriched gas is valid for a large variety of astrophysical problems. For example, for $T \\sim 10^5\\,\\K$ and $T\\sim 10^6\\,\\K$ the reduction of metal-line cooling rates is significant as long as the dimensionless ionization parameter\\footnote{$U \\equiv \\Phi_{\\rm H}/(n_{\\rm H}c)$, where $\\Phi_{\\rm H}$ is the flux of hydrogen ionizing photons (i.e.,\\ photons per unit area and time), $n_{\\rm H}$ is the total hydrogen number density and $c$ is the speed of light.} $U \\ga 10^{-3}$ and $U \\ga 10^{-1}$, respectively. We will focus on the temperature range $10^4 - 10^8\\,\\K$ because gas in this temperature range is usually optically thin and because the effects of photo-ionization are generally unimportant at higher temperatures. Tables containing cooling rates and several other useful quantities as a function of density, temperature, redshift, and composition, appropriate for gas exposed to the models for the evolving meta-galactic UV/X-ray background of \\cite{Haardt2001} are available on the following web site: \\texttt{http://www.strw.leidenuniv.nl/WSS08/}. The web site also contains a number of videos that illustrate the dependence of the cooling rates on various parameters. This paper is organized as follows. In Section~\\ref{sec:method} we present our method for calculating element-by-element cooling rates including photo-ionization and we compare the limiting case of CIE to results taken from the literature. Section~\\ref{sec:photometals} shows how metals and ionizing radiation affect the cooling rates. Section~\\ref{sec:whim} demonstrates the importance for the low-redshift shock-heated IGM, which is thought to contain most of the baryons. In this section we also illustrate the effect of changing the intensity and spectral shape of the ionizing radiation. We investigate the effect of photo-ionization on the relative contributions of individual elements in Section~\\ref{sec:relabund} and we summarize and discuss our conclusions in Section~\\ref{sec:discussion}. Throughout this paper we use the cosmological parameters from \\cite{Komatsu2008}: ($\\Omega_{\\rm m}, \\Omega_{\\Lambda}, \\Omega_{\\rm b}, h) = (0.279, 0.721, 0.0462, 0.701$) and a primordial helium mass fraction $X_{\\rm He} = 0.248$. Densities will be expressed both as proper hydrogen number densities $n_{\\rm H}$ and density contrasts $\\delta \\equiv \\rho_{\\rm b}/\\left <\\rho_{\\rm b}\\right >-1$, where $\\left < \\rho_{\\rm b}\\right >$ is the cosmic mean baryon density. The two are related by \\begin{equation} n_{\\rm H} \\approx 1.9 \\times 10^{-7}\\,\\cm^{-3} ~\\left (1+\\delta\\right )\\left (1+z\\right )^3 \\left ({X_{\\rm H} \\over 0.752}\\right ). \\end{equation} ", "conclusions": "\\label{sec:discussion} Radiative cooling is an essential ingredient of hydrodynamical models of a wide range of astrophysical objects, ranging from the IGM to (proto-)galaxies and molecular clouds. While numerical simulations of objects with a primordial composition often compute non-equilibrium radiative cooling rates explicitly and sometimes even include the effect of ionizing background radiation, the treatment of cooling of chemically enriched material is typically much more approximate. For example, simulations of galaxy formation typically either ignore metal-line cooling altogether or include it assuming pure CIE. In addition, the abundances of all heavy elements are typically scaled by the same factor (the metallicity) (but see \\citealt{Martinez2008} and \\citealt{Maio2007} for recent exceptions). In this simplified treatment metal-line cooling depends only on temperature and metallicity, allowing straightforward interpolation from pre-computed two-dimensional tables. We have used \\textsc{cloudy} to investigate the effects of heavy elements and ionizing radiation on the radiative cooling of gas with properties characteristic of (proto-)galaxies and the IGM, i.e., optically thin gas with densities $n_{\\rm H} \\la 1\\,\\cm^{-3}$ and temperatures $T\\ga 10^4\\,\\K$, assuming ionization equilibrium. We presented a method to incorporate radiative cooling on an element-by-element basis including photo-ionization by an evolving UV/X-ray background, using precomputed tables, which for heavy elements are functions of density, temperature, and redshift and for H\\&He (which must be considered together because they are important contributors to the free electron density) depend additionally on the He/H ratio. Using the 11 elements H, He, C, N, O, Ne, Mg, Si, S, Ca, and Fe, the redshift $z=0$ median absolute errors in the net cooling rate range from 0.33\\%, at $Z = 0.1 Z_{\\odot}$ to 6.1\\% for the extreme metallicity $Z= 10 Z_{\\odot}$, and the errors are smaller for higher redshifts. The tables as well as some scripts that illustrate how to use them are available from the following web site: \\texttt{http://www.strw.leidenuniv.nl/WSS08/}. We also include tables for solar relative abundances which can be used if metallicity, but not the abundances of individual elements are known, as in equation (\\ref{eq:Zmethod}). This web site also contains a number of videos that may be helpful to gain intuition on the importance of various parameters on the cooling rates. We confirmed that, assuming CIE, heavy elements greatly enhance the cooling rates for metallicities $Z \\ga 10^{-1}~Z_\\odot$ and temperatures $T \\la 10^7\\,\\K$. We demonstrated that this remains true in the presence of photo-ionization by the meta-galactic UV/X-ray background. The background radiation removes electrons that would otherwise be collisionally excited, thus reducing the cooling rates. The effect is stronger for higher ionization parameters (i.e.,\\ higher radiation intensities or lower densities) and if the spectral shape of the radiation field is harder. Considering only the meta-galactic radiation field, which provides a lower limit to the intensity of the radiation to which optically thin gas may be exposed, the reduction of the metal-line cooling rates becomes important below $10^6\\,\\K$ for ionization parameters $U \\ga 10^{-1}$ and below $10^5\\,\\K$ for $U \\ga 10^{-3}$ (note that for the HM01 background $U=9\\times 10^{-7}/n_{\\rm H}$ and $2\\times 10^{-5}/n_{\\rm H}$ at $z=0$ and $z=3$, respectively). As an example of the potential importance of including the effects of both photo-ionization and heavy elements, we considered the so-called warm-hot intergalactic medium (WHIM), which is thought to contain a large fraction of the baryons at redshifts $z < 1$. We demonstrated that the overdensities for which gas at typical WHIM metallicities ($Z\\sim 10^{-1}~Z_\\odot$) and temperatures ($T\\sim 10^5 - 10^7\\,\\K$) can cool within a Hubble time, can shift by an order of magnitude depending on whether photo-ionization and metal-line cooling are taken into account. Hence, photo-ionization of heavy elements may have important consequences for predictions of the amount of matter contained in this elusive gas phase. Because chemical enrichment happens in a number of stages, involving a number of processes with different timescales, the relative abundances of the heavy elements varies with redshift and environment by factors of a few. Hence, computing cooling rates on an element-by-element basis rather than scaling all elements by the metallicity, will change the cooling rates by factors of a few. The difference is therefore typically somewhat smaller than the effect of neglecting metals or photo-ionization altogether, but still highly significant. While it was known that different elements dominate the cooling for different temperatures in CIE, we showed that photo-ionization both shifts the peaks due to individual elements to smaller temperatures and reduces their amplitude. Note that since photo-ionization over-ionizes the gas, this effect is similar (but not equivalent to) that found in non-equilibrium calculations without ionizing radiation \\cite[e.g.,][]{Sutherland1993, Gnat2007}. Because the importance of photo-ionization depends on the ionization parameter, the relative contributions of individual elements exposed to a fixed ionizing radiation field depends also on the gas density. Would dropping our assumption of ionization equilibrium have a large effect on the cooling rates? Ionizing radiation results in a plasma that is overionized relative to its temperature. Its effect is therefore similar to that of non-equilibrium ionization following rapid cooling (i.e., if the cooling time is shorter than the recombination times of the ions dominating the cooling, see e.g.,\\ \\citealt{Kafatos1973,Shapiro1976}). We therefore anticipate that the effect of non-equilibrium ionization will be much smaller for our cooling rates than for those that assume CIE. The assumptions that the gas is optically thin and exposed only to the meta-galactic background radiation are likely to be more important than the assumption of ionization equilibrium, particularly since non-equilibrium collisional cooling rates only differ from those assuming CIE by factors of a few or less \\cite[e.g.,][]{Schmutzeler1993,Sutherland1993,Gnat2007}. For column densities $N_{\\rm HI} > 10^{17}~\\cm^{-2}$ self-shielding becomes important and only part of the H-ionizing radiation will penetrate the gas cloud, which would particularly affect the cooling rates for $T\\la 10^5\\,\\K$. At higher temperatures line cooling is dominated by heavier ions, which can only be ionized by higher energy photons and which therefore remain optically thin up to much higher column densities. This is because the photo-ionization cross sections of H and He drop rapidly with increasing frequency for energies exceeding their ionization potentials. Moreover, for $T\\gg 10^4~\\K$ hydrogen is collisionally ionized to a high degree and consequently the optical depth for ionizing radiation will be significantly reduced. It is, however, far from clear that high column densities would reduce the effect of radiation. For self-shielded clouds the cooling radiation may itself be trapped, providing a source of ionizing radiation even in the absence of an external one \\cite[e.g.,][]{Shapiro1976,Gnat2007}. Moreover, gas clouds with columns that exceed $10^{17}~\\cm^{-2}$ are on average expected to be sufficiently close to a galaxy that local sources of ionizing radiation dominate over the background \\citep{Schaye2006,Miralda2005}. Ultimately, these issues can only be resolved if non-equilibrium cooling rates are computed including radiative transfer and if the locations of all relevant sources of ionizing radiation are known. It will be some time before it is feasible to carry out such a calculation in, say, a cosmological hydrodynamical simulation. In the mean time, we believe that our element-by-element calculation of the equilibrium cooling rates for an optically thin gas exposed to the CMB and an evolving UV/X-ray background provides a marked improvement over earlier treatments. In future publications we will present cosmological, hydrodynamical simulations using these cooling rates." }, "0807/0807.4442_arXiv.txt": { "abstract": "We use a combination of analytic tools and an extensive set of the largest and most accurate three-dimensional field theory numerical simulations to study the dynamics of domain wall networks with junctions. We build upon our previous work and consider a class of models which, in the limit of large number $N$ of coupled scalar fields, approaches the so-called `ideal' model (in terms of its potential to lead to network frustration). We consider values of $N$ between $N=2$ and $N=20$, and a range of cosmological epochs, and we also compare this class of models with other toy models used in the past. In all cases we find compelling evidence for a gradual approach to scaling, strongly supporting our no-frustration conjecture. We also discuss the various possible types of junctions (including cases where there is a hierarchy of them) and their roles in the dynamics of the network. Finally, we revise the Zel'dovich bound and provide an updated cosmological bound on the energy scale of this type of defect network: it must be lower than $10 \\, {\\rm keV}$. ", "introduction": "Introduction} As the early universe expanded and cooled down, it is believed to have gone through a series of phase transitions, at which networks of topological defects must necessarily have formed \\cite{KIBBLE}. The type of defect that forms and its specific properties depend on the particular details of each symmetry breaking, and hence there is a wide range of possibilities, which will lead to correspondingly very different cosmological consequences \\cite{VSH}. Domain walls are known to be pathological except if they are very light \\cite{ZEL}, but it has been claimed \\cite{SOLID} that if a domain wall network is frozen in co-moving coordinates (or `frustrates', as is often colloquially put) then it can naturally explain the observational evidence that points to a recent acceleration of the universe. It is clear that this scenario is subject to a number of observational constraints. For example the cosmic microwave background (CMB) data \\cite{WMAP5} severely constrains the characteristic scale of the network, $L$, which needs to be tiny in order not to give rise to exceedingly large CMB fluctuations. Also, recalling that the equation of state of a domain wall gas is given by \\begin{equation} w \\equiv \\frac{p}{\\rho} = -\\frac{2}{3} + v^2\\,, \\end{equation} with $v$ being the root-mean squared (RMS) velocity of the walls and that we require that \\begin{equation} w < -\\frac{1}{3}\\left(1+\\frac{\\Omega_m^0}{\\Omega_{DE}^0}\\right) \\lsim -\\frac{1}{2} \\end{equation} in order to accelerate the universe at the present time it is clear that only a non-relativistic value of the velocity (in fact, basically $v\\sim0$) would have any chance of working. The simplest domain wall models are known to reach a scaling regime (until they dominate the energy density of the universe), as first pointed out in \\cite{PRESS} and recently studied in detail in \\cite{AWALL,SIMS1,SIMS2}, and hence are obviously unable to satisfy these constraints. Nevertheless, it was thought that more complicated models, notably those having junctions, would eventually frustrate and therefore might conceivably be able to satisfy them. In previous work \\cite{IDEAL1,IDEAL2} we have studied the dynamics of domain wall networks with junctions, and investigated in detail energy, geometrical and topological constraints on their properties. This led us to develop an ideal class of models which includes, in the large $N$ limit, what we called the \\textit{ideal model} (that is, the best possible candidate for frustration). A series of analytic and numerical arguments then led us to a \\textit{no frustration conjecture}: even though one can build (purely by hand, as was done in \\cite{BATTYE,CARTER,IDEAL2}) special lattices that would be locally stable against small perturbations, no such configurations are expected to ever emerge from any realistic cosmological phase transition. (A much simpler analysis of the evolution of non-interacting and entangled cosmic string networks using a velocity-dependent one-scale model had already led to a similar result, at least in four space-time dimensions \\cite{NONINT}.) Our high-resolution numerical simulations of various realizations of the ideal class and other models showed clear evidence of a gradual approach to scaling, which was subsequently confirmed in \\cite{BMFLAT} (though note that the latter only reports on numerical simulations in Minkowski space thus neglecting the role of the expansion). Still, a fair criticism of our earlier work \\cite{IDEAL1,IDEAL2} would be that we only considered numerical simulations in two spatial dimensions, and the behavior could well be different in the three spatial dimensions of our universe. More recently \\cite{THIRD} we endeavored to eliminate this shortcoming and presented results of a series of massively parallel domain wall network numerical simulations of the `ideal' class of models in three spatial dimensions, which confirmed our earlier work, providing conclusive evidence for a gradual approach to scaling and hence strongly supporting our no frustration conjecture. The present article is the fourth in this series of papers, and its aims are to present the results of \\cite{THIRD} more thoroughly, to extend them in a number of different ways, and to discuss in some detail the corresponding cosmological implications. In Sect. \\ref{swal} we introduce various approaches to the study of domain walls---the microscopic description, the analytic modeling and the building of phenomenological models---that we will be using in the rest of the paper. In particular we highlight some relevant scaling solutions for domain wall networks and describe the ideal class of models which will be our primary focus. In Sect. \\ref{sfield} we discuss several technical details of our massively parallel field theory numerical code (which uses an improved version of the algorithm of Press, Ryden and Spergel \\cite{PRESS}, henceforth referred to as the PRS algorithm). In particular we describe in some detail how we identify the domain walls and how we measure their velocities. Sect. \\ref{ssim} contains our main numerical results: we describe the outcome of our various sets of three-dimensional high-resolution simulations of the ideal class of models. We also discuss the relevance of the key findings and contrast the results with those of other models. This then leads us to discuss, in Sect. \\ref{smod}, the analytic modeling of our results, including the cosmologically important asymptotic limit of the ideal class of models. In Sect. \\ref{sjun} we present an analysis of the possible hierarchies of junctions in the ideal and other models, as well as a brief discussion of the role of the junctions on the dynamics of the networks. Finally Sect. \\ref{scon} has our conclusions, as well as some comments on the implications of our results for cosmological scenarios involving domain walls and a brief outline of future endeavors in this area. ", "conclusions": "" }, "0807/0807.1098_arXiv.txt": { "abstract": "We present a study of the individual super star clusters (SSCs) in the low-metallicity galaxy SBS 0335-052 using new near-infrared and archival optical {\\it Hubble Space Telescope} observations. The physical properties of the SSCs are derived from fitting model spectral energy distributions (SEDs) to the optical photometry, as well as from the H$\\alpha$ and Pa$\\alpha$ nebular emission. Among the clusters, we find a significant age spread that is correlated with position in the galaxy, suggesting successive cluster formation occurred in SBS 0335-052 triggered by a large-scale disturbance traveling through the galaxy at a speed of $\\sim 35$ km s$^{-1}$. The SSCs exhibit $I$-band ($\\sim 0.8~\\mu$m) and near-IR ($\\sim 1.6-2.1~\\mu$m) excesses with respect to model SEDs fit to the optical data. We hypothesize that the $I$-band excess is dominated by a photoluminescent process known as Extended Red Emission; however, this mechanism cannot account for the excesses observed at longer near-IR wavelengths. From the cluster SEDs and colors, we find that the primary origin of the near-IR excess observed in the youngest SSCs ($\\lesssim 3$ Myr) is hot dust emission, while evolved red supergiants dominate the near-IR light in the older ($\\gtrsim 7$ Myr) clusters. We also find evidence for a porous and clumpy interstellar medium (ISM) surrounding the youngest, embedded SSCs: the ionized gas emission underpredicts the expected ionizing luminosities from the optical stellar continuum, suggesting ionizing photons are leaking out of the immediate vicinity of the clusters before ionizing hydrogen. The corrected, intrinsic ionizing luminosities of the two SSCs younger than $\\sim 3$~Myr are each $\\sim 5 \\times 10^{52}~{\\rm s}^{-1}$, which is equivalent to each cluster hosting $\\sim 5000$ O7.5 V stars. The inferred masses of these SSCs are $\\sim 10^6 M_\\odot$. ", "introduction": "Ancient globular clusters are among the oldest objects known, almost as old as the universe itself with ages estimated in excess of 12~Gyr \\citep{Vandenberg96, Freeman02}. As such, these clusters are valuable relics of the earlier universe, when violent and intense star formation was common. Indeed, globular clusters are ubiquitous around massive galaxies today \\citep{Harris91, Brodie06}, and they must have been formed prodigiously in the primordial universe given that $\\gtrsim 90\\%$ may have been subsequently destroyed \\citep{Fall01,Whitmore07}. However, the conditions required for the creation of globular clusters have puzzled astronomers for decades. For many years the prevailing belief was that globular clusters were simply formed by the gravitational collapse of density inhomogeneities in the early universe \\citep{Peebles68, Fall85}, and little was known about their early evolution. However, in the mid 1990's, observations using the {\\it Hubble Space Telescope} discovered extremely young, massive, and compact clusters in the local universe, the so-called ``super star clusters'' (SSCs) that have properties consistent with those expected for adolescent globular clusters \\citep[e.g.][]{Whitmore03}. The discovery of SSCs precipitated a major shift in our understanding of the conditions required for globular cluster formation. Although the evidence connecting young SSCs and ancient globular clusters is compelling, a critical issue remains: modern day SSCs and ancient globular clusters were formed in environments with very different metallicities. We have not yet observationally constrained what effect low metal abundances had on the formation of massive star clusters. Much effort has gone into understanding the formation of the {\\it first} stars in the universe \\citep[e.g. those stars formed out of truly primordial material,][]{Bromm03,Santoro06,Tumlinson07}, and it is clear that metallicity has a critical role in this regime. Even moderately low metal abundances may affect star formation in a number of ways, including dust formation, cooling and pressure, and the hardness of the resulting stellar spectra. \\begin{figure*}[!t] \\begin{center} \\includegraphics[scale=0.61]{f1.jpg} \\caption{{\\it HST} images of the observations used in this work. \\label{allims}} \\end{center} \\end{figure*} In an effort to investigate massive star cluster formation in an environment similar to that which might be found in primordial galaxies during the time ancient globular clusters were prolifically formed, we have targeted the galaxy SBS~0335-052 (E) for a detailed multi-wavelength study. SBS~0335-052 is a remarkable blue compact dwarf galaxy at a distance of 54 Mpc (NED\\footnote{NASA/IPAC Extragalactic Database}) that is well-known for its extremely low oxygen abundance of $\\sim 1/40Z_\\odot$ \\citep{Izotov90,Melnick92,Izotov01}. Unlike I~Zw~18 and SBS 0335-052W, which have slightly lower metallicities, SBS~0335-052 is undergoing a vigorous starburst with a star formation rate $\\gtrsim 1~M_\\odot$~yr$^{-1}$ and hosts extremely massive young clusters, making it ideal for this study \\citep{Thuan97, Thuan99, Hunt01, Dale01, Plante02, Hunt04, Houck04}. Here we present new near-infrared and archival optical high-resolution observations of SBS 0335-052 with the goal of studying SSC formation and evolution at low metallicity. Near-IR observations are particularly important for probing the dusty birth environments of embedded massive star clusters, and the wavelength range of $\\sim 1-2~\\mu$m is well-suited to assessing a cluster's evolutionary state because it samples the spectral energy distribution (SED) at the nexus of emission between dust and stellar light. The {\\it HST} observations and photometry are presented in \\S\\ref{hstsec} and \\S\\ref{photsec}. In \\S\\ref{propsec}, we discuss various properties of the SSCs. Specifically, we derive the physical properties of the SSCs from fitting model SEDs to the optical photometry, as well as from the H$\\alpha$, Pa$\\alpha$, and thermal radio nebular emission \\citep{Johnson08}. In addition, we present evidence for a porous and clumpy ISM in the youngest embedded clusters in SBS 0335-052 which can account for the apparently discrepant extinction estimates found in the literature. We also provide explanations for the observed {\\it I}-band ($\\sim 0.8~\\mu$m) and near-IR ($\\sim 1.6-2.1~\\mu$m) excesses. Evidence for successive cluster formation is given in \\S\\ref{suc_sec}, and finally, our conclusions are summarized in \\S\\ref{conclusions}. ", "conclusions": "We have presented a multi-wavelength study of the super star clusters in SBS 0335-052, the lowest metallicity galaxy with a star formation rate $\\gtrsim 1~M_\\odot$~yr$^{-1}$ known. We use new near-IR and archival optical {\\it HST} observations as well as free-free radio continuum measurements \\citep{Johnson08} of the galaxy to probe the stellar populations and the gas and dust components of the clusters. The main results of our work are summarized below: \\begin{enumerate} \\item{There is a significant age spread between the SSCs in SBS 0335-052 that is correlated with position in the galaxy. The ages are in the range $\\lesssim 3$ Myr to $\\sim 15$ Myr with the youngest clusters located in the south and the oldest clusters in the north. A large-scale disturbance, with a velocity of $\\sim 35$ km s$^{-1}$, appears to have triggered the successive cluster formation that we observe.} \\item{We find evidence for a porous and clumpy ISM surrounding the youngest SSCs in SBS 0335-052 (1 and 2). The measured ionizing luminosities from H$\\alpha$, Pa$\\alpha$, {\\it and} optically-thin, free-free radio emission \\citep{Johnson08} are lower than expected compared to the optical SEDs, suggesting a fraction of ionizing photons from the stellar continuum are escaping the strict confines of the clusters before ionizing hydrogen, that would have otherwise contributed to the measured ionized gas emission (i.e. H$\\alpha$, Pa$\\alpha$, and radio). The $A_V$'s of SSCs 1 and 2 derived from SED fitting ($\\sim 0.5$ mag) also indicate the existence of low-extinction regions that provide relatively unobscured lines-of-sight into the embedded clusters. The corrected, intrinsic ionizing luminosities of SSCs 1 and 2 are each $\\sim 5 \\times 10^{52}~{\\rm s}^{-1}$ (the equivalent of $\\sim 5000$ O7.5 V stars) and they each have total stellar masses of $\\sim 10^6 M_\\odot$.} \\item{An $I$-band (F791W, $\\sim 0.8~\\mu$m) excess with respect to model SEDs is observed for all of the SSCs in SBS 0335-052. The $I$-band magnitudes are $\\sim 0.5-0.7$ mag brighter than the model predictions. A similar result was found in the young massive clusters in NGC 4449 \\citep{Reines08} and attributed to a photoluminescent process known as Extended Red Emission. We hypothesize that the same mechanism dominates the $I$-band excess in SBS 0335-052.} \\item{All of the SSCs have red near-IR ($\\sim 1.6-2.1~\\mu$m) colors and large near-IR excesses with respect to model SEDs fit to the optical photometry. The red near-IR colors, however, are quantitatively different (redder) in the youngest SSCs (1 and 2) compared to the older SSCs (3-6), clearly indicating a different origin for the red colors. We have demonstrated that the near-IR light from the older clusters is dominated by evolved red supergiants, whereas the near-IR light from the youngest clusters is dominated by hot dust emission. The young SSCs 1 and 2 are also the only radio-detected clusters in the galaxy \\citep{Johnson08}.} \\item{We have proposed a scenario that can account for the apparently discrepant extinction estimates found in the literature for the starburst region encompassing the youngest SSCs (1 and 2) in SBS 0335-052. Our picture is consistent with all of the emission (at all wavelengths) being associated with the optically visible SSCs 1 and 2, and is based on our calculated absorptions of the optical and IR recombination lines relative to the (thermal) free-free radio emission \\citep{Johnson08}. We find that an ISM containing dense dust clumps harbouring large grains ($\\sim 1-2~\\mu$m), with a dust clump covering factor of $\\sim 60\\%$, can account for the absorption of the optical/near-IR recombination lines relative to the radio, as well as the large extinctions derived from previous mid-IR observations. Interclump regions containing diffuse dust can account for the low {\\it measured} extinctions derived from optical/near-IR observations.} \\end{enumerate} The main goal of this case study was to investigate the formation of massive star clusters in an extremely low metallicity environment that might be analogous to conditions in which the now ancient globular clusters were born. In general, the formation of massive star clusters in SBS 0335-052 appears to be similar to that found in higher metallicity counterparts, e.g. Henize~2-10 and Haro~3 \\citep{Johnson03, Johnson04}. These galaxies are all blue compact dwarfs (BCDs), and each of them currently hosts radio-detected massive star clusters with ages $\\lesssim 3$~Myr and masses of $\\sim 10^6 M_\\odot$, all of which are consistent with evolving into globular clusters in several billion years. Perhaps even the low metal abundance in SBS 0335-052 is insufficiently extreme to reflect primordial star formation; in fact, models predict that the transition from metal-free massive Population III stars to ``normal'' (less massive) Population II occurs at $Z/Z_\\odot\\lesssim 10^{-5}$ \\citep{Mackey03,Bromm04}. On the other hand, large metallicity fluctuations because of radiative feedback effects may enable ``hidden'' Population III star formation at slightly higher metal abundances \\citep{Tornatore07}. Clearly more detailed studies need to be carried out on metal-poor star-forming galaxies in the Local Universe, and comparative samples need to be obtained in order to investigate interactions and trends among star cluster and dust formation, environment, and metal enrichment." }, "0807/0807.1321_arXiv.txt": { "abstract": "The highly amplified magnetic fields suggested by observations of some supernova remnant (SNR) shells are most likely an intrinsic part of efficient particle acceleration by shocks. This strong turbulence, which may result from cosmic ray driven instabilities, both resonant and non-resonant, in the shock precursor, is certain to play a critical role in \\SC, \\NL\\ models of strong, \\CR\\ modified shocks. Here we present a Monte Carlo model of nonlinear diffusive shock acceleration (DSA) accounting for magnetic field amplification through resonant instabilities induced by accelerated particles, and including the effects of dissipation of turbulence upstream of a shock and the subsequent precursor plasma heating. Feedback effects between the plasma heating due to turbulence dissipation and particle injection are strong, adding to the \\NL\\ nature of efficient DSA. Describing the turbulence damping in a parameterized way, we reach two important results: first, for conditions typical of supernova remnant shocks, even a small amount of dissipated turbulence energy ($\\sim 10\\%$) is sufficient to significantly heat the precursor plasma, and second, the heating upstream of the shock leads to an increase in the injection of thermal particles at the subshock by a factor of several. In our results, the response of the non-linear shock structure to the boost in particle injection prevented the efficiency of particle acceleration and magnetic field amplification from increasing. We argue, however, that more advanced (possibly, non-resonant) models of turbulence generation and dissipation may lead to a scenario in which particle injection boost due to turbulence dissipation results in more efficient acceleration and even stronger amplified magnetic fields than without the dissipation. ", "introduction": "Observations of young supernova remnants (SNRs) suggest that strong collisionless shocks can simultaneously place a large fraction of the shock ram kinetic energy into \\rel\\ protons \\citep[e.g.,][]{BE87,MD2001, WarrenEtal2005} and amplify the ambient turbulent magnetic field by large factors \\citep[e.g.,][]{Cowsik80, RE92, BambaEtal2003, BKV2003, VL2003, UA2008}. This coupling of diffusive shock acceleration (DSA) and magnetic field amplification (MFA) is critically important because the self-generated magnetic field largely determines the efficiency of DSA, the maximum particle energy a given shock can produce, and the \\syn\\ emission from radiating electrons. The generation and dissipation of strong MHD turbulence in collisionless, multi-fluid plasmas is a complex process. Different \\NL\\ approaches to the modeling of the large scale structure of a shock undergoing efficient cosmic ray acceleration \\citep[e.g.,][]{MV82,AB86,BL2001,AB2006,VEB2006,PLM2006,ZPV2008} have predicted the presence of strong MHD turbulence in the shock precursor. However, an exact modeling of the shock structure in a turbulent medium, including nonthermal particle injection and acceleration, requires a nonperturbative, self-consistent description of a multi-component and multi-scale system including the strong MHD-turbulence dynamics. While a number of analytic models describing resonant and non-resonant amplification and damping of magnetic fluctuations have been proposed, these generally rely on the quasi-linear approximation that the fluctuations are small compared to the background magnetic field, i.e., $\\Delta B \\ll B_0$ \\citep[e.g.,][]{Wentzel74,BT2005,Kulsrud2005}. No consistent analytic description of magnetic turbulence generation with $\\Delta B \\gtrsim B_0$ exists. For these reasons, numerical models with varying ranges of applicability have been proposed which offer a compromise between completeness and speed \\citep[e.g.,][]{Bell2004,AB2006,VEB2006,ZPV2008}. In principle, the problem can be solved completely with few assumptions and approximations with particle-in-cell (PIC) simulations \\citep[e.g.,][]{Bell2004,Spitkovsky2008,NPS2008} or, in the assumption that electrons are not dynamically important, by hybrid models \\citep[e.g.,][]{WO1996, Giacalone2004}. However, modeling the \\NL\\ generation of \\rel\\ particles and strong magnetic turbulence in collisionless shocks is computationally challenging and PIC simulations will not be able to fully address this problem in \\nonrel\\ shocks for some years to come even though they can provide critical information on the plasma processes producing injection that can be obtained in no other way. In Appendix~\\ref{estimatesforpic} we outline the requirements that a PIC simulation must fulfill in order to tackle the problem of efficient DSA with non-linear MFA in SNR shocks. In the \\mc\\ approximation we use here, the plasma interactions are parameterized allowing us to study coupled \\NL\\ effects between the extended shock precursor and the gas subshock. In particular, we investigate the \\NL\\ effects caused by upstream plasma heating due to magnetic field dissipation. The importance of the dissipation of turbulence in the shock precursor can be illustrated by the following estimate. Suppose that in a shock wave of speed $u_0$, the turbulence is generated by the resonant cosmic ray (CR) streaming instability, so the energy density of the turbulence, $U_w$, evolves approximately as $u_0\\,dU_w(x)/dx=v_A\\,d\\Pcr(x)/dx$ \\citep[e.g.,][]{BL2001}, where $\\Pcr(x)$ is the CR pressure at position $x$ and $v_A$ is the \\Alf\\ speed. Ignoring all non-linear effects, the turbulence energy density at the shock positioned at $x=0$ is $U_w(0)=\\rho_0 u_0 \\vazero \\cdot \\Pcr(0)/(\\rho_0 u_0^2)$. The ratio $\\AccEff=\\Pcr(0)/(\\rho_0 u_0^2)$ characterizes the efficiency of acceleration and is typically assumed to be on the order of ten percent or more. In the above, $\\rho$ is the fluid density and the subscript ``0'' always indicates far upstream values. Suppose a fraction, $\\heatpar$, of this energy goes into heating of the thermal gas in the shock precursor so the energy density of the thermal plasma increases by $\\Delta U_{H}(0) = \\heatpar U_w(0)$ at $x=0$. Comparing $\\Delta U_{H}(0)$ with the internal energy density of the far upstream plasma, $\\epsilon_0$, we find \\begin{equation} \\label{upsilonequation} \\eta_H\\ = \\frac{\\Delta U_H(0)}{\\epsilon_0} \\approx \\heatpar \\AccEff \\frac{M_s^2}{M_A} \\ , \\end{equation} where $M_s$ is the sonic, and $M_A$ the \\Alf, Mach number. If $\\eta_H$ is large, the thermal gas in the shock precursor is strongly heated and this influences the subshock strength and the particle injection efficiency. In a non-linearly modified shock, a change in the injection efficiency causes the whole shock structure to change. For typical supernova remnant (SNR) parameters (e.g., $u_0 \\sim 3000$ \\kmps, $T_0 \\sim 10^4$ K, $n_0 \\sim 0.3$ \\pcc, and $B_0 \\sim 3$ \\muG), the ratio $M_s^2/M_A \\approx 250$, and values of $\\heatpar$ as low as a few to ten percent may be important. Because existing analytical descriptions of MHD wave damping rely on the quasi-linear approximation $\\Delta B \\ll B_0$, which is inapplicable for strong turbulence, and because an exact description of this process in the framework of non-linear DSA is currently impossible (see Appendix~\\ref{estimatesforpic}), we propose a parameterization of the turbulence damping rate. In doing this, we are pursuing two goals. First, we make some predictions connecting cosmic ray spectra, turbulent magnetic fields and plasma temperatures, which, in principle, can be tested against high resolution X-ray observations in order to estimate the heating of the thermal gas by turbulence dissipation. And second, once heating is included in our simulation in a parameterized fashion, we will be ready to implement more realistic models of turbulence generation and dissipation as they are developed. Our \\mc\\ simulation can be briefly summarized as follows \\citep[see][ and references therein for more complete details]{EJR90,JE91,VEB2006}. We describe particle transport in a plane shock by Bohm diffusion in a plasma flowing in the $x$-direction with speed $u(x)$. Particles move in small time steps as their local plasma frame momenta are `scattered' at each step in a random walk process on a sphere in momentum space. Some shock heated thermal particles are injected into the acceleration process when their history of random scatterings in the downstream region takes them back upstream. These particles gain energy and some continue to be accelerated in the first-order Fermi mechanism. This form of injection is generally called `thermal leakage' and was first used in the context of DSA in \\citet{EJE1981} \\citep[see also][]{Ellison82}. The magnetic field determining the random walk properties through the diffusion coefficient is the `seed' (interstellar) magnetic field after it has been amplified by a large factor by the CR streaming instability and compressed and advected downstream with the flow. The streaming instability in the non-linear regime ($\\Delta B \\gg B_0$) is described by the traditional quasi-linear equations, where the instability driving term is the CR pressure gradient. These quasi-linear equations are extrapolated into the non-linear regime in a parameterized fashion for lack of a more complete analytic description. The magnetic turbulence generated by the instability is assumed to dissipate at a rate proportional to the turbulence generation rate, and the dissipated energy is pumped directly into the thermal particle pool. An iterative scheme is employed to ensure the conservation of mass, momentum, and energy fluxes, thus producing a self-consistent solution of a steady-state, plane shock, with particle injection and acceleration coupled to the bulk plasma flow modification and to the magnetic field amplification and damping. Our results show that even a small rate of turbulence dissipation can significantly increase the precursor temperature and that this, in turn, can increase the rate of injection of thermal particles. The \\NL\\ feedback of these changes on the shock structure, however, tend to cancel so that the spectrum of high energy particles is only modestly affected. ", "conclusions": "\\label{conclusions} We have parameterized magnetic turbulence dissipation as a fraction of turbulence energy generation and included this effect in our \\MC\\ model of strongly \\NL\\ shocks undergoing efficient DSA. The energy removed from the turbulence goes directly into the thermal particle population in the shock precursor. The \\mc\\ simulation \\SCly\\ reacts to the changes in precursor heating and adjusts the injection of thermal particles into the DSA mechanism, as well as other \\NL\\ effects of DSA, accordingly. Our two most important results are, first, that even a small rate ($\\sim 10$\\%) of turbulence dissipation can drastically increase the precursor temperature, and second, that the precursor heating boosts particle injection into DSA by a large factor. The increase in particle injection modifies the low-energy part of the particle spectrum but, due to \\NL\\ feedback effects, does not significantly change the overall efficiency or the high energy part of the spectrum. Both the precursor heating and modified spectral shape that occur with dissipation may have observable consequences. The fact that the shock back-reaction to the increased injection prevents the acceleration efficiency from changing significantly is a clear consequence of the non-linear structure of the system. The boosted particle injection additionally smoothes the flow speed close to the subshock, which makes it harder for particles returning upstream to gain energy. As a result, the population of the high energy particles is not much changed (except for the decrease in the maximum particle momentum due to the reduction in the effective amplified magnetic field from dissipation) and, because those particles carry the bulk of the CR pressure $\\Pcr$ which drives the streaming instability, the amplification of the magnetic field is not strongly affected by the heating-boosted particle injection. The parameterization we use here is a simple one and a more advanced description of the turbulence damping may change our results. In our model the energy drained from the magnetic turbulence, at all wavelengths, is directly `pumped' into the thermal particles. Superthermal particles only gain extra energy due to heating because the thermal particles were more likely to return upstream and get accelerated. In a more advanced model of dissipation, where energy cascades from large-scale turbulence harmonics to the short-scale ones, the low energy CRs might gain energy directly from the dissipation. It is conceivable that cascading effects might increase the overall acceleration efficiency, the magnetic field amplification, and increase the maximum particle energy a shock can produce. It is also possible that non-resonant turbulence instabilities play an important role in magnetic field amplification \\citep[e.g.,][]{PLM2006}. This opens another possibility for the turbulence dissipation to produce an increase in the magnetic field amplification. For instance, \\citet{BT2005} proposed a mechanism for generating long-wavelength perturbations of magnetic fields by low energy particles. If such a mechanism is responsible for generation of a significant fraction of the turbulence that confines the highest energy particles, then the increased particle injection due to the precursor heating may raise the maximum particle energy and, possibly, the value of the amplified magnetic field. While our model is for the most part phenomenological as far as particle transport, injection and acceleration, magnetic field generation, and dissipation are concerned, it allows us to investigate the coupled nonlinear effects in a shock undergoing efficient DSA and MFA. The more efficient DSA is, the more basic considerations of momentum and energy conservation determine the shock structure and our model describes these effects fully." }, "0807/0807.3597.txt": { "abstract": "With only six known examples, M-dwarf debris disks are rare, even though M dwarfs constitute the majority of stars in the Galaxy. After finding a new M dwarf debris disk in a shallow mid-infrared observation of NGC~2547, we present a considerably deeper \\textsl{Spitzer}-MIPS image of the region, with a maximum exposure time of 15~minutes per pixel. Among sources selected from a previously published membership list, we identify nine new M dwarfs with excess emission at 24~$\\mu$m tracing warm material close to the snow line of these stars, at orbital radii of less than 1~AU. We argue that these are likely debris disks, suggesting that planet formation is under way in these systems. Interestingly, the estimated excess fraction of M stars appears to be higher than that of G and K stars in our sample. ", "introduction": "Even though about 80~\\% of the stars in the Galaxy are M~dwarfs \\citep{lad06}, only very few of these are known to harbor debris disks. This is even more surprising since many M~dwarfs do show signs of protoplanetary disks in earlier stages of evolution with similar frequency as more massive stars (e.g., \\citealp{and05}, \\citealp{muz06}, \\citealp{lad06b}) and we now know seven M dwarfs that harbor extrasolar planets (most recently \\citealp{end08}, see references therein). While protoplanetary disks contain large amounts of primordial gas and dust out of which planets form, debris disks are older and entirely made of collisionally evolving planetesimals and dust particles. Debris disks, thus, are secondary products of the planet formation process and their existence in a given system implies previous formation of protoplanets. Compared to optically thick protoplanetary disks, debris disks are fainter and optically thin. For a recent observational and theoretical review of debris disks, see \\citet{mey07} and \\citet{mor07}. Since the cold dust in debris disks is best studied at far-infrared (FIR) or submillimeter wavelengths, the \\textsl{Spitzer} Space Telescope, given its unprecedented FIR sensitivity, has revolutionized the research on the role of debris disks in star and planet formation. Most of the \\textsl{Spitzer} studies of debris disks have been carried out using the Multiband Imaging Photometer for Spitzer \\citep[MIPS,~][]{rie04}, operating at wavelengths of 24~$\\mu$m, 70~$\\mu$m, and 160~$\\mu$m. Prior to our study, only six debris disks around M dwarfs were known. Early studies based on data obtained by IRAS, the Infrared Astronomical Satellite, reported far-infrared excess emission towards a small number of field M stars \\citep{tsi88,mul89,mat91}, but few of these results could be confirmed subsequently \\citep[e.g.,][]{son02,ria06}. The first firm detection of an excess was the disk of AU~Mic \\citep{kal04,liu04b}, a member of the $\\beta$~Pic moving group with an age of $\\approx$12~Myr. The discovery observation was made with a stellar coronograph detecting scattered light in near-infrared broad-band filters. AU~Mic was also found to have excess emission at 70~$\\mu$m, but not at 24~$\\mu$m \\citep{che05}. Subsequently, in submillimeter radio observations at 850~$\\mu$m, \\citet{liu04a} identified GJ~182, a member of the Local Association Group with an age of $\\approx$50~Myr, as harboring a debris disk. GJ~182 does not show a far-infrared excess in the \\textsl{Spitzer}-MIPS study of \\citet{che05}. In \\textsl{Spitzer}-MIPS data of NGC~2547, a cluster at an age of 30--40~Myr, \\citet{you04} detected the third M~star debris disk (source 23 in our Table~\\ref{tabsigexcl}, see also Teixeira et al., \\textsl{in prep.}). TWA~7 was the fourth object reported; it is a member of the TW Hya association with an age of $<6$~Myr and excess emission reported at 24~$\\mu$m, 70~$\\mu$m, and submillimeter radio wavelengths \\citep{low05,mat07}. One field M dwarf (GJ~842.2) was found to have a debris disk \\citep[in submillimeter observations,][]{les06}. Finally, \\citet{gor07} identified one additional M~dwarf debris disk in NGC~2547 at 24~$\\mu$m, listed as source 22 in our Table~\\ref{tabsigexcl}. Only three sources (TWA~7, source 22, and source 23) have been identified as far-infrared excess sources at 24~$\\mu$m, while the other sources have been found at either 70~$\\mu$m (AU Mic) or submillimeter radio wavelengths. Since material of quite different distances from the central star is probed with these two techniques, debris disks identified in the mid-infrared and in submillimeter radio are not necessarily directly comparable to one another. Studying a sample of 62 field M dwarfs estimated to be older than 1~Gyr, \\citet{gau07} did not find any indication of excess emission at 24~$\\mu$m or 70~$\\mu$m indicative of debris disks (see also \\citealp{pla05,ria06}). In contrast to these \\textsl{Spitzer} results, \\citet{les06} discuss two sets of (sub-)millimeter data, including previously published observations, and find three detections of excess emission in a sample of 23 M~dwarfs with ages ranging from 20--200~Myr, corresponding to an excess fraction of $13^{+6}_{-8}$~\\%, which may not be significantly different from the disk fractions of earlier spectral types. After identifying an M~dwarf debris disk in earlier \\textsl{Spitzer} observations of NGC~2547 (source 23, see above), we performed ten times longer \\textsl{Spitzer}-MIPS 24~$\\mu$m observations of the same region in order to find out whether this was a singular object or part of a larger population. In the remainder of the introduction, we summarize the current knowledge about NGC~2547 before describing the observations in Section~\\ref{sec_obs}. We present the results in Section~\\ref{sec_results}, including remarks on the 70~$\\mu$m and 160~$\\mu$m data as well as variability at 24~$\\mu$m, discuss these results in Section~\\ref{sec_disc} and close with a summary in Section~\\ref{sec_summ}. ", "conclusions": "\\label{sec_disc} There are three different possibilities for the interpretation of the observed excess emission \\citep[e.g.,][]{stru06}. It could be due to remnant primordial material, due to a debris disk with continuously replenished dust, or it could be debris produced by a recent catastrophic collision of planetesimals, i.e., debris not in (quasi-) equilibrium. The dust in debris disks is produced and replenished by collisions, and then removed by processes such as the Poynting-Robertson (PR) drag, radiation pressure and stellar winds. Direct radiation pressure does not play a role for the low-luminosity stars that we consider \\citep{pla05}. The balance of dust production and these removal processes determines the amount of observable dust. %===> Which radius? In order to discuss the physical processes, we first need to determine which region around a central star is probed by the 24~$\\mu$m (excess) emission. As an estimate, we consider for a given stellar luminosity the orbital radius of a blackbody which is at a temperature that puts the peak of its SED at 24~$\\mu$m. For a dwarf star with a luminosity of $L$=0.01~$L_\\odot$, the corresponding radius is only 0.5~AU. %===> PR drag and timescale -> debris disk With the orbital radius of the dust particles constrained, we can now determine the lifetime of dust particles under the influence of the PR drag, which should yield an upper limit of the dust removal timescale \\citep[e.g.,][]{dom03}. The PR lifetime of a dust particle in orbit around a central star scales with particle size and density as well as the square of its orbital radius and is inversely proportional to the luminosity of the central star \\citep{bur79}. For micron-sized particles with a density of $\\rho=2.5$~g\\,cm$^{-3}$ \\citep{div93,che01} located at a distance of 1~AU from a central star with a luminosity of $L=0.01\\,L_\\odot$, the lifetime is still on the order of $1.7\\times10^5$ years. This value is ten times longer for 10~$\\mu$m-sized particles. Since the M~dwarfs discussed here have an age of 30--40~Myr, the particles would need to be replenished and thus cannot be primordial in order to still be observable at that age, indicating the debris disk nature of these systems and ruling out the first possibility mentioned in the beginning of this section. The lifetimes do not approach the cluster age until an orbital radius of 13~AU. Since other dust removal processes like stellar winds potentially operate at the same time, the derived timescale is in fact an upper limit. This result thus suggests that at an age of 30--40~Myr, M~dwarfs can have warm inner debris disks. Given that we know already seven M dwarfs that harbor extrasolar planets orbiting close to their host stars, this finding reinforces the idea that planets form around M stars. With M~star debris disks more frequent than previously thought, chance occurrences like catastrophic collisions of big planetesimals may not have to be invoked for their explanation, ruling out the remaining third possibility. %===> Remarks on radius It is interesting to note that a blackbody with a peak of its SED at 24~$\\mu$m has a temperature of 120~K, not far from the temperature of water sublimation, 153~K, defining the ``snow line'' \\citep{hay81}, separating the inner region of rocky planet formation from the outer region of icy planet formation and playing a role in the definition of ``habitable zones''. \\footnote{We can estimate a permissable range of dust temperatures for the dust we detect at 24\\,$\\mu$m. Comparing the ratio of an assumed 3$\\sigma$ excess at 8\\,$\\mu$m and the observed 24\\,$\\mu$m excess with that expected for blackbody emission suggests an approximate upper limit of $\\sim260$~K for a typical source without an 8\\,$\\mu$m excess. As a strict lower limit, we note that the M-dwarf debris disk around AU~Mic which does not exhibit 24\\,$\\mu$m excess emission has a dust temperature of $\\sim40$~K \\citep{ria06}.} For M stars, the habitable zone lies at orbital radius of about 0.1~AU \\citep{sca07}. Indeed, seven M dwarfs are now known to harbor one or several extrasolar planets at orbital radii of 0.02 to 2.3~AU (most recently \\citealp{end08}, see references therein), with masses spanning 0.016--2.1~$M_J$. The five planets with the smallest semi-major axes ($<0.1$~AU) have masses of $<0.07$~$M_J$. During the early evolution of an M dwarf, the snow line moves inwards due to the decreasing luminosity of the central star, enabling the formation of icy protoplanets within the first 1~Myr at orbital radii of 1--4~AU, not taking into account subsequent migration \\citep{ken07}. The snow lines of M~dwarfs in NGC~2547 have already moved further in when compared to their initial location, so that rocky protoplanets may have previously formed at orbital radii larger than the current snow line. %===> Comp to submm Given that some of the few previously known M dwarf debris disks have been found at submillimeter wavelengths, it is important to keep in mind that material traced in the far infrared is quite different from material traced in the submillimeter radio range. The latter corresponds to material at temperatures of $<20$~K at orbital radii of several hundred AU \\citep[e.g., ][]{les06}. Thus, different regions are probed by far-infrared and submillimeter observations and they cannot be directly compared. %===> SEDs We show the photometric spectral energy distributions of the five M~dwarfs with the strongest excess emission at 24~$\\mu$m in Fig.~\\ref{sedplots}. Their 24~$\\mu$m emission is about an order of magnitude stronger than the expected photospheric flux. To obtain approximate physical properties from the crudely constrained photometric SEDs, we performed fits using disk models by \\citet{rob07}. For one of our best-constrained disks, source 23, the fits suggest that slightly more than 1\\% of the total luminosity is due to the disk, even though the central object is a star with a luminosity of only $L=0.028$~$L_\\odot$. To within an order of magnitude, the resulting dust disk masses are $\\sim10^{-9}$~$M_\\odot$, corresponding to only a few percent of one lunar mass, but in some cases are barely constrained. As a major source of uncertainty, these mass estimates scale with the assumed dust properties which would be much better constrained with (sub-) millimeter data. For all but the densest regions of the disk, the model assumes small dust particles slightly larger than those in the diffuse ISM, and therefore, the mass estimates are a lower limit. Reviewing the submillimeter detections of M~dwarf debris disks, \\citet{les06} quote masses of one to 13 lunar masses for three debris disks in the M0 range. In marked contrast, \\citet{che05} use \\textsl{Spitzer}-MIPS data to derive a minimum dust mass for the debris disk of AU~Mic of only 10$^{-4}$ lunar masses, assuming a particle size of 0.2~$\\mu$m. This mass difference may indicate that submillimeter and far-infrared--detected debris disks are indeed quite different. However, a comparison of both methods towards the same objects would be needed to confirm this conjecture. %===> age effect? All of the M~dwarf members of NGC~2547 can be assumed to be of the cluster age, 30--40~Myr, possibly indicating that at this age, the evolutionary stage of M~dwarfs makes it easier to detect such debris disks. For debris disks around stars of earlier spectral types, there have already been several studies concerning age trends. In a study of main-sequence A stars older than 20~Myr, \\citet{rie05} find that the magnitude of excess at 24~$\\mu$m declines with a $t_0/t$ dependence, with $t_0\\approx 150$~Myr. Studying the double cluster h and $\\chi$ Persei in context with this result and other observations, \\citet{cur08a} conclude that the debris disk excess emission of A~stars peaks at about the age of this cluster, 13~Myr. On longer timescales, the evolution is less clear. While \\citet{tri08} did not find a clear age trend studying FIR excess emission of solar-type field stars with ages ranging from 200~Myr to 10~Gyr, \\citet{mey08}, also studying solar-type stars, found an excess rate decreasing from 18\\% at ages 3--30~Myr to 2\\% at ages 0.3--3~Gyr. \\citet{hil08} find that the 70~$\\mu$m excess emission of 328 FGK stars, tracing cooler dust compared to 24~$\\mu$m, peaks at ages of 30--100~Myr. For a solar-mass star, the debris for rocky planet formation peaks after a few million years for material at 1~AU \\citep{ken04,ken05}. To scale this result for an M dwarf, we can use a simple relation for the disk mass, $M_{\\rm disk} \\propto M_{\\rm star}$, and the collision timescale, $t_{\\rm coll} \\propto M_{\\rm star}^{-1/2}$ \\citep{ken08}. So, if the M dwarf is three times less massive than the Sun, we would expect planet formation to take about five times longer around an M dwarf, i.e., 10--20 Myr, not taking into account minor complications due to the moving snow line. This indicates that the M dwarfs in NGC~2547 are beyond but still close to the maximum predicted debris production. Especially with regard to stars of earlier spectral types, it is important to keep in mind that this argument only allows to compare the evolutionary timescale of debris at an orbital radius of 1~AU. For A stars, the radius probed by 24~$\\mu$m observations is beyond 20~AU, and the timescales at these radii may be very different. %===> known constraints on evolution Currently, the early debris-disk evolution of M dwarfs is poorly constrained from an observational point of view. \\citet{wei04} did not find significant mid-infrared excess (12 and 18~$\\mu$m) towards 16 members of the young, nearby TW Hydrae association and conclude that apparently any planet formation in the terrestrial planet region was rapidly completed. In a study targeting stars at age similar to the age of NGC~2547, \\citet{mam04} observed members of the 30-Myr-old Tucana-Horologium Association, including two M dwarfs, at a wavelength of $\\sim$10~$\\mu$m (\\textsl{N} band). They do not find excess emission towards the two M dwarfs, and their quoted uncertainties would allow excess emission at only $\\sim$10\\% of the expected photospheric flux level. Looking at our SEDs in Fig.~\\ref{sedplots}, if the excess emission is modeled correctly, it would be far more difficult to detect at 10$\\mu$m. The M dwarfs studied by \\citet{pla05}, and not found to exhibit excess emission at 11.7~$\\mu$m, include one young source and nine more objects with ages of $>$600~Myr. The single young star is GJ~3305, a member of the $\\beta$~Pic moving group at an age of $\\sim$12~Myr. \\citet{gau07} found that field M~dwarfs older than 1~Gyr do not show far-infrared excess emission. Due to the nearby location of these sources ($d<20$~pc), the absolute sensitivity of these observations is much higher than that of our dataset. This upper limit for the evolutionary timescale of M~dwarf debris disks indicates that planet formation around M dwarfs has ceased by this time. %===> earlier spectral types We finally consider the excess fractions of stars with earlier spectral types. While \\citet{gor07}, restricting the discussion to their completeness limit, found 30\\ -- 45\\% of the B--F members to show excess emission at 24~$\\mu$m, we interestingly find virtually no G~star excess emission at this wavelength. As pointed out above, while we find excess emission towards K stars, we may not be complete throughout their \\textsl{R-K} range, and many of the ``highly probable'' K members may be background giants. Even so, the 24~$\\mu$m excess rate for both G and K stars, using our conservative criterion for excess sources, is at most a few percent. In the most comprehensive sample of field stars to date, \\citet{tri08} find statistically indistinguishable excess rates for A, F, G, and K stars, with an average age of the solar-type sample of 5~Gyr. The fact that we do not find excess emission towards G~stars can possibly be understood using an argument similar to the one used in our discussion of M dwarf debris disks: given the cluster age, the solar-type stars are far beyond their peak in dust production. In marked contrast, the excess rate for M dwarfs is surprisingly high, even though it is a lower limit for two different reasons: first, our 24~$\\mu$m detections of M dwarfs are not complete and second, we use a conservative criterion to define excess sources. It remains unclear, however, to what degree the excess rates of G, K, and M stars are skewed due to contamination by non-members." }, "0807/0807.2578_arXiv.txt": { "abstract": "{The rate at which galaxies grow via successive mergers is a key element to understand the main phases of galaxy evolution.} {We measure the evolution of the fraction of galaxies in pairs and the merging rate since redshift $z \\sim 1$ assuming a ($H_0=70 km s^{-1} Mpc^{-1}, \\Omega_{M}=0.3$ and $\\Omega_{\\Lambda}=0.7$) cosmology.} {From the VIMOS VLT Deep Survey we use a sample of 6464 galaxies with $I_{AB} \\leq 24$ to identify 314 pairs of galaxies, each member with a secure spectroscopic redshift, which are close in both projected separation and in velocity.} {We estimate that at $z\\sim 0.9$, $10.9 \\pm 3.2\\%$ of galaxies with $M_B(z) \\leq -18-Qz$ ($Q=1.11$) are in pairs with separations $\\Delta r_p \\leq 20h^{-1}\\ kpc$, $\\Delta v \\leq 500$ km/s, and with $\\Delta M_B \\leq 1.5$, significantly larger than $3.8 \\pm 1.7 \\%$ at $z \\sim 0.5$; thus, the pair fraction evolves as $(1+z)^m$ with $m = 4.73 \\pm 2.01$. For bright galaxies with $M_B(z=0) \\leq -18.77$, the pair fraction is higher and its evolution with redshift is flatter with $m=1.50 \\pm 0.76$, a property also observed for galaxies with increasing stellar masses. Early-type pairs (dry mergers) increase their relative fraction from $3\\%$ at $z \\sim 0.9$ to $12\\%$ at $z \\sim 0.5$. The star formation rate traced by the rest-frame [OII] $EW$ increases by $26 \\pm 4\\%$ for pairs with the smallest separation $r_p \\leq 20h^{-1}kpc$. Following the prescription to derive merger timescales of Kitzbichler \\& White (2008) we find that the merger rate of $M_B(z) \\leq -18-Qz$ galaxies evolves as $N_{mg}=(4.96 \\pm 2.07)\\times 10^{-4}) \\times (1+z)^{2.20 \\pm 0.77} mergers\\ Mpc^{-3} Gyr^{-1}$.} {The merger rate of galaxies with $M_B(z) \\leq -18-Qz$ has significantly evolved since $z \\sim 1$ and is strongly dependent on the luminosity or stellar mass of galaxies. The major merger rate increases more rapidly with redshift for galaxies with fainter luminosities or stellar mass, while the evolution of the merger rate for bright or massive galaxies is slower, indicating that the slow evolution reported for the brightest galaxies is not universal. The merger rate is also strongly dependent on the spectral type of galaxies involved. Late-type mergers were more frequent in the past, while early-type mergers are more frequent today, contributing to the rise in the local density of early-type galaxies. About $20\\%$ of the stellar mass in present day galaxies with $log(M/M_{{\\odot}}) \\geq 9.5$ has been accreted through major merging events since $z = 1$. This indicates that major mergers have contributed significantly to the growth in stellar mass density of bright galaxies over the last half of the life of the Universe.} ", "introduction": "In the current hierarchical structure formation paradigm, the mass assembly in galaxies proceeds via a process of coalescence between increasingly more massive dark matter halos. This halo merging tree history can be quantified by a halo merger rate, measuring the growth of mass per average mass in a representative volume of the Universe. However, these models do not directly predict a growth of galaxy mass via mergers (\\cite{moore}), and the actual contribution of mergers to the evolution of galaxies remains poorly predicted. Merging two galaxies is potentially a very powerful process. It is possible that during major merger events, i.e. mergers where the two components have more or less the same mass, disks could be transformed into spheroidals, as predicted using detailed simulations (\\cite{combes}; \\cite{mihos}; \\cite{conselice06}). It is also expected that major merger events profoundly modify the spectrophotometric properties of the galaxies involved, for instance triggering a burst of star formation (e.g. \\cite{P05}). Galaxies in the process of merging are observed, however the contribution of this process to the evolution of the global galaxy population is not yet precisely constrained. Indirect evidence for merging is also inferred from other galaxy properties like the luminosity or mass function. The luminosity of the red bulge dominated population of galaxies is measured to increase since $z\\sim1$, part of which could be produced by mergers (\\cite{ilbert06}). it seems possible that the increase in the density of intermediate mass early-type galaxies since $z\\sim1$ may be happening at the expense of late-type galaxies involved in merging events (\\cite{tresse}). Merging is therefore potentially a very important physical phenomenon which could drive the evolution of galaxies along cosmic time. The average numbers of merger events needed to build a typical $M_*$ galaxy, the contribution of mergers to the mass growth of galaxies, or the identification of a prefered time in the life of the Universe when mergers were more frequent, are all important elements to help towards our understanding of galaxy evolution. It is then crucial to quantify the contribution of merging to the evolution process and its impact on important quantities like the cosmic star formation rate (e.g. \\cite{bouwens}; \\cite{tresse}; \\cite{woods}) or the global stellar mass density (e.g. \\cite{Arnouts}; \\cite{bundy}; \\cite{pozzetti}). To estimate the contribution of mergers to the formation and evolution of galaxies is not a trivial task. In the nearby Universe merger events can be identified aposteriori from perturbed morphologies, wisps, tails, and other peculiar signatures seen at low surface brightness. Only recently volume complete measurements of the merger rate in the nearby Universe are becoming available. In the Millennium catalogue, especially tailored to a volume complete identification of merging events, de Propris et al. (2007) use the relative velocity measured from spectroscopic redshifts to confirm true galaxy pairs in the process of merging. They find that the merger fraction is 2\\% at a mean redshift of 0.06, refining earlier estimates based on pair fraction (\\cite{P00}; \\cite{P02}). At higher redshifts, searching for evidence for past mergers becomes increasingly difficult, because the residual signatures of mergers often have a too low surface brightness. At redshifts $z\\geq0.3$, it is therefore easier to search for 'apriori mergers', encounters that are likely to lead to a merger event, rather than to look for 'aposteriori' signs of past mergers. When two galaxies are close together in space, and depending on their relative velocities, gravity is acting to bring them closer for a bound system that will merge. A measure of the merging frequency is then to count galaxy pairs with a separation and velocity difference such as they are likely to be gravitationally bound and destined to merge. By selecting pairs of galaxies with similar magnitudes and hence approximately with similar masses, one can focus on major merger events. They are able to significantly contribute to the mass assembly, to modify morphologies, as well as to significantly alter the star and gas content of the incoming galaxies. Assuming that a dynamically bound system of two galaxies will most likely evolve into one more massive galaxy, one can then derive the merger rate from the pair count. A major uncertainty of this estimator is the timescale upon which a merger will be completed. N-body simulations are then used to provide reasonable estimates of the merger timescales (\\cite{conselice06}; Kitzbichler \\& White, 2008). Cold dark matter simulations show that the evolution of the dark matter halos merger rate follows a power law $N_{mg} = N_{mg,0}(1+z)^m$ where $N_{mg,0} = N_{mg}(z=0)$ is the local value, and $m$ parameterizes the evolution. While some simulations predict that $m$ should have $2.5 \\leq m \\leq 3.5$ (Gottl\\\"\\o ber, 2001), measuring $m$ directly from galaxy samples is an important step to understand the evolution of galaxies. Many observational attempts have been carried out to track the evolution of the merger rate as a function of redshift (e.g. \\cite{burkey}; \\cite{carlberg94}; \\cite{yee}; \\cite{P97}; \\cite{lefevre00}; \\cite{P00}; \\cite{P02}; \\cite{conselice03}; \\cite{lin04}; \\cite{kartaltepe07a}; \\cite{lin08}; \\cite{lotz08}; \\cite{kampczyk}). Even though, $m$ remains poorly constrained with $0 \\leq m \\leq 6$, meaning either no evolution of the merger rate with cosmic time, or a strong evolution. Part of this large range of values can be understood as coming from the different criteria used to identify merger candidates, or the photometric band used to identify pairs (\\cite{bundy}). Furthermore, comparing measurements at low and high redshifts from different surveys is complicated due to the different selection functions used. At redshifts $z>0.3$, most pair counts so far have been performed from a measurement of the number of pairs observed on deep images, with either a photometric redshift of the galaxies (e.g. \\cite{conselice03}; \\cite{lotz08}), or a spectroscopic redshift of one of the galaxies in the pair (e.g. \\cite{P97}; \\cite{lefevre00}). The effect of contamination by galaxies projected along the line of sight producing false pairs is then estimated from galaxy counts, and the observed pair fraction is corrected to get an estimate of the true pair fraction. As redshift increases, projection effects become increasingly important making it difficult to estimate the true pair fraction, creating a fondamental uncertainty in the measurement of $m$. At $z\\sim1$ a galaxy with a luminosity $L_*$ has a 40\\% probability to have a galaxy with a similar magnitude but at a different redshift projected within an apparent radius of $20 h^{-1} kpc $ (\\cite{lefevre00}). To overcome these limitations, the most secure method to identify a physical pair of galaxies is to obtain a velocity measurement of each galaxy in the pair, enabling to identify pairs of galaxies which are most likely to be gravitationnaly bound. Only recently samples with spectroscopic redshifts for both galaxies in a pair are becoming available (\\cite{lin07}, \\cite{lin08}). In this paper we use for the first time a complete redshift survey to $z \\sim 1$ and as faint as $I_{AB}=24$ to securely identify pairs with {\\it both} galaxies having a spectroscopic redshift. We use the VIMOS VLT Deep Survey (VVDS) (\\cite{lefevre05a}), to search for galaxy pairs and to derive the pair fraction and the merger rate evolution. We present the galaxy sample and the methodology to build a pair sample in Section 2, we derive the pair fraction evolution in Section 3, and we examine the spectrophotometric properties of galaxies in pairs in Section 4. We compute the merger rate in Section 5. We evaluate the fraction of the stellar mass involved in mergers since $z\\sim1$ in Section 6, and conclude in Section 7. We adopt a $H_0=70 km s^{-1} Mpc^{-1}$, $\\Omega_{\\lambda}=0.7$ and $\\Omega_{m}=0.3$ cosmology throughout this work and magnitudes are given in the AB system. ", "conclusions": "\\label{discuss} Our results can be summarized as follows: (i) We find that $3.8 \\pm 1.7$, $9.4 \\pm 2.8$, and $10.9 \\pm 3.2\\ \\%$ of galaxies with $M_B(z) < -18 - Q(z)$ at $z \\sim 0.5, 0.7$ and $0.9$ respectively, are in pairs of galaxies with luminosities $\\Delta M_B \\leq 1.5$ and separations less than $20h^{-1}\\ kpc$. (ii) The evolution of the pair fraction with redshift is strongly dependent on the absolute luminosity or stellar mass of the brighter galaxy in the pair: it evolves more slowly for brighter or more massive galaxies than for faint galaxies. Using the VVDS alone, the pair fraction of galaxies with $M_B(z) < -18 -Q(z)$ is found to strongly evolves with redshift as $\\propto (1+z)^m$ with $m=4.46 \\pm 0.81$ for separations of $(100h^{-1}\\ kpc,500 km/s)$, while for brighter galaxies with $M_B(z) < -18.77 -Q(z)$, we find a slower evolution with $m=3.18 \\pm 1.34$. Combining VVDS data with low redshift measurements from de Propris et al. (2007), Patton et al. (2000) and Patton et al. (2002), and taking $r_p^{max}=\\ 20h^{-1}\\ kpc$, we similarly find $m=1.50 \\pm 0.76$ for bright galaxies with $M_B(z=0) \\leq -18 + 5log(h) \\sim -18.77$ and $m=4.73 \\pm 2.01$ for the fainter $M_B(z=0) \\leq -18$ sample. In addition, the evolution of the pair fraction is found to be stronger with $m = 3.13 \\pm 1.54$ for less massive galaxies with $log(M/M_{\\odot}) \\geq 9.5$, than for more massive galaxies with $log(M/M_{\\odot}) \\geq 10$ for which we find $m = 2.04 \\pm 1.65$. Low mass pairs are therefore contributing more to the evolution of the pair fraction than high mass pairs. (iii) The star formation rate of close pairs is enhanced at separations $r_p \\leq 150h^{-1}\\ kpc$. We find that the mean $EW(OII)$ in close pairs are larger by $26 \\pm 4\\%$ than the one derived for galaxies with larger separations. (iv) The evolution of the pair fraction is stronger for late-type pairs with $m_{late} = 4.74 \\pm 0.81$, than for early-type pairs with $m_{early} = 1.44 \\pm 0.93$. Late-type pairs are therefore contributing significantly more to the observed evolution of the pair fraction than early-type pairs in our $I_{AB} \\leq 24$ sample. (v) Using the merging timescale from Kitzbichler \\& White (2008), we find that the merger rate increases from $\\sim 12.3 \\times 10^{-4}$ to $\\sim 19.4\\times 10^{-4}\\ $mergers$\\ h^{3}\\ Mpc^{-3}\\ Gyr^{-1}$ from $z=0.5$ to $z=0.9$. The merger rate of galaxies with $M_B(z) < -18 -Q(z)$ evolves as $N_{mg}=(4.96 \\pm 2.07) \\times 10^{-4}) \\times (1+z)^{2.20 \\pm 0.77}$. Similarly to the pair fraction, we find that the merger rate evolves faster for fainter or less massive galaxies, with $m_{mg}=2.20 \\pm 0.77$ and $2.38 \\pm 1.57$ respectively, than for brighter or more massive galaxies with $m_{mg}=1.57 \\pm 0.44$ and $1.27 \\pm 1.67$ respectively. The merger rate is evolving more strongly for late-type mergers than for early-type mergers. We conclude that the observed evolution of the pair fraction and merger rate in our $I_{AB} \\leq 24$ sample is mostly driven by low mass late-type galaxies, while the pair fraction and merger rate of high mass early-type galaxies remains roughly constant since $z \\sim 1$. Therefore, the pair fraction or the merger rate are not universal numbers but rather are dependent on the luminosity or stellar mass, and on the spectral type of galaxies involved. Our finding that bright or massive galaxies experience a lower merger rate and a lower evolution of the merger rate extends to higher redshifts the results found in the local Universe by Patton \\& Atfield (2008). Taking into account this pair fraction and merger rate dependancy on galaxy luminosity and spectral type offers a first step to reconcile apparently inconsistent observations.Lotz et al. (2008) find a slow or no evolution of the merger rate and claim that they disagree with previous studies. When taking into account that their result is derived from bright $M_B \\leq -19.94 - 1.3 \\times z$ galaxies, their result is consistent with other studies like Conselice et al. (2003) or Le F\\`evre et al. (2000) which have been analysing fainter samples. The dependency of the merger rate and its evolution on luminosity or stellar mass is indeed a prediction from the latest simulations using advanced semi-analytic models as described in Kitzbichler and White (2008). At the limiting magnitudes or stellar masses of our sample, Kitzbichler and White (2008) predict that the merger rate decreases and evolves more slowly for galaxy samples with increasing luminosity or stellar mass, similar to the trend observed in our sample. The star formation rate is significantly enhanced in merging pairs with a net star formation increase of $\\sim 25\\%$ for these galaxies. Nevertheless, it accounts for only $12\\%$ to $3\\%$ of the global galaxy population from redshift $z \\sim 1$ to $z \\sim 0$ which is not sufficient to counteract the strong fading of the global star formation rate observed since $z\\sim1$. This may indicate that the gas reservoir of massive and intermediate mass galaxies has already been depleted at redshifts $z \\sim 1$, in agreement with their observed peak in star formation at $z\\sim 3.5$ (e.g. Tresse et al., 2007). It is then apparent that the decreasing SFR since $z\\sim 1$ is regulated by other physical processes like gas availability in the intergalactic medium, or feedback. Major merging events are largely dominated by pairs of late or mixed type galaxies, but while early-type mergers represent about $15\\%$ of the merging events of bright galaxies at $z \\sim 1$, they become approximately $25\\%$ of all mergers at $z \\sim 0.5$, which is in good agreement with previous results on dry mergers (e.g. \\cite{lin08}). This indicates that major mergers are efficient in lowering the number density of intermediate mass late-type galaxies to build up more early-type galaxies. We confirm that merging is one of the important physical processes driving galaxy evolution, with the observed galaxy merger rate undoubtedly closely linked to the hierarchical build up of dark matter galaxy halos, with a rapid mass accretion phase of massive halos since $z\\sim1$ (Abbas et al., 2008). Our finding that $\\sim20\\%$ of the stellar mass in present day massive galaxies has experienced a major merger since $z \\sim 1$ is an indication that major mergers are significantly contributing to the observed evolution of the stellar mass density since $z\\sim1$ (Bundy et al., 2005; Arnouts et al., 2007)." }, "0807/0807.0057_arXiv.txt": { "abstract": "We present three-dimensional magnetohydrodynamic (MHD) simulations of superbubbles, to study the importance of MHD effects in the interpretation of images from recent surveys of the Galactic plane. These simulations focus mainly on atmospheres defined by an exponential density distribution and the \\citet{1990ARA&A..28..215D} density distribution. In each case, the magnetic field is parallel to the Galactic plane and we investigate cases with either infinite scale height (constant magnetic field) or a constant ratio of gas pressure to magnetic pressure. The three-dimensional structure of superbubbles in these simulations is discussed with emphasis on the axial ratio of the cavity as a function of magnetic field strength and the age of the bubble. We investigate systematic errors in the age of the bubble and scale height of the surrounding medium that may be introduced by modeling the data with purely hydrodynamic models. Age estimates derived with symmetric hydrodynamic models fitted to an asymmetric magnetized superbubble can differ by up to a factor of four, depending on the direction of the line of sight. The scale height of the surrounding medium based on the Kompaneets model may be up to 50\\% lower than the actual scale height. We also present the first ever predictions of Faraday rotation by a magnetized superbubble based on three-dimensional MHD simulations. We emphasize the importance of MHD effects in the interpretation of observations of superbubbles. ", "introduction": "\\label{intro} The combined stellar wind and supernova ejecta of groups of O and B stars blow large bubbles in the interstellar medium. The largest of these bubbles, with size scales of 100 pc to 1 kpc are commonly referred to as superbubbles. The basic structure of a superbubble consists of a hot low-density interior, the cavity, surrounded by a cool shell of swept-up interstellar medium. The continuous formation and dissipation of superbubbles is an important factor in the energy balance of the interstellar medium, and determines the locations of different phases of the interstellar medium on large scales \\citep{mckee1977}. Compression of the interstellar medium in the shell may increase cooling and trigger the formation of a new generation of stars. Also, the ability of large superbubbles to break out of the disk of a galaxy and initiate an outflow of chemically enriched plasma and ionizing radiation from the disk into the halo has a profound influence on the evolution of galaxies. Several examples of well-defined superbubbles have been identified in the Galaxy (e.g. Heiles 1984; Maciejewski et~al. 1996; Normandeau et al. 1996; Heiles 1998; Ehlerov\\'a \\& Palous 1999; Callaway et~al. 2000; Reynolds et~al. 2001; McClure-Griffiths et al. 2002, Pidopryhora et~al. 2007). These have been studied in detail, thanks to their relative proximity. Observations with parsec-scale resolution of neutral and ionized gas reveal important details about the interaction between the hot ejecta and the interstellar medium. New high-resolution surveys of Galactic atomic hydrogen (\\HI) emission \\citep{taylor2003,mcclure2005,stil2006} have provided unprecedented images with morphological and kinematic information of Galactic superbubbles \\citep{1996Natur.380..687N,mcclure2003}. Physically interesting parameters are usually derived from the observations by means of analytic models that assume spherical symmetry \\citep{1975ApJ...200L.107C,1977ApJ...218..377W} or axial symmetry \\citep{Kompaneets1960,1999ApJ...516..843B}. In this paper we investigate the importance of MHD effects on physical quantities derived from observed superbubbles. The effect of the Galactic magnetic field is difficult to model because it introduces anisotropy in the medium, which requires three-dimensional numerical simulations. The first three-dimensional magnetohydrodynamic simulations of superbubbles were presented by \\citet{1998MNRAS.298..797T}, who discussed the importance of the magnetic field in the break-out of superbubbles from the Galactic disk. \\citet{1998MNRAS.298..797T} also described significant departures from spherical and axial symmetry in the shape of a magnetized bubble resulting from the interaction of the expanding superbubble with the Galactic magnetic field. The shape and the size of a superbubble depend on the strength and the geometry of the Galactic magnetic field as much as they depend on the density distribution of the ambient interstellar medium. The expanding superbubble in turn redefines the geometry of the interstellar medium and the Galactic magnetic field in a volume several hundred parsecs across. \\citet{korpi1999} and \\citet{deavillez2005} performed thee-dimensional MHD simulations that include the evolution of a superbubble in a supernova-driven turbulent multi-phase interstellar medium. The super bubbles in these simulations show significant departures form symmetry because of inhomogeneities in the medium in which the super bubble expands. \\citet{korpi1999} and \\citet{deavillez2005} found that a super bubble can break out into the halo in such a medium. Although these simulations provide valuable insight in the dynamics of superbubbles in a multi-phase interstellar medium, it is in general difficult to relate these simulations to specific observed super bubbles (see however Fuchs et~al. 2006 for a simulation of the Local Bubble). Astrophysical parameters derived from observations, such as the age and the energy of a superbubble, or the density distribution of the ambient medium have relied on symmetric analytic models that do not include a magnetic field. The departure from axial symmetry imposed by the magnetic field introduces systematic errors that have not been considered before. In this paper we present three-dimensional MHD simulations of superbubbles, and we explore the errors introduced by commonly used methods to determine basic parameters from observations. In particular, we study the axial ratio of the wind-blown cavity for different times and magnetic field configurations. The smaller simulation volume and time span used in our simulations compared to \\citet{1998MNRAS.298..797T} are more tailored to the latitude coverage of the Galactic plane surveys. We also calculate the first images of Faraday rotation by a magnetized superbubble derived from our simulations, and emphasize the importance of such simulations to make meaningful predictions in this area. We describe frequently used analytic models \\S 2 and detail our numerical setup and methods in \\S 3. Numerical results and the analysis of magnetic effects on derived parameters are presented \\S 4. The specific case of the W4 supperbubble is discussed in \\S 5. Faraday rotation by magnetized superbubbles is discussed in \\S 6, and conclusions are presented in \\S 7. ", "conclusions": "\\label{conc} We present three-dimensional MHD simulations of a superbubble evolving in a magnetized medium. In these simulations, we assume two different atmospheric models, exponential and Dickey \\& Lockman (1990), and two different magnetic field configurations, constant magnetic field and constant $\\beta$, for varying values of the magnetic field strength. With these simulations we aim to study the importance of MHD effects on the interpretation of observed superbubbles. As noted before by \\citet{1998MNRAS.298..797T}, a superbubble in a magnetized medium becomes significantly elongated along the magnetic field. We find that the axial ratio of the bubble in the Galactic plane is $\\sim$ 0.8 after 5 Myr and $\\sim 0.6$ after 12.5 Myr, depending on the age of the bubble and the strength of the magnetic field but not on the vertical structure of either the magnetic field or the gas. The elongated shape of the bubble in the Galactic plane may lead to significant errors in the age determination of magnetized superbubbles from observations using symmetric hydrodynamic models. The derived age depends on the location of the observer. The ratio of the smallest age to the largest age derived by observers assuming axial or spherical symmetry is found to be proportional to the square of the axial ratio of the bubble in the Galactic plane, which creates a discrepancy in the age of up to a factor $\\sim$4. We have analyzed systematic errors in the age of the bubble and the scale height of the ambient medium introduced by fitting the Kompaneets model to a magnetized superbubble looking along the magnetic field lines. The scale height may be underestimated by 30\\% to 50\\% and the age by 50\\%. In particular, we investigated the curiously small scale height of the interstellar medium near the W4 superbubble found by \\citet{1999ApJ...516..843B}. We re-analyzed HI data from the Canadian Galactic Plane Survey and found that the density of the ambient medium $n_{H}\\approx 2 \\rm{cm}^{-3}$ which is smaller than previously thought. This lower density helps to diminish the apparent age discrepancy between the W4 bubble and the star cluster OCL 352. However, our analysis of the systematic errors introduced by fitting the Kompaneets model shows that they are not big enough to explain the scale height of 25 pc found by \\citet{1999ApJ...516..843B} for the W4 region. We use the MHD simulations to predict the rotation measure distribution of superbubbles based on three-dimensional MHD simulations, and emphasize the importance of such simulations to make these predictions. As expected, the appearance of a magnetized superbubble depends on the perspective of the observer. If the observer looks along the magnetic field lines, the largest rotation measures are seen at the intersection of the shell with the Galactic plane. The rotation measure is increased at larger distances from the Galactic plane. If an observer in the Galactic plane looks perpendicular to the magnetic field lines, the rotation measures are much smaller, but most importantly most of the structure in rotation measure appears in projection on the low-density cavity, and not on the shell surrounding it. The simulations and analysis presented in this paper highlight the importance of three-dimensional MHD simulations of superbubbles evolving in the Galactic magnetic field to the interpretation of new high-resolution images of the Galactic plane at radio wavelengths from the International Galactic Plane Survey." }, "0807/0807.2863_arXiv.txt": { "abstract": "We perform a set of high-resolution, fully self-consistent dissipationless $N$-body simulations to investigate the influence of cold dark matter (CDM) substructure on the dynamical evolution of thin galactic disks. Our method combines cosmological simulations of galaxy-sized CDM halos to derive the properties of substructure populations and controlled numerical experiments of consecutive subhalo impacts onto initially-thin, fully-formed disk galaxies. We demonstrate that close encounters between massive subhalos and galactic disks since $z \\sim 1$ should be common occurrences in {\\LCDM} models. In contrast, extremely few satellites in present-day CDM halos are likely to have a significant impact on the disk structure. One typical host halo merger history is used to seed controlled $N$-body experiments of subhalo-disk encounters. As a result of these accretion events, the disk thickens considerably at all radii with the disk scale height increasing in excess of a factor of $2$ in the solar neighborhood. We show that interactions with the subhalo population produce a wealth of distinctive morphological signatures in the disk stars, many of which resemble those being discovered in the Milky Way (MW), M31, and in other disk galaxies, including: conspicuous flares; bars; low-lived, ring-like features in the outskirts; and low-density, filamentary structures above the disk plane. We compare a resulting dynamically-cold, ring-like feature in our simulations to the Monoceros ring stellar structure in the MW. The comparison shows quantitative agreement in both spatial distribution and kinematics, suggesting that such observed complex stellar components may arise naturally as disk stars are excited by encounters with subhalos. These findings highlight the significant role of CDM substructure in setting the structure of disk galaxies and driving galaxy evolution. ", "introduction": "The currently favored cold dark matter (CDM) paradigm of hierarchical structure formation (e.g., \\cite[Blumenthal et al. 1984]{Blumenthal_etal84}), predicts significant dark matter halo substructure in the form of small, dense, self-bound {\\it subhalos} orbiting within the virialized regions of larger host halos (e.g., \\cite[Klypin et al. 1999]{Klypin_etal99}). Observational probes of substructure abundance thus constitute fundamental tests of the CDM model. Due to the fact that most subhalos associated with galaxy-sized host halos lack of a significant luminous component, a constraint on the amount of substructure in these systems may be obtained via their gravitational influence on galactic disks. If there is a considerable subhalo population, it may produce strong tidal effects and induce distinctive gravitational signatures which might be imprinted on the structure and kinematics of the host galactic disk. Thus, establishing the role of substructure in shaping the fine structure of galactic disks may prove fundamental in informing our ideas about global properties of galaxy formation and evolution. Significant theoretical effort has been devoted to quantifying the resilience of galactic disks to infalling satellites (e.g., \\cite[Quinn \\& Goodman 1986; Velazquez \\& White 1999; Font et al. 2001; Gauthier et al. 2006; Read et al. 2008; Villalobos \\& Helmi 2008]{Quinn_Goodman86,Walker_etal96,Velazquez_White99,Font_etal01, Gauthier_etal06,Read_etal08,Villalobos_Helmi08}). Despite their usefulness, most earlier investigations suffered basic shortcomings that limited their applicability. For example, some considered encounters of single satellites with galactic disks, a set-up which is at odds with CDM predictions of multiple, nearly contemporaneous accretion events. Other studies made ad hoc assumptions about the orbital parameters and internal structures of the infalling systems. Consequently, it remains uncertain whether these earlier investigations faithfully captured the responses of galactic disks to halo substructure in a cosmological context. Here we address this issue using a hybrid approach that combines cosmological simulations to derive the merger histories of galaxy-sized CDM halos with controlled numerical experiments of consecutive subhalo impacts onto $N$-body realizations of fully-formed disk galaxies. We demonstrate that accretion histories of the kind expected in {\\LCDM} models are capable of severely perturbing galactic disks and generating a wealth of distinctive signatures in the structural and kinematic properties of disk stars. The resulting morphological features are similar to those being discovered in the Milky Way (MW), M31, and in other disk galaxies. We also show that satellite-disk interactions produce dynamically-cold, ring-like features around galactic disks that are quantitatively similar to the Monoceros ring in the MW. This suggests that such observed stellar structures may arise naturally as a result of subhalo-disk encounters, which can excite perturbations in {\\it disk stars}. These findings imply that detailed observations of galactic structure may be able to distinguish between competing cosmological models by determining whether the detailed structure of galactic disks is as excited as predicted by the CDM paradigm. ", "conclusions": "" }, "0807/0807.3674_arXiv.txt": { "abstract": "High-redshift, dust-obscured galaxies -- selected to be luminous in the radio but relatively faint at 850\\,$\\mu$m -- appear to represent a different population from the ultra-luminous submillimeter- (submm-) bright population. They may be star-forming galaxies with hotter dust temperatures or they may have lower far-infrared luminosities and larger contributions from obscured active galactic nuclei (AGN). Here we present observations of three $z\\sim2$ examples of this population, which we term {\\it submm-faint radio galaxies -- SFRGs} in CO(3--2) using the IRAM Plateau de Bure Interferometer to study their gas and dynamical properties. We estimate the molecular gas mass in each of the three SFRGs ($8.3\\times10^{9}$M$_\\odot$, $<5.6\\times10^{9}$M$_\\odot$ and $15.4\\times10^{9}$M$_\\odot$, respectively) and, in the case of \\rga, a dynamical mass by measurement of the width of the CO(3--2) line ($8\\times10^{10} \\csc^2i$\\,M$_\\odot$). While these gas masses are substantial, on average they are 4$\\times$ lower than submm-selected galaxies (SMGs). Radio-inferred star formation rates ($<{\\rm SFR_{\\rm radio}}>=970$\\,M$_{\\odot}\\,$yr$^{-1}$) suggest much higher star-formation efficiencies than are found for SMGs, and shorter gas depletion time scales ($\\sim$11\\,Myr), much shorter than the time required to form their current stellar masses ($\\sim$160\\,Myr; $\\sim$10$^{11}$\\,M$_\\odot$). By contrast, SFRs may be overestimated by factors of a few, bringing the efficiencies in line with those typically measured for other ultraluminous star-forming galaxies and suggesting SFRGs are more like ultraviolet- (UV-)selected star-forming galaxies with enhanced radio emission. A tentative detection of \\rga\\ at 350\\,$\\mu$m suggests hotter dust temperatures -- and thus similar gas-to-dust mass fractions -- as the SMGs. We conclude that SFRGs' radio luminosities are larger than would naturally scale from local ULIRGs given their gas masses or gas fractions. ", "introduction": "Submm surveys have provided an efficient probe of star-formation activity in ultraluminous infrared (IR) galaxies (ULIRGs, $>$10$^{12}$\\,L$_\\odot$) in the distant Universe (e.g., Smail, Ivison \\& Blain 1997; Hughes et al.\\ 1998; Barger et al.\\ 1998), with bright submm emission providing unambiguous evidence of massive quantities of dust, heated predominantly by young stars rather than AGN (e.g., Chapman et al.\\ 2003a; Alexander et al.\\ 2005; Menendez-Delmestre et al.\\ 2007; Valiante et al.\\ 2007; Pope et al.\\ 2008). Before the availability of the Atacama Large Millimeter Array (ALMA), confusion will continue to limit the sensitivity of current submm surveys. As a result, many ULIRGs fall below the detection limits due to variations in their spectral energy distributions (SEDs) -- usually parameterized in terms of dust temperature ($T_{\\rm d}$) -- meaning that entire populations of star-forming galaxies may have been missed by submm surveys. For a fixed far-infrared luminosity (FIR), a galaxy with a higher $T_{\\rm d}$ will be weaker in the submm at 850$\\mu$m than a galaxy with a lower $T_{\\rm d}$. Specifically, raising $T_{\\rm d}$ from the canonical $\\sim$35~K for SMGs to 45~K will result in a factor $\\sim$10$\\times$ drop in 850$\\mu$m flux density (Blain 1999; Chapman et al.\\ 2004). These galaxies should, though, be accessible in the radio waveband, regardless of their specific SEDs, since the radio correlates with the integrated FIR emission (Helou et al.\\ 1985) with a small $\\sim0.2$dex dispersion and no observable dependence on SED type. However, there is potential for large AGN contaminations in the radio, as has often been the case with mid-IR selection of $z>1$ ULIRGs (e.g., Houck et al.\\ 2005; Yan et al.\\ 2005, 2007; Sajina et al.\\ 2007; Weedman et al.\\ 2006a, 2006b; Desai et al.\\ 2006) and the facilities required to provide the high-resolution, multi-frequency radio data needed to decontaminate the samples (e.g., Ivison et al.\\ 2007a) are not yet available. Substantial populations of apparently star-forming galaxies at $z\\sim2$ have been uncovered through deep 1.4-GHz radio continuum observations, many of which are not detected at submm wavelengths with the current generation of instruments (Barger et al.\\ 2000; Chapman et al.\\ 2001, 2003b, 2004a -- hereafter C04). These galaxies are luminous in the radio and spectroscopy suggests that star formation is powering their bolometric output (there is little or no sign of high-ionization emission lines, characteristic of AGN in their UV/optical spectra). These galaxies could, in principle, span a range in properties from deeply obscured AGN to far-IR-luminous starbursts. In the latter case, one would expect a different SED from a typical SMG -- a higher $T_{\\rm d}$ for instance. These {\\it submm-faint radio sources} have a large volume density at $z\\sim2$, even larger than the SMGs (Haarsma et al.\\ 1998; Richards et al.\\ 1999; Chapman et al.\\ 2003a, C04; Barger et al.\\ 2007). There are $\\rho = 2 \\times 10^{-5}$\\,Mpc$^{-3}$ radio sources with L$_{\\rm 1.4 GHz}>10^{31}$ ergs\\,s$^{-1}$\\,Hz$^{-1}$ at $z\\sim2$ compared with $\\rho = (6.2\\pm2.3) \\times 10^{-6}$~Mpc$^{-3}$ for SMGs brighter than 5\\,mJy at 850\\,$\\mu$m at the same epoch (Chapman et al.\\ 2003b, 2005). As essentially all of these SMGs form a subset of these radio sources (Chapman et al.\\ 2005; Pope et al.\\ 2006; cf.\\ Ivison et al.\\ 2002) this implies $\\sim14\\times10^{-6}$~Mpc$^{-3}$ luminous radio sources remain undetected at submm wavelengths. Understanding the exact properties of these galaxies is therefore of great importance. If they are all forming stars at the rates implied by their radio luminosities, they would triple the observed SFR density (SFRD) at $z\\sim2$. By contrast, if their radio luminosity comes from a mix of star formation and AGN, they have less impact on the global SFRD but they increase the highly obscured AGN fraction at these epochs (e.g., Daddi et al.\\ 2007a; Casey et al.\\ 2008) and contribute substantially to black hole growth. Together with other observations, the redshifted cooling emission lines of CO allow us to assess and compare the energy source of SFRGs with that of SMGs and other distant star-forming galaxies via measurements of their gas and dynamical masses. In this paper, we present the results of a pilot study with the \\pdbilong\\ to detect molecular gas in SFRGs through the rotational CO(3--2) line emission. In \\S~2 we describe the sample properties and observations both with \\pdbishort\\ and other facilities. Section \\S~3 presents the CO(3--2) detections and limits obtained from the \\pdbishort\\ observations, \\S~4.1 estimates gas properties, star-formation rates and efficiencies, and \\S~4.2 compares the SFRGs to other galaxy populations. Finally \\S~5 discusses the results and places them in a broader galaxy evolution context. Throughout we assume a cosmology with $h=0.7, \\Omega_\\Lambda = 0.72, \\Omega_M = 0.28$ (e.g., Hinshaw et al.\\ 2008). \\begin{figure} \\centering \\includegraphics[width=5.1cm,angle=0]{f1a.ps} % \\includegraphics[width=4.7cm,angle=-90]{f1c.ps} % \\includegraphics[width=5.1cm,angle=0]{f1b.ps} % \\includegraphics[width=4.7cm,angle=-90]{f1d.ps} % \\vskip0.4cm \\caption{{\\bf top panels:} CO(3--2) spectra for the two candidate detections. The spectra are shown smoothed with a 50~km/s boxcar filter, and with respect to the zero velocity offsets defined from the H$\\alpha$ emission line redshift (red dashed line). The best-fit Gaussian profile is shown for the emission line in \\rga, along with the Uv-Inferred redshift from inter-stellar absorption lines (blue dashed line). {\\bf bottom panels:} velocity-averaged spatial maps of CO emission, from $-$1500 to $-$800\\,km\\,s$^{-1}$ (\\rga) where contours are from $-$1 to 5$\\sigma$ in steps of $\\sigma$ (0.05\\,mJy\\,beam$^{-1}$). \\rgc\\ does not represent a formal CO detection. We measure a significance of 3.2$\\sigma$ integrating over the full band). The field of view, on a side, is 35\\arcsec, with the size of the beam shown to the lower left. Both CO emitters lie exactly at the radio source position to within 1\\arcsec.} \\label{Pic} \\end{figure} ", "conclusions": "$\\bullet$ We conclude that the radio luminosities of these SFRGs are higher for their overall mass (gas plus stellar) than for the SMGs (given that if the radio-SFRs are over-estimated for one class, they could well be for both). $\\bullet$ We note that SMGs in general, and also these SFRGs, are outliers of the stellar mass-SFR correlation (Daddi et al.\\ 2007a), probably due to the higher efficiency in forming stars for a similar stellar mass and CO luminosity. Lower gas masses in the SFRGs would imply even higher SFEs than the SMGs. By contrast, if the SFRs are significantly over-estimated by the radio, or the CO-to-H$_2$ conversion were significantly different from SMGs, then SFRGs could have similar efficiencies to typical ULIRGs. Together with the apparent low-efficiency star-forming (U)LIRGs from Daddi et al.\\ (2008), the SMGs and SFRGs with SFRs several to 10$\\times$ larger than the Daddi et al.\\ galaxies suggest roughly equal numbers of stars being formed in both high- and low-efficiency modes at $z\\sim2$. Massive galaxies are being built in an impressive variety of modes in the $z=1-3$ peak star-formation period. $\\bullet$ If the radio-inferred SFRs are correct, then these SFRGs are more efficient star formers than SMGs, and cannot obviously be interpreted as {\\em scaled up} versions of local ULIRGs as Tacconi et al.\\ (2006, 2008) have argued is the case for SMGs. The SFRGs' radio luminosities are larger than would naturally scale from local ULIRGs given the gas masses or gas fractions. These observed gas masses and star-formation properties may be typical of the SFRG population and further work is justified to explore this population with improved statistics. $\\bullet$ Our results underscore the fact that ultraluminous galaxies in the high-redshift Universe have been discovered with a wide range of star-forming efficiencies, the SFRGs apparently being one extreme. Massive galaxies are likely being built in a variety of modes in the $z=1-3$ peak star-formation period." }, "0807/0807.1447_arXiv.txt": { "abstract": "{} {To investigate the chemical relations between complex organics based on their spatial distributions and excitation conditions in the low-mass young stellar objects IRAS~16293-2422 ``A'' and ``B''.} {Interferometric observations with the Submillimeter Array have been performed at 5$\\arcsec\\times3\\arcsec$\\ (800$\\times$500~AU) resolution revealing emission lines of HNCO, CH$_3$CN, CH$_2$CO, CH$_3$CHO and C$_2$H$_5$OH. Rotational temperatures are determined from rotational diagrams when a sufficient number of lines are detected.} {Compact emission is detected for all species studied here. For HNCO and CH$_3$CN it mostly arises from source ``A'', CH$_2$CO and C$_2$H$_5$OH have comparable strength for both sources and CH$_3$CHO arises exclusively from source ``B''. HNCO, CH$_3$CN and CH$_3$CHO have rotational temperatures $>$200~K implying that they arise from hot gas. The $(u,v)$-visibility data reveal that HNCO also has extended cold emission, which could not be previously determined through single dish data. } {The relative abundances of the molecules studied here are very similar within factors of a few to those found in high-mass YSOs. This illustrates that the chemistry between high- and low-mass objects appears to be relatively similar and thus independent of luminosity and cloud mass. In contrast, bigger abundance differences are seen between the ``A'' and ``B'' source. For instance, the HNCO abundance relative to CH$_3$OH is $\\sim$4 times higher toward ``A'', which may be due to a higher initial OCN$^-$ ice abundances in source ``A'' compared to ``B''. Furthermore, not all oxygen-bearing species are co-existent, with CH$_3$CHO/CH$_3$OH an order of magnitude higher toward ``B'' than ``A''. The different spatial behavior of CH$_2$CO and C$_2$H$_5$OH compared with CH$_3$CHO suggests that successive hydrogenation reactions on grain-surfaces are not sufficient to explain the observed gas phase abundance of the latter. Selective destruction of CH$_3$CHO may result in the anti-coincidence of these species in source ``A''. These results illustrate the power of interferometric compared with single dish data in terms of testing chemical models.} ", "introduction": "The envelopes of some low-mass protostars contain many complex organic molecules\\footnote{In this paper molecules are considered complex if they contain more than four atoms.} \\citep{blake1994,dishoeck1995,cazaux2003,bottinelli2004b,bottinelli2004,bottinelli2007,jorgensen2005b,jorgensen2005a,sakai2006,sakai2007}. This raises the question whether these are low-mass versions of ``hot cores'', chemically very rich environments in high-mass star forming regions that are thought to have their origin in grain-mantle evaporation and the subsequent rapid gas phase reactions. The low-mass counterpart is sometimes called a ``hot corino''. The presence of warm material has long been suggested from modeling of the SEDs of these sources \\citep[e.g.,][]{adams1987,jorgensen2002,shirley2002} and has been firmly established by molecular excitation studies \\citep{blake1994,dishoeck1995,ceccarelli2000} and by resolved interferometric imaging \\citep[see e.g.,][]{chandler2005,bottinelli2004,jorgensen2005}. However, in some sources the emission peaks offset from the continuum source \\citep[e.g.,][]{chandler2005}. This offset is in disagreement with the ``hot corino'' hypothesis for low-mass stars in which complex molecules evaporate through passive heating and is more in favor of other explanations such as the presence of disks or outflows that create shocks in the envelope. Currently, there is an ongoing debate on whether the emission of complex organics comes from ``hot corinos'' or from other types of regions. Also, the extent to which the observed organics are first generation molecules created in the ice or second generation produced in the gas is still an open question. The aim of this work is to map the emission of complex organics to address the latter question, namely to determine their most likely formation mechanism. One method to study chemical links between species is to study molecular abundances through single-dish surveys in a large number of sources \\citep{vdtak2000b,vdtak2003,ikeda2001,bisschop2007a}. An alternative method is to look for spatial correlations by interferometric observations of a single source through which it is possible to distinguish compact and extended emission as well as the exact location of the compact emission. The species that are studied here are the nitrogen-bearing species, HNCO and CH$_3$CN and the oxygen-bearing species CH$_2$CO, CH$_3$CHO, and C$_2$H$_5$OH that are commonly found toward or surrounding hot cores. The molecules CH$_2$CO, CH$_3$CHO and C$_2$H$_5$OH are proposed to be linked by successive hydrogenation on the surfaces of grains \\citep{tielens1997}. The source studied in this paper, IRAS~16293-2422, is a very well studied and chemically rich low-mass YSO. It is a binary with its main components, named ``A'' and ``B'', separated by 5\\arcsec \\citep[corresponding to 800~AU at 160~pc][]{mundy1992}. From single dish observations and modeling, \\citet{dishoeck1995}, \\citet{ceccarelli1999,ceccarelli2000} and \\citet{schoier2002} concluded that emission from several organic molecules arises from a compact region. The data for some species such as H$_2$CO and CH$_3$CN are significantly better fitted if a ``jump'' in the abundance at 80--90~K due to grain-mantle evaporation is assumed in a spherical circum-binary envelope \\citep{ceccarelli2000,schoier2002}. Through interferometric observations it is possible to distinguish between the emission from the cold extended envelope and more compact emission as well as the peak location: is the emission coming from source ``A'' or ``B''? Previous observations have shown that some complex species are much more abundant toward one source than the other, such as CH$_3$OCHO which is more prominent toward the ``A'' source \\citep[][]{bottinelli2004,huang2005,kuan2004,remijan2006}. The nitrogen-bearing species HNCO and CH$_3$CN have previously been detected toward IRAS~16293-2422 through single dish observations \\citep{dishoeck1995,cazaux2003}. They often have high rotational temperatures in star forming regions implying that they are present in hot gas \\citep{olmi1993,zinchenko2000,cazaux2003,bisschop2007a}. This is also the case for the oxygen-bearing molecule C$_2$H$_5$OH \\citep{ikeda2001,bisschop2007a}. In contrast, CH$_2$CO and CH$_3$CHO are often detected in high-mass sources with low rotational temperatures \\citep{ikeda2002,bisschop2007a}. For CH$_3$CHO higher rotational temperatures are also found, but these can be due to the {\\it b}-type transitions that are radiatively pumped and do not represent the actual kinetic temperature of the gas \\citep{nummelin2000,turner1991}. Recent laboratory experiments by \\citet{bisschop2007c} have shown that it is possible to explain the gas phase abundances of C$_2$H$_5$OH if a solid state formation route through CH$_3$CHO hydrogenation is assumed. The observed absence of CH$_3$CHO in hot gas combined with its detection in cold ices \\citep{keane2001,gibb2004}, however, implies that it must be destroyed at higher ice temperatures, before evaporation commences, or directly after evaporation in the gas phase. A comparison of the excitation properties as well as interferometric observations in which the spatial distribution of the species can be determined are good tools to further elucidate the chemical relations between these species. This paper is structured as follows: \\S~\\ref{obs} presents the Submillimeter Array (SMA) observing strategy as well as maps and spectra; \\S~\\ref{an} explains the analysis of the interferometric observations in the $(u,v)$-plane and the fitting of rotational diagrams; \\S~\\ref{results} presents the results of the rotational diagram and flux analysis; \\S~\\ref{disc} discusses the relative abundances of the different species with respect to each other as well as the implications for the chemistry; \\S~\\ref{sum} summarizes the main conclusions. ", "conclusions": "\\label{sum} We have performed an interferometric study of the complex organic species HNCO, CH$_3$CN, CH$_2$CO, CH$_3$CHO and C$_2$H$_5$OH toward the low-mass protostar IRAS~16293-2422 with the SMA. Previously published data from \\citet{kuan2004} are used to determine abundances relative to CH$_3$OH. The main conclusions of this work are: \\begin{itemize} \\item The emission from both HNCO and CH$_3$CN is compact and is seen toward both sources in the binary, with only 10--20\\% arising from source ``B''. Additionally, the lowest excitation line of HNCO shows extended emission suggestive of its presence in a cold extended envelope. The relatively higher abundances with respect to CH$_3$OH in source ``A'' may originate from higher initial abundances of OCN$^-$ in the ice. \\item For CH$_2$CO and C$_2$H$_5$OH only one line is detected due to compact emission. For CH$_2$CO these lines are detected with similar strength toward both sources and C$_2$H$_5$OH is clearly detected in source ``B'', but due to line-blending it is difficult to determine the flux for source ``A''. Compact hot emission for CH$_3$CHO is detected only toward source ``B''. Comparison with previous single dish observations by \\citet{cazaux2003} suggests that a cold extended component is present as well. If CH$_2$CO, CH$_3$CHO and C$_2$H$_5$OH are related through successive hydrogenation on the surfaces of grains, the same spatial behavior is expected for all three species. Since this is not observed, it suggests that hydrogenation reactions on grain surfaces alone cannot account for the observed gas phase abundance ratios. The difference between the two IRAS~16292-2422 sources can be explained if CH$_3$CHO would be selectively destroyed in source ``A'' right before or after grain-mantle evaporation. \\end{itemize} The discussion in this paper demonstrates the strength of clearly resolved interferometric observations for studies of the chemistry in star forming regions. First, molecules that have both extended and compact emission can easily be identified based on $(u,v)$-visibility analysis. In single dish observations these components will be averaged together giving rotational temperatures in between that of the hot and cold component. Additionally, compact emission arising from different components in the same single-dish beam can be separated, which leads in some cases to abundance ratios that can vary over an order of magnitude for different sources as exemplified in Table~\\ref{comp} for HNCO and CH$_3$CHO. For HNCO the presence of a cold extended component could not be derived from the single dish observations, whereas for CH$_3$CHO it is the hot component that is not detected. Further interferometric studies are needed to elucidate the chemical relations between complex organics." }, "0807/0807.3442_arXiv.txt": { "abstract": "We estimate the local number density of sources of ultra-high-energy cosmic rays (UHECRs) based on the statistical features of their arrival direction distribution. We calculate the arrival distributions of protons above $10^{19}$ eV taking into account their propagation process in the Galactic magnetic field and a structured intergalactic magnetic field, and statistically compare those with the observational result of the Pierre Auger Observatory. The anisotropy in the arrival distribution at the highest energies enables us to estimate the number density of UHECR sources as $\\sim 10^{-4}~{\\rm Mpc}^{-3}$ assuming the persistent activity of UHECR sources. We compare the estimated number density of UHECR sources with the number densities of known astrophysical objects. This estimated number density is consistent with the number density of Fanaroff-Reily I galaxies. We also discuss the reproducability of the observed {\\it isotropy} in the arrival distribution above $10^{19}$ eV. We find that the estimated source model cannot reproduce the observed isotropy. However, the observed isotropy can be reproduced with the number density of $10^{-2}$-$10^{-3}~{\\rm Mpc}^{-3}$. This fact indicates the existence of UHECR sources with a maximum acceleration energy of $\\sim 10^{19}$ eV whose number density is an order of magnitude more than that injecting the highest energy cosmic rays. ", "introduction": "\\label{introduction} The origin of ultra-high-energy cosmic rays (UHECRs) has been highly unknown in spite of prolonged effort to construct larger UHECR observatories and to detect more events \\cite{nagano00}. In cosmic ray spectrum, a sharply spectral steepening at around $10^{20}$eV has been predicted theoretically by interactions with photopion production with cosmic microwave background (CMB) photons, known as Greisen-Zatsepin-Kuz'min (GZK) steepening \\cite{greisen66,zatsepin66}. The feature of this steepening observed by the High Resolution Fly's Eye (HiRes) \\cite{abbasi04a,abbasi08} and the Pierre Auger Observatory (PAO) \\cite{yamamoto07,abraham08} implies that astrophysical sources are much more dominant than top-down sources (for a review, see Ref. \\cite{bhattacharjee00}) at the highest energies, though physical reasons for the extension of the energy spectrum beyond the GZK energies reported by the Akeno Giant Air Shower Array (AGASA) have been not understood yet \\cite{takeda98,takeda03}. Several objects to accelerate particles up to $10^{20}$ eV have been suggested, but there has been little observational evidence which is a UHECR source (see Ref. \\cite{bhattacharjee00,torres04} and reference therein). The arrival distribution of UHECRs has information on the distribution of their sources, also including that on intergalactic and the Galactic magnetic field (IGMF and GMF). The AGASA and PAO reported statistically significant anisotropy at small angular scale in the arrival distribution \\cite{takeda99,pao07,pao08}. The small-scale anisotropy implies point-like sources, and enables us to constrain the number of nearby UHECR sources. The source number density, $n_s$, is one of important parameters to investigate the nature of UHECR sources. Comparing the estimated number density with the number densities of known astrophysical objects, we can constrain the object-type of UHECR sources. Several authors have estimated as $n_s \\sim 10^{-5}$-$10^{-6}~{\\rm Mpc}^{-3}$ using the published AGASA data above $4 \\times 10^{19}$ eV assuming the same injection rate over all sources \\cite{yoshiguchi03,blasi04,kachelriess05,takami06,takami07}, and we also constrained the source number density as $10^{-4}~{\\rm Mpc}^{-3}$, assuming a model that the injection rate is proportional to the luminosity of galaxies \\cite{takami07}. The PAO reported the correlation between the arrival directions of UHECR above $5.7 \\times 10^{19}$ eV and the positions of nearby active galactic nuclei (AGNs) listed in the 12th Veron-Cetty \\& Veron catalog \\cite{veron06} within the angular scale of $3.1^{\\circ}$ \\cite{pao07,pao08}. However, most of the PAO-correlated AGNs are Seyfert galaxies and LINERs, which have much weaker activity than radio-loud AGNs like Fanaroff-Reily II (FR II) galaxies \\cite{moskalenko08}. To understand whether these galaxies with weak activity are really UHECR sources or not, another information on UHECR sources is required. The PAO estimated a lower limit of the UHECR source number as 61 based on anisotropy in the arrival distribution of their detected events, simply assuming Poisson statistics and not taking into account UHECR propagation \\cite{pao08}. This is certainly a lower limit of the number of the sources, but is not quantity which can be compared with astronomical observables because we do not know what radius they are included in. Their number density is an observable. Taking into account UHECR propagation, we can also estimate plausible value of it, not {\\it limit}. In this study, we simulate the arrival distribution of UHECRs above $10^{19}$ eV, taking into account their propagation process in intergalactic and the Galactic space. We extract an anisotropy signal from the simulated arrival distribution and then compare this signal with that observed by the PAO above $5.7 \\times 10^{19}$ eV to estimate the source number density. The reports for the correlation by the PAO with nearby large-scale structure \\cite{pao07,pao08,kashti08,takami08c} show that it plays a crucial role on the arrival distribution of UHECRs. Thus, we adopt the models of the IGMF and source distribution which reproduce the local structure actually observed around the Milky Way, developed in our previous work \\cite{takami06}. We also discuss the isotropy in the arrival distribution at $\\sim 10^{19}$ eV and estimate the source number density for the lower energies. The composition is assumed to be pure protons. This paper is structured as follows. In section \\ref{methods}, our calculation method for calculating UHECR arrival distribution and a statistical method to estimate the source number density are explained. In section \\ref{results}, the calculation results are shown and we discuss the results and conclude in section \\ref{discussion}. ", "conclusions": "\\label{discussion} In this study, we estimated the number density of UHECR sources based on the statistical features of the arrival distribution of UHECRs observed by the PAO. We simulated the arrival distributions of protons above $10^{19}$ eV, taking into consideration their propagation process in the Galactic and intergalactic space. The IGMF model adopted was associated with the observed large-scale structure. Comparing the simulated arrival distributions with the data observed by the PAO statistically, we estimated the number density which can best reproduce the observational result. The anisotropy signal above $5.7 \\times 10^{19}$ eV led to $n_s \\sim 10^{-4}~{\\rm Mpc}^{-3}$ which is consistent with that of FR I galaxies, which are a candidate to accelerate protons up to $10^{20}$ eV. We also focused on isotropy in the arrival distribution at around $10^{19}$ eV, and then found $n_s \\sim 10^{-2}$-$10^{-3}~{\\rm Mpc}^{-3}$, which is one or two order of magnitude larger than that estimated from the anisotropy. In this calculation, cosmological evolution of UHECR sources is not considered. Since protons with $\\sim 10^{19}$ eV can reach the earth from sources at the distance of 1Gpc ($z \\sim 0.25$), which is comparable with the energy-loss length of such protons by Bethe-Heitler pair creation, it might be possible that the cosmological evolution could reproduce the isotropy even if the local number density of UHECR sources is comparable with $\\sim 10^{-4}~{\\rm Mpc}^{-3}$. However, a typical evolution factor is $(1+z)^3$, and therefore the number density at $z \\sim 0.25$ is only twice more than the local one. Such a small factor could not change the estimated number density by two order of magnitude less. The difference between the two estimated number densities is significant. The difference can be interpreted as the evidence of the existence of UHECR sources which can accelerate protons up to $\\sim 10^{19}$ eV assuming the persistent activity of UHECR sources. $n_s \\sim 10^{-2}$-$10^{-3}~{\\rm Mpc}^{-3}$ is comparable with the number densities of bright galaxies or Seyfert galaxies. The proton acceleration up to $10^{19}$ eV might be common in the universe. We could also interpret this difference as transient activity to emit UHECRs. If bursting sources, like GRBs, are assumed, apparent number density, which corresponds to $n_s$ estimated in this study, depends on the threshold of UHECR energies because the dispersion of the time-delay is larger at lower energies \\cite{miralda96}. However, quantitative discussion on this possibility exceeds the scope of this study. It is next study of ours. In order to reproduce isotropy in the arrival distribution of UHECRs above $10^{19}$ eV, sources within 5 Mpc were artificially neglected. In these sources, Cen A, the nearest radio-loud AGN, is involved. The isotropy at around $10^{19}$ eV implies that it is not UHECR sources in our persistent source model. However, radio-loud AGNs have been plausible site for particle acceleration up to the highest energy \\cite{torres04}, and the two of the highest energy events of the PAO are correlated with the position of Cen A within $\\sim 3^{\\circ}$. Whether Cen A is really a UHECR source or not is a key to understand the mechanism to generate the highest energy cosmic rays. If Cen A is not a UHECR source, we can interpret that the source of cosmic rays arriving in the direction of Centaurus is behind Cen A and more distant, so that the strong anisotropy is not generated. The 2 events towards Cen A are also positionally correlated with NGC 5090 with $\\sim 3^{\\circ}$, a radio galaxy with the distance of $\\sim 40$ Mpc. This might be a real source though the activity is weak like the other PAO-correlated AGNs. These events are also towards Centaurus cluster whose distance is about 40 Mpc. An idea to generate UHE particles is cluster accretion shock \\cite{kang96,inoue05,inoue07}. However, this scenario can accelerate protons up to $\\sim 10^{19}$ eV. If Cen A is a UHECR source, we propose several possibilities not to generate the anisotropy. One is UHECR composition. In the case of heavy dominated composition at energies above $10^{19}$ eV, the trajectories of UHECRs are deflected $Z$ times more than those of protons, and the discussion on the isotropy in this paper is not applied. Several composition measurements imply the existence of some fraction of heavy elements for all large uncertainty on hadronic interaction models in extensive air shower \\cite{unger07,glushkov07}. The other is that the UHECR production is transient. For a bursting source, the energy of cosmic rays observed at present is in narrow energy range since the time-delay of cosmic rays depends on their energies \\cite{miralda96}. In this brief picture, the 2 events are arrived from Cen A, and the lower energies will come in the future. The detailed discussion on the two possibilities is our near future plan. The estimated number densities have some uncertainty because of the small number of detected events. The uncertainty to determine the number density at the highest energies can be reduced by increasing observed events \\cite{takami07} and significant anisotropy would be observed in the arrival distribution at $\\sim 10^{19}$ eV in the near future \\cite{waxman97,kashti08}. The dramatically increase of detected events in the near future will provide us more useful information on UHECR sources. \\subsubsection*{Acknowledgements:} The work of H.T. is supported by Grants-in-Aid from JSPS Fellows. The work of K.S. is supported by Grants-in-Aid for Scientific Research provided by the Ministry of Education, Science and Culture of Japan through Research Grants S19104006." }, "0807/0807.1196.txt": { "abstract": "%context heading (optional) {The Cepheid period-luminosity (PL) relation is unquestionably one of the most powerful tools at our disposal for determining the extragalactic distance scale. While significant progress has been made in the past few years towards its understanding and characterization both on the observational and theoretical sides, the debate on the influence that chemical composition may have on the PL relation is still unsettled.} %aims heading (mandatory) {With the aim to assess the influence of the stellar iron content on the PL relation in the $V$ and $K$ bands, we have related the $V$-band and the $K$-band residuals from the standard PL relations of Freedman et al. (2001) and Persson et al. (2004), respectively, to [Fe/H].} %method {We used direct measurements of the iron abundances of 68 Galactic and Magellanic Cepheids from FEROS and UVES high-resolution and high signal-to-noise spectra.} %results heading (mandatory) {We find a mean iron abundance ([Fe/H]) about solar ($\\sigma$ = 0.10) for our Galactic sample (32 stars), $\\sim$ -0.33 dex ($\\sigma$ = 0.13) for the Large Magellanic Cloud (LMC) sample (22 stars) and $\\sim$ -0.75 dex ($\\sigma$ = 0.08) for the Small Magellanic Cloud (SMC) sample (14 stars). Our abundance measurements of the Magellanic Cepheids double the number of stars studied up to now at high resolution. The metallicity affects the $V$-band Cepheid PL relation and metal-rich Cepheids appear to be systematically fainter than metal-poor ones. These findings depend neither on the adopted distance scale for Galactic Cepheids nor on the adopted LMC distance modulus. Current data do not allow us to reach a firm conclusion concerning the metallicity dependence of the $K$-band PL relation. The new Galactic distances indicate a small effect, whereas the old ones support a marginal effect. } %Conclusions {Recent robust estimates of the LMC distance and current results indicate that the Cepheid PL relation is not Universal. } { ", "introduction": "Since the dawn of modern astronomy the Cepheid Period-Luminosity (PL) relation is a key tool in determining Galactic and extragalactic distances. In spite of its fundamental importance, the debate on the role played by the chemical composition on the pulsation properties of Cepheids is still open, with different theoretical models and observational results leading to markedly different conclusions. From the theoretical point of view pulsation models by different groups lead to substantially different results. Linear models (e.g. Chiosi et al. 1992; Sandage et al. 1999; Baraffe \\& Alibert 2001), based on nonadiabatic radiative models, suggest a moderate dependence of the PL relation on the metallicity. The predicted change at log(P) = 1 is less than 0.1 mag for metal abundances ranging from the SMC (Z=0.004) to the Galaxy (Z =0.02), independent of wavelength. Nonlinear convective models (e.g. Bono et al. 1999; Caputo et al. 2000; Caputo 2008) instead predict a larger dependence on the same interval of metallicity: the change is 0.4 mag in $V$, 0.3 mag in $I$ and 0.2 mag in $K$, again at log(P) = 1. Moreover, the predicted change in these latter models is such that metal-rich Cepheids are fainter than metal-poor ones, at variance with the results of the linear models. Fiorentino et al. (2002) and, more recently, Marconi, Musella \\& Fiorentino (2005) investigations, also based on nonlinear models, suggest that there may be also a dependence on the helium abundance. On the observational side, the majority of the constraints comes from indirect measurements of the metallicity, mostly in external galaxies, such as oxygen nebular abundances derived from spectra of H II regions at the same Galactocentric distance as the Cepheid fields (e.g. Sasselov et al. 1997; Kennicutt et al. 1998; Sakai et al. 2004). These analyses indicate that metal-rich Cepheids are brighter than metal-poor ones (hence at variance with the predictions of nonlinear convective models), but it is important to note that the results span a disappointingly large range of values (see Table~1 and Fig.~1). More recently Macri et al. (2006) found, by adopting a large sample of Cepheids in two different fields of NGC~4258 and the [O/H] gradient based on H II regions provided by Zaritsky et al. (1994), a metallicity effect of $\\gamma=-0.29\\pm0.09$ mag/dex. This galaxy has been adopted as a benchmark for estimating the metallicity effect, since an accurate geometrical distance based on water maser emission is available (Herrnstein et al. 2005). However, in a thorough investigation Tammann et al. (2007) suggested that the flat slope of the Period-Color relation of the Cepheids located in the inner metal-rich field could be due to a second parameter, likely helium, other than the metal abundance. Furthermore, Bono et al. (2008) found, using the new and more accurate nebular oxygen abundances for a good sample of H II region in NGC~4258 provided by D\\`iaz et al. (2000), a shallower abundance gradient. In particular, the new estimates seem to suggest that both the inner and the outer field might have a mean oxygen abundance very similar to LMC. They also found a very good agreement between predicted and observed Period-Wesenheit ($V,I$) relation. Nonlinear convective models predict for this relation a metallicity effect of $\\gamma=+0.05\\pm0.03$ mag/dex. On the basis of independent distances for 18 galaxies based on Cepheid and on the Tip of the Red Giant Branch, Tammann et al. (2007) found a small metallicity effect ($\\gamma=-0.017\\pm0.113$ mag/dex). On the other hand, Fouqu\\'e et al. (2007) using a sample of 59 Galactic Cepheids whose distances were estimated using different methods -- HST trigonometric parallaxes (Benedict et al. 2007), revised Hipparcos parallaxes (van Leeuwen et al. 2007), infrared surface brightness method (Fouqu\\'e \\& Gieren 1997), and interferometric Baade-Wesselink method (Kervella et al. 2004), zero-age-main-sequence fitting of open clusters (Turner \\& Burke 2002) -- found no significant difference between optical and Near-Infrared (NIR) slopes of Galactic and LMC Cepheids (Udalski et al. 1999; Persson et al. 2004). \\begin{figure}[h] \\centering \\includegraphics[height=0.35\\textheight, width=0.40\\textwidth]{5661f1.ps} \\vspace*{0.5truecm} \\caption{ Comparison of recent results for the metallicity sensitivity of Cepheid distances. FM1990: Freedman \\& Madore (1990); G1994: Gould (1994); Ko1997: Kochanek (1997); S1997: Sasselov et al (1997); Ke1998: Kennicutt et al (1998); F2001: Freedman et al (2001); U2001: Udalski et al (2001); C2002: Ciardullo et al (2002); Sa2004: Sakai et al (2004); Sto2004: Storm et al (2004); Gro2004: Groenewegen et al (2004); M2006: Macri et al. (2006); Sah2006: Saha et al. (2006); Be2007: Benedict et al. (2007); Fo2007: Fouqu\\'e et al. (2007); Tam2007: Tammann et al. (2007); B2008: Bono et al. (2008). See Table~1. } \\label{Fig1} \\end{figure} An alternative approach is to measure directly the metal content of Cepheid stars, which, so far, has been attempted only by few studies, primarily focused on stars of our own Galaxy (Luck \\& Lambert, 1992; Fry \\& Carney 1997; Luck et al. 1998; Andrievsky et al. 2002a,b,c; Luck et al. 2003; Andrievsky et al. 2004). Fry \\& Carney (1997, hereafter FC97), for instance, have derived iron and $\\alpha$-element abundances for 23 Galactic Cepheids from high resolution and high signal-to-noise spectra. They found a spread in [Fe/H] of about 0.4 dex, which they claim is real. Using approximately half of their sample, the stars belonging to clusters or associations, they have made a preliminary evaluation of metallicity effects on the zero point of the PL relation, finding that metal-rich Cepheids are brighter than metal-poor ones. Thus, finding a result similar to the studies based on indirect measurements of the metallicity. The impressive observational effort carried out by Andrievsky and collaborators (Andrievsky et al. 2002a, 2002b, 2002c; Luck et al. 2003; Andrievsky et al. 2004; Kovtyukhet al. 2005b) has, instead, taken advantage of high resolution spectra of 130 Galactic Cepheids (collected with different instruments at different telescopes) in order to determine their chemical composition and study the Galactic abundance gradient. The sample covers a range of Galactocentric distances from 4 to 14 kpc. The emerging picture can be best described by a relatively steep gradient (about -0.14 dex kpc$^{-1}$) for Galactocentric distances less than 7 kpc, followed by a much shallower slope ($\\approx$ -0.03 dex kpc$^{-1}$) between 7 and 10 kpc, a discontinuity at approximately 10 kpc and a nearly constant metallicity of about -0.2 dex towards larger Galactocentric distances, out to about 14 kpc. In relation to our work, it is important to note that Andrievsky and collaborators did not investigate the effects of the chemical composition on the Cepheid PL relation: on the contrary, they used the PL relation to determine the distances of their stars. Outside our Galaxy, Luck \\& Lambert (1992, hereafter LL92) have studied 10 Cepheids in the Magellanic Clouds (MCs). Five are in the Large MC (LMC) and five in the Small MC (SMC). For the former sample, they found a mean [Fe/H] of -0.36 dex with a dispersion of 0.3 dex, while for the latter one the mean [Fe/H] is -0.60 dex with a rather small dispersion of less than 0.15 dex. A more recent study by Luck et al. (1998, hereafter L98) on 10 LMC Cepheids and 6 SMC Cepheids, 4 of which in common with LL92, confirmed the mean [Fe/H] value in the LMC (-0.30 dex), found very little evidence of a significant metallicity dispersion in the LMC (contrary to LL92, but similarly to the SMC), and slightly revised downwards the mean [Fe/H] of the SMC (-0.74 vs -0.60 found by LL92). \\begin{table*}[!ht] %\\begin{sidewaystable*} %\\begin{minipage}[t][180mm]{\\textwidth} \\label{table:1} \\caption{Overview of recent results for the metallicity sensitivity of Cepheid distances. In the first column is listed the variation of the distance modulus $\\mu$ per dex of metallicity, the negative sign indicates that the true distance is longer than the one obtained neglecting the effect of the metallicity. In the second column is listed the elemental abundance used as reference for the metallicity. The third and fourth columns give the method and the reference of the different studies. See also Fig.~1.} %\\centering \\begin{tabular}{ l c p{9cm} l} \\hline $\\delta \\mu / \\delta [M/H]$ & ~ & Method & Reference \\\\ (mag/dex) & ~ & ~ \\\\ \\hline\\hline -0.32 $\\pm$ 0.21 & [Fe/H] & Analysis of Cepheids in 3 fields of M31 ($BVRI$ bands) & Freedman \\& Madore (1990) \\\\ -0.88 $\\pm$ 0.16 & [Fe/H] & Comparison of Cepheids from 3 fields of M31 and LMC ($BVRI$ bands) & Gould (1994) \\\\ -0.40 $\\pm$ 0.20 & [O/H] & Simultaneous solution for distances to 17 galaxies ($UBVRIJHK$ bands) & Kochanek (1997) \\\\ -0.44$_{-0.20}^{+0.10}$ & [O/H] & Comparison of EROS observations of SMC and LMC Cepheids ($VR$ bands) & Sasselov et al. (1997) \\\\ -0.24 $\\pm$ 0.16 & [O/H] & Comparison of HST observations of inner and outer fields of M101 & Kennicutt et al. (1998) \\\\ -0.12 $\\pm$ 0.08 & [O/H] & Comparison of 10 Cepheid galaxies with Tip of the Red Giant Branch distances & Kennicutt et al. (1998) \\\\ -0.20 $\\pm$ 0.20 & [O/H] & Value adopted for the HST Key Project final result & Freedman et al. (2001) \\\\ 0 & [Fe/H] & OGLE result comparing Cepheids in IC1613 and MC ($VI$ bands) & Udalski et al. (2001) \\\\ 0 & [O/H] & Comparison of Planetary Nebula luminosity function distance scale and Surface Brightness fluctuation distance scale & Ciardullo et al. (2002) \\\\ -0.24 $\\pm$ 0.05 & [O/H] & Comparison of 17 Cepheid galaxies with Tip of the Red Giant Branch distances & Sakai et al. (2004) \\\\ -0.21 $\\pm$ 0.19 & [Fe/H] & Baade-Wesselink analysis of Galactic and SMC Cepheids ($VK$ bands) & Storm et al. (2004) \\\\ -0.23 $\\pm$ 0.19 & [Fe/H] & Baade-Wesselink analysis of Galactic and SMC Cepheids ($I$ band) & Storm et al. (2004) \\\\ -0.29 $\\pm$ 0.19 & [Fe/H] & Baade-Wesselink analysis of Galactic and SMC Cepheids ($W$ index) & Storm et al. (2004) \\\\ -0.27 $\\pm$ 0.08 & [Fe/H] & Compilation from the literature of distances and metallicities of 53 Galactic and MC Cepheids ($VIWK$ bands) & Groenewegen et al. (2004) \\\\ -0.39 $\\pm$ 0.03 & [Fe/H] & Cepheid distances to SNe Ia host galaxies & Saha et al. (2006)\\\\ -0.29 $\\pm$ 0.09 & [O/H] & Cepheids in NGC~4258 and [O/H] gradient from Zaritsky et al. (1994) & Macri et al. (2006)\\\\ -0.10 $\\pm$ 0.03 & [Fe/H] & Weighted mean of Kennicutt, Macri and Groenewegen estimates & Benedict et al. (2007)\\\\ -0.017 $\\pm$ 0.113 & [O/H] & Comparison between Cepheid and TRGB distances for 18 galaxies & Tammann et al. (2007)\\\\ 0 & [Fe/H] & Comparison between the slopes of Galactic and LMC Cepheids & Fouqu\\'e et al. (2007)\\\\ +0.05 $\\pm$ 0.03 & [Fe/H] & Predicted Period-Wesenheit ($V,I$) relation & Bono et al. (2008)\\\\ \\hline \\hline \\end{tabular} %\\vfill %\\end{minipage} %\\end{sidewaystable*} \\end{table*} Finally, we mention two studies that followed slightly different approaches. Groenewegen et al. (2004) have selected from the literature a sample of 37 Galactic, 10 LMC and 6 SMC Cepheids for which individual metallicity estimates and $BVIK$ photometry were known. Their work aimed at investigating the metallicity dependence of the PL relation using individual metallicity determinations as well as good individual distance estimates for Galactic Cepheids. They inferred a metallicity effect of about -0.27 $\\pm$ 0.08 mag/dex in the zero point in $VIWK$, in the sense that metal-rich Cepheids are brighter than the metal-poor ones (see Table~1 and Fig.~1, for a comparison with other studies). Also Storm et al. (2004) discussed the effect of the metallicity on the PL relation using 34 Galactic and 5 SMC Cepheids, for which they determined accurate individual distances with the Baade-Wesselink method. Assuming an average abundance for the SMC Cepheids of [Fe/H]= -0.7 and solar metallicity for the Galactic ones, they determined, in a purely differential way, the following corrections: -0.21 $\\pm$ 0.19 for the $V$ and $K$ bands, -0.23 $\\pm$ 0.19 for the $I$-band and -0.29 $\\pm$ 0.19 for the Wesenheit index W. These agree well with Groenewegen et al. (2004). Despite these ongoing observational efforts, it is important to underline that none of the observational studies undertaken so far has directly determined elemental abundances of a large sample of Cepheids in order to explicitly infer the metallicity effect on the PL relation, taking advantage of a sample that has been homogeneously analysed. The novelty of our approach consists exactly in this, i.e. in the homogeneous analysis of a large sample of stars (68) in three galaxies (the Milky Way, the LMC and the SMC) spanning a factor of ten in metallicity and for which distances and $BVJK$ photometry are available. Preliminary results based on a sub-sample of the data discussed here were presented by Romaniello et al. (2005). Here we present the results about the iron content for the complete sample. In a forthcoming paper we will discuss the $\\alpha$-elements abundances. The paper is organised as follows. The data sample is presented in Section 2. In Section 3 we thoroughly describe the data analysis and how we determine the metallicity of our stars. We compare our iron abundances with previous results in Section 4. The dependence of the PL relation on [Fe/H] is discussed in Section 5. Finally, Section 6 summarizes our concluding remarks. \\begin{table*} \\label{table:2} \\caption{Pulsation phases ($\\phi$) and intrinsic parameters of the Galactic Cepheids. Both AP Pup and AX Vel were not included in the analysis of the metallicity effect because accurate distance estimates are not available in literature. In particular, for AP Pup we only listed the apparent mean magnitudes. In the last column is listed the duplicity status according to Szabados (2003): B - spectroscopic binary, Bc - spectroscopic binary that needs confirmation, O - spectroscopic binary with known orbit, V - visual binary} \\centering \\renewcommand{\\footnoterule}{} \\begin{tabular}{l l c c c c c c c c c} \\hline\\hline ID & $\\log P$ & $\\phi$ & $\\mu_{Old}$ & $E(B-V)_{Old}$ & $\\mu_{New}$ & $E(B-V)_{New}$ & $M_{B}$ & $M_{V}$ & $M_{K}$ & Duplicity \\\\ \\hline l Car & 1.5509 & 0.580 & 8.99$^{d}$ & 0.170$^{d}$ & 8.56$^{b}$ & 0.147$^{a}$ & -4.17\t& -5.28 \t & -7.53 & \\ldots \\\\ U Car & 1.5891 & 0.490 & 10.97$^{d}$ & 0.283$^{d}$ & 10.87$^{a}$ & 0.265$^{a}$ & -4.50\t& -5.41 \t & -7.44 & B \\\\ V Car & 0.8259 & 0.375 & 9.84$^{e}$ & 0.174$^{h}$ & 10.09$^{a}$ & 0.169$^{a}$ & -2.54\t& -3.24 \t & -4.86 & B \\\\ WZ Car & 1.3620 & 0.745 & 12.92$^{d}$ & 0.384$^{d}$ & 12.69$^{a}$ & 0.370$^{a}$ & -3.80\t& -4.58 \t & -6.52 & \\ldots \\\\ V Cen & 0.7399 & 0.155 & 9.18$^{d}$ & 0.289$^{c}$ & 8.91$^{a}$ & 0.292$^{a}$ & -2.41\t& -2.99 \t & -4.49 & \\ldots \\\\ KN Cen & 1.5319 & 0.867 & 13.12$^{d}$ & 0.926$^{d}$ & 12.84$^{a}$ & 0.797$^{a}$ & -4.63\t& -5.46 \t & -7.59 & B \\\\ VW Cen & 1.1771 & 0.967 & 12.80$^{d}$ & 0.448$^{d}$ & 12.76$^{a}$ & 0.428$^{a}$ & -2.93\t& -3.85 \t & -6.08 & B \\\\ XX Cen & 1.0395 & 0.338 & 11.11$^{d}$ & 0.260$^{d}$ & 10.90$^{a}$ & 0.266$^{a}$ & -3.19\t& -3.91 \t & -5.58 & B \\\\ $\\beta$ Dor & 0.9931 & 0.529 & 7.52$^{c}$ & 0.040$^{c}$ & 7.50$^{b}$ & 0.052$^{a}$ & -3.16\t& -3.91 \t & -5.57 & \\ldots \\\\ $\\zeta$ Gem & 1.0065 & 0.460 & 7.78$^{c}$ & 0.010$^{c}$ & 7.81$^{b}$ & 0.014$^{a}$ & -3.16\t& -3.94 \t & -5.72 & V \\\\ GH Lup & 0.9675 & 0.031 & 10.05$^{e}$ & 0.364$^{h}$ & 10.25$^{a}$ & 0.335$^{a}$ & -2.77\t& -3.66 \t & -5.54 & B \\\\ T Mon & 1.4319 & 0.574 & 10.82$^{d}$ & 0.209$^{c}$ & 10.71$^{a}$ & 0.181$^{a}$ & -4.16\t& -5.15 \t & -7.25 & O \\\\ S Mus & 0.9850 & 0.266 & 9.81$^{e}$ & 0.147$^{h}$ & 9.57$^{a}$ & 0.212$^{a}$ & -3.48\t& -4.10 \t & -5.62 & O \\\\ UU Mus & 1.0658 & 0.865 & 12.59$^{d}$ & 0.413$^{d}$ & 12.41$^{a}$ & 0.399$^{a}$ & -3.12\t& -3.86 \t & -5.70 & \\ldots \\\\ S Nor & 0.9892 & 0.343 & 9.91$^{d}$ & 0.189$^{c}$ & 9.87$^{a}$ & 0.179$^{a}$ & -3.23\t& -4.00 \t & -5.77 & B \\\\ U Nor & 1.1019 & 0.422 & 10.72$^{d}$ & 0.892$^{d}$ & 10.46$^{a}$ & 0.862$^{a}$ & -3.14\t& -3.90 \t & -5.72 & \\ldots \\\\ X Pup & 1.4143 & 0.232 & 12.36$^{e}$ & 0.443$^{h}$ & 11.64$^{a}$ & 0.402$^{a}$ & -3.57\t& -4.38 \t & -6.34 & \\ldots \\\\ AP Pup & 0.7062 & 0.109 & \\ldots$^{f}$ & \\ldots$^{f}$ &\\ldots$^{f}$ &\\ldots$^{f}$ & 7.37\t& 6.78\t & 5.26 & B \\\\ AQ Pup & 1.4786 & 0.436 & 12.52$^{d}$ & 0.512$^{c}$ & 12.41$^{a}$ & 0.518$^{a}$ & -4.53\t& -5.35 \t & -7.27 & B \\\\ BN Pup & 1.1359 & 0.397 & 12.95$^{d}$ & 0.438$^{c}$ & 12.93$^{a}$ & 0.416$^{a}$ & -3.55\t& -4.33 \t & -6.14 & \\ldots \\\\ LS Pup & 1.1506 & 0.012 & 13.55$^{d}$ & 0.478$^{d}$ & 13.39$^{a}$ & 0.461$^{a}$ & -3.60\t& -4.37 \t & -6.18 & B \\\\ RS Pup & 1.6174 & 0.944 & 11.56$^{d}$ & 0.446$^{c}$ & 11.30$^{a}$ & 0.457$^{a}$ & -4.71\t& -5.69 \t & -7.81 & \\ldots \\\\ VZ Pup & 1.3649 & 0.816 & 13.08$^{d}$ & 0.471$^{c}$ & 12.84$^{a}$ & 0.459$^{a}$ & -3.93\t& -4.63 \t & -6.31 & \\ldots \\\\ KQ Sco & 1.4577 & 0.446 & 12.36$^{c}$ & 0.839$^{c}$ & 12.23$^{g}$ & 0.869$^{a}$ & -4.05\t& -5.11 \t & -7.55 & \\ldots \\\\ EU Tau & 0.3227 & 0.414 & 10.27$^{c}$ & 0.170$^{c}$ & 10.27$^{c}$ & 0.170$^{d}$ & -2.26\t& -2.74 \t & -4.05 & Bc\\\\ SZ Tau & 0.4981 & 0.744 & 8.73$^{c}$ & 0.290$^{c}$ & 8.55$^{a}$ & 0.295$^{a}$ & -2.38\t& -2.93 \t & -4.33 & B \\\\ T Vel & 0.6665 & 0.233 & 9.80$^{d}$ & 0.281$^{c}$ & 10.02$^{a}$ & 0.289$^{a}$ & -2.24\t& -2.88 \t & -4.47 & B \\\\ AX Vel & 0.5650 & 0.872 & 10.76$^{f}$ & 0.224$^{h}$ &\\ldots$^{f}$ & 0.224$^{a}$ & \\ldots& \\ldots \t & \\ldots & \\ldots \\\\ RY Vel & 1.4496 & 0.704 & 12.02$^{d}$ & 0.562$^{c}$ & 11.73$^{a}$ & 0.547$^{a}$ & -4.23\t& -5.05 \t & -6.96 & \\ldots \\\\ RZ Vel & 1.3096 & 0.793 & 11.02$^{d}$ & 0.335$^{c}$ & 10.77$^{a}$ & 0.299$^{a}$ & -3.78\t& -4.61 \t & -6.56 & \\ldots \\\\ SW Vel & 1.3700 & 0.792 & 11.00$^{d}$ & 0.349$^{c}$ & 11.88$^{a}$ & 0.344$^{a}$ & -4.02\t& -4.83 \t & -6.75 & \\ldots \\\\ SX Vel & 0.9800 & 0.497 & 11.44$^{e}$ & 0.250$^{h}$ & 11.41$^{g}$ & 0.263$^{a}$ & -3.33\t& -3.95 \t & -5.49 & \\ldots \\\\ \\hline \\hline \\multicolumn{8}{l}{$^a$ Fouqu\\'e et al. (2007).}\\\\ \\multicolumn{8}{l}{$^b$ Benedict et al. (2007).}\\\\ \\multicolumn{8}{l}{$^c$ Groenewegen et al. (2004).}\\\\ \\multicolumn{8}{l}{$^d$ Storm et al. (2004).}\\\\ \\multicolumn{8}{l}{$^e$ Laney \\& Stobie (1995) and Groenewegen (2004).}\\\\ \\multicolumn{8}{l}{$^f$ Not included in the analysis of the metallicity effect.}\\\\ \\multicolumn{8}{l}{$^g$ Groenewegen (2008).}\\\\ \\multicolumn{8}{l}{$^h$ Fernie et al. (1995).}\\\\ \\end{tabular} \\end{table*} \\begin{table*} \\label{table:3} \\caption{Pulsation phases ($\\phi$) and intrinsic parameters of the Magellanic Cepheids. Periods ($\\log P$), apparent mean magnitudes and reddenings come from Laney \\& Stobie (1994). The mean $K$-band magnitudes were transformed into the 2MASS photometric system using the transformation provided by Koen et al. (2007).} \\centering \\begin{tabular}{l r r r r r c} \\hline\\hline ID & $\\log P$ & $\\phi$ & $B_0$ & $V_0$ & $K_0$ & $E(B-V)$\\\\ \\hline \\multicolumn{7}{c}{LMC} \\\\ \\hline HV 877 & 1.654 & 0.682 & 14.06 & 12.98 & 10.77 & 0.12 \\\\ HV 879 & 1.566 & 0.256 & 14.12 & 13.15 & 11.03 & 0.06 \\\\ HV 971 & 0.968 & 0.237 & 14.86 & 14.24 & 12.68 & 0.06 \\\\ HV 997 & 1.119 & 0.130 & 14.94 & 14.19 & 12.37 & 0.10 \\\\ HV 1013 & 1.382 & 0.710 & 14.39 & 13.46 & 11.41 & 0.11 \\\\ HV 1023 & 1.425 & 0.144 & 14.48 & 13.51 & 11.45 & 0.07 \\\\ HV 2260 & 1.112 & 0.144 & 15.19 & 14.43 & 12.67 & 0.13 \\\\ HV 2294 & 1.563 & 0.605 & 13.19 & 12.45 & 10.74 & 0.07 \\\\ HV 2337 & 0.837 & 0.861 &\\ldots & \\ldots& 13.27 & 0.07 \\\\ HV 2352 & 1.134 & 0.201 & 14.49 & 13.84 & 12.25 & 0.10 \\\\ HV 2369 & 1.684 & 0.136 & 13.15 & 12.29 & 10.38 & 0.10 \\\\ HV 2405 & 0.840 & 0.037 & \\ldots& \\ldots& 13.43 & 0.07 \\\\ HV 2580 & 1.228 & 0.119 & 14.33 & 13.67 & 11.92 & 0.09 \\\\ HV 2733 & 0.941 & 0.411 & 14.85 & 14.34 & 13.00 & 0.11 \\\\ HV 2793 & 1.283 & 0.917 & 14.49 & 13.58 & 11.75 & 0.10 \\\\ HV 2827 & 1.897 & 0.880 & 13.19 & 12.03 & 9.80 & 0.08 \\\\ HV 2836 & 1.244 & 0.059 & 14.85 & 14.02 & 12.04 & 0.18 \\\\ HV 2864 & 1.041 & 0.055 & 15.16 & 14.42 & 12.77 & 0.07 \\\\ HV 5497 & 1.997 & 0.321 & 12.73 & 11.63 & 9.43 & 0.10 \\\\ HV 6093 & 0.680 & 0.024 & 15.74 & 15.16 & 13.71 & 0.06 \\\\ HV 12452 & 0.941 & 0.860 & 15.25 & 14.60 & 12.83 & 0.06 \\\\ HV 12700 & 0.911 & 0.342 & 15.62 & 14.87 & 13.12 & -0.01 \\\\ \\hline \\multicolumn{7}{c}{SMC} \\\\ \\hline HV 817 & 1.277 & 0.298 & 14.13 & 13.59 & 12.12 & 0.08 \\\\ HV 823 & 1.504 & 0.873 & 14.46 & 13.60 & 11.58 & 0.05 \\\\ HV 824 & 1.818 & 0.315 & 13.06 & 12.27 & 10.33 & 0.03 \\\\ HV 829 & 1.931 & 0.348 & 12.61 & 11.81 & 9.92 & 0.03 \\\\ HV 834 & 1.866 & 0.557 & 12.95 & 12.14 & 10.20 & 0.02 \\\\ HV 837 & 1.631 & 0.822 & 13.95 & 13.10 & 11.11 & 0.04 \\\\ HV 847 & 1.433 & 0.500 & 14.40 & 13.66 & 11.83 & 0.08 \\\\ HV 865 & 1.523 & 0.108 & 13.55 & 12.93 & 11.21 & 0.06 \\\\ HV 1365 & 1.094 & 0.184 & 15.39 & 14.79 & 13.20 & 0.07 \\\\ HV 1954 & 1.223 & 0.847 & 14.13 & 13.62 & 12.12 & 0.07 \\\\ HV 2064 & 1.527 & 0.279 & 14.28 & 13.50 & 11.61 & 0.07 \\\\ HV 2195 & 1.621 & 0.135 & 13.85 & 13.07 & 11.09 & -0.02 \\\\ HV 2209 & 1.355 & 0.822 & 13.99 & 13.42 & 11.84 & 0.04 \\\\ HV 11211 & 1.330 & 0.516 & 14.36 & 13.64 & 11.83 & 0.06 \\\\ \\hline \\hline \\end{tabular} \\end{table*} ", "conclusions": "We have directly measured the iron abundances for 68 Galactic and Magellanic Cepheids from FEROS and UVES high resolution and high signal-to-noise spectra. We have used these measurements to assess the influence of the stellar iron content on the Cepheid PL relation in the $V$ and in the $K$ band. In order to do this we have related the $V$-band and the $K$-band residuals from the standard PL relations of Freedman et al. (2001) and Persson et al. (2004), respectively, to [Fe/H]. Differently from previous studies, we can constrain the PL relation using Cepheids with known distance moduli and chemical abundances, homogeneously measured, that cover almost a factor of ten in metallicity. For our Galactic sample, we find that the mean value of the iron content is solar ($\\sigma$ = 0.10, see Fig.~6), with a range of values between -0.18 dex and +0.25 dex. For the LMC sample, we find that the mean value is about $\\sim$ -0.33 dex ($\\sigma$ =0.13, see Fig.~6), with a range of values between -0.62 dex and -0.10 dex. For the SMC sample, we find that the mean value is about $\\sim$ -0.75 dex ($\\sigma$ = 0.08, see Fig.~6), with a range of values between -0.87 and -0.63. We have compared our results with the analyses of FC97, Andrievsky et al. (2002a, 2002b, 2002c) and Luck et al. (2003) for the Galactic Cepheids and LL92 and L98 for the Magellanic Clouds. Regarding the Galactic sample, our results are marginally more in agreement with Andrievsky's values than with FC97 and the differences, on average, appear rather small. Considering the Magellanic Cepheids, we have a poor agreement with LL92, which could be in part accounted for by different analytical tools and data quality. Our data are in better agreement with L98 results and the spread of our iron abundances in the LMC and SMC is similar to the one they reported. We note that the mean metallicity that they found with their complete sample (-0.30 dex and -0.74 dex) is in very good agreement with our results. Our main results concerning the effect of the iron abundance on the PL relation are summarized in Fig.~8 and Fig.~9 (bottom panels) and they hold for a LMC distance modulus of 18.50. In Fig.~10 is also showed the comparison in the $V$-band with the empirical results of Kennicutt et al. (1998) in two Cepheid fields of M101 (open circles and solid line). The main findings can be summarized as follows: \\begin{itemize} \\item The $V$-band PL relation does depend on the metal abundance. This finding is marginally affected by the adopted distance scale for the Galactic Cepheids and by the LMC distance. \\item Current data do not allow us to reach a firm conclusion concerning the dependence of the $K$-band PL relation on the metal content. The use of the most recent distances for Galactic Cepheids (Benedict et al. 2007; Fouqu\\`e et al. 2007; van Leeuwen et al. 2007) indicates a mild metallicity effect. On the other hand, the use of the old distances (Storm et al. 2004) suggest a vanishing effect. \\item Residuals based on a canonical LMC distance ($\\mu_{LMC}=18.5$) and on the most recent distances for Galactic Cepheids present a well defined effect in the $V$-band. The metal-poor and the metal-rich bin are $\\approx 2 \\sigma$ and $\\approx 9 \\sigma$ from the null hypothesis. Moreover, the two metallicity bins differ at the $3\\sigma$ level. \\item By assuming a \"short\" LMC distance ($\\mu_{LMC}=18.3$) the residuals present a strong metallicity dependency in the zero-point of the $V$-band PL relation. By assuming a \"long\" LMC distance ($\\mu_{LMC}=18.7$) we found a strong metallicity effect when moving from metal-poor to metal-rich Cepheids. This indicates a significant change in the slope and probably in the zero-point. The findings based on the \"short\" and on the \"long\" LMC distance are at odds with current empirical and theoretical evidence, suggesting a smaller metallicity effect. \\item Metal-rich Cepheids in the $V$-band are systematically fainter than metal-poor ones. This evidence is strongly supported by the canonical, the \"short\" and the \"long\" LMC distance. \\end{itemize} The above results together with recent robust LMC distance estimates indicate that the behaviours based on the canonical distance appear to be the most reliable ones. In order to constrain on a more quantitative basis the metallicity dependence of both zero-point and slope of the optical PL relations is required a larger number of Cepheids covering a broader range in metal abundances. Moreover, for each metallicity bin Cepheids covering a broad period range are required to reduce the error on the residuals and to constrain on a quantitative basis the fine structure of the PL relation in optical and NIR photometric bands." }, "0807/0807.0731_arXiv.txt": { "abstract": "In this paper we present preliminary results from cosmological simulations of modified gravity in the dark matter sector. Our results show improvements over standard cold dark matter cosmology. The abundance of low-mass haloes in the modified gravity model fit observations better than the conventional theory, while the differences of the modified density fluctuation power spectrum differs from the standard, $\\Lambda$CDM power spectrum are small enough to make these two models observationally indistinguishable at large scales. ", "introduction": "Over the past 30 years cosmology became a successful and mature empirical science. The growing accuracy of astronomical observations gave the possibility to test theoretical models. Past decade has established the \"standard cosmological model\" dubbed also as the $\\Lambda$CDM cosmology, or the \"concordance model\". In this picture the evolution of the expanding universe, dominated by non-relativistic (cold) dark mater particles is governed by Einstein's gravity with a positive cosmological constant. Although the standard model is very successful in explaining many observations it still has some difficulties. The physical nature of the dark matter (DM) sector is still unknown. Moreover, the mass distribution in the real universe (in particular, the emptiness of voids) differs significantly from the predictions of the $\\Lambda$CDM model. Modifications of the standard gravity theory have been proposed to deal with this impasse. In this paper we show preliminary results of our work based on one of such modifications. ", "conclusions": "Our results show that the modification of standard gravity we consider can close the gap between some properties of standard $\\Lambda$CDM cosmology simulations and today's high resolution astronomical observations. Deviations from the standard power spectrum are negligible, and they agree with large scale observations. On the other hand, our simulations of scalar interaction show desired clustering properties on small scales. While the court on the physical nature of dark matter particles is still out it is scientifically justified to probe various theoretical possibilities. In this paper we have shown that the long-range scalar DM interaction has the potential to improve the standard cosmological model." }, "0807/0807.0441_arXiv.txt": { "abstract": "We calculate the rate of in-fall of stellar matter on an accretion disk during the collapse of a rapidly rotating massive star, and estimate the luminosity of the relativistic jet that results from accretion on to the central black hole. We find that the jet luminosity remains high for about $10^2$ s, at a level comparable to the typical luminosity observed in gamma-ray bursts (GRBs). The luminosity then decreases rapidly with time for about $\\sim10^3$ s, roughly as $\\sim t^{-3}$; the duration depends on the size and rotation speed of the stellar core. The rapid decrease of the jet power explains the steeply declining x-ray flux observed at the end of most long duration GRBs. Observations with the Swift satellite show that, following the steep decline, many GRBs exhibit a plateau in the x-ray lightcurve (XLC) that lasts for about 10$^4$ s. We suggest that this puzzling feature is due to continued accretion in the central engine. A plateau in the jet luminosity can arise when the viscosity parameter $\\alpha$ is small, $\\sim10^{-2}$ or less. A plateau is also produced by continued fall-back of matter -- either from an extended stellar envelope or from material that failed to escape with the supernova ejecta. In a few GRBs, the XLC is observed to drop suddenly at the end of the plateau phase, while in others the XLC declines more slowly as $\\sim t^{-1}-t^{-2}$. These features arise naturally in the accretion model depending on the radius and mean specific angular momentum of the stellar envelope. The total energy in the disk-wind accompanying accretion is found to be about 10$^{52}$ erg. This is comparable to the energy observed in supernovae associated with GRBs, suggesting that the wind might be the primary agent responsible for the explosion. The accretion model thus provides a coherent explanation for the diverse and puzzling features observed in the early XLC of GRBs. It might be possible to use this model to invert gamma-ray and x-ray observations of GRBs and thereby infer basic properties of the core and envelope of the GRB progenitor star. ", "introduction": "Observations of gamma-ray bursts (GRBs) carried out by the NASA Swift satellite in the last two years have shown that the $\\gamma$-ray prompt emission turns off abruptly after about a minute. The abrupt shutoff is evidenced by the rapidly declining x-ray flux (t$^{-3}$ or faster) from about 80s to 300s, which joins smoothly the prompt GRB lightcurve when extrapolated back in time and spectral band (Tagliaferri et al. 2005; Nousek et al. 2006; O'Brien et al. 2006). The abrupt decline suggests that the activity at the center of the explosion declines very rapidly with time after about a minute of more or less steady activity. At the same time, Swift observations have provided overwhelming evidence that the GRB central engine continues operating for hours and perhaps even days. There are two independent indications for this phenomenon. First, we often see a phase in the early x-ray lightcurve of long-GRBs, from $\\sim 10^3$s to 10$^4$s, during which the flux declines slowly with time. We will refer to this as a ``plateau'' in the lightcurve. There are many proposals to explain the plateau. The models are, however, highly constrained by a lack of correlation between x-ray and optical features for many GRBs. Therefore, a successful proposal must invoke continuing activity of the central engine to produce the x-ray plateau (e.g., Zhang 2006; Panaitescu 2007, and references therein), whereas the simultaneous optical emission may be produced in the afterglow. Second, in roughly a third of the observed GRBs, the x-ray flux is seen to increase suddenly and then to drop precipitously in what are referred to as ``flares.'' In some cases the flux during these flares increases by a factor of $\\sim 10^2$ on a time scale $\\delta t\\ll t$ (Burrows et al. 2005; Chincarini et al. 2007), which cannot be explained in terms of a density inhomogeneity in the external medium (Nakar \\& Granot, 2006). Variable activity in the central engine is a more natural explanation. A sudden drop in the flux --- e.g., in the case of GRB 070110 the x-ray lightcurve was nearly flat for 20 ks and then fell off as $t^{-8}$ (Troja et al. 2007) --- is also not possible to understand other than as the result of highly variable central engine activity. The above conflicting requirements, viz., (i) a sudden drop in activity at the end of the main GRB ($t\\ \\lta 10^2$ s), (ii) continued steady activity during an extended plateau ($t\\sim10^4$ s), and (iii) occasional dramatic flares, are challenging for models of the central engine. In this paper we consider the currently most popular model of GRBs, which postulates ultra-rapid accretion of gas on to a newly-formed black hole or neutron star (Narayan, Paczy\\'nski \\& Piran 1992; Popham, Woosley \\& Fryer 1999; Narayan, Piran \\& Kumar 2001). The accretion disk may be the result of (i) gas fall-back after a hypernova, as in the collapsar model (Woosley 1993; Paczy\\'nski 1998; MacFadyen \\& Woosley 1999), which is considered relevant for long duration GRBs, or (ii) during the merger of a double neutron star or neutron star-black hole binary (Eichler et al. 1989; Narayan et al. 1992), which is currently considered the most likely explanation for short duration GRBs. We attempt to reconcile the accretion model for collapsars with Swift observations of long duration GRBs. In \\S~2, we make use of a reasonably realistic model of the progenitor of a collapsar and estimate the rate at which gas is added to the accretion torus at the center. The calculation is based on a crude free-fall model for the collapsing star and the results are to be taken in the spirit of an order of magnitude estimate. Then, in \\S~3, we use this model to study the variation of accretion power with time. We show that the model naturally reproduces both the sudden shutoff of the prompt GRB emission and the extended plateau in the lightcurve. We speculate on possible scenarios for producing flares and suggest an explanation for the hypernova explosion associated with some GRBs. We conclude in \\S~4 with a discussion and summary. ", "conclusions": "This work was motivated by two basic questions posed by the extensive and excellent Swift observations of GRBs: (1) Why does the $\\gamma$-ray/x-ray flux undergo a sharp decline (flux decreasing as $t^{-3}$ or faster) about one minute after the start of the burst, even though the central engine itself is apparently active for hours? (2) After the sharp decline, how does the power from the engine remain nearly constant for a period of $\\sim10^4$ s to produce the long plateau that is observed in the x-ray lightcurves of many GRBs? We have attempted to answer these questions in the context of an accretion model of GRBs in which the central engine is powered by accretion on to a BH, and the GRB luminosity is proportional to the power in the resulting relativistic jets. We employ the pre-collapse stellar model of Woosley \\& Heger (2006), with the density structure and rotation profile shown in Fig. 2. We investigate the post-collapse accretion activity of this model from $\\sim$1 s to $\\sim10^5$ s and compare the model predictions with GRB observations. We compute the mass in-fall rate during the collapse of the massive star, and the viscous accretion rate on to the central BH using ideas from Narayan et al. (2001) and Kohri et al. (2005). We then estimate the relativistic jet luminosity from the mass accretion rate and the BH spin, using a prescription proposed by McKinney (2005). We find that the jet has a luminosity of about $10^{51} ~{\\rm erg\\,s^{-1}}$, lasting for about 10-20s (Fig. 4), which is of order the power and duration of a typical long GRB. The luminosity arises from the fall-back and accretion of the outer half of the core of the progenitor star; the material here has sufficient angular momentum to go into orbit, whereas the material from the inner half of the core collapses directly to form the BH. When the outer layers of the stellar core -- where the density falls off rapidly with $r$ -- are accreted via the ADAF process, the jet luminosity drops rapidly. This provides a straightforward and natural explanation for the steep decline of the early x-ray lightcurve observed by Swift. Fluence of gamma-ray bursts can be used to constrain the rotation rate in the core of their progenitor stars to within a factor of 10 or better; too large or too small $\\Omega$ in the core results in a small jet luminosity (\\S3.4), and causes the temporal decay of the lightcurve at the end of the prompt emission, for a few hundred seconds, to be either too shallow or too steep (fig. 5). In our model, the plateau in the x-ray lightcurve requires continued accretion to power the jet. One possibility is that the viscous time scale in the disk is so long that it takes a time $\\sim10^4$ s for the material in the disk to accrete. A viscosity parameter $\\alpha\\ \\lta\\ 10^{-2}$ is required. In this scenario, although mass fall-back ceases in a few hundred seconds, accretion continues on for a few hours (Fig. 6). While the model succeeds in producing an extended plateau, it causes the fall-off of the early GRB lightcurve to be less steep -- more like $t^{-2}$ than $t^{-3}$ -- (Fig. 6) in conflict with observations. The decline at the end of the plateau is also fairly shallow, inconsistent with observations of some GRBs. A second and, in our opinion, more likely scenario is that there is continued mass fall-back for the entire duration of the plateau (Fig. 7). In this case, the accretion time scale itself is short (i.e., $\\alpha$ is large), so the lightcurve reflects the mass fall-back rate. One possibility is that the fall-back is due to material that fails to be ejected by the supernova explosion. This idea has two problems. First, it is hard to see why there should be an extended period of a fairly constant fall-back rate as required by the observations, whereas we expect the fall-back rate to vary much more steeply as $t^{-5/3}$ (Chevalier 1989). Second, it is hard to see how we can have a sudden cutoff in the fall-back at the end of the plateau as seen in several GRBs. Another possibility is that the progenitor star has a core-envelope structure, with the core producing the early GRB and the envelope producing the plateau. The core must have a radius of $\\sim {\\rm few}\\times10^{10}$ cm to explain the duration of the GRB and the envelope must have a radius of $\\sim {\\rm few}\\times10^{11}$ cm to explain the duration of the plateau. There should be a large density (or $j$) contrast between the core and the envelope, in order to explain the sharp cutoff of the prompt emission, and the density profile in the envelope must be fairly shallow, $\\rho\\sim r^{-2}$, in order to obtain a shallow plateau. Depending on the rotation profile of the star, various kinds of lightcurves --- including ones in which we have a very rapid cutoff of the x-ray luminosity at the end of the plateau --- are possible. This scenario is therefore capable of explaining almost all observed cases. GRBs that do not have a plateau in their lightcurve are also easily explained; these presumably had progenitors with only a core and no envelope (like the model shown in Fig. 2). A nice feature of this model is that we could, in principle, use GRB observations to deduce the density and rotation structure of the progenitor star. On the other hand, it is not clear that evolved massive stars do have the kind of core-envelope structure we need to explain a typical GRB x-ray plateau. (We are not aware of any pre-supernova models in the literature with the required properties.) An important implication of the accretion model of GRB central engines is that the accretion flows are advection-dominated and thus have strong outflows/winds. We have estimated the total energy in the wind for the pre-collapse stellar model of Fig. 2 and find it to be about 2x10$^{52}$ erg. This is sufficient to explode the star, and might explain the observed energetics of supernovae associated with GRBs. A generic prediction of the late fall-back model for the x-ray plateau is that brighter GRBs should have a weaker (lower luminosity) and shorter duration plateau. The reason is that a stronger GRB, with its stronger jet and wind, is likely to eject more of the stellar envelope during the main burst. Indeed, recent simulations of relativistic jet-induced supernovae support this prediction, with more luminous explosions expelling more of the stellar envelope and leaving less material available for accretion (eg. Tominaga 2007). In order to test this prediction, we have looked at a sample of bursts with known redshift and isotropic equivalent luminosities (Butler \\& Kocevski 2007). From this set of bursts, we have selected two subsets, those with distinct x-ray plateaus (GRBs 070110, 060614, 050315, 060607A, 060729, 070810A) and those clearly lacking x-ray plateaus (GRBs 071020, 070318, 061007, 050922C, 050826, 070411). Consistent with the prediction of the fall-back model, we find that the average peak isotropic equivalent luminosity (per frequency interval) of the subset of bursts with distinct x-ray plateaus is $\\sim$ 4 times lower than that of the subset of bursts lacking an x-ray plateau. Moreover, using the same sample of bursts, we find that the average peak isotropic equivalent luminosity (per frequency interval) of bursts with an x-ray plateau is lower than that of bursts without an x-ray plateau at the $>$ 10\\% level of significance, assuming normal distributions for the luminosities of each of these populations of bursts. A related prediction is that the plateau should be absent, or at least weak, in those cases where we see a bright supernova event associated with a GRB. The idea is that a bright supernova implies powerful ejection, and there should be less material available for accretion. We have only one well observed case of a GRB-supernova association in the Swift sample of bursts (GRB 030329), and the x-ray lightcurve did not have a plateau. While this observation is consistent with the late fall-back model, we note that it is just a single object and therefore the result is not very significant. It is interesting to note that x-ray plateaus are not seen for short duration GRBs (see Nakar, 2007, for an excellent review). This is consistent with our model. If short GRBs are the result of the merger of double neutron star binaries (the currently popular model), then there is no material in an extended envelope in the progenitor to produce late fall-back. The model described in this paper is obviously incomplete. We have not provided any quantitative explanation for the x-ray flares seen during the plateau phase (and even later) in many GRBs. The only qualitative idea we have offered is that the flares reflect an instability in the accretion disk. We have also simplified the model considerably by postulating a direct proportionality between the jet power at the point where it is launched from the BH and the observed luminosity. Several factors could seriously modify this relation. First, the tunneling of the jet through the stellar material may be inefficient, and so the power that escapes from the surface of the star may be a small (and variable) fraction of the jet power at the base. Second, the efficiency with which the escaping jet power is converted to radiation (the physics of which is poorly understood) may be variable. Finally, the beaming of the jet may be different for the prompt GRB and the x-ray plateau, and may even vary during the plateau. We have ignored these complications in the interests of simplicity. \\noindent{\\bf Acknowledgment:} We thank Stan Woosley for providing his stellar model and for very useful comments on the draft. We are grateful to Craig Wheeler for a number of excellent suggestions that improved the paper. This work is supported in part by a NSF grant (AST-0406878), and NASA Swift-GI-program." }, "0807/0807.5052_arXiv.txt": { "abstract": "{ {\\em Context:} There is increasing need for good algorithms for modeling the aggregation and fragmentation of solid particles (dust grains, dust aggregates, boulders) in various astrophysical settings, including protoplanetary disks, planetary- and sub-stellar atmospheres and dense molecular cloud cores. Here we describe a new algorithm that combines advantages of various standard methods into one. \\\\ {\\em Aims:} The aim is to develop a method that 1) can solve for aggregation and fragmentation, 2) can easily include the effect and evolution of grain properties such as compactness, composition, etc., and 3) can be built as a coagulation/fragmentation module into a hydrodynamics simulation where it 3a) allows for non-`thermalized' non-local motions of particles (e.g.~movement of particles in turbulent flows with stopping time larger than eddy turn-over time) and 3b) focuses computational effort there where most of the mass is. \\\\ {\\em Methods:} We develop a Monte-Carlo method in which we follow the ``life'' of a limited number of representative particles. Each of these particles is associated with a certain fraction of the total dust mass and thereby represents a large number of true particles which all are assumed to have the same properties as their representative particle. Under the assumption that the total number of true particles vastly exceeds the number of representative particles, the chance of a representative particle colliding with another representative particle is negligibly small, and we therefore ignore this possibility. This now makes it possible to employ a statistical approach to the evolution of the representative particles, which is the core of our Monte Carlo method. \\\\ {\\em Results:} The method reproduces the known analytic solutions of simplified coagulation kernels, and compares well to numerical results for Brownian motion using other methods. For reasonably well-behaved kernels it produces good results even for moderate number of swarms.} ", "introduction": "Dust particle aggregation is a very common process in various astrophysical settings. In protoplanetary disks the aggregation of dust particles forms the very initial step of planet formation (see e.g.\\ Dominik et al.~\\citeyear{dominikblum:2007}). It also modifies the optical properties of the disk, and it has influence on the chemistry and free electron abundance in a disk (Sano et al.~\\citeyear{sanomirama:2000}; Semenov et al.~\\citeyear{semenovwiebe:2004}; Ilgner \\& Nelson \\citeyear{ilgnernelson:2006a}). The appearance and evolution of a protoplanetary disk is therefore critically affected by the dust aggregation process. In sub-stellar and planetary atmospheres the aggregation of dust particles and the coagulation of fluid droplets can affect the structure of cloud layers. It can therefore strongly affect the spectrum of these objects and influence the local conditions within these atmospheres. The process of aggregation/coagulation and the reverse process of fragmentation or cratering are therefore important processes to understand, but at the same time they are extremely complex. Traditional methods solve the Smoluchowski equation for the particle mass distribution function $f(m)$, where $f(m)$ is defined such that $f(m)dm$ denotes the number of particles per cubic centimeter with masses in the interval $[m,m+dm]$. This kind of method has been used in many papers on dust coagulation before (e.g.\\ Nakagawa et al.~\\citeyear{nakanakahayashi:1981}; Weidenschilling \\citeyear{weidenschilling:1984,weid1997}; Schmitt et al.~\\citeyear{schmitthenningmucha:1997}; Suttner \\& Yorke \\citeyear{suttneryorke:1999}; Tanaka et al.~\\citeyear{tanakahimemoida:2005}; Dullemond \\& Dominik \\citeyear{duldom:2005}; Nomura \\& Nakagawa \\citeyear{nomuranaka:2006}). Methods of this kind are efficient, but have many known problems. First of all a coarse sampling of the particle mass leads to systematic errors such as the acceleration of growth (Ohtsuki et al.~\\citeyear{ohtsuki:1990}). High resolution is therefore required, which may make certain problems computationally expensive. Moreover, if one wishes to include additional properties of a particle, such as porosity, charge, composition etc, then each of these properties adds another dimension to the problem. If each of these dimensions is sampled properly, this can quickly make the problem prohibitively computationally expensive. Finally, the traditional methods are less well suited for modeling stochastic behavior of particles unless this stochastic behavior can be treated in an averaged way. For instance, in protoplanetary disks if the stopping time of a particle is roughly equal to the turbulent eddy turn-over time, then the velocity of a particle with respect to the gas is stochastic: at the same location there can exist two particles with identical properties but which happen to have different velocities because they entered the eddy from different directions (see e.g.\\ the simulations by Johansen et al.~\\citeyear{johansen:2006}). To circumvent problems of this kind Ormel et al.~(\\citeyear{ormelmonte:2007}) have presented a Monte Carlo approach to coagulation. In this approach the particles are treated as computational particles in a volume which is representative of a much larger volume. The simulation follows the life of $N$ particles as they collide and stick or fragment. The collision rates among these particles are computed, and by use of random numbers it is then determined which particle collides with which. The outcome of the collision is then determined depending on the properties of the two colliding particles and their relative impact velocity. This method, under ideal conditions, provides the true simulation of the process, except that random numbers are used in combination with collision rates to determine the next collision event. This method has many advantages over the tradiational methods. It is nearly trivial to add any number of particle properties to each particle. There is less worry of systematic errors because it is so close to a true simulation of the system, and it is easy to implement. A disadvantage is that upon coagulation the number of computational particles goes down as the particles coagulate. Ormel et al.~solve this problem by enlarging the volume of the simulation and hence add new particles, but this means that the method is not very well suited for modeling coagulation within a spatially resolved setting such as a hydrodynamic simulation or a model of a protoplanetary disk. It is the purpose of the present paper to present an alternative Monte Carlo method which can quite naturally deal with extremely large numbers of particles, which keeps the number of computational particles constant throughout the simulation and which can be used in spatially resolved models. ", "conclusions": "We have shown that our representative particle method for aggregation of particles in astrophysical settings works well for standard kernels. It has the usual advantages of Monte Carlo methods that one can add particle properties easily and without loss of computational speed. Moreover, it naturally conserves the number of computational elements, so there is no need to ``add'' or ``remove'' particles. Each representative particle represents a fixed portion of the total mass of solids. Our method may have various possible interesting extensions and applications. Here we speculate on a few of these. For instance, the fact that each representative particle corresponds to a fixed amount of solid mass makes the method ideal for implementation into spatially resolved models such as hydrodynamic simulations of planetary atmospheres or protoplanetary disks. We can then follow the exact motion of each representative particle through the possibly turbulent environment, and thereby automatically treat the stochastic nature and deviation from a Boltzmann distribution of the motion of particles with stopping times of the same order as the turbulent eddy turn-over time. It is necessesary, however, to assure that a sufficiently large number of representative particles is present in each grid cell of the hydrodynamic simulation. For large scale hydrodynamic simulations this may lead to a very large computational demand for the coagulation computation, as well as for tracking the exact motion of these particles. If strong clumping of the particles happens, however, much of the ``action'' anyway happens in these ``clumps'', and it may then not be too critical that other grid cells are not sufficiently populated by representative particles. This, however, is something that has to be experimented. Our representative particle method can in principle also be used to model the sublimation and condensation of dust grains. If a particle sublimates then the representative particle becomes simply an atom or molecule of the vapor of this process. It will then follow the gas motion until the temperature becomes low enough that it can condense again. Other representative particles which are still in the solid phase may represent physical particles that can act as a condensation nucleus. Finally, in our method the properties of the particle can not only change due to collisions, but we can easily implement other environmental factors in the alteration of particle properties. There are two main drawbacks of the method. First, it only works for large particle numbers, i.e.\\ it cannot treat problems in which individual particles start dominating their immediate environment. Ormel's method and its expected extension do not have this problem. Secondly, the method cannot be accelerated using implicit integration, while Brauer's method can. All in all we believe that this method may have interesting applications in the field of dust aggregation and droplet coagulation in protoplanetary disks and planetary atmospheres. \\begin{acknowledgement} We wish to thank Frithjof Brauer, Anders Johansen, Patrick Glaschke, Thomas Henning, J\\\"urgen Blum, Carsten G\\\"uttler, Carsten Dominik and Dominik Paszun for useful comments. We also thank the anonymous referee for very useful comments that helped to improve the paper and for pointing us to a problem in our presentation of the solution to the particle number conservation paradox. \\end{acknowledgement}" }, "0807/0807.2499_arXiv.txt": { "abstract": "{\\mbox{BD$-6\\degr$1178} identified with the infrared source IRAS\\,05238$-$0626 is shown for the first time to be a spectroscopic binary (SB2) by analyzing the high-resolution spectra taken with the NES echelle spectrograph of the 6-m telescope. The components of the binary have close spectral types and luminosity classes: F5\\,IV--III and F3\\,V. The heliocentric radial velocities are measured for both components at four observing moments in 2004--2005. Both stars have close rotation velocities, which are equal to 24 and 19\\,km/s. We do not confirm the classification of BD$-6\\degr$1178 as a supergiant in the transition stage of becoming a planetary nebula. BD$-6\\degr$1178 probably is a young pre-MS stars. It is possibly a member of the 1c subgroup of the Ori\\,OB1 association.} \\titlerunning{\\it Spectroscopic binary BD${ -6\\degr}$1178} \\authorrunning{\\it Klochkova \\& Chentsov} ", "introduction": "In this paper we continue to publish the results of our spectroscopy of stars with IR excesses (see [\\cite{Klochkova1995, KSPV, IRAS23304, AFGL2688, IRAS20000}] and references therein for the main results). BD$-6\\degr$1178 is an optical component of the infrared source IRAS\\,05238$-$0626 (with galactic coordinates l=208.9$\\degr$, b=$-21.8\\degr$). This object is considered to be a candidate to protoplanetary nebula (PPN) according to the observed excess of radiation in the 12--60\\,$\\mu$m wavelength region and its position on the IR colour--colour diagram [\\cite{Garcia-1990, Reddy, Fujii}]. Recall that according to modern concepts (see, e.g., [\\cite{Block}]), objects observed at the short-lived evolutionary stage of a young planetary (protoplanetary) nebula are intermediate-mass stars evolving away from the asymptotic giant branch (AGB) toward the stage of a planetary nebula. The initial main-sequence (MS) masses of these stars lie in the 3--8\\,${\\mathcal M}_{\\odot}$ mass interval. During the AGB stage these stars have lost much mass in the form of a powerful stellar wind, and as a result, at the PPN stage the stars have the form of degenerate carbon-oxygen nuclei with typical masses of about 0.6\\,${\\mathcal M}_{\\odot}$ surrounded by expanding gas-and-dust shell. The astronomers are interested in studying PPNs, first because they allow one to study stellar-wind driven mass loss and second, because they offer a unique opportunity of observing the result of the stellar nucleosynthesis, mixing, and dredge-up of products of nuclear reactions that occurred during the preceding evolution of the star. About a dozen objects overabundant in heavy metals synthesized via neutronization of iron nuclei under the conditions of low neutron density (\\mbox{$s$-process}) have been found among the PPN-candidate studied. An analysis of the properties of PPNs showed that the expected overabundances of $s$-process elements are observed only the atmospheres of \\mbox{C-rich} stars whose IR spectra contain an emission at 21\\,$\\mu$m [\\cite{Klochkova1995, IRAS20000, Klochkova1997, Winckel}]. However, the overwhelming majority of PPNs exhibit neither carbon (\\mbox{O-rich} stars) nor heavy-element overabundance (see, e.g., [\\cite{Klochkova1995, QYSge, IRAS19475}]). The correlation found between the excess of heavy elements in the star's atmosphere and the peculiarity of the IR spectrum of the envelope of the star remains unexplained and hence a further increase of the sample of PPN objects studied is needed. Currently, we know little about BD$-6\\degr$1178. Its sky coordinates for the epoch of 2000 are: $\\alpha$=05$^{\\rm h} 26^{\\rm m} 19.8^{\\rm s}$, $\\delta$=$-6\\degr 23^{\\rm '} 57^{\\rm ''}$. The $V$- and $B$-band apparent magnitudes are equal to $V$\\,=\\,$10.52^m$ and \\mbox{$B$\\,=\\,$10.96^m$ [\\cite{Fujii}]}. Some evidence for the photometric variability of the star was found: according to the NSVS catalog [\\cite{Wozniak}], the mean magnitude of the star in a close-to-$R$-filter passband varies in the interval 10.78--$10.87^m$ and its standard deviation is about $0.01^m$. Modeling of the spectral energy distribution based on the multicolor photometry in the visual and near IR yields effective-temperature values T$_{eff}$ ranging from 8000\\,K [\\cite{Garcia-1990}] to 7400\\,K [\\cite{Fujii}], which corresponds to late A --- early F-subclasses. As for spectroscopic observations, only low-resolution ($\\approx$5\\,\\AA/pixel) spectra have been published so far for BD$-6\\degr$1178. These spectra yielded the following estimates for the spectral type: F2\\,II [\\cite{Reddy}], F4 [\\cite{Suarez}], and F5 [\\cite{Torres}]. In view of the aforesaid, it becomes evident from these results that further detailed study of the optical spectrum of the star is needed. In this paper we report the results of our numerous high-resolution spectroscopic observations of BD$-6\\degr$1178 made with the 6-m telescope of the Special Astrophysical Observatory of the Russian Academy of Sciences (SAO RAS). The aim of this study is to perform two-dimensional spectral classification, search for spectroscopic variability, analyze the velocity field in the star's atmosphere and envelope, and refine its evolutionary status. In Section\\,\\ref{observ} the methods of observation and reduction are described; in Section\\,\\ref{results} we present and analyze the observational data obtained, and in Section\\,\\ref{conclus} we briefly sums up the main results. ", "conclusions": "The results of our quantitative spectral classification of BD$-6\\degr$1178 led us to conclude that it is a double-line spectroscopic binary. Both components are F-type stars: F5\\,IV--III\\,+\\,F3V. We measured the heliocentric velocities of both components of the binary at four time moments. We found no grounds to classify BD$-6\\degr$1178 as a post-AGB star. The coordinates of BD$-6\\degr$1178 and its heliocentric distance 450\\,pc allow it to be suspected of a membership in the Ori\\,OB1 association. Thus BD$-6\\degr$1178 may be a young pre-MS object of the Galactic disk." }, "0807/0807.2450_arXiv.txt": { "abstract": "We present Keck adaptive optics imaging of the L4+L4 binary \\hdbin\\ along with archival \\HST\\ and Gemini-North observations, which together span $\\approx$70\\% of the binary's orbital period. From the relative orbit, we determine a total dynamical mass of 0.109$\\pm$0.002~\\Msun\\ (114$\\pm$2~\\Mjup). The flux ratio of \\hdbin\\ is near unity, so both components are unambiguously substellar for any plausible mass ratio. An independent constraint on the age of the system is available from the primary \\hdprim\\ (G2V, [M/H]~=~0.0). The ensemble of available indicators suggests an age comparable to the Hyades, with the most precise age being \\hdage\\ based on gyrochronology. Therefore, \\hdbin\\ is now a unique benchmark among field L and T~dwarfs, with a well-determined mass, luminosity, and age. We find that substellar theoretical models disagree with our observations. (1) Both components of \\hdbin\\ appear to be overluminous by a factor of $\\approx$2--3$\\times$ compared to evolutionary models. The age of the system would have to be notably younger than the gyro age to ameliorate the luminosity disagreement. (2) Effective temperatures derived from evolutionary models for HD~130948B and C are inconsistent with temperatures determined from spectral synthesis for objects of similar spectral type. Overall, regardless of the adopted age, evolutionary and atmospheric models give inconsistent results, which indicates systematic errors in at least one class of models, possibly both. The masses of \\hdbin\\ happen to be very near the theoretical mass limit for lithium burning, and thus measuring the differential lithium depletion between B and C will provide a uniquely discriminating test of theoretical models. The potential underestimate of luminosities by evolutionary models would have wide-ranging implications; therefore, a more refined age estimate for \\hdprim\\ is critically needed. ", "introduction": "More than a decade after their discovery, brown dwarfs continue to offer key insights into the astrophysics governing of some of the lowest temperature products of star formation. Brown dwarfs in the field are particularly useful as probes of very cold atmospheres. For instance, the atmospheres of extrasolar planets are very difficult to study directly due to their intrinsic faintness and proximity to very bright stars. However, brown dwarfs are typically found in relative isolation, and their atmospheres are subject to the same processes (e.g., dust formation and sedimentation, and non-equilibrium molecular chemistry) that are at work in their much less massive plantary counterparts. Despite the broad relevance of brown dwarfs, their fundamental properties remain poorly constrained by observations. In particular, very few direct mass measurements are available for brown dwarfs. To date, a total of six objects have been identified as unambiguously substellar \\citep[$M<0.072$~\\Msun\\ at solar metallicity;][]{2000ARA&A..38..337C} via dynamical mass measurements with precisions ranging from 6--9\\%: both components of the T5.0+T5.5 binary 2MASS~J15344984$-$2952274AB \\citep{2008arXiv0807.0238L}, both components of the young M6.5+M6.5 eclipsing binary 2MASS~J05352184$-$0546085 in the Orion Nebula \\citep{2006Natur.440..311S}, and two tertiary components of hierarchical triples in which the primaries are M stars, GJ~802B \\citep{gl802b-ireland} and Gl~569Bb \\citep{2001ApJ...560..390L, 2004astro.ph..7334O, 2006ApJ...644.1183S}. Direct mass measurements of brown dwarfs are critical for empirically constraining substellar evolutionary models. Since brown dwarfs have no sustainable source of internal energy, they follow a mass--luminosity--age relation, rather than the simpler mass--luminosity relation for main-sequence stars. Thus, mass measurements alone cannot fully constrain theoretical models, although mass and luminosity measurements of brown dwarfs in coeval binary systems can offer stringent tests of theoretical models \\citep[e.g.,][]{2008arXiv0807.0238L}. To fully constrain evolutionary models, systems with independent measurements of the mass, age, and \\Lbol\\ (or one of the much less observationally accessible quanities \\Teff\\ or $R$) are required. Such systems are quite rare, but they represent the gold standard among ``benchmark'' brown dwarfs. \\citet{2002ApJ...567L.133P} discovered the L~dwarf binary \\hdbin\\ in a hierarchical triple configuration with the young solar analog \\hdprim\\ (G2V) using the curvature adaptive optics (AO) system Hokupa`a on the Gemini North Telescope on 2001 February 24 UT. The L~dwarfs are separated from each other by $\\lesssim$0$\\farcs$13, and they lie 2$\\farcs$6 from the primary G star. \\hdbin\\ has been the target of AO-fed slit spectroscopy with NIRSPEC on the Keck II Telescope (1.15--1.35~\\micron) and IRCS on the Subaru Telescope (1.5--1.8~\\micron, 1.95--2.4~\\micron). \\citet{2002ApJ...567L..59G} used the latter spectra to determine the spectral types of the B and C components, both L4$\\pm$1, via spectral template matching. These are consistent with the less precise NIRSPEC $J$-band spectral types of dL2$\\pm$2, which are on the \\citet{1999AJ....118.2466M} system, found by \\citet{2002ApJ...567L.133P} for both HD~130948B and C. We present here a dynamical mass measurement for \\hdbin\\ based on Keck natural guide star adaptive optics (NGS AO) imaging of \\hdbin, as well as an analysis of {\\sl Hubble Space Telescope} (\\HST) and Gemini archival images. In addition to an independent age estimate (\\hdage, see \\S \\ref{sec:age}), the primary star provides a wealth of information about the system. \\citet{2005ApJS..159..141V} measured a solar metallicity for \\hdprim\\ ([M/H]~=~0.00, [Fe/H]~=~0.05), which is important since metallicity can play a significant role in shaping the spectra of brown dwarfs \\citep[e.g.,][]{2005astro.ph..9066B,2006liu-hd3651b}. Most importantly, the distance to \\hdprim\\ has been measured very precisely by \\Hipparcos, with a revised parallax of 55.01$\\pm$0.24~mas \\citep{2007hnrr.book.....V}, corresponding to a distance of $d$~=~18.18$\\pm$0.08~pc. Thus, the distance is measured to an exquisite precision of 0.44\\%, which is invaluable since the error in the dynamical mass scales as 3$\\times$ the distance error (i.e., the 0.44\\% error in distance translates into a 1.3\\% error in mass). HD~130948BC can thus serve as both an ``age benchmark'' and ``mass benchmark'' system in studying brown dwarfs. In the literature, the term benchmark is often applied to any readily observable unique or extreme objects, but here we specifically use the term to refer to systems for which fundamental properties may be directly determined. \\citet{2006MNRAS.368.1281P} highlighted the value of systems where the age and composition of substellar objects can be independently determined, e.g., from a stellar or white dwarf companion, and \\citet{2008arXiv0807.0238L} described an equivalent use of systems with dynamical mass measurements. Essentially, since brown dwarfs follow a mass--luminosity--age relation, the measurement of either mass or age in addition to the measured luminosity allows any other quantity to be fully specified using evolutionary models. This approach can be extremely useful, for example, by offering precise determinations of \\Teff\\ and \\logg, which can then be compared directly to atmospheric models. Of course, the measurement of mass, age, {\\em and} luminosity offers a direct test of evolutionary models, which is possible with \\hdbin. ", "conclusions": "We have determined the orbit of the young L4+L4 binary \\hdbin\\ using relative astrometry of the system spanning 7 years of its 10-year orbital period. The astrometric measurements and their uncertainties were extensively tested through Monte Carlo simulations. The fitted orbital parameters and revised \\Hipparcos\\ parallax give a total dynamical mass of 0.109$\\pm$0.002~\\Msun. The precision in mass is 2\\%, with nearly equal contributions to the uncertainty from the 1.7\\% error in the best-fit orbit and the 1.3\\% error in mass from the \\Hipparcos\\ parallax error. For any plausible mass ratio, both components of \\hdbin\\ are unambiguously substellar. \\hdbin\\ has the most precise mass determination for a brown dwarf binary to date. The primary star \\hdprim\\ offers an independent constraint on the age of the system from various indicators: rotation, chromospheric activity, isochrone fitting, X-ray emission, and lithium depletion. The ensemble of all available age indicators is consistent with an age for \\hdprim\\ similar the Hyades (625~Myr). For example, its rotation period is inconsistent with ages much younger than the Hyades and its chromospheric activity is inconsistent with ages much older than the Hyades. Our preferred age estimate is \\hdage, derived from the gyrochronology formalism of \\citet{2007ApJ...669.1167B} and \\citet{mam08-ages}. With a measured mass, luminosity, and age, \\hdbin\\ provides the first direct test of the luminosity evolution predicted by theoretical models for substellar field dwarfs. Both the Tucson models \\citep{1997ApJ...491..856B} and Lyon models \\citep[DUSTY;][]{2000ApJ...542..464C} underpredict the luminosities of HD~130948B and C given their masses and age. The discrepancy is quite large, about a factor of 2 for the Lyon models and a factor of 3 for the Tucson models. In order to explain this discrepancy entirely, model radii would have to be underpredicted by 30--40\\%. The age of \\hdprim\\ would need to be $\\approx$0.4~Gyr younger than we have estimated in order to resolve this discrepancy. This is inconsistent with the preferred gyro age but can be accommodated by other age indicators; a more refined age estimate for \\hdprim\\ is critically needed. Since the mass of \\hdbin\\ is more precisely determined than its age, we have used the mass with the individual bolometric luminosities to infer all other properties (age, \\Teff, etc.) from evolutionary models. We use a Monte Carlo approach to compute model-inferred quantities, and we are careful to account for covariance between the observational errors, the most notable of which is the correlation of the luminosities of the two components through their measured flux ratio. Because we use mass and \\Lbol\\ to derive model-inferred properties, any potential systematic errors in luminosity evolution will be reflected in the model-inferred quantities. For example, the very precise model-inferred ages for \\hdbin\\ (0.41$^{+0.04}_{-0.03}$~Gyr from Tucson models; 0.45$^{+0.05}_{-0.04}$~Gyr from Lyon models) are self-consistent, but they are inconsistent with the independent age estimate for \\hdprim\\ (\\hdage). Lacking measured radii for \\hdbin, we have used evolutionary models to derive effective temperatures. Given the mass and luminosity of each component, evolutionary models predict effective temperatures of $\\approx$1900--2000~K. Alternatively, given the mass of each component and age of the primary star, evolutionary models predict effective temperatures of $\\approx$1600--1700~K. (The disagreement between these two temperature ranges is just a reflection of the systematic errors in luminosity evolution.) Spectral synthesis using atmospheric models gives temperatures of 1700--1800~K for objects of similar spectral type to \\hdbin\\ \\citep[L3--L5;][]{2008ApJ...678.1372C}. Using evolutionary models and the measured luminosity ratio gives $\\Delta$\\Teff~=~90~K. Resolved spectroscopy of HD~130948B and C has previously shown that they have indistinguishable spectral types, so this rather large temperature difference may indicate that spectral type does not hold a one-to-one correspondence with \\Teff\\ mid-L~dwarfs, even for two coeval objects. Better spectral types for the two components of \\hdbin\\ are needed to address this apparent discrepancy. Comparing the different effective temperature determinations for \\hdbin\\ on the H-R diagram shows that the evolutionary models, atmospheric models, and observational data cannot be simultaneously brought into consistency with each other, regardless of the adopted age of the system. Thus, systematic errors in some combination of the atmospheric and/or evolutionary models are needed to explain the observed discrepancy. The best current age estimate indicates that both evolutionary and atmospheric models harbor systematic errors. Further evaluation of the disagreement between models and the data requires a refined age estimate for \\hdprim. Resolved multi-band spectroscopy of \\hdbin\\ is also needed to reduce the uncertainties in the atmospheric model effective temperatures by direct spectral synthesis fitting. We also find large discrepancies when comparing the observed near-infrared colors of \\hdbin\\ to the Lyon models. This suggests that using color-magnitude diagrams to infer the properties of field L~dwarfs from evolutionary models will lead to large errors in the resulting quantities (e.g., mass and/or age). For example, if we inferred the ages and masses of the components of \\hdbin\\ from the model $J-K$ or $H-K$ color-magnitude diagrams, we would derive masses that are $\\approx$20--30\\% smaller than observed and ages $\\approx$2$\\times$ younger than the age of the primary star (\\hdage). One novel aspect of using \\hdbin\\ to constrain theoretical models is the application of the ``binary lithium test'', originally proposed by \\citet{2005astro.ph..8082L}. This is made possible by the fortuitous circumstance that the components of \\hdbin\\ are very near the mass limit for lithium burning. As a consequence, the Lyon and Tucson evolutionary models, which are almost indistinguishable in their predictions of substellar bulk properties, give very different predictions for the amount of primordial lithium remaining in the B and C components. Thus, resolved optical spectroscopy to detect the lithium doublet at 6708~\\AA\\ would provide a very discriminating test of the evolutionary models. Such a constraint is significant in that it directly tests the properties of fully convective substellar interiors (e.g., the core temperature) and/or the lithium reaction rates. \\hdbin\\ is the only system currently known for which such an empirical calibration of lithium burning is possible. Substellar theoretical models are in sore need of empirical validation as they have been employed for more than a decade to interpret observations of field dwarfs. Given the independent constraints on the age and composition provided by a stellar companion, dynamical mass measurements for triple systems like HD~130948ABC provide the most challenging tests of substellar theoretical models. However, substellar companions to stars are quite rare \\citep[$\\approx$1$\\pm$1\\%, e.g.,][]{2001AJ....121.2189O, 2004AJ....127.2871M, 2005AJ....130.1845L, 2007ApJS..173..143B, 2007ApJ...670.1367L}, and even more rare are substellar binary companions that yield dynamical mass measurements in a reasonable time frame. When the stellar companion is a bright star like \\hdprim, a wealth of additional information is available, the most important of which is a very precise \\Hipparcos\\ distance measurement since this is the limiting factor in the precision of the dynamical mass. Stars bright enough to enable seismological measurements can yield the most stringent (10--20\\%) age determinations possible \\citep[e.g.,][]{2005A&A...434.1085C,2008ApJ...673.1093B}. Thus, \\hdbin\\ represents a rare class of benchmark systems for which the most precise mass and age determinations are possible. Our observations of \\hdbin\\ indicate that substellar models currently harbor significant systematic errors. The potential underestimation of \\Lbol\\ by evolutionary models has far-reaching implications. For example, such models have been used to determine the low-mass end of the intial mass function and to predict the radii of extrasolar planets. Obtaining measurements for more systems like \\hdbin\\ over a broad range of mass, luminosity, and age will be critical in understanding and resolving the discrepancies that have been revealed between observations and theoretical models." }, "0807/0807.0039_arXiv.txt": { "abstract": "We discuss the cosmological degeneracy between the Hubble parameter $H(z)$, the age of the universe and cosmological parameters describing simple variations from the minimal $\\Lambda$CDM model. We show that independent determinations of the Hubble parameter $H(z)$ such as those recently obtained from ages of passively evolving galaxies% , combined with Cosmic Microwave Background data (WMAP 5-years), provides stringent constraints on possible deviations from the $\\Lambda$CDM model. In particular we find that this data combination constrains at the 68\\% (95\\%) \\textit{c.\\,l.}\\,\\,the following parameters: sum of the neutrino masses $\\sum m_\\nu < 0.5\\,(1.0)$ eV, number of relativistic neutrino species $N_{\\rm rel} = 4.1^{+0.4}_{-0.9}\\,(^{+1.1}_{-1.5})$, dark energy equation of state parameter $w = -0.95 \\pm 0.17\\,(\\pm\\,0.32)$, and curvature $\\Omega_k = 0.002 \\pm 0.006\\,(\\pm\\,0.014)$\\,, in excellent agreement with dataset combinations involving Cosmic Microwave Background, Supernovae and Baryon Acoustic Oscillations. This offers a valuable consistency check for systematic errors. ", "introduction": "The recent measurements of cosmic microwave background (CMB) anisotropies and polarization \\cite{spergelwmap2,dunkleywmap08}, alone or in combination with other cosmological data sets, have provided confirmation of the standard cosmological model and an accurate determination of some of its key parameters. In particular, the new determination of the age of the Universe $13.68 \\pm 0.13$ Gyrs improves by an order of magnitude previous determinations from, e.g., cosmochronology of long-lived radioactive nuclei \\cite{thorium} and population synthesis of the oldest stellar populations \\cite{Jimenez96,dunlop96,chaboyerkrauss03} and by a factor of 2 previous determinations from CMB data. With cosmological parameters so tightly constrained within the framework of the standard flat-$\\Lambda$CDM model, it is important however to constrain possible deviations from the standard cosmological model. Beyond the primordial parameters describing the shape of the primordial power spectrum and late-time parameters such as the optical depth to the last scattering surface, CMB observations so far constrain directly parameters such as \\cite{kosowsky} the angular size distance to last scattering combined with the sound horizon at decoupling, the baryon-to-photon ratio and the redshift of matter radiation equality. This implies that, for models beyond the standard flat-$\\Lambda$CDM, CMB data alone still show large degeneracies among ``derived\"\\footnote{Note that the name \"derived parameters\" has been sometimes used in the literature with a slightly different emphasis, denoting parameters such as the bias $b$, $\\sigma_8$, etc..} cosmological parameters such as the matter density parameter $\\Omega_m$, the curvature $\\Omega_k$, the dark energy equation of state paramenter $w$, the effective number of relativistic neutrino species $N_{\\rm eff}$, the sum of neutrino masses $\\sum m_{\\nu}$ and the Hubble parameter $H_0$. For example (see e.g., \\cite{Eisensteinwhite04, debernardisetal08, de Bernardis:2007bu}), departures from the standard model described by a deviation from $3$ neutrino species, can arise from the decay of dark matter particles \\cite{bonometto98, lopez98, hannestad98, kaplinghat01}, quintessence \\cite{bean}, exotic models \\cite{unparticles}, and additional hypothetical relativistic particles. This affects the matter-radiation equality yielding, even for a flat, cosmological constant-dominated model, a degeneracy between $N_{\\rm eff}$, $H_0$ and $\\Omega_m$. A departure from dark energy being described by a cosmological constant (i.e. a component with equation of state $w\\neq-1$), yields a different angular size distance to last scattering, and thus degeneracy between $w$, $H_0$ and $\\Omega_m$ even for a flat universe. Finally, relaxing the flatness assumption yields the so-called ``geometric degeneracy\" (between age or $H_0$ and $\\Omega_m$ and $\\Omega_{\\Lambda}$). In order to go beyond the concordance $\\Lambda$CDM model parameters determination, one needs extra data sets that probe different physics and are affected by different systematics. In this work we concentrate on the measurements of $H(z)$ using passively evolving red-envelope galaxies and how using them helps to constrain cosmological parameters dropping the assumption of the concordance model. In particular we show that the recent determinations of $H_0$ from the HST key project \\cite{hstkey} and $H(z)$ provided by \\cite{svj} (SVJ; based on \\cite{data} and references therein) can provide, when combined with CMB and other cosmological data, new and tighter constraints on deviations from the standard $\\Lambda$CDM model, as first shown in \\cite{de Bernardis:2007bu, data}. This approach of combining different data-sets to constrain parameters that are otherwise poorly constrained, is called ``concordance approach\". While it is a very powerful approach, the same ``concordance\" approach is used to test data sets for systematic errors. It is therefore important to consider enough data sets to have an over-constrained problem and as diverse data sets as possible, relying on different physics and affected by different systematics. Only in this case, if all data sets agree, one can be confident that the systematic errors are safely below the statistical errors and that the cosmological constraints are robust. After obtaining constraints on deviations from the simple $\\Lambda$CDM model obtained with WMAP 5-years data and $H(z)$ measurements, we compare them with those obtained from the combination of WMAP 5-years with Baryon Acoustic Oscillations and Supernovae data. We find good agreement between the two approaches. We conclude that any possible systematic effect in the non-CMB data sets is below the statistical errors, and that there is no evidence for a deviation from the % flat-$\\Lambda$CDM model, thus offering support to the standard cosmological model. ", "conclusions": "Measurements of $H(z)$ constrain the age of the Universe at different redshifts and thus break the CMB-only degeneracy between the age of the universe today ($t_0$) and the parameters describing deviations from the $\\Lambda$CDM model. As shown in table~\\ref{tab:ages}, the age of the Universe as constrained by WMAP5-only data, is very sensitive to the presence of some of these parameters, especially to the possibility of having a background of $N_{\\rm eff}$ relativistic particles (with $N_{\\rm eff}$ not fixed to $3.04$) or allowing a non-zero curvature (see first row, 3rd and 5th columns of table~\\ref{tab:ages}). This is because many models which deviates from the standard $\\Lambda$CDM but are consistent with CMB data, are not a good fit to the $H(z)$ data. Some illustrative examples are shown in Figure 1. The combination WMAP5+H significatively reduces the degeneracy between $t_0$ and some of the ``extra\" parameters, thus improving the constraints on the age of the Universe in those models, for example, by almost a factor of 3 for the case with $N_{\\rm eff}$ not fixed to 3.04, and almost a factor of 5 for the case of non-zero curvature (see same columns but second row, in table~\\ref{tab:ages}). \\begin{table}[htb] \\caption{\\label{tab:ages} Determination of the age of the Universe ($68.3 \\%$ c.l.) in several different cosmological models for WMAP 5-years data alone and WMAP5+H data. } \\begin{center} \\begin{tabular}{|c||c|c|c|c|c|} \\hline AGE (Gyr) & $\\Lambda$CDM & $\\Lambda$CDM+$N_{\\rm eff}$ & $\\Lambda$CDM +$\\sum m_{\\nu} $ & $\\Lambda$CDM+$\\Omega_k$ & wCDM \\\\ \\hline & & & & &\\\\ WMAP 5-years & $13.69 \\pm 0.13$ & $12.08 \\pm 1.29$ & $14.06\\pm 0.27$ &$16.32 \\pm 1.76$ & $13.74 \\pm 0.34$ \\\\ & & & & &\\\\ WMAP5+H & $13.65^{+0.14}_{-0.10}$ & $12.87^{+0.61}_{-0.31}$ & $13.81^{+0.24}_{-0.14}$ & $13.61^{+0.29}_{-0.44}$ & $13.67^{+0.24}_{-0.08}$ \\\\ & & & & &\\\\ \\hline \\end{tabular} \\end{center} \\end{table} In Figure \\ref{lcdm_mnu} we explore the resulting constraints on the neutrino properties. In all cases the dotted line is the WMAP 5-years only result, the dashed line is WMAP5+HST and the solid line is WMAP5+H. We find that the combination WMAP5+H constrains the sum of neutrino masses to be $\\sum m_{\\nu} < 0.48$ eV and $< 0.93$ eV at the 68\\% and 95\\% confidence levels, respectively, thus improving the WMAP-only constraints by 50\\%. The constraints on the effective number of neutrino species is $N_{\\rm rel} = 4.1^{+0.4}_{-0.9}\\,(^{+1.1}_{-1.5})$ at the 68\\% (95\\%) confidence level and $N_{\\rm rel} > 2.2 $ at better than the 99\\% confidence level. For comparison, note that Ref.~\\cite{de Bernardis:2007bu} obtained $N_{\\rm rel} = 4.0 \\pm 1.2$ at the 68\\% confidence level and $N_{\\rm rel} > 1.8$ at better than 99\\% confidence level, from WMAP 3-years data combined with the same $H(z)$ measurements we use in this paper. Therefore, the improvement we obtain in the constraints of $N_{\\rm rel}$, simply reflects the improvement of the CMB alone constraints in the WMAP 5-years data (compare e.g. the age constraints in table 1 of \\cite{de Bernardis:2007bu} and our table 1). Note that our results are compatible (within the 1-2$\\sigma$ regions) with the constraints on $N_{\\rm eff}$ obtained~\\cite{SimhaSteigman} from BBN alone or combined with other data sets, as well as have statistical errors of the same order (see table 1 from ~\\cite{SimhaSteigman}). \\begin{figure}% \\hspace{-0.55cm}\\includegraphics[width=8cm, angle=0]{lcdm+curv_WMAP5.eps} \\hspace{-0.1cm}\\includegraphics[width=8cm, angle=0]{cdm+wlam_WMAP5.eps} \\caption{\\textit{Left}: Constraints on the curvature of a $\\Lambda$CDM model from WMAP 5-years alone (dotted line), WMAP5+HST (dashed line) and WMAP5+H (solid line). With the $H(z)$ measurements, the curvature is constrained to 0.002 $\\pm$ 0.006 ($\\pm$ 0.014) at the $68\\%$ ($95\\%$) confidence level. The WMAP 5-years only line shows the well known geometric degeneracy. \\textit{Right}: Constraints on the dark energy equation of state parameter from WMAP 5-years alone (dotted line), WMAP5+HST (dashed line) and WMAP5+H (solid line). With the $H(z)$ measurements we obtain $w = -0.95 \\pm 0.17\\,(\\pm\\,0.32)$ at the 68\\% (95\\%) confidence level.} \\label{fig:olcdm} \\end{figure} The left panel of Figure \\ref{fig:olcdm} shows the constraints on the geometry of the Universe. The WMAP5+H combination yields $\\Omega_k=0.002 \\pm 0.006\\,(\\pm\\,0.014)$ at the 68\\% (95\\%) confidence levels, thus breaking the geometric degeneracy. In the right panel of Figure \\ref{fig:olcdm}, we report the constraints on the dark energy equation of state parameter (asssumed constant). The WMAP5+H combination yields $w = -0.95 \\pm 0.17\\,(\\pm\\,0.32)$ at the 68\\% (95\\%) confidence level, which improves the WMAP 5-years only constraints by a factor $\\sim 70\\%$. While the WMAP 5-years constraint has a hard prior on the Hubble constant $H_0<100$ km/s/Mpc which imposes a lower limit on $w$, the WMAP5+HST and WMAP5+H combinations are insensitive to this prior. In table \\ref{tab:compare} we compare the WMAP5+H constraints on deviations from the $\\Lambda$CDM model, with those obtained by the combination WMAP 5-years with Baryon Acoustic Oscillation data (BAO) \\cite{percivaletal07} and with Supernovae as obtained by \\cite{komatsuwmap08}. \\begin{table}[htb] \\caption{\\label{tab:compare} Cosmological constraints at $68$\\% ($95 \\%$) c.l.\\,\\,on the extra parameters characterizing deviations of the standard $\\Lambda$CDM model% , comparing their values as extracted from WMAP 5-years only, WMAP5+BAO+SN and WMAP5+H.} \\begin{center} \\begin{tabular}{|c||c|c|c|c|c|c|} \\hline Parameter & WMAP 5-years only & WMAP5+BAO+SN & WMAP5+H\\\\ \\hline & & &\\\\ $N_{\\rm eff}$ & $> 2.3\\,\\, (95 \\%)$ & $4.4^{+1.5}_{-1.5}\\,^{(*)}$ & $4.10^{+0.37}_{-0.94} (^{+1.12}_{-1.50})$\\\\ & & &\\\\ $\\sum m_\\nu$ & $< 1.3$ eV (95 \\%) & $< 0.61$ eV (95 \\%) & $< 0.93$ eV (95 \\%) \\\\ & & &\\\\ $w$ & $^{> -2.37}_{< -0.68}\\,\\, (95 \\%)$ & $-0.972_{-0.060}^{+0.061} (_{-0.138}^{+0.112})$ & $-0.945_{-0.155}^{+0.194} (_{-0.350}^{+0.311})$\\\\ & & &\\\\ $\\Omega_k$ & $^{< +0.017}_{> -0.063}\\,\\, (95 \\%) $ & $-0.0052^{+0.0064}_{-0.0064} (^{+0.0137}_{-0.0123})$ & $0.002^{+0.0059}_{-0.0059} (^{+0.012}_{-0.018})$\\\\ & & & \\\\ \\hline \\end{tabular} \\end{center} $^{(*)}$with HST prior \\end{table} \\begin{figure}% \\hspace{-0.0cm}\\vspace{0.2cm}\\includegraphics[width=7.5cm, height=6.2cm, angle=0]{contour_w_omega_k.eps} \\hspace{0.2cm}\\includegraphics[width=7.5cm, height=5.8cm, angle=0]{wmapfig.eps} \\caption{\\textit{Left}: Constraints in the $\\Omega_k$-$w$ plane from WMAP 5-years alone (purple), WMAP5+H (blue 68\\% and 95\\% c.l.). For comparison we show (right) the wmap team's WMAP 5-years only constraints (black), WMAP5+BAO (yellow) and WMAP5+Supernovae (red) (see \\cite{komatsuwmap08}). The differences in the WMAP 5-years only constraints are due to different choice of priors. Most notably, different boundaries on the $H_0$ prior are used: $0.41000$ is strongly uncompatible with any mass-follows-light model} \\citep{gil}. Conversely, a realistic NFW model reasonably reproduces the observed velocity profile and, in principle, does not conflict with the presence of an overdensity of baryons at its center, like Sgr,N. However, it has to be recalled that the NFW model does not make any definite prediction on the SB profile of the embedded stellar nucleus and the compatibility of the observed SB profile of Sgr,N with the NFW model that that fits the velocity profile is not established. \\item M54 and Sgr,N have definitely different kinematical properties. In particular, the velocity dispersion profiles are very different: in the radial range $1.5\\arcmin < r < 6.5\\arcmin$ the statistical significance of the difference in the velocity distribution is very large. \\item We have provided observational evidence that the velocity dispersion profile of Sgr remains flat from $r\\simeq 1\\arcmin$ to $r\\simeq 100\\arcmin$ and that there is no apparent transition in the velocity profile corresponding to the onset of the stellar nucleus, at $r\\la 10\\arcmin$. This fact as well as those listed above strongly suggest that Sgr,N and M54 had independent origins, as we would expect that, if the cluster provided the mass seed to collect the Sgr gas that later formed the metal-rich nucleus, Sgr,N stars would have shown a declining velocity dispersion profile, compatible with a mass-follows-light distribution. \\item Our N-body simulations that follow the orbits of massive clusters (1-2 $\\times 10^6 M_{\\sun}$, representing M54) within different NFW halos (representing the Sgr galaxy) show that for a large range of initial distances and relative velocities, the orbit of the cluster decays completely by dynamical friction within 3 Gyr, at most. Moreover, at the end of the simulations, the cluster is perfectly concentric with the cusp of the host halo (within the resolution of the simulation) and the difference in average velocity is always less than $\\sim 2$ km/s. Hence, the observed phase-space coincidence between M54 and Sgr,N can be naturally explained by the ``dynamical friction hypothesis'' (M05a). \\item According to FC06 the mass of Central Massive Objects, independently if they are Black Holes or stellar nuclei, is $\\simeq \\frac{3}{1000}$ of the mass of the host galaxy ($\\frac{M_{CMO}}{M_{gal}}=0.003$). Using our estimates for the total mass of Sgr,N and M54 we obtain the following estimates for the total mass of the Sgr galaxy, depending on whether we assume is the M54, Sgr,N or the sum of the two (as we would do if we observed the system at the distance of Virgo cluster) is the Central Massive Object: $M_{Sgr}=1.0\\times10^9~M_{\\sun}$, $M_{Sgr}=2.1\\times10^9 ~M_{\\sun}$, and $M_{Sgr}=3.0\\times10^9~M_{\\sun}$, respectively. These values are in very good agreement with the independent estimates by \\citet{maj} that ranges from $M_{Sgr}=5.8\\times10^8 ~M_{\\sun}$ to $M_{Sgr}=6.9\\times10^9~M_{\\sun}$. Hence both Sgr,N and M54 singularly or taken together have a mass compatible with being the CMO of the Sgr galaxy. \\item Both Sgr,N and M54, as well as the combination of the two, when placed in the $M_V$ vs. log$r_h$ diagram lie in a region that is populated by globular clusters {\\em and} galactic stellar nuclei. They are also compatible with the Color-Magnitude relation of nuclei shown by FC06 and with the $M_V$ vs. $\\sigma$ relation satisfied by globular clusters, nuclei and UCDs \\cite[see][]{geha,evsti}, as shown in Fig.~\\ref{mvsigma} Hence the structure and dynamics of Sgr,N, M54 and their combinations are fully compatible with other galactic nuclei. \\item A detailed study of the mass profile of Sgr,N and M54 using the Schwarzschild method \\cite[see][and references therein]{rix,ven} is currently ongoing and the results will be presented elsewhere (Ibata et al., in preparation). \\end{enumerate} These findings lend very strong support to the scenario proposed by M05a to explain the M54/Sgr,N system: the nucleus of the galaxy formed {\\em in situ}, at the bottom of the potential well of the Sagittarius galaxy; the {\\em globular cluster} M54 was independently driven to the same site by dynamical friction. As a complement of the above conclusions, it must be recalled that the stellar population that dominates Sgr,N formed several Gyrs later than M54 \\citep[see][and references therein]{paul,sl97,ls00,mic_con,siegel}. In the present context this does not appear particularly relevant, but it should be kept in mind that M54 can have reached the very center of Sgr {\\em after} Sgr,N formed its stars \\cite[within the last $\\sim 5-9$ Gyr][]{mic_con}, or even {\\em before} this, depending on the birthplace of the cluster within the Sgr galaxy. In any case it is very likely that both the cluster and the processed Sgr gas were {\\em independently} driven - by different mechanisms - to the bottom of the potential well of the Sgr galaxy, i.e. to the center of its Dark Matter halo. A conclusive proof that M54 was driven to its present position by dynamical friction could be provided by the successful detection of genuine extra-tidal stars at large distances from the cluster center as envisaged in Sect.~4.3, while, as said their non-detection will not disprove the above scenario. As a possible alternative to this view (or as an extreme version of it) it can be conceived that M54 formed at the bottom of the overall potential well of the Sgr galaxy since the beginning. In this framework M54 and Sgr,N are just the results of two subsequent episodes of star formation both occurring at the very center of Sgr, the second from enriched gas that was infalling on very different orbits with respect to the gas that formed M54, thus resulting in different final stellar kinematics. We regard this possibility as much less likely with respect to the ``dynamical-friction'' scenario depicted above. First, if M54 is considered as part of the (metal-poor) field population of Sgr, its presence would be at odd with the global metallicity gradient observed by \\citet{alard}, \\citet{sdgs2}, \\citet{sl97} and others in Sgr, as well in all dwarf spheroidal galaxy studied to date \\cite[more metal rich and younger populations are preferentially found at the center of dSph's][]{harbeck}. While dE nuclei have been indicated as a possible exceptions to this trend, the ingestion of large metal poor globular clusters in pre-existing metal rich nuclei seems one possible natural way to reconcile the generally observed metallicity gradients and the presence of nuclei that are bluer than their parent galaxy \\citep{lotz4}. Second, as both M54 and Sgr,N would have {\\em formed} from gas falling into the same potential well, the reason for the different kinematics remains to be explaned, while it is a natural outcome if M54 formed elsewhere as a classical globular cluster. \\begin{figure} \\plotone{mvsigma.ps} \\caption{Sgr,N and M54 (filled triangles) in the $M_V$ vs. $\\sigma$ plane \\cite[see][]{evsti}. Both systems lie on the locus common to globular clusters, dE nuclei, Dwarf-Globular Transition Objects \\cite[DGTO][]{hase}, and, possibly, UCDs. The same is true for their combination (Sgr,N+M54, open triangle), i.e. for a system having luminosity equal to the sum of the luminosities of M54 and Sgr,N and having $\\sigma$ as estimated from an integrated spectrum obtained with a $1\\arcsec$ slit from the distance of the Virgo cluster of galaxies ($\\sigma\\simeq 12.4$ km/s). Data for Galactic globulars are from \\citet[][$M_V$]{macsyd} and \\citet[][$\\sigma$]{tad}; for M31 globulars we took integrated magnitudes from the Revised Bologna Catalog \\citep{rbc} and $\\sigma$ from \\citet{djor}; data for dE and dE,N are from \\citet{geha,geha2}; UCDs data are from \\citet{ucd} and \\citet{evsti}; data for DGTO are from \\citet{hase}. Some remarkable bright clusters have been labeled, for reference \\cite[see][]{lucky}. \\label{mvsigma}} \\end{figure} \\subsection{The process of galaxy nucleation} If we take for demonstrated the above concluding remarks, we can ask what we have learned about the process of galaxy nucleation from the case studied here. Concerning the two main mechanisms that have been put forward in the literature, i.e. (a) formation of the nucleus by infall of globular cluster(s) to the center of the galaxy, or (b) {\\em in situ} formation by accumulation of gas at the center of the potential well and its subsequent conversion into a stellar overdensity \\citep[see Sect. 1. and][]{grant,cote6}, the main conclusion that can be drawn from the case of Sgr is that both channels are viable and actually both have been at work ``simultaneously'' in Sgr. The present analysis has shown that a stellar nucleus made of the typical material that dominates the baryonic mass budget of Sgr is present in this galaxy, independent of M54, as it display the same flat velocity dispersion profile as the whole core of Sgr, much different from that of the cluster. In a likely scenario, the enriched gas from previous generations of Sgr stars accumulated at the bottom of the overall potential well of the galaxy, until star formation transformed it into a stellar nucleus whose surface brightness is $\\ga 100$ times larger than in the surrounding Sgr core, a substructure within a larger galaxy. On the other hand, M54 is a (relatively) ordinary massive, old and metal poor globular cluster. Independent of its birthplace within the early Sgr galaxy, its mass and the density of the surrounding medium of the host galaxy drove it to the bottom of the Sgr potential well, by dynamical friction. During its trip to the densest central region of Sgr, the dense cluster managed to survive the tidal force of the host galaxy, hence it reached the present position as a (partially?) self-bound stellar system. While the mass budget at the center of Sgr is probably dominated by the underlying DM halo, M54 dominates the overall light distribution : an observer taking photometry and/or spectra of the unresolved nucleus of Sgr from a distant galaxy (say, a galaxy in the Virgo cluster) would find that the object looks like a bright and blue globular cluster; the integrated velocity dispersion would not show any peculiar feature revealing the composite nature of the observed nucleus (see Fig.~\\ref{mvsigma}, above). Probably, it would be impossible to disentangle the two systems from the integrated light\\footnote{It would be interesting to check if there is some spectral feature that may reveal the composite nature of the ``system'', in the present case. We plan to do this in the future by combining properly scaled synthetic spectra representing the light output of M54 and Sgr,N.}. Finally, the Sgr case seems to support the observed ubiquity of nuclei (see Sect.~1. and C06). The Sgr galaxy was able to form a sizable nucleus ``twice'' and with two different formation channels: if either of the two channels had not been viable for some reason, the galaxy would have ended up with a nucleus in any case. The fact that Sgr is the only case of galaxy with a clear nucleus among those classified as dwarf spheroidals in the Local Group may suggest that the progenitor of Sgr was in fact a brighter dE or disc galaxy that has been transformed into a dSph by the interaction with the Milky Way \\cite[][M05b, and references therein]{maj,mayer} \\subsection{Suggestions for further investigations} The results presented in this paper suggest several interesting lines of research that we did not follow up for practical reasons. However we feel that it is worth briefly mentioning some of them here, as a possible starting point for future studies. \\begin{itemize} \\item The results presented by \\citet{read}, \\citet{goerdt} and \\citet{sanchez}, suggest that the complete decay of a massive cluster to the very center of the host galaxy is much easier and faster if a central density cusp is present. It is even possible that a central cusp is actually {\\em required} to bring a cluster down to the very bottom of the overall potential well. If this is the case, the position of M54 within Sgr,N would support the existence of a central cusp in actual DM halos, a point that has been questioned by several authors \\cite[see][and references therein]{sanchez}. The ``complete infall'' of a massive cluster may need a NFW cusp and, simultaneously, it may transform the cusp into a core by transferring orbital energy and momentum to the surrounding ``medium'', thus possibly providing a self-regulating mechanism that simultaneously prevent the further decay of other clusters \\citep[that, in general cases, would be quite difficult, given the expected decay times, see][and references therein]{herna,oh,read,sanchez,goerdt}. In this context, it may be worth to recall also the work by \\citet{stri}, whose results militates against the presence of a large-size core in the Fornax dSph, and by \\citet{boy} on the resilience of cuspy halos in major mergers. These ideas seems worthy of detailed theoretical follow up. \\item It has been suggested several times that bright and metal poor globular clusters may be of cosmological origin \\citep[see][and references therein]{BS,sills}. If this is the case they may be very intimately linked to the earliest phases of formation of galaxies and there may have been plenty of opportunities for most of them to become the nuclei of some dwarf galaxy. The tidal stress that they suffered during their infall to the center of their host galaxies may be at the origin of the larger half-light sizes that are observed in those of them that have been suggested as possible nuclear remnants of ancient dwarf galaxies \\citep{frebland,lucky,macsyd}. This kind of scenario may be explored with dedicated N-body simulations, possibly including gas and stars. It would be interesting also to consider in detail the results presented here in relation with the scenario for the origin of globular clusters recently discussed by \\citet{bok07}. \\item All the analysis presented in this paper have been performed within the standard Newtonian gravitation theory and dynamics. It may be worth to consider the observational scenario emerged from this study also in the framework of Modified Newtonian Dynamics paradigms \\citep[MOND, see][and references therein]{mil,sand}, even if it may not necessarily be an ideal case. The transition between ordinary Newtonian regime and MOND regime occurs around $r\\simeq 4.5\\arcmin - 6.5\\arcmin$, depending on the actual stellar mass of M54+Sgr,N, that is in a range covered by our data. \\item There is little doubt that the final fate of the Sgr galaxy will be its complete tidal disruption. Once the large scale stellar body of Sgr will be completely dispersed into the Galactic halo, the final remnant of this (once) relatively large galaxy would be a faint nucleus embedding a bright globular cluster. An observer lacking any knowledge of the origin of this object would conclude that it is a very bright and peculiar globular cluster, dominated by a metal poor population ($[Fe/H]\\sim -1.5$, possibly with some spread) but also including a small fraction ($\\sim$ 10\\%) of metal rich stars (with average $[Fe/H]\\sim -0.4$). Also the abundance patterns would appear different: for instance, the metal poor (M54) stars would appear as moderately $\\alpha$-enhanced \\citep{brown}, while metal rich stars (Sgr,N) would have solar or sub-solar $[\\alpha/Fe]$ ratios \\citep{luves}. The radial velocities would reveal that the metal rich and metal poor stars have {\\em the same systemic velocity}, but {\\em different velocity dispersion profiles} and {slightly different density profiles}, possibly with some rotation in the metal poor component. However the dominance of metal-poor (M54) stars would be so high that there would be no hint of the presence of a Dark Matter component. The half-light radius of the system would appear slightly larger with respect to ordinary globulars \\citep{macsyd,lucky}. Most of these likely characteristics of the future remnant of Sgr seem to have a counterpart in the widely studied and mysterious stellar system $\\omega$ Centauri \\citep[see][and references therein]{norris,omega,ele1,ele2,ele3,ven}, that has been proposed as the possible remnant of a nucleated dwarf elliptical since \\citet{freeman}. While there are also noticeable differences between the two cases, the analogy seems very intriguing and potentially powerful\\footnote{It is interesting to recall that $\\omega$ Cen shares with Sgr some remarkable chemical peculiarities. In particular, stars of comparable metallicity in the two systems are strongly enhanced in s-process elements and the ratio of heavy s-process to light s-process elements [hs/ls] is very similar. Furthermore, $\\omega$ Cen and Sgr are, up to now, the only stellar systems known to have deficient [Cu/Fe] ratios with respect to the trend in the Galactic disk and halo \\citep{mcw}}. It is possible that at least some of the observational features of $\\omega$ Centauri that appear so difficult to explain may find their natural place within a scenario like the one described above. \\end{itemize}" }, "0807/0807.4510_arXiv.txt": { "abstract": "The energy region spanning from $\\sim 10^{17}$ to $\\lesssim 10^{19}$ eV is critical for understanding both, the Galactic and the extragalactic cosmic ray fluxes. This is the region where the propagation regime of nuclei inside the Galactic magnetic environment changes from diffusive to ballistic, as well as the region where, very likely, the most powerful Galactic accelerators reach their maximum output energies. In this work, a diffusion Galactic model is used to analyze the end of the Galactic cosmic ray spectrum and its mixing with the extragalactic cosmic ray flux. In particular, we study the conditions that must be met, from the spectral and composition points of view, by the Galactic and the extragalactic fluxes in order to reproduce simultaneously the total spectrum and elongation rate measured over the transition region by HiRes and Auger. Our analysis favors a mixed extragalactic spectrum in combination with a Galactic spectrum enhanced by additional high energy components, i.e., extending beyond the maximum energies expected from regular supernova remnants. The two additional components have mixed composition, with the lowest energy one heavier than the highest energy one. The potential impact on the astrophysical analysis of the assumed hadronic interaction model is also assessed in detail. ", "introduction": " ", "conclusions": "" }, "0807/0807.1932_arXiv.txt": { "abstract": "A growing number of observations indicate that magnetic fields are present among a small fraction of massive O- and B-type stars, yet the origin of these fields remains unclear. Here we present the results of a VLT/FORS1 spectropolarimetric survey of 15 B-type members of the open cluster NGC 3766. We have detected two magnetic B stars in the cluster, including one with a large field of nearly 2 kG, and we find marginal detections of two additional stars. There is no correlation between the observed longitudinal field strengths and the projected rotational velocity, suggesting that a dynamo origin for the fields is unlikely. We also use the Oblique Dipole Rotator model to simulate populations of magnetic stars with uniform or slightly varying magnetic flux on the ZAMS. None of the models successfully reproduces our observed range in $B_\\ell$ and the expected number of field detections, and we rule out a purely fossil origin for the observed fields. ", "introduction": "Magnetic fields have long been inferred in solar and late-type stars and are produced by a dynamo mechanism in the outer convection zones. Studies of hot star structure and evolution have long neglected magnetic fields since the radiative envelopes of hot stars are not expected to produce them. Yet, an increasing number of observations have revealed that magnetic fields are present in $\\sim 5$\\% of hot stars \\citep{petit2008}, even though the origin of such fields is uncertain \\citep{neiner2007}. Various models have proposed that they may be generated in the convective cores, although no known mechanism can transport them to the surface in a timescale consistent with observations \\citep{charbonneau2001, macgregor2003}. Other models have proposed that the radiative envelopes somehow maintain a dynamo process near the stellar surface (e.g.\\ \\citealt{spruit2002}; \\citealt{macdonald2004}). Finally, magnetic fields in hot stars may be a relic from stellar formation, hence the present day field is a ``fossil'' of the primordial field. A growing number of studies have begun to investigate magnetic fields among massive stars in open clusters. Main-sequence (MS) and pre-main sequence (PMS) magnetic OB stars have recently been detected in the Orion Nebula Cluster \\citep{petit2008}, NGC 6611, and NGC 2244 \\citep{alecian2008} with a variety of ages, rotational periods, and chemical abundances. Currently, the most widescale investigation of magnetic, early-type cluster members considers 258 A- and B-type stars in more than 40 open clusters \\citep{bagnulo2006, landstreet2008}. These studies offer strong potential to infer the origin of the magnetic fields in massive stars. If the magnetic fields are relics of the molecular cloud from which the stars formed, then a population of coeval OB stars should offer widespread evidence of the initial field. A past or present dynamo likely depends on other physical conditions within the stars, especially rotation, and would likely be less widespread among cluster members. Observations currently suggest a fossil origin for the magnetic fields in massive stars. \\citet{landstreet2008} find that the strongest fields are associated with stars near the zero-age main-sequence (ZAMS) and that the field strength decays over time. For all but the strongest fields, the decay is consistent with magnetic flux conservation as the stars increase in radius during the MS lifetime. \\citet{landstreet2008} also find that some of the strongest fields are found in the most slowly rotating stars, and they find no correlation between rotation and magnetic field strength as expected for a magnetic dynamo in hot stars. However, there is no evidence that widespread magnetic braking contributes to the loss of angular momentum in B-type stars. The rotational velocity of B stars has been observed to decline with age \\citep{huang2006}, but no more than expected for their increase in radius during the MS lifetime and angular momentum losses due to stellar winds. The open cluster NGC 3766 is between 14.5 and 25 Myr in age (WEBDA; \\citealt{moitinho1997}; \\citealt{tadross2001}), with a reddening $E(B-V) = 0.22 \\pm 0.03$ and at a distance of 1.9 to 2.3 kpc \\citep{mcswain2008}. In our spectroscopic study of this cluster \\citep{mcswain2008}, we measured projected rotational velocities, $V \\sin i$, for 37 members. We also measured the effective temperatures, $T_{\\rm eff}$, and polar surface gravities, $\\log g_{\\rm polar}$, for 42 cluster members. More recently, we have analyzed spectra of additional cluster members using the same technique, and these will be presented in a forthcoming work (McSwain et al., in prep). In these works, we have identified several B-type members with unusually high or low helium abundances that are excellent candidates for the presence of magnetic fields. Since NGC 3766 offers a number of likely magnetic field candidates as well as a chance to study the rotational and evolutionary dependence of hot star magnetic fields, it is an excellent target to search for magnetic B stars. For these reasons, we have conducted a spectropolarimetric survey of 15 B-type stars in NGC 3766. In this work, we present the definite detection of two magnetic stars in the cluster, including one strong field of almost 2 kG. We also identify two stars with marginal magnetic field detections that are worth further investigation. The observations and magnetic field measurements are discussed in Section 2. In Section 3, we discuss the possibility that the stellar magnetic field strength is correlated with rotation and/or MS evolution. We also present a method to test whether the observed magnetic field distribution is consistent with a fossil field origin. Our conclusions are summarized in Section 4. ", "conclusions": "We have detected 2 definite and 2 possible magnetic B-type stars in our survey of 15 members of the open cluster NGC 3766. Both of our definite $B_\\ell$ detections are found in stars with abnormal He abundances, and we find a marginal $B_\\ell$ signature in two other cluster members that should be confirmed by higher resolution spectropolarimetric observations. Our observations provide only a single snapshot of $B_\\ell$ among the 15 stars in our survey. Further time-resolved observations of these B-type stars are necessary to determine the intensity and topology of the fields and investigate the spatial distribution of orientation angles throughout the region. We do not observe any trends with the detection of $B_\\ell$ with $V \\sin i$ or $\\log g_{\\rm polar}$, suggesting neither rotation nor evolution along the MS are major factors influencing the presence of magnetic fields. If the magnetic fields are produced by a magnetic dynamo, a correlation between $B_\\ell$ and $V \\sin i$ would be expected. To examine the possibility that the detected fields have a fossil origin, we used the Oblique Dipole Rotator model to simulate populations having a uniform or slightly varying magnetic flux on the ZAMS. Assuming conservation of the field strength with MS evolution, we compared our 4 definite and marginal detections to the modeled fossil field populations. None of our simple fossil models successfully reproduces both our observed range in $B_\\ell$ and the expected number of field detections. Therefore we tentatively rule out a purely fossil origin for the magnetic stars in NGC 3766. Our results are also consistent with the bimodal distribution of magnetic field strengths found by \\citet{auriere2007}, who proposed a critical magnetic threshold for the stability of large scale magnetic fields among early type stars. This or some other, unknown mechanism may contribute to the magnetic fields detected in NGC 3766." }, "0807/0807.1103_arXiv.txt": { "abstract": "AM-1, at $\\sim$120 kpc, and Pal~14, at $\\sim$70 kpc, are two of the most distant Galactic globular clusters known. We present {\\it Hubble Space Telescope} WFPC2 photometry of AM-1 and Pal~14 that reveals unprecedented depth and detail in the color-magnitude diagrams of these two clusters. Absolute and relative age measurements confirm that both are younger than the inner halo globular cluster M~3 by 1.5--2 Gyr assuming all three clusters have similar compositions. Thus AM-1 and Pal~14 join Pal~3, Pal~4, and Eridanus (studied by Stetson et al.) as distant Galactic globular clusters with red horizontal branches and young ages relative to the inner halo. Within the context of the entire body of research on the ages of second parameter globular clusters, the observed correlation between age and horizontal branch morphology suggests that age is the best candidate for the second parameter. However, this conclusion is tempered by the lack of precise chemical abundance determinations for a significant fraction of second parameter globular clusters. ", "introduction": "The second parameter phenomenon, a long-standing problem in Galactic astronomy, recognizes that the horizontal branch (HB) morphology of Galactic globular clusters (GCs) cannot be explained by the first parameter, metallicity, alone: at least one more factor, a `second parameter,' must be involved. In a landmark study, \\citet{sz78} demonstrated that GCs inside the solar circle (which they deemed the inner halo) exhibit a tight correlation between [Fe/H] and HB morphology while those that lie outside the solar circle (the outer halo) show no such well-defined trend. This in turn led \\citet{sz78} to theorize that the formation of the Galaxy was a prolonged, complex process. While the second parameter phenomenon has informed our understanding of galaxy formation, the identity of the second parameter remains a topic of some debate. A number of candidates have been suggested including age, helium abundance, CNO abundances, cluster central density, and stellar rotation (see \\citet{fp96} for a review). It is clear that its influence grows with distance from the Galactic center \\citep{sz78,zi80,ldz94}. To gain a better understanding of the second parameter it is logical to carefully examine the most distant GCs. Only six known GCs inhabit the Galaxy at distances greater than 50 kpc \\citep{ha96} \\footnote{References to the Harris catalog actually refer to the February 2003 revision that can be found at http://www.physics.mcmaster.ca/$\\sim$harris/mwgc.dat.} and five of the six are prime examples of second parameter GCs, i.e. metal poor GCs with red HBs. All six of the GCs with $\\rgc >$ 50 kpc: Pal~14, NGC~2419, Eridanus, Pal~3, Pal~4, and AM-1 (listed in order of increasing $\\rgc$) have been observed by the Hubble Space Telescope (HST) using the Wide Field Planetary Camera 2 (WFPC2). \\citet{ha97} presented a HST/WFPC2 color-magnitude diagram (CMD) of NGC~2419---the only one with $\\rgc >$ 50 kpc and a blue HB---and concluded that it is roughly coeval with the old, metal poor cluster M~92. In contrast, \\citet[hereafter S99]{st99} presented HST/WFPC2 CMDs of Pal~3, Pal~4, and Eridanus and concluded that these clusters are younger than relatively nearby clusters of similar metallicities (M~3 and M~5) by $\\sim$1.5--2 Gyr assuming the inner and outer halo GCs have similar abundances. The two remaining GCs, AM-1 and Pal~14, are the subject of this paper. AM-1 was classified as a GC by \\citet{am79} though it had been observed previously and named ESO 201-10 by \\citet{ho75}. \\citet{am79} estimated the distance to AM-1 at 300 kpc and, although it has since been revised downward to $\\sim$120 kpc \\citep{aa84,or85,ma89,hi06}, it is still the most distant Galactic GC known. Age estimates for AM-1 have proved uncertain to date because none of the published photometry has reached the main sequence. Pal~14, at $\\rgc \\sim$70 kpc, was discovered by van den Bergh in 1958 and later classified as a GC by \\citet{ar60}; it is also known as AvdB. \\citet{sa97} published the deepest CMD of Pal~14 to date, reaching more than a magnitude fainter than the main sequence turn off (MSTO), and concluded that Pal~14 is 3--4 Gyr younger than inner halo GCs with similar metallicities. We describe the observations and data reduction in $\\S$2 and present the WFPC2 CMDs of AM-1 and Pal~14 in section $\\S$3. Available information regarding the reddening and metallicity of both clusters is discussed in $\\S$4 followed by relative and absolute age analyses in $\\S$5. The results are discussed and put into context in $\\S$6 and the paper concludes with a summary in $\\S$7. ", "conclusions": "Color magnitude diagrams of AM-1 and Pal~14 from HST/WFPC2 F555W--F814W photometry were presented. The CMDs reveal unprecedented depth, reaching $\\sim$4 mag below the MSTO in both clusters. AM-1 and Pal~14 have very similar CMD morphologies. AM-1 shows a more densely populated RGB and HB, albeit with very few bright red giants, and an impressive display of blue stragglers. Pal~14 has a sparsely populated RGB and HB which is likely due to its large size (compared to the area covered by WFPC2) and low density. Three different age measurement techniques reveal that AM-1 and Pal~14 are $\\sim$1.5--2 Gyr younger than M~3 assuming all three GCs have similar compositions. Isochrone fitting suggests that if AM-1 and Pal~14 have similar abundances, then Pal~14 is about 0.5 Gyr younger than AM-1. S99 found that Pal~3, Pal~4 and Eridanus are also $\\sim$1.5--2 Gyr younger than inner halo GCs of comparable metal abundance. Thus all five of the GCs with $\\rgc >$ 50 kpc and red HBs are 1.5--2 Gyr younger than GCs with similar metallicities in the inner halo (to the extent that assumptions regarding similar abundances are valid). The variation of the second parameter phenomenon with $\\rgc$ has been known for some time, but the identity of the second parameter itself remains a topic of active debate in the literature, even some 40 years after it was discovered. When combined with previously published results on the ages of clusters that exhibit the second parameter effect, the present study provides further evidence that age is a strong candidate for the second parameter. However, there are a significant number of globular clusters important to identifying the second parameter that have not been accurately age-dated. Of equal importance is the precise determination of abundances in these clusters. Improvements to both the total metal abundance and the abundance ratios in these clusters will lead to more appropriate comparisons with well-studied, nearby GCs and theoretical models and, ultimately, to a better understanding of the second parameter phenomenon and the formation history of the Galactic halo." }, "0807/0807.1335_arXiv.txt": { "abstract": "{We show that expanding or contracting Kasner universes are unstable due to the amplification of gravitational waves (GW). As an application of this general relativity effect, we consider a pre-inflationary anisotropic geometry characterized by a Kasner-like expansion, which is driven dynamically towards inflation by a scalar field. We investigate the evolution of linear metric fluctuations around this background, and calculate the amplification of the long-wavelength GW of a certain polarization during the anisotropic expansion (this effect is absent for another GW polarization, and for scalar fluctuations). These GW are superimposed to the usual tensor modes of quantum origin from inflation, and are potentially observable if the total number of inflationary e-folds exceeds the minimum required to homogenize the observable universe only by a small margin. Their contribution to the temperature anisotropy angular power spectrum decreases with the multipole $\\ell$ as $\\ell^{-p}$, where $p$ depends on the slope of the initial GW power-spectrum. Constraints on the long-wavelength GW can be translated into limits on the total duration of inflation and the initial GW amplitude. The instability of classical GW (and zero-vacuum fluctuations of gravitons) during Kasner-like expansion (or contraction) may have other interesting applications. In particular, if GW become non-linear, they can significantly alter the geometry before the onset of inflation.} ", "introduction": "The inflationary stage of the very early universe explains the dynamical origin of the observed isotropic and homogeneous FRW geometry. The patch of the FRW geometry covers the cosmological horizon and beyond if inflation lasted \\begin{equation}\\label{efold} N=62-\\ln \\left(\\frac{10^{16}GeV}{V^{1/4}} \\right)+\\Delta \\ , \\end{equation} e-folds or longer. Here $V$ is the potential energy of the inflation, and $\\Delta$ is a correction from the (p)reheating stage after inflation, which is not essential for our discussion. Chaotic inflationary models, associated with a large energy ($\\sim$ GUT scale) of $V^{1/4} \\sim 10^{16}$GeV, predict a very large number of inflationary e-folds, $N \\gg 62$. Long-lasting inflation erases all classical anisotropies and inhomogeneities of the pre-inflationary stage. However, scalar and tensor vacuum fluctuations during inflation lead to almost scale free post-inflationary scalar and tensor metric inhomogeneities around our smooth observable FRW patch. In particular, the amplitude of the gravitational waves generated from the vacuum fluctuations during inflation is proportional to $V^{1/2}$, $h_k \\simeq\\frac{H}{2\\pi M_p} \\sim \\frac{V^{1/2}}{M_p^2}$ (where $M_p$ is the reduced Planck mass). There are significant efforts to measure the $B$-mode of $\\Delta T/T$ polarizations, since this will provide a direct probe of the scale of inflation. The current $95 \\%$ C.L. limits on $r$ (ratio of the tensor to scalar amplitudes of cosmological fluctuations) $r \\lta 0.43$ (WMAP-only) and $ r \\lta 0.2$ (WMAP plus acoustic baryon oscillation, plus supernovae) \\cite{WMAP} shall be improved to $r \\lta 0.1$ by the Planck mission \\cite{planck}, to $r \\lta 0.05$ by the ${\\rm C}_\\ell$over \\cite{clover}, EBEX \\cite{ebex}, and Spider \\cite{spider} experiments (see \\cite{epic} for the study of a mission that can improve over these limits). While these limits imply a detection in the case of high energy inflation, a number of other inflationary models, including many of the string theory constructions have lower energy, and therefore lead to GW of much smaller amplitude, which are virtually unobservable through $B$ mode polarization \\footnote{Future gravitational waves astronomy may allow to probe $r$ up to the level $10^{-6}$ with BBO \\cite{BBO} or ultimate DECIGO \\cite{decigo} direct detection experiments.}. In anticipation of the null signal observation of the primordial GW from inflation, it is worth thinking about other implementations of this result for the theory of inflation, besides putting limits on the energy scale $V^{1/4}$. There are models of inflation (including many string theory inflationary models) where the total number of e-folds, $N$, does not exceed the minimum (\\ref{efold}) by a large number. If the extra number of e-folds $\\Delta N$ beyond (\\ref{efold}) is relatively small then pre-inflationary inhomogeneities of the geometry are not erased completely, and their residuals can be subject to observational constraints. In the context of this idea, in this paper we suggest an additional mechanism to have observable gravitational waves associated with inflation. These gravitational waves are very different from the GW generated from the vacuum fluctuations during inflation. Firstly, they are the residual tensor inhomogeneities from the pre-inflationary stage. Secondly, they can be of a classical, rather than quantum, origin. Thirdly, while their initial amplitude and spectrum are given by the initial conditions, they are significantly affected by the number of ``extra'' e-folds $\\Delta N$. Therefore, observational limits on gravity waves result in constraints on a combination of $\\Delta N$ and of the initial amplitude. The choice of the initial geometry of the universe before inflation is wide open. In principle, one may assume an arbitrary geometry with significant tensor inhomogeneities component, and much smaller scalar inhomogeneities. This choice is, however, very artificial. A much more comfortable choice of the pre-inflationary stage will be a generic anisotropic Kasner-like geometry with small inhomogeneities around it. The origin of the anisotropic universe with the scalar field can be treated with quantum cosmology, or can be embedded in the modern context of the tunneling in the string theory landscape. In fact, a Kasner-like (Bianchi I) space was a rather typical choice in previous papers on pre-inflationary geometry, see e.g. \\cite{pre}. Most of the works on an anisotropic pre-inflationary stage aimed to investigate how the initial anisotropy is diluted by the dynamics of the scalar field towards inflation \\cite{starob83}. The formalism of linear fluctuations about an anisotropic geometry driven by a scalar field toward inflation was constructed only recently \\cite{Gumrukcuoglu:2006xj,Pereira:2007yy,Gumrukcuoglu:2007bx,Pitrou:2008gk}. Besides the technical aspects of calculations of cosmological fluctuations, there is a substantial conceptual difference between computations in the standard inflationary setting and in the anisotropic case. For an isotropic space undergoing inflationary expansion, all the modes have an oscillator-like time-dependence at sufficiently early times, when their frequency coincides with their momentum. One can therefore use quantum initial conditions for these modes. This is no longer the case for an expansion starting from an initial Kasner singularity. In this case, a range of modes, which can potentially be observed today (if $\\Delta N$ is not too large), are not oscillating initially and therefore cannot be quantized on the initial time hyper-surface; as a consequence, there is an issue in providing the initial conditions for such modes. For this reason we will adopt another perspective, namely, we will consider generic small classical inhomogeneities around the homogeneous background, as an approximation to the more generic anisotropic and inhomogeneous cosmological solution. Equipped with this philosophy, we consider an anisotropic expanding universe filled up by the scalar field with a potential $V(\\phi)$ which is typical for the string theory inflation. We add generic linear metric fluctuations about this geometry. The evolution of these fluctuations is by itself an interesting academic subject. However, it acquires a special significance in the context of the GW signals from inflation, because of a new effect that we report here of amplification of long-wavelength GW modes during the Kasner expansion. This growth terminates when a mode enters the ``average'' Hubble radius (the average of that for all the three spatial directions), or, for larger wavelength modes, when the background geometry changes from anisotropic Kasner to isotropic inflationary expansion. We perform explicit computations in the case of an isotropy of two spatial directions. In this case the computation becomes much more transparent and explicitly $k$ dependent. Fluctuations for arbitrary $a ,\\, b ,\\, c$ were considered in the formalism of \\cite{Pereira:2007yy,Pitrou:2008gk}, where the $k$ dependence is not explicit. We verified that our results agree with \\cite{Pereira:2007yy,Pitrou:2008gk} in the axisymmetric $b=c$ limit. We find that only one of the two GW polarizations undergoes significant amplification. Therefore, even if we assume for simplicity equi-partition of the amplitudes of the three inhomogeneous physical modes of the system (the scalar and the two GW polarization) at the initial time, the final spectra that will be frozen at large scales in the inflationary regime will be very different from each other, in strong contrast to what is obtained in the standard inflationary computations. This result can have different consequences, that we explore in the present work. Suppose that the growing GW mode is still linear (but significantly exceeds other modes) when the space becomes isotropic. Then, we can have significant yet linear classical GW fluctuations at the beginning of inflation, say of amplitude $\\lta {\\cal O}(1)$. If the modes which correspond to the largest scales that we can presently observe left the horizon after the first $\\Delta N$ e-folds of inflation, their amplitude decreased by the factor $e^{-\\Delta N}$ in this period. If $\\Delta N$ is relatively small, say $\\sim {\\cal O}(10)$ the freeze out amplitude of these GW modes would be $\\sim {\\cal O}(10^{-6})$. The angular spectrum of these GW will rapidly decrease as the multipole number $\\ell$ grows, since smaller angular scales are affected by modes which spend more time inside the horizon during the inflationary stage. Suppose instead that the growing GW mode becomes non-linear before the onset of inflation. In this case the background geometry departs from the original onset. Besides the phenomenological signatures, it is interesting to study the origin of the amplification of the GW mode. It turns out that the effect of GW amplification is related to the anisotropic Kasner stage of expansion. Therefore we will separately study GW in the pure expanding Kasner cosmology. For completeness, we also include the study of GW in a contracting Kasner universe, which is especially interesting due to the universality of anisotropic Kasner approach to singularity. The plan of the paper is the following. In Section~\\ref{sec:background} we discuss the evolution of the anisotropic universe driven by the scalar field towards inflation. In Section~\\ref{sec-linear} we briefly review the formalism of the linear fluctuations in the case of a scalar field in an anisotropic geometry, paying particular attention to the GW modes. In Section~\\ref{sec:decoupled} we compute the amplification of one of the two GW modes that takes place at large scales in the anisotropic era. In Section~\\ref{sec:pair} we discuss instead the evolution of the other two physical modes of the system. In Section~\\ref{sec:kasner} we study the evolution of the perturbations in a pure Kasner expanding or contracting Universe. In Section~\\ref{sec:obs} we return to the cosmological set-up, and we compute the contribution of the GW polarization amplified during the anisotropic stage to the CMB temperature anisotropies. In particular, by requiring that the power in the quadrupole does not exceed the observed one, we set some limits on the initial amplitude of the perturbations vs. the duration of the inflationary stage. In Section~\\ref{sec:sum} we summarize the results and list some open questions following from the present study, which we plan to address in a future work. ", "conclusions": "\\label{sec:sum} We found a new effect of instability of the gravitational waves in an expanding and contracting Kasner geometries. We demonstrated the effect for a particular choice of the Kasner exponents $(-1/3, 2/3, 2/3)$, but we expect it is rather generic. This particular choice allows to simplify the description of GW in such a way that their wave equations is manifestly depending on the momenta. For the contracting Kasner geometry, we found that our unstable GW mode is identical to the unstable large-scale inhomogeneous mode first identified in \\cite{LK63}-\\cite{BKL82} (for arbitrary Kasner exponents ($p_1,p_2, p_3$)). Backreaction of this unstable mode is conjectured to alter the Kasner exponents, differently in different spatial patches, depending on the spatial profile of the initial GW. The Kasner geometry is a rather universal asymptotic solution for many interesting situations: it describes approach towards black holes or cosmological singularities (including higher dimensional and supersymmetric cases \\cite{Damour:2000hv}), and it describes generic anisotropic expansion from singularity prior to inflation. All of these situations are inevitably accompanied by quantum fluctuations of the gravitons), and, possibly, also by classical gravitational waves. There is a long list of questions arising in connection with a new effect of the gravity waves instability in anisotropic geometry. We have to investigate how the instability growth depends on the Kasner exponents $p_1,p_2, p_3$. It will be interesting to understand if the instability of the classical GW also results in the instability of the graviton zero vacuum fluctuations. It will also be worth to understand the impact of the GW instability on the structure of the singularity inside the black hole and cosmological singularity. It is also interesting to investigate the backreaction of the GW instability on the contracting and expanding anisotropic geometries depending on the initial GW profile. In this paper we considered the effect of GW instability in the context of the anisotropic pre-inflationary stage. The transition from Kasner expansion to inflation terminates the effect of GW instability but leaves classical GW signal as the initial conditions for inflation. If inflation does not last very long, then the residual GW can contribute to the CMB temperature anisotropies. Since GW polarizations and the scalar mode of cosmological fluctuations behave differently during anisotropic pre-inflation, we can consider only the leading contribution, namely, $H_\\times$ mode of the GW polarization. In this paper we calculated its contribution to the total $\\Delta T/T$ anisotropy angular power spectrum. The angular power spectrum of the signal decreases with $\\ell$ in a power-law manner and depends on the initial spectrum of the classical GW fluctuations, $C_\\ell \\sim 1/ \\ell^p$, where the exponent $p$ depends on the power-spectrum of the classical GW at the initial hypersurface. It is interesting to note that this result is qualitatively similar to the result of \\cite{linde08}, where the impact of GW from the our-universe-bubble nucleation was considered. The signal from $H_\\times$ is rather anisotropic, and there is an interesting question about the anisotropy of its multipole structure $\\langle a_{\\ell m} \\, a_{\\ell' m'}^* \\rangle$, which is intriguing in connection with an apparent alignment of the low multipoles of $\\Delta T/T$. While such an analysis is beyond the aims of the present work, we have estimated the initial conditions (initial amplitude of the GW, versus the duration of inflation) which can lead to potentially observable effects. We leave a more extended analysis to future investigation. \\vspace{1cm} {\\bf \\large Acknowledgements} \\bigskip \\noindent We thank J.R.~Bond, C.R.~Contaldi, T.~Damour, I.~Khalatnikov, A.~Linde, C.~Pitrou, M.~Sasaki, A.~Starobinsky, J.P.~Uzan and J.~Weinwright for useful discussions and correspondence. The work of A.E.G. and M.P. was partially supported by the DOE grant DE-FG02-94ER-40823. LK was supported by NSERC and CIFAR. \\appendix" }, "0807/0807.4456_arXiv.txt": { "abstract": "{}% {We identify and characterize low-mass stars in the $\\sim$\\,3\\,Myr old \\tr region by means of a deep \\chandra X-ray observation, and study their optical and near-IR properties. We compare X-ray activity of \\tr stars with known characteristics of Orion and Cygnus\\,OB2 stars.}{We analyzed a 88.4 ksec \\chandra\\,ACIS-I observation pointed at the center of \\tr. Because of diffuse X-ray emission, source detection was performed using the PWDetect code for two different energy ranges: 0.5-8.0\\,keV and 0.9-8.0\\,keV. Results were merged into a single final list. We positionally correlate X-ray sources with optical and 2MASS catalogues. Source events were extracted with the IDL-based routine ACIS-Extract. X-ray variability was characterized using the Kolmogorov-Smirnov test and spectra were fitted by using XSPEC. X-ray spectra of early-type, massive stars were analyzed individually.}{Our list of X-ray sources consists of 1035 entries, 660 of which have near-IR counterparts and are probably associated with \\tr members. From near-IR color-color and color-magnitudes diagrams we compute individual masses of stars and their \\Av values. The cluster median extinction is \\Av=3.6 mag, while OB-type stars appear less absorbed, having \\Av=2.0 mag. About 15\\% of the near-IR counterparts show disk-induced excesses. X-ray variability is found in 77 sources, and typical X-ray spectral parameters are N$_{\\rm H}\\sim$5.37$\\times$10$^{21}$ cm$^{-2}$ and kT$\\sim$1.95 keV, with 1$\\sigma$ dispersions of 0.45 dex and 0.8 keV, respectively. OB stars appear, softer with a median kT$\\sim$0.65 keV. The median X-ray luminosity is 6.3$\\times$10$^{30}$ \\ergs\\,, while variable sources show a larger median \\Lx value of 13$\\times$10$^{30}$ \\ergs. OB-stars have an even higher median \\Lx of 80$\\times$10$^{30}$ \\ergs, about 10 times that of the low-mass stars.}{The \\tr region has a very rich population of low-mass X-ray emitting stars. An important fraction of its circumstellar disks survive the intense radiation field of its massive stars. Stars with masses 1.5-2.5 M$_\\odot$ display X-ray activity similar to that of stars in Cyg\\,OB2 but much less intense than observed for Orion Nebula Cluster members.} ", "introduction": "\\label{intro} The Carina nebula region (NGC\\,3372) is one of the most massive star formation regions of the Galaxy. It is associated with a giant H{\\sc ii} region spanning about 4 deg$^2$ of the sky and being bisected by a prominent V-shaped dark gas and dusty lane. This prominent young structure is not as compact as some of the other young galactic clusters, but seemingly related to a spiral feature. In this direction, we are looking almost tangentially to the now recognized Carina-Sagittarius spiral arm, at the edge of a giant molecular cloud extending over about 130 pc and with a content in excess of 5$\\times$10$^5$ solar masses \\citep{1988ApJ...331..181G}. The concentration of massive stars (i.e. M$\\geq$20 M$_\\odot$) interacts with the parent giant molecular cloud of the region, leading to triggered star formation events on intermediate to lower masses \\cite[e.g.][]{2004MNRAS.351.1457S}.\\\\ This region harbors several open clusters and/or star concentrations (Trumpler 14, 15 and 16; Collinder 228 and 232; Bochum 10 and 11) containing more than 60 known O-type stars \\citep{1995RMxAC...2...57F}. Large cavities within the giant molecular cloud are supposed to be carved out by the Tr\\,14 and 16 open clusters, which contain most of massive stars of the region. In particular Tr\\,16 includes three rare main-sequence O3 stars, the Wolf-Rayet (WR) star HD93162 and the famous luminous blue variable (LBV) \\etacar. There is a historical controversy about the distance and age of Tr\\,14 and Tr\\,16 \\citep{1995RMxAC...2...51W}. For instance, from extensive spectroscopy and photometry \\cite{1993AJ....105..980M} find 3.2 kpc for both clusters. However, photometric studies are strongly affected by differential extinction in the region and peculiar reddening, and so the derived distance are different. An example is the \\cite{2004A&A...418..525C} work, who for different R=\\Av/E(B-V) values (3.48 and 4.16 for Tr\\,16 and Tr\\,14 regions), compute distances of 4.0 kpc and 2.5 kpc, respectively. A more reliable distance (2250$\\pm$180 pc), was derived from proper motion and Doppler velocities of the expanding \\etacar Homunculus using HST-STIS\\footnote{Data from Hubble Space Telescope (HST) with the Space Telescope Imaging Spectrograph (STIS)} observations \\citep{1997ARA&A..35....1D}. Recent work \\citep{2003MNRAS.339...44T} derives a common distance DM=12.14 (2.7 kpc) and an age between $\\sim$1\\,Myr and 3\\,Myr, for Tr14 and Tr16, respectively. For this study, we adopt for \\tr a distance of 2250 pc and an age of 3\\,Myr. This young age agrees with the \\cite{2000ApJ...532L.145S} results, who report the existence of several embedded IR sources where star formation might be active. Also, \\cite{2001ApJ...549..578D} confirm clear evidence of pre-main sequence (PMS) stars in the region, while \\cite{2001MNRAS.327...46B} have identified two compact H{\\sc ii} regions possibly linked to very young O-type stars. Finally, \\cite{2004MNRAS.355.1237H} report a compact cluster of infrared PMS-stars in Tr\\,16. Of the existing methods to identify young stellar populations, the use of X-ray emission is perhaps the least biased \\citep{2002ApJ...574..258F}. While in main-sequence (MS) stars, from late A to M dwarfs, X-rays are believed to originate from the hot coronal gas that is heated by stellar dynamo magnetic fields \\citep{1987ApJ...315..687M}, for late type Pre-MS stars (T-Tauri stars (TTSs)) X-ray emission is attributed to solar-like coronal activity but elevated by a factor of 10$^3$-10$^4$ \\citep{1999ARA&A..37..363F}. Several authors suggested the possibility of detecting early pre-main sequence (PMS) objects through their hard X-ray emission escaping the highly obscured regions (see \\cite{1992AJ....104..758W, 1997PASJ...49..461K, 1997ApJ...486L..39H, 2002ApJ...579L..95H}). Recently, X--ray surveys have been successful in identifying the young and pre-MS population in star-forming regions, including: $i-$ deeply embedded Class I young stellar objects (YSOs), $ii-$ low-mass T-Tauri PMS stars, $iii-$ intermediate-mass Herbig Ae/Be PMS stars, $iv-$ zero-age MS stars. Moreover, X-ray emission from low-mass pre-MS stars usually exhibits a strong variability that helps to confirm membership. On last decade, X-ray observations of young stars on star-forming regions were intensified thanks to the high spatial resolution and the improved broad-band ([0.2-12.0] and [0.5-10.0] keV) effective area of the \\xmm\\, and \\chandra\\, satellites. A first X-ray survey in the Carina region by \\cite{2003MNRAS.346..704C} was performed on the basis of two early \\xmm\\, observations (rev\\,\\#115 and \\#116) centered on \\etacar. Because of the spatial resolution of the EPIC\\footnote{European Photon Image Camera has about six times less spatial resolution than \\chandra ACIS-I camera.} camera and relatively short exposure time of the observations ($\\sim$35 ksec), they detected only 80 X-ray sources, most of them related to the massive OB-type stars with \\Lx$\\sim$10$^{32}$-10$^{34}$ \\ergs. Before the observation used here, three \\chandra\\, observations were obtained on this region, two (obsId\\,50 and 1249) in the timed exposure mode, and the third (obsId\\,51) in the continuous clocking mode, which produces no image. Using only observation obsId\\,1249, \\cite{2003ApJ...589..509E} presented luminosities and hardness ratios of the hot stars in Tr\\,16, and part of Tr\\,14. Low-resolution X-ray spectra of luminous sources were discussed by \\cite{2004ApJ...612.1065E}. However the short exposure time of such observation ($\\sim$ 9.5 ksec) was a serious limitation for the study of intermediate- and low mass stellar population of the region. This limitation exists even if obsId. 50 and obsId. 1249 are combined \\citep{2007ApJ...656..462S}, reaching completeness just at X-ray luminosity (\\Lx) of $\\sim$7$\\times$10$^{31}$ \\ergs, i.e. the X-ray emission level typical of single O- and early B-type stars. In this paper we present results of the analysis of the deepest X-ray observation ever done in this region ($\\sim$90 ksec). Section \\ref{obs} gives details on the observation and data reduction procedures. Section\\,3 explains the method used to detect the sources, photon extraction and the construction of the catalog. In section\\,4 we present results of the cross-correlation with existing near-IR and optical catalogs of objects and their characterization based on their color-color (CC) and color-magnitude (CM) diagrams. Section\\,5 presents a statistical study of variability in the X-ray domain. Section\\,6 is dealing with results of the analysis of extracted X-ray spectra. In section\\,7 we discuss X-ray luminosities of stars and compare them statistically with the X-ray source population of ONC and Cygnus\\,OB2 star forming regions. In section\\,8 we discuss X-ray and stellar parameter of O- and early B- type stars. Finally, in section\\,9 we give a summary and draw conclusions of the paper. ", "conclusions": "\\label{concl} We report here results of a deep {\\em Chandra} X-ray observation pointed toward the $\\sim$3\\,Myr old star forming region \\tr. Source detection was performed using the PWDdetect code, identifying 1035 X-ray sources in the 17'$\\times$17' ACIS-I FOV. Most of these seem to be outside the obscured V-shaped region of dust and gas. Star formation in this part of the masked region has probably been disrupted and/or diminished as the stellar winds are blocked inside the cloud, due to the efficiency of photo-evaporation processes caused by interactions of the nearby hot massive stars with the dense dust and gas structures. Data extraction was performed using the semi-automated IDL-based {\\sc ACIS Extract} package, which is well suited to the analysis of observations of crowded fields such as ours. The X-ray source list was cross-identified with optical and near-IR (2MASS) catalogs: 28 X-ray sources (of 44 within the FOV) were identified with optically characterized OB members of \\tr and 760 with 2MASS sources. Among these latter sources almost all are believed to be \\tr members. About 90 X-ray sources without optical/NIR counterparts are estimated to be of extragalactic nature (AGNs), while the remaining X-ray sources with no counterpart are likely associated with members that are fainter than the 2MASS completeness limit. In order to characterize the previously unidentified likely cluster members with NIR counterparts, we placed them on NIR color-magnitude (K$_{\\rm s}$\\,vs.\\,H-K$_{\\rm s}$) and color-color (H-K$_{\\rm s}$\\,vs.\\,J-H) diagrams. A first estimate of interstellar extinction was obtained adopting a 3\\,Myr isochrone for the low- and intermediate-mass stars and assuming that O- and early B-type stars lie on the MS. We find a median visual absorption for OB stars of \\Av$\\sim$2.0 mag, while low mass likely members seem to be slightly more absorbed, \\Av$\\sim$3.6 mag. We also use the 3\\,Myr isochrone and the J magnitude to estimate masses of likely members assuming that they share the same distance and absorption. Our sample of X-ray selected members with near-IR counterparts reaches down to M=0.5-0.6\\,M$_\\odot$, and is likely complete down to $\\sim$1.5\\,M$_\\odot$. From the H-K$_{\\rm s}$\\,vs.\\,J-H diagram we estimate that $\\sim$15\\% (51/339) of low-mass stars have NIR excesses, finding it to be quite a high percentage with respect to the 2 Myr old Cyg\\,OB2. We believe that the disk fraction in young SFRs is more dependent on spatial morphology of gas and dust around massive stars, which may enhance photo-evaporation, and thus shorten disk lifetimes, than on the total number of massive stars in the region. At least 77 sources, i.e. $\\sim$7.4\\%, were found to be variable within our observation with a confidence level greater than 99.9\\%. Only three of the 28 detected O- and early B-type stars were detected as variable during our 90-ksec observation, in spite of the high statistics of the OB stars' light curves. These exceptions are the known binary O3.5V+O8V star HD93205 (our source \\#242), showing a rather linear decay of the count rate during the observation plus a short-term variability, the B1.5\\,V star Tr16-11 (source \\#136) and the B1V star Tr16-5 (source\\#489). The latter two show a flare-like variability probably related to unresolved low-mass companions. We modeled the ACIS X-ray spectra of sources with more than 20 photons and where f$_{\\rm cont}\\,<$1. We assumed an absorbed single-component thermal emission model. The median log(N$_{\\rm H}$) of the sources is 21.73 (cm$^{-2}$). This value agrees well with the median \\Av computed from the near-IR diagram. The median kT of low-mass stars is 2.6 keV. Sources associated with O- and early B-type stars are instead quite soft (median kT: 0.60 keV). Absorption corrected X-ray luminosities of OB stars were calculated from the best-fit spectral models. O and B-type stars are the most luminous, with L$_{\\rm x}$= 2.5$\\times10^{30}$-6.3$\\times10^{33}$ erg\\,s$^{-1}$. Their X-ray and bolometric luminosities are in rough agreement with the relation L$_{\\rm x}$/L$_{\\rm bol}=10^{-7}$, albeit with an order of magnitude dispersion. Low mass stars have L$_{\\rm x}$ ranging between $10^{30}$ and $10^{31}$ \\ergs (median \\Lx = $2.8\\times10^{30}$). Variable low mass stars are on average 0.5 dex brighter (log(\\Lx)$\\sim$31.0 \\ergs). These X-ray luminosities are consistent with those of similar mass stars in the slightly younger (2\\,Myr) Cyg~OB2 region. However, in the mass range 1.5-2.5 M$_\\odot$, the ONC\\,(1~Myr) shows higher X-ray activity level than observed in \\tr stars in the same mass range. We believe that the age-\\Lx activity connection is an acceptable explanation of this result." }, "0807/0807.2647_arXiv.txt": { "abstract": "We report the results of the first self-consistent three-dimensional adaptive mesh refinement magnetohydrodynamical simulations of Population III star formation including the Biermann Battery effect. We find that the Population III stars formed including this effect are both qualitatively and quantitatively similar to those from hydrodynamics-only (non-MHD) cosmological simulations. We observe peak magnetic fields of $\\simeq 10^{-9}$~G in the center of our star-forming halo at $z\\simeq 17.55$. The magnetic fields created by the Biermann Battery effect are predominantly formed early in the evolution of the primordial halo at low density and large spatial scales, and then grow through compression and by shear flows. The fields seen in this calculation are never large enough to be dynamically important (with $\\beta \\geq 10^{15}$ at all times), and should be considered the minimum possible fields in existence during Population III star formation, and may be seed fields for the stellar dynamo or the magnetorotational instability at higher densities and smaller spatial scales. ", "introduction": "The nature of the first generation of stars, as well as their influence on later structure formation, is a fundamental problem in cosmology. A great deal of theoretical progress has been made (see recent reviews by \\citet{Bromm04},~\\citet{2005SSRv..117..445G}, and~\\citet{Ciardi05}. In the past decade, cosmological hydrodynamic simulations of Population III star formation have achieved great success, and significantly different numerical methods have produced results that agree quite well \\citep{Abel02, 2002ApJ...564...23B,Yoshida03, O'Shea07}. These calculations have given a reasonably clear picture of the formation process of Population III stars, and have provided some constraints on many of their important properties. These calculations, while useful, largely ignore an important issue: the relevance of magnetic fields in Population III star formation. Magnetic fields are widely observed in our galaxy, in other galaxies, and in galaxy clusters, and the origin of these fields is one of the most fundamental and challenging problems in astrophysics \\citep{Carilli02, Widrow02}. One possibility is that magnetic fields are created and amplified in the first generation of stars and are spread throughout the IGM when these stars explode, providing seed fields for later generations of stars and for further amplification by dynamo effects. If some seed magnetic fields exist before Population III stars form, they may help to remove angular momentum from the star-forming clouds, significantly changing the ultimate mass range of these stars \\citep{Pudritz89, Davies00}. Several groups have examined the importance of magnetic fields on the evolution of Population III protostellar disks using analytic or semi-analytic models. These include \\citet{silk06}, \\citet{tan04a} and \\citet{tan04b}, who model (among other aspects of primordial star formation and evolution) dynamos in primordial accretion disks. Other authors, including \\citet{flower03} and \\citet{maki07}, use one-zone calculations to examine the collapse of the primordial star-forming cloud and the assumptions of flux-freezing. While useful, these models do not self-consistently include the effects of both magnetic fields and cosmological structure formation. In this Letter, we report the results of the first magnetohydrodynamic simulations of Population III star formation including the Biermann Battery effect \\citep{Biermann50} within the context of cosmological structure formation. We describe our numerical methods and the simulation used in Section~\\ref{sec:methods}, present our key results in Section~\\ref{sec:results}, and discuss and summarize our results in Section~\\ref{sec:discussion}. ", "conclusions": "\\label{sec:discussion} In this Letter, we have shown results from the first cosmological magnetohydrodynamical simulation of Population III star formation including the Biermann Battery effect. This effect is one of the most robust methods of generating magnetic fields in the Universe \\citep{Biermann50}, and thus provides useful constraints on the minimum magnetic field expected in situations where Population III star formation will take place. We find from our simulation that small magnetic fields are primarily generated via the Biermann Battery at relatively low overdensities, and are then amplified to values of nearly $10^{-9}$~G at the center of the cosmological halo via gravitational collapse. While significant, this magnetic field is still quite small -- the plasma $\\beta$ is never smaller than $10^{15}$ at any point during the simulation. This suggests strongly that the magnetic fields do not play a significant dynamical role up to n$_B \\sim 10^{10}$~cm$^{-3}$, when the simulation is terminated. As a result, the evolution of the primordial gas within the cosmological halo is very close to the hydrodynamic results obtained by \\citet{O'Shea07}. At later times, however, the magnetic fields may be amplified to dynamically relevant values via a process such as the magnetorotational instability or a dynamo \\citep[][]{silk06,tan04a,tan04b}. The simulation of further evolution of our evolving protostellar core is currently beyond the capability of our numerical tools. However, there are multiple avenues that we can follow to more thoroughly explore the relevance of magnetic fields in Population III star formation. We can begin our calculation with seed fields having strengths based on limits from observation and theory~\\citep{Widrow02}. We can examine the effects of magnetic fields at higher densities by implementing more chemical and cooling processes \\citep[as discussed by ][]{turk08}. Both of these projects are underway, and we will report on the results in an upcoming paper. To summarize, we have performed the first calculations that incorporate the Biermann Battery in cosmological magnetohydrodynamic simulations of Population III star formation. Our key results are as follows: 1. From an initial state with no magnetic fields, a combination of the Biermann Battery and compressional amplification can result in fields with strengths of $|B| \\simeq 10^{-9}$~G at n$_B \\simeq 10^{10}$~cm$^{-3}$ at the center of a cosmological halo where a Population III star will form. 2. The Biermann Battery creates fields predominantly at low density (n$_B \\leq 10$~cm$^{-3}$) and large spatial scales in Population III star-forming halos. 3. The magnetic fields created by the Biermann Battery are dynamically unimportant at all densities below n$_B \\simeq 10^{10}$~cm$^{-3}$ -- $\\beta \\equiv P_{th} / P_{B} \\geq 10^{15}$ at all times during the evolution of the halo." }, "0807/0807.0532_arXiv.txt": { "abstract": "The present work aims at performing a comprehensive census and characterisation of the pre-main sequence (PMS) population in the cometary cloud L1615/L1616, in order to assess the significance of the triggered star formation scenario and investigate the impact of massive stars on its star formation history and mass spectrum. Our study is based on $UBVR_CI_C$ and $JHKs$ photometry, as well as optical multi-object spectroscopy. We performed a physical parametrisation of the young stellar population in L1615/L1616. We identified 25 new T~Tauri stars mainly projected on the dense head of the cometary cloud, almost doubling the current number of known members. We studied the spatial distribution of the cloud members as a function of the age and H$\\alpha$ emission. The star formation efficiency in the cloud is $\\sim$~7--8\\,\\%, as expected for molecular clouds in the vicinity of OB associations. The slope of the initial mass function (IMF), in the mass range $0.1{\\le}M{\\le}5.5$~$M_{\\odot}$, is consistent with that of other T and OB associations, providing further support of an universal IMF down to the hydrogen burning limit, regardless of environmental conditions. The cometary appearance, as well as the high star formation efficiency, can be explained in terms of triggered star formation induced by the strong UV radiation from OB stars or supernovae shockwaves. The age spread as well as both the spatial and age distribution of the PMS objects provide strong evidence of sequential, multiple events and possibly still ongoing star formation activity in the cloud. ", "introduction": "\\label{sec:Intro} The Lynds clouds L1615 and L1616 \\citep{Lynds62} are both located at an angular distance of about $6\\degr$ West of the Orion OB1 associations. These clouds actually form a cometary-shaped single cloud with a ``head-tail'' distribution subtending about $40\\arcmin$ \\citep[5.2~pc assuming a distance of 450~pc;][]{Alcala04} roughly in the East-West direction. The dense head of the cloud complex (i.e. L1616 only), pointing toward East and facing the bright Orion-belt stars, harbours the IRAS Small Scale Structure X0504-034 \\citep{Helou88,Ramesh95} and a bright reflection nebula, NGC\\,1788 \\citep[= DG\\,51, Ced\\,40, vdB\\,33, RNO\\,35, LBN\\,916;][]{Lynds65}, which is illuminated by a small cluster of stars \\citep{Stanke02}. The two intermediate-mass stars HD\\,293815 and Kiso\\,A-0974~15 are the brightest visible members of this cluster. L1615/L1616, like many other cometary clouds off the main Orion star forming regions, clearly shows evidence of ongoing star formation activity which might have been triggered by the strong impact of the UV radiation from the massive, luminous stars in the Orion complex \\citep{Maddalena86,Stanke02, Alcala04,Kun04,Lee05,Lee07}. The illumination of dense clumps in molecular clouds by OB stars could be responsible for their collapse and subsequent star formation. The UV radiation from the OB stars may sweep the molecular material of the cloud into a cometary shape with a dense core located at the head of the cometary cloud. The radiation and wind of OB stars may also have an important impact on the mass accretion during the star formation process. While in a T~association a protostar may accumulate a significant fraction of mass, the mass accretion of a low-mass protostar in a region exposed to the wind of OB stars can be terminated earlier because of the photo-evaporation of the circumstellar matter \\citep{Kroupa01,Kroupa02}. Therefore, many low-mass protostars may not complete their accretion and hence can result as brown dwarfs (BDs). This mechanism might affect the low-mass end of the initial mass function (IMF). Recent studies have provided some information about the shape of the IMF in the very low-mass and sub-stellar regimes. While the IMF in the Orion Nebula Cluster appears to rise below 0.1 M$_{\\odot}$ \\citep{Hillenbrand00}, in T~association like Taurus-Auriga and Cha\\,II there is some indication of a deficit of sub-stellar objects \\citep{Luhman00,Briceno02,Spezzi08}. Other studies of the young cluster IC\\,348, which is devoid of very massive stars, have also revealed a deficit of BDs relative to the Orion Nebula Cluster \\citep{Preibisch03,Muench03,Lada06}. Because of its vicinity to the Orion OB associations the L1615/L1616 cometary cloud constitutes an ideal laboratory to investigate the triggered star formation scenario. The most recent census of the pre-main sequence (PMS) stars in L1616 was provided by \\citet{Alcala04}, who presented a multi-wavelength study of the region, from X-ray to near-infrared wavelengths. They found 22 new low-mass PMS stars distributed mainly to the East of L1616, in about 1~square degree field. By adding the 22~new PMS stars to the previously confirmed members of the cloud \\citep{Cohen79,Sterzik95,Nakano95,Stanke02} and counting the millimeter radio-source, i.e. the Class-0 protostar MMS1A found by \\citet{Stanke02}, \\citet{Alcala04} ended up with a sample of 33~young stellar objects associated with L1616. However, the latter work could not investigate important aspects of the star formation in this cloud, like the IMF and the impact of environmental conditions on the mass spectrum, because their sample was rather incomplete. Therefore, the aim of the present study is to perform a comprehensive census and characterisation of the PMS population in L1615/L1616, in order to investigate the star formation history, the relevance of the triggered scenario, and to study the IMF. To this aim, we report both optical and near-infrared observations, as well as multi-object optical spectroscopy in L1615/L1616. An important goal of our study is to compare the mass function in L1615/L1616 with that of other T and OB~associations. The outline of the paper is as follows: we describe our observations and data reduction in Section~\\ref{sec:PhotObs} and \\ref{sec:Spec-Obs}; the results are reported in Section~\\ref{sec:PMS-Ident}, while the data analysis and the physical proprieties of the new PMS stars are presented in Section~\\ref{sec:Data analysis}. Our discussion and conclusions are developed in Section~\\ref{sec:Discussion}. Some details on the spectral-type classification as well as reddening, radius, and luminosity determination are given in the Appendix~\\ref{App-A} and \\ref{App-B}, respectively. ", "conclusions": "\\label{sec:Discussion} Once the physical parameters of the PMS objects are known, we can study the star formation in L1615/L1616. In the next sub-sections we discuss the history, rate and efficiency of star formation, and some issues related to the mass-spectrum. \\subsection{The star formation history} \\label{sec:Star-Form-Hist} \\clearpage \\begin{figure*} \\centering \\resizebox{\\hsize}{!}{\\includegraphics[draft=false]{f4.eps}}% \\caption{IRAS 100$\\mu$m dust emission map covering a sky-area of $5^\\circ\\times5^\\circ$ around L1615/L1616 (left panel). The Classical and Weak T Tauri stars are represented with filled dots and squares, respectively. The upper arrow shows the direction to the Orion OB1 associations, located at about 7.5$^\\circ$ ($\\sim60$ pc at a distance of 450 pc) to the North-East of L1615/L1616. The lower arrow points toward the Orion A Giant Molecular Cloud, also located at about 7.5$^\\circ$ to the South-East of L1615/L1616. The dashed square defines the area surveyed with WFI. This region is zoomed in the right panel, where the two intermediate-mass members HD\\,293815 and Kiso\\,A-0974~15 are marked with five-pointed star symbols.} \\label{Fig:IRAS-Image} \\end{figure*} \\clearpage Based on the most complete census of the L1615/L1616's population performed in this work, we further investigated the triggered star formation scenario suggested by \\citet{Ramesh95}, \\citet{Stanke02}, and \\citet{Alcala04}. In left panel of Figure~\\ref{Fig:IRAS-Image} the IRAS~100~$\\mu$m dust emission map of a $5^\\circ\\times5^\\circ$ sky-area around L1615/L1616 is shown. About $64$\\,\\% of the CTTSs are clustered, in the densest part of L1615/L1616, i.e.~within the boundaries of the NGC\\,1788 reflection nebula and to the West of the bright rim of the cloud (Figure~\\ref{Fig:IRAS-Image}, right panel). Such rim is located to the East of NGC\\,1788, at $\\sim6.5{\\arcmin}$ from the head of the cometary cloud (about 0.85 pc at a distance of 450 pc) and it is directly exposed to the UV radiation from the Orion OB stars. On the other hand, the WTTSs are more scattered (only $\\sim22$\\,\\% is projected on NGC\\,1788) and mainly occupy the side of the cloud facing the OB1 associations and to the East of the bright rim. By applying the Kolmogorov-Smirnov test to the WTTS and CTTS age distributions we found that the two populations are similar at a confidence level of $\\sim 80$~\\%. Thus, it seems there are no age differences between CTTSs and WTTSs in our PMS sample. This result is consistent with previous PMS populations studies \\citep[see][and reference therein]{Feigelson99}; CTTSs are predicted to have ages between about 0.5 and 3 Myr, although some stars retain CTTS characteristics even at ages as old as 20~Myr. On the other hand, many WTTSs occupy the same region on the H-R diagram as CTTSs do, whereas some of them are approaching the zero-age main sequence. The age distribution of the L1615/L1616 population peaks between 1 and 3 Myr, depending on the adopted evolutionary tracks (Figure~\\ref{Fig:Hist_Ages_Masses}, right panels). This is in agreement with the findings by \\citet{Alcala04}, but the age spread found by us is significantly higher than what these authors claim, exceeding the value expected on the basis of the uncertainties on luminosity and temperature. The L1615/L1616 members span a wide range in age, from less than 0.1~Myr up to about 30~Myr. This might suggest multiple events of star formation in the cloud, which would further support the hypothesis of triggered star formation. In order to investigate a possible age difference between ``on-cloud'' and ``off-cloud'' PMS objects, we divided the sample in two groups, fixing as dividing line the bright rim of the cloud. Twenty of 56 objects in our sample are located within the boundaries of NGC\\,1788, i.e. to the West of the bright rim, while 36 are off-cloud members. The age distributions of the two groups are shown in Figure~\\ref{Compare_Age}. The on-cloud PMS stars are statistically younger than those located to the East of the bright rim, regardless of the adopted evolutionary tracks. Applying a Kolmogorov-Smirnov test we found that the probability that the two sets are sub-sample of the same statistical population is very low; in particular, we found confidence levels of 1.29, 0.18, and 0.19\\,\\% when using the PMS evolutionary tracks by \\citet{Baraffe98} \\& \\citet{Chabrier00}, \\citet{Dantona97}, and \\citet{Palla99}, respectively. We thus concluded that there is a clear age difference between the two groups of stars. The above findings further support the scenario of triggered star formation in L1615/L1616, as proposed by \\citet{Stanke02}. In this context, the spatial dispersion and older age of the ``off-cloud'' members can be explained as a consequence of the ``rocket acceleration'' effect. As recently pointed out by \\citet{Kun04}, this acceleration continues after the onset of star formation and the parental cloud is further accelerated with respect to the newly formed objects. As a consequence, the cloud is soon swept off the newly formed stars. This hypothesis, together with the rapid dispersion typical of small clouds, causes the spatial displacement of the oldest cloud members. The conspicuous number of PMS stars found apparently isolated from classical star forming regions \\citep[e.g.][]{Alcala95,Covino97,Guillout98a,Guillout98b,Frasca03, Zickgraf05} might be a consequence of this mechanism as well. \\clearpage \\begin{figure} \\centering \\resizebox{\\hsize}{!}{\\includegraphics[draft=false]{f5.eps}} \\caption{Mass (left panel) and age (right panel) distributions of the L1615/L1616 PMS population derived by using the evolutionary tracks by \\citet{Baraffe98} \\& \\citet{Chabrier00} (upper panels), \\citet{Dantona97} (central panels) and \\citet{Palla99} (lower panels).} \\label{Fig:Hist_Ages_Masses} \\end{figure} \\clearpage \\thispagestyle{empty} \\setlength{\\voffset}{-15mm} \\begin{figure} \\centering \\resizebox{\\hsize}{!}{\\includegraphics[draft=false]{f6.eps}} \\caption{Age distributions of the on-cloud and off-cloud PMS stars in L1615/L1616 derived by using the evolutionary tracks by \\citet{Baraffe98} \\& \\citet{Chabrier00} (upper panel), \\citet{Dantona97} (central panel) and \\citet{Palla99} (lower panel).} \\label{Compare_Age} \\end{figure} \\clearpage \\setlength{\\voffset}{0mm} \\subsection{The star formation efficiency} \\label{sec:SFE} Based on $^{12}$CO and $^{13}$CO column density maps, \\citet{Ramesh95} estimated that the mass of L1616 alone is in the range $169-193$~$M_{\\sun}$, depending on the used tracer. \\citet{Yonekura99} mapped both L1615 and L1616 in the CO J=1-0 transition line. They inferred that the mass of the cloud system as a whole is $\\sim 530$~$M_{\\sun}$, $\\sim 350$~$M_{\\sun}$ and $\\sim 440 ~M_{\\sun}$ based on $^{12}$CO, $^{13}$CO and $^{18}$CO observations respectively. They also derived the mass of the $^{13}$CO and $^{18}$CO cores (i.e. L1616 only) finding a value of $\\sim 146$~$M_{\\sun}$ and $\\sim 161$~$M_{\\sun}$, in good agreement with the \\citet{Ramesh95}'s determination. \\citet{Alcala04} estimated the SFE in L1616 to be $\\sim14$\\,\\%, i.e. higher than the average value measured in other low-mass star forming region ($\\la3$\\,\\%). The SFE reported by \\citet{Alcala04} for L1616 is based on the cloud mass reported by \\citet{Ramesh95} and the total stellar mass of the 32 members of the cloud investigated by the authors (i.e. $\\sim30~M_{\\odot}$). Since the star formation history of L1615 and L1616 is intimately connected, we have re-calculated the SFE considering the system L1615 plus L1616 as a whole. By using the NICER color excess method by \\citet{Lombardi01}, we mapped the dust extinction of the cloud complex and derived its mass. In the NICER technique the $J-H$ and $H-Ks$ colours of the field stars are compared to the colours of stars in a nearby reference field. The colour-excesses of the field stars are then combined and transformed to the visual extinction $A_{\\mathrm V}$, fixing the form of the reddening curve. The normal interstellar extinction law derived by \\citep{Cardelli89} has been adopted by us for this purpose. We retrieved the 2MASS colours of all the point-like sources located in a $50{\\arcmin}\\times50{\\arcmin}$ region centered on NGC\\,1788 and in a reference field close to the L1615/L1616 cloud\\footnote{The SOFI survey is only centered on the densest part of the cloud and can not be used to map the cloud as a whole.}. The latter field was selected by using the nearby low-intensity regions of the IRAS~100$\\mu$m dust emission map, according to prescription by \\citet{Kainulainen06}. Finally, the visual extinction map resulting from the NICER method was converted to cloud mass by assuming the standard gas-to-dust ratio \\citep{Bohlin78}. We found a value of $\\sim550$~$M_{\\odot}$, in good agreement with the one inferred by \\citet{Yonekura99} on the basis of $^{12}$CO observations. Based on this value and on the total mass of the 56 present-day known members of the complex (i.e. 41--46~$M_{\\odot}$, depending on the adopted evolutionary tracks), we derived a SFE in L1615/L1616 of 7--8\\,\\%, i.e. significantly lower than the previous estimate by \\citet{Alcala04}, but still in good agreement with the one generally found in giant molecular clouds hosting OB associations (5--10\\,\\%) and predicted by theoretical calculations on the formation of OB associations \\citep[see][and reference therein]{Clark05}. \\subsection{Density of star formation and star formation rate} \\label{sec:SFR} An interesting question is whether L1615/L1615 can be considered as a cluster. According to the definition suggested by \\citet{Lada03}, a cluster is a group of some 35 members with a total mass density larger than 1.0~M$_{\\odot}$\\,pc$^{-3}$. To estimate the density of PMS objects in L1615/L1615, we considered the region spectroscopically surveyed by us (Figure~\\ref{Fig:SpecSurvey}). This region covers an area of approximately 0.25 square degrees and includes about 40 PMS objects. Assuming a distance of 450~pc, the resulting area is about 4~pc$^2$, which means a surface density of about 10-11 PMS objects per pc$^2$ and a volume density on the order of 7 PMS objects per pc$^3$. In the latter calculation we estimated the volume as $V = 0.752{\\times}Area^{1.5}$ \\citep[see][]{Jorgensen07}, assuming a locally spherical distribution of sources. The average mass of the 40 PMS objects, as determined from the results in Section~\\ref{sec:Masses-Ages}, is on the order of 0.7~M$_{\\odot}$, which implies a volume density of about 4-5~M$_{\\odot}$\\,pc$^{-3}$. Therefore, the group of 40 PMS objects confined in the spectroscopically surveyed area can be considered as a cluster according to the criterion by \\citet{Lada03}. Now, it is interesting to estimate the rate at which the stars in this small cluster are formed. According to the results of Section~\\ref{sec:Masses-Ages}, we estimated that the total mass in PMS objects in that region is on the order of 41--46~M$_\\odot$. Therefore, considering the average age of 2~Myr for these objects (see Figure~\\ref{Fig:Hist_Ages_Masses}), we found that the cometary cloud is turning some 20-23~$M_{\\odot}$ into PMS objects every Myr, which is lower than the star formation rate in other clusters like those in Serpens \\citep{Harvey07}, but higher than in other T~associations like Chamaeleon~II \\citep{Alcala07} and Lupus \\citep{Merin07}. Thus we concluded that L1615/L1616 is a small cluster with a moderate star formation rate. \\subsection{The Initial Mass Function} Though based on a low-number statistics, the first clue on the Initial Mass Function in L1615/L1616 was given by \\citet{Alcala04}. Based on this study, the IMF in this region appears roughly consistent with that of the solar neighbourhood in the mass range $0.31.7$, where the use of the isochrones in the Cousins system would produced many spurious member candidates. This consideration, together with the contamination by background/foreground stars in the field taken into account by the authors, could explain why \\citet{Alcala04} overestimated the number of BD candidates. The use of the adequate isochrones for the WFI-Cousins system would produce much less BD candidates, in agreement with the number of PMS objects with mass close to the Hydrogen burning limit confirmed by our spectroscopic follow-up (see Table~\\ref{Tab:mass_age}). Though the $R_{SS}$ estimated by us for L1615/L1616 needs further refinements, the region seems to be poor of BDs, with a $R_{SS}$ ratio close to that observed in T~associations or, perhaps, even lower. Though in L1615/L1616 the radiation from the massive stars in the Orion OB1 associations is likely responsible for the star formation (Section~\\ref{sec:Star-Form-Hist}), this primordial process does not seems to favour the formation of BDs. As reported in \\citet{Alcala04}, the radial velocity (RV) distribution of the stars in L1615/L1616 shows a well defined peak at $22.3$~Km/s with the standard deviation of $4.6$~Km/s which is consistent with the average RV error (i.e. $\\sim5$~Km/s). This may indicate that the velocity dispersion of the stars must be less than $5$~Km/s. Considering a velocity dispersion of a few Km/s, $2$~Myr old objects would disperse over a distance of about $2$~pc, which is the approximate projected size of the head of the cometary cloud and has been completely covered by our photometric and spectroscopic observations. Thus, the results of our study in L1615/L1616 does not play in favour of dynamical ejection or photoevaporation by ionising radiation from massive stars being triggering factors of the BDs formation mechanism." }, "0807/0807.1702_arXiv.txt": { "abstract": "A joint spectral analysis of some \\chandra\\ ACIS X-ray data and Molonglo Observatory Synthesis Telescope radio data was performed for 13 small regions along the bright northeastern rim of the supernova remnant \\snr. These data were fitted with a synchrotron radiation model. The nonthermal electron spectrum used to compute the photon emission spectra is the traditional exponentially cut off power law, with one notable difference: The power-law index is not a constant. It is a linear function of the logarithm of the momentum. This functional form enables us to show, for the first time, that the synchrotron spectrum of \\snr\\ seems to flatten with increasing energy. The effective power-law index of the electron spectrum is 2.2 at 1~GeV (\\ie, radio synchrotron--emitting momenta) and 2.0 at about 10~TeV (\\ie, X-ray synchrotron--emitting momenta). This amount of change in the index is qualitatively consistent with theoretical models of the amount of curvature in the proton spectrum of the remnant. The evidence of spectral curvature implies that cosmic rays are dynamically important instead of being ``test'' particles. The spectral analysis also provides a means of determining the critical frequency of the synchrotron spectrum associated with the highest-energy electrons. The critical frequency seems to vary along the northeastern rim, with a maximum value of $1.1^{+1.0}_{-0.5} \\times 10^{17}$~Hz. This value implies that the electron diffusion coefficient can be no larger than a factor of $\\sim$4.5--21 times the Bohm diffusion coefficient if the velocity of the forward shock is in the range 2300--5000~km~s$^{-1}$. Since the coefficient is close to the Bohm limit, electrons are accelerated nearly as fast as possible in the regions where the critical frequency is about $10^{17}$~Hz. ", "introduction": "\\label{int} Several observational and theoretical clues support the suggestion that Galactic cosmic rays, up to an energy of 100~TeV \\citep{lag83} or more \\citep{jok87,vol88,bel01}, are accelerated predominantly in the shocks of supernova remnants. If Galactic supernovae occur at an average rate of one event every 30 years, then an average supernova remnant must transfer about 10\\% \\citep{dru89} of the initial $10^{51}$~ergs of kinetic energy of the ejecta to cosmic rays. In this case, the cosmic-ray energy density may be large enough to affect the structure of the shock (\\ie, cosmic rays may not be mere ``test'' particles). Three consequences of a large cosmic-ray pressure are potentially observable. One consequence is that the ambient material is slowed before it crosses the subshock. This effect reduces the temperature of the shocked gas \\citep{che83,ell00,dec00}. An upper limit on the temperature is provided by the well-known relation between the shock speed $v_{s}$ and the postshock temperature of a gas whose ratio of specific heats $\\gamma = \\frac{5}{3}$: \\begin{equation} kT_{i} \\le \\frac{3}{16} m_{i} v_{s}^2, \\label{eqn1} \\end{equation} where $k$ is Boltzmann's constant and $m_{i}$ and $T_{i}$ are the mass and immediate postshock temperature, respectively, of particle species $i$. Since kinetic energy is transferred to cosmic rays at the expense of the thermalization of the shocked gas, the equality is appropriate only in the limit that the cosmic-ray energy density is negligible. \\cite{hug00} analyzed the transverse motion of the X-ray--emitting material (mostly reverse-shocked oxygen and neon) in the supernova remnant 1E~0102.2$-$7219 and inferred a forward-shock velocity $v_{s} = 6200^{+1500}_{-1600}$~km~s$^{-1}$. In this case, the temperature of the shocked protons must be lower than the right-hand side of equation~(\\ref{eqn1}) or else the fitted electron temperature will be too low to be explained by Coulomb heating \\citep{hug00}. However, the proton temperature is sensitive to the shock velocity ($T \\propto v_{s}^{2}$), and a variety of speeds are reported for 1E~0102.2$-$7219. \\cite{fla04} report that the \\chandra\\ High Energy Transmission Grating \\ion{Ne}{10} line emission data are consistent with a model that includes radial \\ion{Ne}{10} velocities up to $1800 \\pm 450$~km~s$^{-1}$. \\citet{fin06} analyzed several features in \\ion{O}{3} images. The mean transverse velocity of the features is $2000 \\pm 200$~km~s$^{-1}$. \\citet{eri01} report that the motion of some \\ion{O}{3}--emitting material is best described by a radial \\ion{O}{3} velocity of about 1800~km~s$^{-1}$. Since these velocity results differ and since none of the measurements provides a direct measure of the velocity of the forward shock, the claim that a significant fraction of the internal energy in 1E~0102.2$-$7219 has been transferred to cosmic rays requires additional support. A second consequence is that the total compression ratio is larger than 4 \\citep{ell91,ber99}. Since cosmic rays slow the upstream material before it crosses the subshock, the velocity of the shocked material relative to the subshock is reduced. As a result, the contact discontinuity is closer to the subshock than it would be in the absence of a large cosmic-ray pressure. \\citet{war05} report that in Tycho's supernova remnant the mean ratio of the radius of the contact discontinuity to the radius of the forward shock is $0.93 \\pm 0.02$. This separation corresponds to a total compression ratio $r = 5.1_{-1.0}^{+1.9}$. Similarly, \\citet{cas08a} report that for the southeastern rim of \\snr\\ the ratio varies from a value of $0.91^{+0.03}_{-0.02}$ (\\ie, $r = 4.0^{+2.2}_{-0.5}$) between the bright X-ray synchrotron--emitting filaments to a value of 1 (\\ie, $r = \\infty$) along the filaments. The mean value between the filaments is $0.96 \\pm 0.03$ (\\ie, $r = 9.0^{+25}_{-3.6}$). These authors explore ways in which the location of the contact discontinuity can approach the location of the forward shock, but they cannot explain why these two features curiously appear to be coincident with one another along the filaments. \\citet{kse05a}, who also studied \\snr, infer a mean compression ratio $r = 5.2$. A third consequence of a large cosmic-ray pressure is that the shock transition region is broadened or ``smeared out'' \\citep{ell91,ber99}. Only the subshock has a short transition length. In this case, low-energy cosmic rays, which have relatively small diffusion lengths, experience only a portion of the velocity gradient as they scatter back and forth across the subshock. Higher energy particles, which have relatively large diffusion lengths, experience a larger portion (or all) of the velocity jump. Since the rate of energy gain increases as the velocity difference increases, higher energy particles gain energy faster than lower energy particles. As a result, cosmic-ray spectra do not have power-law distributions. The spectra flatten with increasing energy \\citep{bel87,ell91,ber99}. \\cite{jon03} and \\cite{vin06} report evidence of curvature in the synchrotron spectra of Cas~A and RCW~86, respectively. One possible explanation for such curvature is that the cosmic-ray electrons producing the synchrotron emission have curved spectra. Here we describe the results of an analysis of some radio and X-ray synchrotron data, which suggest that the synchrotron spectrum of \\snr\\ might be curved. The data, assumptions, and analysis techniques are described in \\S\\ \\ref{dat}. The results of the analysis are discussed in \\S\\ \\ref{dis}. Our conclusions are summarized in \\S\\ \\ref{con}. ", "conclusions": "\\label{con} We have performed a joint spectral analysis of some {\\sl Chandra} ACIS X-ray data and MOST radio data for 13 small regions along the bright northeastern rim of the supernova remnant \\snr. The data were fitted with a model that includes a synchrotron emission component. This component is based on an electron spectrum that has a momentum-dependent spectral index. The rate of change in the index for each decade in momentum is a free parameter of the fit. If the assumptions described in \\S\\ \\ref{curv} are valid, then the results of the spectral analysis, which show that the synchrotron spectra of \\snr\\ are curved, can be interpreted as evidence of curvature in the GeV-to-TeV electron spectra. The mean amount of curvature in the electron spectra is qualitatively consistent with predictions of the amount of curvature in the proton spectrum of \\snr\\ \\citep{ell00}. The best-fit power-law index at 1~GeV (\\ie, at radio synchrotron--emitting momenta) is $2.221^{+0.013}_{-0.012}$. Including the effect of curvature, the effective spectral index at about 10~TeV (\\ie, at X-ray synchrotron--emitting momenta) is $2.005 \\pm 0.027$ (90\\% confidence level uncertainties). This effective index is consistent with the predictions of \\citet{ber02}. The evidence of curved electron spectra suggests that cosmic rays are not ``test'' particles. The cosmic-ray pressure at the shock is large enough to modify the structure of the shock. Since nonthermal electrons contain only about 0.1\\% (\\ie, $10^{48}$~ergs) or less of the total internal energy, the results provide indirect evidence of a much more energetic population of cosmic-ray protons. Collectively, the evidence of (1) spectral curvature in Cas~A \\citep{jon03}, RCW~86 \\citep{vin06}, and \\snr, (2) an unusually low electron temperature in 1E~0102.2$-$7219 \\citep{hug00}, and (3) a compression ratio greater than 4 in Tycho \\citep{war05} and \\snr\\ \\citep{cas08a} suggests that efficient particle acceleration may be a common feature of young, shell-type supernova remnants. The results of the spectral analysis also determine the ``cutoff critical frequency'' $\\nu_{m}$. This frequency seems to vary from one region to another (see also Rothenflug \\etal\\ 2004), which implies that the exponential cutoff momentum of the electron spectrum and/or the strength of the magnetic field varies. It is not possible to identify the cause of the variation using the synchrotron spectral data alone. As described in the Appendix, the cutoff frequency can be used to set an upper limit on the mean diffusion coefficient $\\bar{\\kappa}$ of the highest-energy electrons. Aside from the cutoff frequency, this limit depends (strongly) on the velocity of the forward shock $u_{1}$ and (weakly) on the compression ratio $r$ and the ratio of the upstream to downstream magnetic field strengths $B_{1} / B_{2}$. If $\\nu_{m} = 1.1 \\times 10^{17}$~Hz, $u_{1} = 2300$--5000~km~s$^{-1}$, $r = 4$, and $B_{1} / B_{2} \\approx 0$, then $\\bar{\\kappa} < (4.5$--$21) \\bar{\\kappa}_{\\rm B}$, where $\\bar{\\kappa}_{\\rm B}$ is the mean Bohm diffusion coefficient. This result implies that at least some of the highest-energy electrons in \\snr\\ diffuse close to the Bohm limit (\\ie, are accelerated about as fast as possible), which provides additional support for the idea that Galactic cosmic rays are predominantly accelerated by the shocks of supernova remnants." }, "0807/0807.1969_arXiv.txt": { "abstract": " ", "introduction": "Since the first map of supernova remnant (SNR) RX J1713.7-3946 in very high-energy gamma-rays obtained by H.E.S.S. (Aharonian et al. \\cite{RX1713-2004nature}), there are few more SNRs with spatial distributions of VHE $\\gamma$-ray emission reported (see TeV Gamma-ray Source Catalog\\footnote{Mori M., TeV Gamma-ray Source Catalog [available at \\url{http://www.icrr.u-tokyo.ac.jp/~morim/TeV-catalog/}]} and reviews Rowell et al. \\cite{Rowell-2005}, Funk \\cite{Funk-2008}). Surface brightness distribution of $\\gamma$-ray emission of astrophysical objects is an important possibility to test models of kinetics of astroparticles as well as dynamics of magnetic field and turbulence in astrophysical plasma. Thus, an actual and important task is the modeling of the respective surface brightness distribution. Inverse Compton (IC) electron-photon interactions is one of the most important processes of gamma-ray production in SNRs. Application of the exact formalism of IC emission to the numerical calculation of SNR images requires quite powerfull CPU resources. In such a situation, an accurate approximation may be of interest because it considerably reduces CPU time and also results in some new formulae in classical analysis (Sect. \\ref{ICdelta} and \\ref{ICThomson}). A common approach is to deal with IC emissivity for a given energy of the initial (monochromatic) photons (e.g. Jones \\cite{jones-68}, Blumenthal \\& Gould \\cite{Blum-Gould-70}) with assumption of a given shape of electron spectrum (power-law as a common choice). The resulting IC photon spectrum is then given by the integration over the energy distribution of the field photons. Such an approach is essential for special cases of the field photon energy distributions. It is known however, that the black-body radiation field may be used for IC emission in many astrophysical objects. In particular, for SNRs under typical conditions, one may consider just black-body photons (either with single or with few different temperatures representing CMB/IR/optical radiation). In SNRs with no clear IR emission assosiated, the contribution from CMB photons dominates the role of infrared and optical photons (see discussion in Appendix in Lazendic et al. \\cite{Lazendic-et-al-04}). The IR/optical photon fields may contribute typically 10\\%-15\\% of the IC flux in such SNRs (Gaisser et al. \\cite{Gaisser-et-al-98}, Baring et al. \\cite{Baring-et-al-99}). In this note, we present an approximation for IC emissivity which may be applied to IC emission originating from the black-body photon field with some temperature $T$. Since $T$ is a parameter in our approach, the approximation may be used for calculation of IC radiation from different photon fields (CMB, IR, optical). The target radiation field around some SNRs (e.g. around Galatic center) may not be black-body and/or the contribution from IR/optical photons may dominate over CMB there (Porter et al. \\cite{Porter-et-al-06}, Hinton \\& Aharonian \\cite{Hinton-Aharonian-07}). In cases when different components of the target radiation field may be approximated by a superposition of multiple Planck distributions with different $T$, our approximation may be used in a similar fashion. Namely, the overall IC emission will be the weigthed sum of single approximations, each with different value of the temperature. In cases when the initial radiation field may not be approximately described by a sum of black-body distributions, our approximation is not applicable. Another assumption in the present paper is the isotropy of the electron and photon fields. A thorough treatment of anisotropic IC scattering from cosmic-ray electrons is done by Moskalenko \\& Strong (\\cite{Moskalenko-Strong-00}). Our approximation is given in terms of an energy of incident electrons rather than in commonly used terms of the field photon energy. Such an approach opens the possibility for accurate modeling of IC emission of electrons with energy spectra being different from power-law. For example, if we are interested in electron energies around maximum possible values. It is known that contribution from electrons accelerated by the shock to $E\\rs{max}\\sim 30-300$ TeV is important in interpretation of the H.E.S.S. observations of shell SNRs. ", "conclusions": "Numerical evaluation of the spatial distribution of the IC emission in SNRs requires essential computational resources because the IC volume emissivity (\\ref{PICdef}) consists in two enclosed integrations (over initial photon and electron energies). We developed the approximation (\\ref{calIappranyeta}) of the spectral distribution of the IC emission power $p(E\\rs{\\gamma})$ of electrons with Lorentz factor $\\gamma$ which are interacting with the isotropical black-body photon field. Namely, Eq.~(\\ref{calIappranyeta}) restores known results (Blumenthal \\& Gould \\cite{Blum-Gould-70}) with high enough accuracy in any regime, from Thomson to extreme Klein-Nishina limits. It may be used for different astrophysical objects. For $kTE\\rs{\\gamma}\\lesssim 100\\left(m\\rs{e}c^2\\right)^{2}$, i.e. in case of IC $\\gamma$-ray emission from electrons accelerated by the forward shock of SNRs, it is suitable to use a bit simple approximation (\\ref{calIappr}). In the Thomson limit, our approach results in a simple expression (\\ref{TomshICemis}). Our approximation may be used in situation when the initial radiation field may be approximated with the Planck function with some temperature $T$ or when it may be represented by a superposition of the black-body distributions with different $T$. In addition, it assumes isotropy of electron and photon fields. The approximation is given in terms of an energy of incident electrons rather than in commonly used terms of the field photon energy (there are known approximation for the latter approach, see e.g. Schlickeiser \\cite{Schlick-book}). Therefore, our approximation may be useful for analysis of the role of the electron spectrum with shapes different from power-law. The main idea behind our approach is the possibility to split initial integral into two parts which, contrary to the original integral, may be scaled. This scaling is the reason of accuracy of the approximation over the wide range of parameters, from Thomson to extreme Klein-Nishina regime. There is well known `monochromatic approximation' where electron is scattered by the monochromatic photons with energy $\\epsilon\\rs{o}$ (e.g. Schlickeiser \\cite{Schlick-book}). Fig.~\\ref{fig-d} shows that the spectral distribution of the radiation power of `single' electron with Lorentz factor $\\gamma$ scattered by photons distributed with the Planck function is peaked at some energy $E\\rs{\\gamma m}(\\gamma)$. This allows us to introduce -- similarly to the case of synchrotron emission -- the `delta-function' approximation for IC emission. In this approximation, Eq.~(\\ref{deltaapprox}), all radiated energy of electron is assumed to be at $E\\rs{\\gamma m}(\\gamma)$. In the Thomson limit, $E\\rs{\\gamma m}(\\gamma)\\approx 4\\epsilon\\rs{c}\\gamma^2$ where $\\epsilon\\rs{c}=kT$. In the classical `monochromatic approximation', the average $\\left\\langle E\\rs{\\gamma}\\right\\rangle=(4/3)\\epsilon\\rs{o}\\gamma^2$ is used as an estimator for the energy of emitted IC photons. Our approach results in some new expressions which represent known results. Namely, Eq.~(\\ref{TomshICemis}) yeilds the spectral distribution of IC radiation power of a \"single\" electron and Eq.~(\\ref{Thomsemiss}) represents the spectrum of IC emission from the power-law spectrum of electrons, in the Thomson limit. These expression account for integration over all possible energies of the seed black-body photons. They can be derived thanks to the possibility of analytical integration of ${\\cal I}$ in the Thomson regime, Eq.~(\\ref{app3})." }, "0807/0807.2067_arXiv.txt": { "abstract": "{ $\\theta$~Carinae belongs to a group of peculiar early-type stars (OBN) with enhanced nitrogen and carbon deficiency. It is also known as a binary system, but it is not clear yet whether the chemical anomalies can be explained by mass transfer between the two components. On the basis of the previously reported spectral variability of a few metal lines it may be expected that \\tcar{} possesses a weak magnetic field. } {A study of the physical nature of this hot massive binary which is furthermore a well-known blue straggler lying $\\sim$2\\,mag above the turnoff of the young open cluster IC~2602 is important to understand the origin of its strong chemical anomalies.} { We acquired high resolution spectroscopic and low resolution spectropolarimetric observations to achieve the following goals: a) to improve the orbital parameters to allow a more in-depth discussion on the possibility of mass transfer in the binary system, b) to carry out a non-local thermodynamic equilibrium (NLTE) abundance analysis, and c) to search for the presence of a magnetic field. } { The study of the radial velocities using CORALIE spectra allowed us to significantly improve the orbital parameters. A comparative NLTE abundance analysis was undertaken for \\tcar{} and two other early B-type stars with recently detected magnetic fields, \\tsco{} and \\xcma{}. The analysis revealed significantly different abundance patterns: a one-order-of-magnitude nitrogen overabundance and carbon depletion was found in \\tcar{}, while the oxygen abundance is roughly solar. For the stars \\xcma{} and \\tsco{} the carbon abundance is solar and, while an N excess is also detected, it is of much smaller amplitude (0.4--0.6\\,dex). Such an N overabundance is typical of the values already found for other slowly-rotating (magnetic) B-type dwarfs. For \\tcar{}, we attribute instead the chemical peculiarities to a past episode of mass transfer between the two binary components. The results of the search for a magnetic field using FORS\\,1 at the VLT consisting of 26 measurements over a time span of $\\sim$1.2\\,h are rather inconclusive: only few measurements have a significance level of 3$\\sigma$. Although we detect a periodicity of the order of $\\sim$8.8\\,min in the dataset involving the measurements on all hydrogen Balmer lines with the exception of the H$\\alpha$ and H$\\beta$ lines, these results have to be confirmed by additional time-resolved magnetic field observations. } {} ", "introduction": "\\label{sect:intro} It has long been assumed that massive stars do not have magnetic fields, as they lack the convective outer mantle prevalent in lower mass stars. However, indirect evidence is supporting that magnetic fields are indeed present in massive O and early B-type stars (e.g., Henrichs et al.\\ \\cite{henrichs}; Rauw et al.\\ \\cite{rauw}; Cohen et al.\\ \\cite{cohen}; Gagn{\\'e} et al.\\ \\cite{gagne}). Yet, only very few direct magnetic field detections have been reported in O-type stars so far (Donati et al.\\ \\cite{donati06a}; Wade et al.\\ \\cite{wade06}; Hubrig et al.\\ \\cite{hubrig07}). Among the hottest B-type stars, a magnetic field has been discovered in the B0.2V star \\tsco{} (Donati et al.\\ \\cite{donati06b}) and in the B0.7IV star \\xcma{}, which is one of the hottest $\\beta$\\,Cephei stars, with a rather large longitudinal magnetic field of up to 300\\,G (Hubrig et al.\\ \\cite{hubrig06}). Walborn (\\cite{walborn06}) listed the hot B0.2V star $\\theta$~Carinae (HD\\,93030, HIP\\,52419, HR\\,4199; $m_V$ = 2.78) among a few other massive stars with unexplained spectral peculiarities or variations for which a magnetic field could be expected. It has a peculiar, variable spectrum in both optical and UV. The spectral peculiarities in the blue-violet are an enhancement of nitrogen and deficiency of carbon, but also definite line-intensity variations as well as other line asymmetries (Walborn \\cite{walborn76,walborn79}). The non-detection of a magnetic field was reported by Borra \\& Landstreet (\\cite{borra_land}), who used a photoelectric Balmer-line magnetograph to measure circular polarization in the wings of the H$\\beta$ line. Although \\tcar{} is a very bright target, easily observable with spectropolarimeters, no other magnetic field measurements have been reported so far in the literature, mainly due to the unavailability of instruments equipped with polarization analyzing optics on telescopes located in the southern hemisphere. \\tcar{} is an SB1 system with one of the shortest orbital periods known among massive stars ($P$ = 2.2\\,d; Lloyd et al.\\ \\cite{lloyd}). A discussion of the possibility of mass transfer in the binary system, which would be a natural explanation for the remarkable spectral peculiarities and for the singular location of this object in the H-R diagram of the 30\\,Myr old open cluster IC\\,2602, was presented by Walborn (\\cite{walborn76}). However, the spectral type of the companion remains unknown. Previously determined orbital elements were rather uncertain (Lloyd et al.\\ \\cite{lloyd}) and other periods of $\\sim$1\\,day and $\\sim$25~days have been suggested due to a systematic difference in the radial velocities from different observers. Randich et al.\\ (\\cite{randich}) reported in their ROSAT/PSPC study of the cluster IC\\,2602 that \\tcar{} is the brightest X-ray object among the studied cluster members, with a soft X-ray luminosity amounting to up to $\\log L_{\\rm X}$ = 30.99\\,ergs\\,s$^{-1}$. { Recently, Naz\\'e \\& Rauw (submitted), using XMM-{\\it Newton} observations, showed that the X-ray flux of \\tcar{} is slightly lower than the flux typically observed in O and early B-type stars and confirmed the unusual softness of the X-ray emission. Further, they noted that X-ray lines appear narrow and unshifted, reminiscent of those of $\\beta$\\,Cru and the magnetic star \\tsco{} (Donati et al.\\ \\cite{donati06b}). } Below, we present the results of our new spectroscopic and spectropolarimetric study of this peculiar massive star and discuss possible origins of its anomalies. ", "conclusions": "\\label{sect:disc} \\subsection{Binarity and evolutionary history} \\begin{figure} \\centering \\includegraphics[height=0.48\\textwidth,angle=270]{9972f8.eps} \\caption{Color-Magnitude diagram of IC\\,2602. Small crosses mark the position of main sequence stars with $T_{\\rm eff}$ = 30\\,000--32\\,000\\,K and $\\log g$ = 4.0--4.4, according to theoretical models. Solid lines represent isochrones for $\\log\\tau$ = 3.0, 7.5, and 8.0. The photometric position of \\tcar{} is marked with a filled square and other IC\\,2602 members with filled circles. Photometric data and cluster parameters [$(m-M)$ = 6.1, $E(B-V)$ = 0.024] were taken from the WEBDA database (Mermilliod \\cite{mermilliod}). } \\label{fig:cmd} \\end{figure} We have presented in Sect.~\\ref{sect:binarity} the first high quality radial velocity curve for \\tcar{}. It is well fitted by a Keplerian orbit with eccentricity $e$ = 0.127, leaving no doubt about the binary origin of these variations. The improved determination of the orbital parameters allows us to make a more sophisticated interpretation of the nature of this system and to review the possibility of a mass transfer between primary and secondary as the cause of the observed chemical anomalies. The physical parameters of \\tcar{} determined in Sect.~\\ref{sect:abundances} suggest that it is located in the H-R diagram close to the ZAMS. According to the Geneva stellar models (Schaller et al.\\ \\cite{schaller}; Lejeune \\& Schaerer \\cite{lejeune_schaerer}) for solar abundances, the observed values of $T_{\\rm eff}$ = 31\\,000\\,K and $\\log g$ = 4.2 correspond to a star with mass $M_1$ = 15.25\\,M$_{\\sun}$, radius $R_1$ = 5.1\\,R$_{\\sun}$, and age $\\log\\tau$ = 6.0. Considering a 1$\\sigma$ uncertainty in $\\log g$ and $T_{\\rm eff}$, we obtain $\\log\\tau \\le 6.7$. This age value is considerably lower than the accepted age for the cluster IC\\,2602 ($\\log\\tau = 7.83$, Kharchenko et al.\\ \\cite{kharchenko}; $\\log\\tau = 7.63$, Eggen \\cite{eggen}; $\\log\\tau = 7.32$, Whiteoak \\cite{whiteoak}; $\\log\\tau = 7.16$, Hill \\& Perry \\cite{hill_perry}), consistent with the classification of this star as a blue straggler. We note that the absolute magnitude and intrinsic colors interpolated in the same theoretical grid are in agreement with the position of \\tcar{} in the Color--Magnitude diagram of the cluster IC\\,2602 (see Fig.~\\ref{fig:cmd}). \\begin{figure} \\centering \\includegraphics[width=0.35\\textwidth]{9972f9.eps} \\caption{Physical parameters as a function of the orbital inclination: orbital semiaxis $a$, critical radius for the primary $R_1^{\\rm crit}$ and the secondary $R_2^{\\rm crit}$, primary radius derived from the rotational velocity $R_1^{\\rm vsini}$, and mass of the secondary star $M_2$. The full lines correspond to $M_1$ = 15.25\\,M$_{\\sun}$ and the dotted lines to $M_1$ = 13.9\\,M$_{\\sun}$ and $M_1$ = 18.2\\,M$_{\\sun}$.} \\label{fig:param} \\end{figure} The uncertainty in the stellar mass and radius interpolated in the stellar model grid was estimated from the adopted errors for $T_{\\rm eff}$ and $\\log g$. The primary mass ranges from 13.9\\,M$_{\\sun}$ (corresponding to a ZAMS star with $T_{\\rm eff} = 30\\,000$\\,K) to 18.2\\,M$_{\\sun}$ (corresponding to $T_{\\rm eff} = 32\\,000$\\,K and $\\log g = 4.0$). From the estimated primary mass, relevant information for the system can be derived from the radial velocity curve, even when the orbital inclination is in principle unknown. We show in Fig.~\\ref{fig:param} the absolute value of the orbital semiaxis and the mass of the secondary star as a function of the orbital inclination. For each parameter plotted in this figure the solid line corresponds to the solution with $M_1 = 15.25$\\,M$_{\\sun}$, while solutions with 13.9\\,M$_{\\sun}$ and 18.2\\,M$_{\\sun}$ are plotted with dotted lines. To evaluate the occurrence of a mass exchange in the system, we calculated the volume radius of the Roche lobe at periastron ($R_1^{\\rm crit}$ and $R_2^{\\rm crit}$), which are plotted in Fig.~\\ref{fig:param}. From this figure it is clear that these parameters are not strongly dependent on the orbital inclination unless the latter is low. However, a low inclination can be ruled out for the following reasons. On the one hand, as argued by Walborn (\\cite{walborn79}), the absence of a pronounced emission at H$\\alpha$ indicates that the equatorial velocity is lower than 300\\,km\\,s$^{-1}$ and consequently, from the observed $v \\sin i$ value we obtain $i \\geq 22^\\circ$. This value is presented by the solid vertical line in Fig.~\\ref{fig:param} and corresponds to an upper limit of about 2.0\\,M$_{\\sun}$ for the secondary component. On the other hand, if we assume that the rotation is synchronized with the orbital motion at periastron, the relatively small radius deduced from the spectroscopic value of $\\log g$ is compatible with the observed $v \\sin i$ only for high inclinations. For the case of synchronous rotation we derive a rotational period of 1.692$\\pm$0.007\\,d and hence $R_1\\sin i$ = 3.78$\\pm$0.27\\,R$_{\\sun}$. The adopted range for $M_1$ (13.9--18.2\\,M$_{\\sun}$) corresponds to $R_1$ = 4.7--7.1\\,R$_{\\sun}$, and therefore we obtain $i = 32^\\circ - 54^\\circ$. These limits are marked with two dashed vertical lines in Fig.~\\ref{fig:param}. Interestingly, this rotational period of 1.692$\\pm$0.007\\,d is in good agreement with the 1.779\\,d period discovered by Walborn (\\cite{walborn79}) who studied intensity variations of \\ion{Si}{IV} $\\lambda$4089. The radius derived from $v \\sin i$ assuming synchronization would coincide with the critical radius for $i \\approx 25^\\circ$. In this case $R_1 \\approx 9.1\\,$R$_{\\sun}$ and $\\log g \\approx 3.7$, which is lower than the spectroscopic value by 2.5$\\sigma$. Therefore, we expect that, at present, the primary star is probably not filling its Roche lobe, although it is close to that point. A star of 15.25\\,M$_{\\sun}$ (13.9--18.5\\,M$_{\\sun}$) with a present radius of 5.1\\,R$_{\\sun}$ (4.7--7.1\\,R$_{\\sun}$) has an evolutionary age of 1.0\\,Myr (0--4.3\\,Myr) and, evolving as a single star, would have at the Terminal Age Main Sequence (TAMS) a radius of about 13.6\\,R$_{\\sun}$ (12.6--16.3\\,R$_{\\sun}$), which is significantly larger than the critical radius. Thus, we can expect mass-transfer to occur from the primary to the secondary during the main-sequence stage, at an evolutionary age of about 11\\,Myr. Since \\tcar{} is a bona-fide blue straggler of IC\\,2602, it is very likely that the system has suffered in the past a mass-transfer in the opposite direction. Eggen \\& Iben (\\cite{eggen_iben}) have discussed extensively this possibility as an explanation for its abnormal position in the Color-Magnitude diagram. The apparent evolutionary youth of \\tcar{} and its peculiar chemical abundances are best understood if the present primary star was originally the less-massive component and has accreted mass transfered from the companion. Its original mass could be less than half of the total mass of the system, similar to other stars near the turnoff point in the cluster. After considerable mass-transfer, it would appear as a non-evolved massive star above the cluster turnoff and near the ZAMS. As was noted by Eggen \\& Iben (\\cite{eggen_iben}), during the evolution of this binary system an important loss of angular momentum has taken place, making it possible to achieve a system with such a short period and low mass-ratio. The orbital angular momentum of a binary system can be written as $$ J_{\\rm orb} = \\frac{2 \\pi}{P} M a^2 \\frac{q}{(1+q)^2} \\sqrt{1-e^2} $$ while for the rotational momentum of each star we have $$ J_{rot} = \\frac{2 \\pi}{P_{rot}} \\beta^2 M R^2 $$ where $\\beta$ is the radius of gyration, which is typically 0.25 for a main-sequence star (Claret \\& Gim\\'enez \\cite{claret_gimenez}). In any case the rotational contribution to the total angular momentum is small. We can use these expressions to calculate the angular momentum of the system and compare the present value with that at the moment when the mass-ratio was close to unity. As illustration, let us assume that the original primary star (presently the secondary) with 9\\,$M_{\\sun}$ filled its Roche lobe when it was still in the main-sequence with a radius of the order of 7--9\\,$R_{\\sun}$. Therefore, at that time the separation between the stars was about 20--25\\,$R_{\\sun}$. Assuming a conservative mass transfer and adopting for the present configuration $M_1$ = 15\\,$M_{\\sun}$, $M_2$ = 1.0\\,$M_{\\sun}$, and a = 19\\,$R_{\\sun}$, we can estimate that the present angular momentum is about 0.22--0.25 of the original value. In conclusion, most of the angular momentum has been lost. \\subsection{Abundances and magnetic field} The binary and blue straggler nature of \\tcar{} leads us to relate the observed chemical anomalies to the past history of the binary system. First, mass transfer of chemically processed material from the initially more massive component has dramatically altered the CNO surface abundances. Second, the spun up phase that ensued may have further enhanced nitrogen and depleted carbon in \\tcar{} as a result of rotational mixing (see e.g.\\ Langer et al.\\ \\cite{langer}). The tendency of magnetic B stars to show a higher incidence of an N excess compared to stars with no field detection (Morel et al.\\ \\cite{morel08}) also indicates that magnetic phenomena could play a role. The results of the previous studies, along with the fact that little is known about the magnetic properties of hot, massive stars, has motivated our search for a magnetic field in \\tcar{}. There are only six massive stars with detected magnetic fields that are hotter than \\tcar{}: $\\theta^1$\\,Ori~C, HD\\,191612, HD\\,155806, HD\\,148937, 9\\,Sgr, and \\tsco{}. The star HD\\,191612 with a rotation period of 538\\,d was observed with ESPaDOnS only during four consecutive nights (Donati et al.\\ \\cite{donati06a}). Among the published magnetic field measurements of $\\theta^1$\\,Ori~C, only four measurements show significance at the 3$\\sigma$ level (Wade et al.\\ \\cite{wade06}). HD\\,155806 and HD\\,148937 were observed only once and 9\\,Sgr three times (Hubrig et al.\\ \\cite{hubrig07}; Hubrig et al., in preparation). The sixth star, \\tsco{}, with physical parameters very similar to those of \\tcar{}, was studied over the rotational period of 41\\,d. It possesses the weakest mean longitudinal magnetic field with a maximum value of 88\\,G and a very complicated geometry. The structure of its magnetic field topology features in particular a significantly warped torus of closed magnetic loops encircling the star and additional smaller networks of closed field lines (Donati et al.\\ \\cite{donati06b}). In our magnetic field study of \\tcar{}, only very few measurements were achieved at a significance level of 3$\\sigma$ and their sporadic appearance is difficult to explain in the framework of the presence of a global large-scale organized magnetic field. On the other hand, the presence of a complex structure of the magnetic field in \\tcar{} may hypothetically be possible in an analogous manner to the magnetic field topology of \\tsco{}. The detected periodicity of the order of 8.8\\,min in the dataset of measurements carried out on hydrogen Balmer lines with the exclusion of H$\\beta$ is surprising, and if it is not spurious, its discovery would give rise to the important question whether the presence of pulsations could cause such a periodicity. No studies of short-time pulsations exist for \\tcar{} so far. B0 main-sequence stars are expected to pulsate in low radial order p- and g-modes with periods of the order of hours (typically 3 to 8 hours). For such massive stars, periods of the order of a few minutes would correspond to high radial order p-modes, but non-adiabatic codes do not predict their excitation. However, since current models do not take into account the presence of a magnetic field, one cannot exclude the possibility that magnetic fields would favour the excitation of these types of modes in analogy to roAp stars which do possess a magnetic field and pulsate in high radial order modes of the order of a few minutes. Our data with the CORALIE spectrograph were taken typically every 7~minutes. This leads to a sampling rate which is too low to allow us to detect variations of spectral line profiles on short time scales. Since the mechanism of the generation and the maintenance of magnetic fields in massive stars is not well understood yet, we are cautious in drawing any conclusion on the presence and the behaviour of a magnetic field in \\tcar{}. Only additional time-resolved magnetic field observations will tell us about the presence and the structure of the magnetic field geometry of \\tcar{} and will help to discriminate among the hypotheses described above." }, "0807/0807.2856_arXiv.txt": { "abstract": "Accurately understanding the interior structure of extra-solar planets is critical for inferring their formation and evolution. The internal density distribution of a planet has a direct effect on the star-planet orbit through the gravitational quadrupole field created by the rotational and tidal bulges. These quadrupoles induce apsidal precession that is proportional to the planetary Love number ($k_{2p}$, twice the apsidal motion constant), a bulk physical characteristic of the planet that depends on the internal density distribution, including the presence or absence of a massive solid core. We find that the quadrupole of the planetary tidal bulge is the dominant source of apsidal precession for very hot Jupiters ($a \\lesssim 0.025$ AU), exceeding the effects of general relativity and the stellar quadrupole by more than an order of magnitude. For the shortest-period planets, the planetary interior induces precession of a few degrees per year. By investigating the full photometric signal of apsidal precession, we find that changes in transit shapes are much more important than transit timing variations. With its long baseline of ultra-precise photometry, the space-based \\emph{Kepler} mission can realistically detect apsidal precession with the accuracy necessary to infer the presence or absence of a massive core in very hot Jupiters with orbital eccentricities as low as $e \\simeq 0.003$. The signal due to $k_{2p}$ creates unique transit light curve variations that are generally not degenerate with other parameters or phenomena. We discuss the plausibility of measuring $k_{2p}$ in an effort to directly constrain the interior properties of extra-solar planets. ", "introduction": "Whether studying planets within our solar system or planets orbiting other stars, understanding planetary interiors represents our best strategy for determining their bulk composition, internal dynamics, and formation histories. For our closest neighbors, we have had the luxury of sending spacecraft to accurately measure the higher-order gravity fields of these objects, yielding invaluable constraints on their interior density distributions. Using these observations, we have been able, for instance, to infer the presence of large cores, providing support for the core-accretion theory of planet formation \\citep{2005AREPS..33..493G}. Study of planets outside our solar system, however, has necessitated the development and usage of more indirect techniques. Nevertheless, as the number of well-characterized extra-solar planets grows, we gain more clues that help us answer the most fundamental questions about how planets form and evolve. Guided by our current understanding of planetary physics, we have begun to study the interiors of extra-solar planets. This endeavor has been dominated by a model-based approach, in which the mass and radius of a planet are measured using radial velocity and transit photometry observations, and the interior properties are inferred by finding the model most consistent with those two observations. This strategy clearly requires a set of assumptions, not the least of which is that the physical processes at work in extra-solar planets are just like those that we understand for our own giant planets. While it does seem that this approach is adequate for explaining most of the known transiting planets, there does exist a group of planets for which the usual set of assumptions are not capable of reproducing the observations \\citep[e.g.,][]{2006A&A...453L..21G,2007ApJ...661..502B}. These are the planets with so-called positive ``radius anomalies'', including the first-discovered transiting planet HD 209458b \\citep{2000ApJ...529L..45C}. Though most of these planets can be explained by adjusting different pieces of the interior physics in the models (including opacities, equations of state, and heat deposition), it is currently impossible to discern which combination of these possible explanations is actually responsible for their observed sizes \\citep{2006A&A...453L..21G}. Additional uncertainties also exist for planets at the other end of the size spectrum. For the group of under-sized extra-solar planets, such as HD 149026b, the canonical approach is to give the planet a massive highly condensed core of heavy elements in order to match the observed radius. This approach also provides a first order estimate of the planet's bulk composition, in terms of its fraction of heavy elements. There is also the added complication of how the assumed state of differentiation affects the inferred composition and predicted structure \\citep{Baraffe2008}. Currently, the most promising approach to modeling the distinctive features of extra-solar planet interiors is to study the known transiting planets as an ensemble. The group can be used to develop either a single consistent model that reproduces all the observations \\citep[e.g.,][]{2006A&A...453L..21G} or to showcase the possible diversity in model parameters \\citep[e.g., opacities, as in][]{2007ApJ...661..502B}. Surely, a model-independent measure of interior structure would be valuable in order to begin disentangling otherwise unconstrained physics. The idea of obtaining direct structural measurements for distant objects is by no means a new one. For decades, the interiors of eclipsing binary stars have been measured by observing ``apsidal motion,'' i.e. precession of the orbit due to the non-point-mass component of the gravitational field \\citep{1928MNRAS..88..641R,1938MNRAS..98..734C,1939MNRAS..99..451S,1939MNRAS..99..662S}. The signal of the changing orbit is encoded in the light curves of these systems by altering the timing of the primary and secondary eclipses. From these eclipse times, it is straightforward to determine the so-called apsidal motion constant which then constrains the allowed interior density distributions. Interior measurements inferred from apsidal precession were among the first indications that stars were highly centrally condensed. While it seems non-intuitive, we show in this paper that we can use a similar technique to measure the interior properties of very hot Jupiters. Most surprisingly, the interior structure signal for very hot Jupiters actually dominates over the signal from the star, yielding an unambiguous determination of planetary interior properties. Our theoretical analysis is also extended to full simulated photometry in order to explore the observability of apsidal precession. We show that this precession is observable by measuring the subtle variations in transit light curves. The photometric analysis is focused on the data expected from NASA's \\emph{Kepler} mission, which successfully launched on March 6, 2009 \\citep{2003SPIE.4854..129B,2006Ap&SS.304..391K}. \\emph{Kepler} will obtain exquisite photometry on $\\sim$100,000 stars, of which about 30 are expected to host hot Jupiters with periods less than 3 days \\citep{2008arXiv0804.1150B}. \\emph{Kepler} has the potential to measure the gravitational quadrupoles of very hot Jupiters though the technique described below. If successful, this will constitute a major step towards an understanding of the diversity of planetary interiors. In Section 2, we describe the background theory that connects interior structure and orbital dynamics and explore which effects are most important. Section 3 applies this theory to the observable changes in the transit photometry, including full \\emph{Kepler} simulated light curves. We show in Section 4 that the signal due to the planetary interior has a unique signature. Other methods for inferring planetary interior properties are discussed in Section 5. The final section discusses the important conclusions of our work. ", "conclusions": "The planetary mass and radius are the only bulk physical characteristics measured for extra-solar planets to date. In this paper, we find that the planetary Love number ($k_{2p}$, equivalent to $J_2$) can also have an observationally detectable signal (quadrupole-induced apsidal precession) which can provide a new and unique probe into the interiors of very hot Jupiters. In particular, $k_{2p}$ is influenced by the size of a solid core and other internal properties. Core sizes can be used to infer the formation and evolution of individual extra-solar planets \\citep[e.g.,][]{2009arXiv0901.0582D,2009arXiv0903.1997H}. The presence of a nearby massive star creates a large tidal potential on these planets, raising significant tidal bulges which then induce non-Keplerian effects on the star-planet orbit itself. The resulting apsidal precession accounts for $\\sim$95\\% of the total apsidal precession in the best cases (Figure \\ref{odotfig}). Hence, we find that the internal density distribution, characterized by $k_{2p}$, has a large and clear signal, not to be confused with any other parameters or phenomena. We urge those modeling the interior structures of extra-solar planets to tabulate the values of $k_{2p}$ for their various models. Encouraged by this result, we calculated full photometric light-curves like those expected from the \\emph{Kepler} mission to determine the realistic observability of the interior signal. We estimate that \\emph{Kepler} should be able to distinguish between interiors with and without massive cores ($\\Delta k_{2p} \\simeq 0.1$) for very hot Jupiters with eccentricities around $e \\sim 0.003$ (Figure \\ref{ek2planet}). Eccentricities this high may occur for some of the very hot Jupiters expected to be found by \\emph{Kepler}, though these planets usually have highly damped eccentricities. Much stronger constraints on apsidal precession can be obtained by combining \\emph{Kepler} photometry with precise secondary transits observed in the infrared. In cases where apsidal precession is not observed, the data can set strong upper limits on planetary eccentricities. In analyzing \\emph{Kepler}'s photometric signal of apsidal precession, we find that transit timing variations are an almost negligible source of signal, though transit timing has been the focus of many observational and theoretical papers to date. The effect of ``transit shaping'' has $\\sim$30 times the photometric signal of transit timing for apsidal precession \\citep[see Figure \\ref{candywrapperpieces},][]{PK08,JB08}). At orientations where transit timing and shaping are weakest, the changing offset between primary and secondary transit times can be used to measure $k_{2p}$ (Figure \\ref{bowtiemain}). It may also be possible to measure $k_{2p}$ from high-precision multi-color photometry by directly detecting the planetary asphericity in transit. Such a measurement does not require a long baseline or an eccentric orbit. Very hot Jupiters are also excellent candidates for detecting tidal semi-major axis decay, where we find that relatively small period changes of $\\dot{P} \\simeq 1$ ms/yr should be detectable. This could constitute the first measurements (or constraints) on tidal $Q_*$ for a variety of individual stars. We note that \\emph{Kepler} measurements of transit timing and shaping for eclipsing binaries should also provide powerful constraints on stellar interiors through apsidal motion and binary orbital decay (due to tides, if the components are asynchronous). Accurately measuring the interior structure of distant extra-solar planets seems too good to be true. Nevertheless, the exquisite precision, constant monitoring, and 3.5-year baseline of the \\emph{Kepler} mission combined with the high sensitivity of transit light curves to small changes in the star-planet orbit make this measurement plausible. Our focus on \\emph{Kepler} data should not be interpreted to mean that other observations will be incapable of measuring $k_{2p}$. In fact, the opposite is true since the size of the apsidal precession signal increases dramatically with a longer baseline. Combining \\emph{Kepler} measurements with future ground and space based observations can create a powerful tool for measuring $k_{2p}$. In the far future, many planets will have measured apsidal precession rates (like eclipsing binary systems have now) and inferred $k_{2p}$ values. Incorporating these measurements into interior models holds promise for greater understanding of all extra-solar planets." }, "0807/0807.2251_arXiv.txt": { "abstract": "In the frame of the collapsar model for long gamma ray bursts (GRBs), we investigate the formation of a torus around a spinning BH and we check what rotational properties a progenitor star must have in order to sustain torus accretion over relatively long activity periods. We also study the time evolution of the BH spin parameter. We take into account the coupling between BH mass, its spin parameter and the critical specific angular momentum of accreting gas, needed for the torus to form. The large BH spin reduces the critical angular momentum which in turn can increase the GRB duration with respect to the Schwarzschild BH case. We quantify this effect and estimate the GRB durations in three cases: when a hyper accreting torus operates or a BH spins very fast or both. We show under what conditions a given progenitor star produces a burst that can last as short as several seconds and as long as several hundred of seconds. Our models indicate that it is possible for a single collapse to produce three kinds of jets: (1) a very short, lasting between a fraction of a second and a few seconds, 'precursor' jet, powered only by a hyper accreting torus before the BH spins up, (2) an 'early' jet, lasting several tens of seconds and powered by both hyper accretion and BH rotation, and (3) a 'late' jet, powered only by the spinning BH. ", "introduction": "The commonly accepted mechanism for a long gamma ray burst (GRB) production invokes a collapsar scenario (Woosley 1995; Paczy\\'nski 1998; MacFadyen \\& Woosley 1999). In this model the material from the collapsing star feeds the accretion disk, then the accretion energy is being transferred to the jet, which in turn produces gamma rays at some distance from the central engine. Therefore the whole event cannot last much longer than the existence of a rotationally supported torus in the collapsar center. Within the collapsar model the jet can also be produced by a rotating black hole (BH) which can be spun up by the accreting torus material. Among the most plausible mechanisms of the energy extraction from the accretion flow are the neutrino-antineutrino annihilation (Mochkovitch et al. 1993), or the magnetic fields (e.g. Blandford \\& Payne 1982; Contopoulos 1995; Proga et al. 2003). The neutrino cooling (e.g. Popham, Woosley \\& Fryer 1999; Di Matteo et al. 2002; Janiuk et al. 2004) is effective only if the accretion rate is large ($\\dot m \\gtrsim 0.01 M_{\\odot}$ sec$^{-1})$). Also, a large BH spin ($A_{\\rm Kerr} \\gtrsim 0.9$) is thought to be a necessary condition for the jet launching: for $A_{\\rm Kerr}\\sim 0.9$, about 1\\% of the accreted rest-mass energy is emitted back as a Poynting jet (Blandford \\& Znajek 1977; McKinney 2005). On the other hand, the rotationally supported torus may form only when the substantial amount of specific angular momentum is carried in the material. In our recent article (Janiuk \\& Proga 2008; hereafter Paper I) we studied the problem of whether the collapsing star envelope contained enough specific angular momentum in order to support the formation of the torus. This condition was parametrized by the so called critical specific angular momentum which in case of a non-rotating BH depends only on its mass. In the present work, we take into account also the BH rotation and the coupling between the specific angular momentum of the accreting material, the BH mass and its spin. We show that as in Paper I, during the collapse the amount of the rotating material, which was initially available for the torus formation, may later become insufficient to support the torus. Moreover, the spin of the BH is changed by accretion (see e.g. recent studies by Gammie et al. 2004; King \\& Pringle 2006; Belczy\\'nski et al. 2007). In our models, depending on the accretion scenario, both the spin-up and spin-down of BH are possible, because part of the infalling material has very small specific angular momentum. The outline of this paper is as follows. In Section \\ref{sec:model}, we briefly describe the model of the evolution of the collapsing star together with the initial conditions. The results are presented in Section \\ref{sec:results}; the GRB durations are estimated in Section \\ref{sec:durations}. In Section \\ref{sec:diss}, we discuss results in the context of a long GRB production mechanisms and conditions for the distribution of the specific angular momentum in a progenitor star. In Appendix \\ref{sec:rafal} we provide some formulae for the description of the mass and spin evolution of the BH. ", "conclusions": "\\label{sec:diss} In this article, we studied the collapsar model for long GRBs, powered by accretion onto a spinning BH, which formed from the core of a massive, rotating Wolf-Rayet star. To describe the rotation of the stellar interior, we adopted two different analytical functions, accounting for either a differential rotation (models {\\bf A1}, {\\bf A2}, {\\bf A3}), or a constant ratio between the gravitational and centrifugal forces (models {\\bf D1}, {\\bf D2}, {\\bf D3}). This study is an important test for the rotation models of the GRB progenitor stars (e.g. Heger et al. 2005; Yoon et al. 2006; Detmers et al. 2008). To describe how the accretion proceeds during the collapse, we adopted three different scenarios: (1) uniform accretion, (2) two phase accretion, first from the poles and then from the torus and (3) only torus accretion. The accretion onto the BH is in our approach a homologous process, in which the subsequent shells of the envelope add their mass to the central object. The angular momentum is also accreted, but the limit for it is the critical angular momentum, to prevent the BH from spinning with $A_{\\rm Kerr} \\ge 1.0$. In this sense, we assume that the whole angular momentum with $l > l_{\\rm crit}$, i.e. in the torus, is transported outwards. We do not invoke any particular mechanism of transport (i.e. the viscosity), and the momentum is taken out by a negligibly small amount of mass (e.g., Pringle 1981). This simplified approach describes well a more realistic situation, in which the matter with small and large angular momentum can be mixed. Therefore some parts of the gas with large $l_{\\rm spec}$ might reach the BH, while some other parts might be blown out with the polar outflow. We focused on the evolution of the BH spin during the collapse. The large BH spin is important for GRB production in two ways: first, to power the jet emission via the Blandford-Znajek (BZ) mechanism, and second, because it alters the condition for the torus formation, i.e. the critical specific angular momentum. We found that the spin of the BH strongly depends on both the model of the $l_{\\rm spec}$ distribution and on the accretion scenario. In the torus accretion (i.e. either the second phase of the scenario 2, models {\\bf A2} and {\\bf D2}, or the scenario 3, models {\\bf A3} and {\\bf D3}), the accreting material has specific angular momentum always $l_{\\rm spec} \\ge l_{\\rm crit}$. This angular momentum must be transported outwards before reaching the BH, so that the gas which is changing the BH spin has the specific angular momentum equal to $l_{\\rm crit}$. Nevertheless, it is enough to spin up the BH to the maximal rotation, $A_{\\rm Kerr}=0.9999$, which happens in most cases at the very beginning of the collapse. The polar accretion, i.e. the first phase of scenario 2. (models {\\bf A2} and {\\bf D2}), leads only to the BH spin-down in all the models. The uniform accretion scenario is the most complex. In the model {\\bf A1} it leads only to a temporary increase of the BH spin, while during the accretion of the outer shells, the BH is spinning down. In the model {\\bf D1}, the BH spin first decreases, while later during the collapse it may increase, provided the stellar envelope contains enough $l_{\\rm spec}$. We found that in model {\\bf A1}, the final BH spin after the collapse is always about $A_{\\rm end} \\sim 0.85$, and it does not depend on the normalization the specific angular momentum contained in the stellar envelope, i.e. on $x$. However, the pattern of the BH spin evolution is very sensitive to this parameter. Therefore for small values of $x$ it may happen that even for a short time during the collapse, the BH never reaches a spin $A_{\\rm Kerr}>0.9$, which we consider necessary to power the jet with the BZ mechanism. However, in the same models, the torus does exist and the accretion rate in the torus is large enough to power the jet via the neutrino annihilation. This might lead to a relatively short living (less than $\\sim$ 7-8 s) GRB central engine without a very rapidly spinning BH. On the other hand, for $A_{\\rm Kerr}>0.9$, the stage of a rapidly spinning BH begins very shortly after the collapse has started, and lasts much longer after the accretion rate in the torus has dropped below $\\dot m=0.01$ M$_{\\odot}$s$^{-1}$. For instance, a GRB powered by the BZ mechanism may last almost $\\sim 120$ s, while that powered by the neutrino annihilation (concurrent with the spinning BH) lasted only $\\sim 40$ s. A very short time required for the BH to spin up, while the collapse proceeds, is of the order of $\\sim 1.5$ s. In model {\\bf D1} the situation is different. Here we do not find any models with only the neutrino-powered bursts, i.e. with a large accretion rate but not accompanied with a rapidly spinning BH. In other words, whenever there exists a torus with a large accretion rate, the BH is spun up to $A_{\\rm Kerr}>0.9$, and the timescale for this spin up is a fraction of a second ($\\sim 0.15$ s). Similarly to model {\\bf A1}, the stage of a large BH spin can last much longer, after the accretion rate has dropped below $\\dot m=0.01$ M$_{\\odot}$s$^{-1}$. For instance, the BZ-powered burst lasting $\\sim 430$ s is accompanied by a $\\sim 100$ s burst powered by both BZ and neutrino mechanisms. Observationally, this behavior may have led to three kinds of jets. The first is a very short, lasting between a fraction of a second and few seconds, 'precursor' jet, powered by only the neutrino annihilation, before the BH spins up. The second is an 'early' jet, lasting several tens of seconds and powered by both neutrino and BZ mechanisms. The third is a 'late' jet, powered by only the spinning BH via the BZ mechanism. In our models, we can have the GRB jets with all the three components, as well as the 'orphan precursor' jets, when the BH failed to spin up. The precursors have been detected by Ginga, BeppoSAX, BATSE, INTEGRAL and Swift in some GRBs (e.g. Murakami et al. 1991; Piro et al. 2005; Lazzati 2005; Romano et al. 2006; McBreen et al. 2006). These GRB precursors are an important observational test for their theoretical models (e.g. Ramirez-Ruiz et al. 2002; Umeda et al. 2005; Morsony et al. 2007; Wang \\& Meszaros 2008). For instance, in the sample of BATSE bursts, studied by Lazzati (2005), about 20\\% of the bursts had a precursor, which was characterized by a non-thermal spectrum and contained less than 1 per cent of the total counts. The main GRB in these events was delayed with respect to the precursor by 10-200 seconds. As argued by Morsony et al. (2007), who in the 2-D numerical MHD simulations identified three distinct phases during the jet propagation, this large gap in the emission might be a selection effect. Because of different opening angles of these three jets, some observers located at large viewing angles may see a 'dead' phase, i.e. the break in the emission, related to the second jet. Another explanation of the gap between the precursor and the main jet could be the development of the instabilities in the hyper-accreting disk (Perna et al. 2006; Janiuk et al. 2007), possibly combined with the viewing angle effects. We therefore conclude that in the present model, the 'dead' phase would refer to an 'early' jet, which is powered by both neutrino and BZ mechanisms and can be collimated to a much narrower angle than the 'late' jet. For the viewing angle larger than the 'early' jet but smaller than the 'late' jet opening angle, the observer should see the precursor, followed by a gap in the emission on the order of 40-150 seconds, and then see the 'main' GRB. We also notice that recently, the observation of the bright, long GRB 080319B (Racusin et al. 2008), seems to have confirmed that the jet's opening angle may vary, indicating for the two types of jets. Finally, comparing our current models with the results presented for a non-rotating BH (Paper I), we notice that the GRB durations are similar in case of model {\\bf A1}, i.e. $\\sim 40$ s vs. $\\sim 50$ for the Schwarzschild and Kerr BH main jet, respectively. In model {\\bf D1}, the discrepancy is more pronounced, namely $\\sim 50$ s vs. $\\sim 100$ s, respectively. On the other hand, in the current work, the model {\\bf D1} produces GRBs powered by neutrino annihilation only for a very narrow range of parameters (i.e. $x$), while in Paper I for this model we found no limitations for $x$." }, "0807/0807.2910_arXiv.txt": { "abstract": "{Some recent observational results impose significant constraints on all the models that have been proposed to explain the Galactic black-hole X-ray sources in the hard state. In particular, it has been found that during the hard state of Cyg X-1 the power-law photon number spectral index, $\\Gamma$, is correlated with the average time lag, $$, between hard and soft X-rays. Furthermore, the peak frequencies of the four Lorentzians that fit the observed power spectra are correlated with both $\\Gamma$ and $$. } {We have investigated whether our jet model can reproduce these correlations.} {We performed Monte Carlo simulations of Compton upscattering of soft, accretion-disk photons in the jet and computed the time lag between hard and soft photons and the power-law index $\\Gamma$ of the resulting photon number spectra.} {We demonstrate that our jet model naturally explains the above correlations, with no additional requirements and no additional parameters. } {} ", "introduction": "The states of Galactic black-hole X-ray sources (GBHs) are determined from the following quantities: i) their disk fraction, which is the ratio of the disk flux to the total flux, both unabsorbed, at $2-20$ keV; ii) the photon number spectral index, $\\Gamma$, of the hard band, power-law component at energies below any break or cutoff; iii) the root-mean-square (rms) power in the power spectral density (PSD) integrated from 0.1--10 Hz; iv) the integrated rms amplitude of any quasi-periodic oscillation (QPO) detected in the range 0.1--10 Hz \\citep{remi06}. In particular, the so-called hard state is characterized by a power-law X-ray spectrum with photon number spectral index $1.5 \\simless \\Gamma \\simless 2.1$, which suffers an exponential cutoff at $\\sim$ 100 or few 100 keV, and contributes more than 80\\% of the 2--20 keV flux. Several models have succeeded in reproducing the observed X-ray spectrum in this state \\citep[see e.g.][]{tita94,esin97,pout99,reig03,gian04}. When a GBH is in the hard state, radio emission is also detected and a jet is either seen or inferred \\citep{fend01}. Again, several models are successful in reproducing the energy spectrum from the radio domain to the hard X-rays \\citep[see e.g.][]{mark01,vada01,corb02,mark03,gian05}. The multiplicity of models that can fit the time average spectrum of GBHs well indicates that this alone is not enough to distinguish the most realistic among them. Timing data of the observed intensity in various energy bands provide a totally different ``dimension\" than the spectral one, and can in principle provide information on the location, size, physical conditions and kinematics of the plasma responsible for the X-ray emission in these objects. For example, the X-ray light curves of GBHs in different energy bands show similar variations, but those in the higher energy band generally lag the variations detected in the lower energy band \\citep{nowa99,ford99}. The time delay, $t_{\\rm lag}$, usually depends on frequency, $\\nu$. For Cyg X-1 the time delays decrease with increasing frequency roughly as $t_{\\rm lag} \\propto \\nu^{-\\beta}$, where $\\beta \\approx 0.7$. As constraining as it is, this relation has also been explained in more than one way \\citep[see e.g.][]{pout99,reig03,kord04}. Even more constraining to models is the fact that the high-frequency power spectrum flattens as the photon energy increases \\citep{nowa99}. Equivalent to this is the fact that the width of the temporal autocorrelation function of the light curves of Cyg X-1 decreases with increasing photon energy \\citep{macc00}. These observational facts are contrary to what one expects from simple Comptonization models, yet possible explanations for them have also been offered \\citep{koto01,gian04}. Significant advances in our understanding of how the X-ray mechanism in GBHs (and compact accreting objects in general) works can be achieved when we combine variability and spectral information. This has been possible in the last few years, due mainly to \\xte. For example, in the case of Cyg X-1, \\citet{pott03} used 130 \\xte\\ observations between 1998 and 2001 to study the long-term evolution of the source's power and energy spectrum in detail. They used 512 s long light curves to estimate the power spectral density (PSD), which they fitted with a model that consisted of the sum of multiple Lorentzian profiles. They generally achieved good fits using four broad Lorentzian components with typical peak frequencies of $\\nu_1\\sim 0.2$ Hz, $\\nu_2\\sim 2$ Hz, $\\nu_3\\sim 6$ Hz, and $\\nu_4\\sim 40$ Hz, which vary in unison with time: their ratios remain constant despite each frequency varies by up to a factor of $\\sim 5$. Up to now, no physical explanation of these Lorentzians has been offered. As for the energy spectrum, they find that a simple model consisting of a power-law component of index $\\Gamma$, a multi-temperature disk black body and a reflection from neutral material component could fit the energy spectrum of the source below 20 keV. They present convincing evidence that the timing and spectral properties of the source are closely linked. In fact, a number of the correlations they present impose stringent constraints on all the models proposed so far. These correlations are: a) The observed photon number spectral index, $\\Gamma$, correlates with the observed average time lag, $$, between 2--4 keV and 8--13 keV, averaged over the 3.2--10 Hz frequency band, in the sense that $$ increases as the spectrum steepens. Therefore, a model that can successfully explain the spectral slope $\\Gamma$ should also predict the right amount of time delay between these specific energy bands and vice versa: a model that can fit well the time lag vs. Fourier frequency relation, should also predict the appropriate value of the index $\\Gamma$ for a given time delay. b) The observed photon number spectral index, $\\Gamma$, also correlates with the Lorentzian peak frequencies: the frequencies increase as the spectrum steepens \\citep[see also][]{shap06,shap07}. In summary, the \\citet{pott03} results show that a specific value of $\\Gamma$, say $\\Gamma = 2$, which is relatively easy to obtain with a simple Comptonization model, must also be associated with {\\it specific} values of the Lorentzian peak frequencies and a {\\it specific} value of $$! These are challenging relations that any model must be able to explain. In this work we investigate whether the jet model that we have proposed can meet these constraints under reasonable assumptions. We think that the jets in GBHs play a central role in all the observed phenomena, not only in the radio emission. We have shown that Compton upscattering in the jet of soft photons from the accretion disk can explain: 1) the X-ray spectrum and the time-lag dependence on Fourier frequency \\citep[][hereafter Paper I]{reig03}; 2) the narrowing of the autocorrelation function and the increase of the rms amplitude of variability with increasing photon energy \\citep[][hereafter Paper II]{gian04}; 3) the energy spectrum from radio to hard X-rays of XTE J1118+480, the only source for which we have simultaneous observations for all energy bands \\citep[][hereafter Paper III]{gian05}. In this paper we show that the jet model proposed and developed in the above work can explain the $\\Gamma$ -- $$ relation of \\citet{pott03}, with no additional parameters or requirements, if we let just two model parameters, namely the radius at the base of the jet and the optical depth along its axis, vary over a reasonable range of values. At the same time, the model can also explain the \"$\\Gamma$ -- Lorentzian peak frequencies $\\nu_i$\" relation qualitatively if we identify $1/\\nu_i$, $i=1, 2, 3, 4$, with quantities proportional to the characteristic timescale $M_{\\rm out}/{\\dot M}_{\\rm out}$, where ${\\dot M}_{\\rm out}$ is the rate of mass outflow in the jet and $M_{\\rm out}$ the available mass to power the jet. Finally, the model parameter variations that are needed to explain the \\citet{pott03} relations also predict that, on long time scales, the radio variations should also correlate with $\\Gamma$, which, as we show, is indeed the case in Cyg X-1. In Sect. 2 we describe briefly our model, in Sect. 3 we present and discuss our results, and we close with our conclusions in Sect. 4. ", "conclusions": "In a series of three papers (Papers I, II, and III) it has been shown that our jet model can explain a number of observations regarding black-hole X-ray binaries in the hard state. These are: a) the energy spectrum from radio to hard X-rays; b) the time-lag of the hard photons relative to the softer ones as a function of Fourier frequency; c) the flattening of the power spectra at high frequencies with increasing photon energy and the narrowing of the autocorrelation function. In this paper we investigate whether our model can explain the long term variations in the spectral and timing properties of Cyg X-1, which are closely linked. The main result of our work is that we can reproduce the $\\Gamma-$ correlation of \\citet{pott03} if we assume reasonable variations in just two model parameters around their mean values: the width of the jet at its base and the optical depth along the jet's axis (or equivalently the electron density at the base of the jet). We find that, if $\\tau_{\\parallel}\\approx 1.5-10$, $R_0\\approx 5-50 r_s$ and $\\tau_{\\parallel}\\propto R_0^{-1}$, then the model predicts a $\\Gamma-$ relation that is almost identical to the one that \\citet{pott03} observed. Furthermore, if the flux variability operates on a time scale which is proportional to $M_{\\rm out}/\\dot{M}_{\\rm out}$ (i.e. the characteristic time needed to eject all the available matter to the jet), then the same $\\tau$ and $R_0$ variations can also explain the {\\it shape} of the ``spectral index - Lorentzian peak frequencies\" correlation of \\citet{pott03}. Finally, given the model parameter variations that are needed to explain the observed \"spectral - timing\" correlations, the model predicts a \"radio flux -- $\\Gamma$\" correlation, which, as we show, does indeed hold in Cyg X-1. The $\\tau_{\\parallel}\\propto R_0^{-1}$ relation, which is necessary to explain the observed $\\Gamma-$ relation, is a natural outcome of the model if simply $\\dot{M}_{\\rm out}\\propto R_0$. We propose that this is the underlying ``fundamental\" relation that governs the {\\it long-term} evolution of the physical characteristics of the jet, hence of the observational spectral and timing properties of the source. The physical reason for this relation is unclear at the moment; but whatever its origin, the relation itself does make sense because as the rate of the mass ejected in the jet increases, it seems reasonable to assume that the jet will become ``bigger\", i.e. $R_0$ will also increase. In fact, in this way, one can explain reasonably well all the long-term observed changes with variations in one fundamental physical parameter of the system, namely the accretion rate. As the accretion rate increases, $\\dot{M}_{\\rm out}$ should also increase. If the jet size increases proportionally to $\\dot{M}_{\\rm out}$, then $n_0 \\propto R_0^{-1}$, and hence the $\\Gamma$ -- $$ relation. We hope that this ``accretion rate -- ejection rate -- jet size\" relation will provide an important and useful hint for all models that try to explain the connection between accretion flows and outflows, hence the formation of jets in regions close to the central engines in accreting compact objects. We find that the $\\dot{M}_{\\rm out}\\propto R_0$ relation predicts that the radio flux should increase linearly with the photon number spectral index. To the best of our knowledge, such a relation has never been suggested or even detected either in Cyg X-1 or in any other GBH in the past. We show that a linear ``radio flux -- spectral index\" relation does exist on long time scales in Cyg X-1. We believe that this result strongly supports our view that a) X-rays in Cyg X-1 are produced by Compton upscattering in the jet of soft photons from the accretion disk and b) the jet's size and optical depth evolve with time because $\\dot{M}_{\\rm out}\\propto R_0$. Furthermore, that the ``radio flux -- spectral index\" relation breaks at index values higher than $\\sim 2.2$, provides interesting hints into what may be happening as the source moves to the intermediate and thermal state. Most probably the magnetic field decreases and/or there is weaker particle acceleration in the jet (i.e. the electron distribution index, $\\alpha$, becomes larger) as the accretion rate increases (as inferred from the increase in the disk thermal emission). The explanation of the short time-scale variability properties of Cyg X-1 is significantly more challenging. Our model cannot explain why four Lorentzian components exist in the source's power spectrum (assuming of course that the Lorentzians are the fundamental building blocks of the power spectrum in Cyg X-1 and other GBHs). However, it can explain, to some extent, the way they vary with the source's photon number spectral index. If indeed the long-term evolution of the source is governed by variations in $R_0$ and $\\dot{M}_{\\rm out}$, it is natural to assume that the short-term variations are caused by short-term, random variations of $\\dot{M}_{\\rm out}$ (and hence of $n_0$ or $\\tau_{\\parallel}$, and finally X-ray flux), around its mean value. We find that, if these variations operate on a time scale that is proportional to the characteristic time needed to eject all the available matter to the jet, then the model can explain the $\\it shape$ of the ``$\\Gamma$ - Lorentzian peak frequencies\" relations of \\citet{pott03}. There is no reason why the model ``$\\Gamma -$ inverse time needed to eject the mass to the jet\" relation should have the same shape as the observed ``$\\Gamma -$ Lorentzian peak frequencies\". Thus we do not find this agreement to be coincidental. Instead, it is one more evidence in favor of the jet model and the parameters' evolution we present in this work, and we find it operates on long time scales in Cyg X-1. Our results suggest that the characteristic frequencies seen in the power spectrum should increase with increasing size of the source, i.e. $R_0$, a result that goes against current views. However, the characteristic frequencies may not scale directly with the source size, but with accretion rate, hence $\\dot{M}_{\\rm out}$. As argued above, it seems reasonable to think that, as the accretion rate increases, so too the mass ejection rate to the jet. Perhaps, then, it is the accretion rate (through $\\dot{M}_{\\rm out}$) that drives the variations in both $R_0$ and on the characteristic time scales. A physical explanation as to how this may happen and as to the value (and number) of characteristic time scales, requires knowledge of the details of the acceleration/ejection process of the jet, and is beyond the scope of the present work. However, as we mentioned above, we hope that the results we present here will help the investigation of the proper mechanism responsible for the launch of the jets and winds in accreting compact objects. We note that our conclusion, presented in Sect. 3.2, that the Lorentzian peak frequencies $\\nu_i$ {\\it increase} with the size $R_0$ of the base of the jet, holds for a given source (in our case Cyg X-1). If one wanted to extend our model to other sources, then the mass of the black hole comes in. The argument goes as follows: Making the reasonable assumption that $M_{\\rm out} \\propto R_0^2$ and using $n_0 \\propto R_0^{-1}$, we find that the $\\nu_i$ scale as $1/R_0$. Since $R_0$ is typically a few times $r_g$, we conclude that {\\it the Lorentzian peak frequencies $\\nu_i$ are inversely proportional to the mass of the black hole}. In closing, we want to point out that the above results and conclusions are consistent with the fact that, in the thermal state of Cyg X-1, the time lags versus Fourier frequency are {\\it identical} to those in the hard state \\citep{pott00}. In the thermal state, the radio emission is very weak and possibly undetectable, {\\it but we think that the jet is there} and that it produces, by Comptonization, both the time lags versus Fourier frequency and the steep power-law energy spectrum above $\\sim 10$ keV. According to our model, the suppression of both the radio and the X-ray emission in this state should be caused by a magnetic field suppression and a weaker particle acceleration, respectively." }, "0807/0807.1122_arXiv.txt": { "abstract": "{Numerical simulations of forced turbulence in elongated shearing boxes are carried out to demonstrate that a nonhelical turbulence in conjunction with a linear shear can give rise to a mean-field dynamo. Exponential growth of magnetic field at scales larger than the outer (forcing) scale of the turbulence is found. Over a range of values of the shearing rate $S$ spanning approximately two orders of magnitude, the growth rate of the magnetic field is proportional to the imposed shear, $\\gamma\\propto S$, while the characteristic spatial scale of the field is $\\lb\\propto S^{-1/2}$. The effect is quite general: earlier results for the nonrotating case by \\cite{}Yousef et~al. (2008) are extended to shearing boxes with Keplerian rotation; it is also shown that the shear dynamo mechanism operates both below and above the threshold for the fluctuation dynamo. The apparently generic nature of the shear dynamo effect makes it an attractive object of study for the purpose of understanding the generation of magnetic fields in astrophysical systems.} ", "introduction": "Turbulence is generally considered to play a fundamental role in the generation and maintenance of magnetic fields found in a wide range of astrophysical systems. It is a numerically well established property of turbulence to amplify magnetic fluctuations at the same or smaller scales than the scales of the turbulent motions \\citep{meneguzzi81,sch04,haugen04,iskakov07,sch07}. This type of dynamo is known as small-scale, or fluctuation, dynamo. It is believed to be a universal property of turbulent systems and, at least in the case of large magnetic Prandtl numbers, a simple theoretical picture exists of the field amplification via random stretching by turbulent motions (\\citealt{moffatt64,zeldovich84}; see also \\citealt{sch04,SC07}). It often turns out to be more difficult to either establish or explain the generation of magnetic fields at much larger scales than the turbulence scale. Such fields are observed or believed to exist in many astrophysical bodies (stars, galaxies, disks), so the question of their origin is an important theoretical challenge. Mechanisms for the generation of such large-scale fields are known as large-scale, or mean-field, dynamos. A motley of such mean-field dynamos has been studied in the literature \\citep[e.g.,][]{moffatt78,krause80,brand05review,raedler07}. The precise way in which they operate often seems to be system dependent and introducing ever more realistic features into one's theoretical model produces ever more complicated behaviour. While such modelling is necessary for quantitative understanding, it is quite interesting to ask what are the generic ingredients required to produce large-scale fields. One such ingredient appears to be the presence of net kinetic helicity in the system: in many mean-field dynamos, the key generation mechanism is the so-called $\\alpha$ effect \\citep{steenbeck66}, whereby an assembly of non-mirror-symmetric velocity fluctuations having a nonzero net helicity are responsible for magnetic-field generation. The existence of a mean-field dynamo in helically forced turbulence is well established numerically \\citep{brand01,maron02,brand08}. However, the requirement of net helicity may somewhat limit the applicability of the $\\alpha$ effect. It is also far from certain that the direct link between kinetic helicity and mean-field generation via the $\\alpha$ effect found in the model case where the helicity is injected by the random forcing, carries over to the cases where the helicity arises naturally (e.g., in a rotating convective layer; see \\citealt{cattaneo06}). Therefore, there is a strong motivation for seeking alternative mean-field dynamo processes driven by nonhelical turbulence. It is clear that a nonhelical turbulence by itself cannot make large-scale fields. In recent years, many authors have argued that large-scale magnetic fields can be generated by nonhelical velocity fluctuations when acted upon by a large-scale shear: theoretical paradigms put forward to support such a dynamo, which we refer to as the shear dynamo, have included the shear-current effect \\citep{RK03,RK04}, the stochastic $\\alpha$ effect (\\citealt{vishniac97,silantev00,fedotov03,fedotov06,proctor07,brand07}; see, however, \\citealt{kleeorin08}), negative-diffusivity type theories \\citep{urpin99a,urpin99b,ruediger01,urpin02,urpin06}, shear amplification of the fluctuation-dynamo-generated small-scale fields \\citep{blackman98}. While there is not yet agreement between theoreticians about the validity or areas of practical applicability of these models, it is clear that they are addressing a fundamental issue. Indeed, shear is an extremely generic property of astrophysical systems, so the idea of a { shear dynamo} gives us a particularly attractive scenario for ubiquitous generation of large-scale magnetic fields. Until recently, a numerical demonstration of this type of dynamo remained elusive. Originally motivated by the predictions of \\citet{RK03,RK04}, we have previously performed numerical simulations of nonhelical turbulence with a superimposed linear shear and demonstrated the existence of the shear dynamo \\citep{yousef08}. Theoretical understanding of these numerical results is still poor. More work and, we believe, more information gathered from numerical experiments are needed in order to make progress towards understanding the properties of this dynamo and the underlying physical processes that produce it. This paper, which is an extension of the work by \\citet{yousef08}, aims at presenting a collection of new numerical results regarding the existence and behaviour of the shear dynamo in various regimes. We focus on three different cases of astrophysical interest, namely shear dynamo in the presence of forced nonhelical turbulence and a linear velocity shear, shear dynamo in the presence of forced nonhelical turbulence and Keplerian differential rotation, and finally shear dynamo in the presence of forced nonhelical turbulence, Keplerian differential rotation and a small-scale fluctuation dynamo. In order to study the effects of shear and rotation, we adopt a local rotating shearing sheet model. This model and the corresponding numerical set-up are presented in \\secref{sec:model}. In \\secref{sec:linear}, we consider the case of linear shear without rotation. Results for the Keplerian regime are presented in \\secref{sec:kepler}. \\Secref{sec:fd} describes some preliminary results on the shear dynamo in the presence of small scale magnetic fluctuations generated by the fluctuation dynamo. A short discussion concludes the paper (\\secref{sec:conc}). ", "conclusions": "\\label{sec:conc} Can nonhelical turbulence in combination with large-scale velocity shear act as a mean-field dynamo and generate magnetic fields with length scales much larger than the outer scale of the turbulence? This has been the subject of considerable recent debate \\citep{vishniac97,urpin99a,urpin99b,silantev00,urpin02,fedotov03,RK03,RK04,brand05,fedotov06,raedler06,ruediger06,raedler07,proctor07,brand07,kleeorin08}. To our knowledge, the paper by \\citet{yousef08} and this paper present the first set of dedicated numerical experiments that demonstrates that such a generation mechanism is feasible. However, in retrospect, one might conjecture that the shear-dynamo might have already been seen in several earlier numerical studies that combined large-scale flows, and consequently large-scale shear, with nonhelical forcing at smaller (or, in some cases, similar) scales and reported generation of magnetic fields at scales larger than the forcing scale \\citep{brand05,ponty05,mininni05,shapovalov06}. We have carried out a suite of numerical experiments on the shear dynamo effect in vertically elongated shearing boxes and for magnetic Reynolds numbers subcritical with respect to the fluctuation dynamo. For the values of the imposed shear $S$ between $1/8$ and $8$ (corresponding to $S\\tau\\sim 0.04 \\dots 3$, where $\\tau$ is the turnover rate of the randomly forced velocity fluctuations), we have found that the dynamo growth rate is $\\gamma\\propto S$ and the characteristic length scale of the generated mean magnetic field is $\\lb\\propto 1/\\sqrt{S}$. The first key result of this paper, compared with the earlier study by \\citet{yousef08}, is that the shear dynamo works both in the nonrotating case and for the case of Keplerian rotation ($\\Omega=-2S/3$). There does not appear to be much difference, qualitative or quantitative, between the rotating and nonrotating cases, although perhaps it would be interesting to look at non-Keplerian cases and try to identify the role of rotation via a parameter scan in $\\Omega$ independent of the one in $S$. The second key result, claimed on the basis of only a preliminary study, is that the shear dynamo works both for situations that are sub- and supercritical with respect to the fluctuation (small-scale) dynamo. In the latter case, the overall magnetic energy grows very quickly due to the fluctuation dynamo effect independent of the presence of the shear. Imposing the shear on the magnetohydrodynamic turbulence resulting from the saturation of the fluctuation dynamo, leads to the emergence of magnetic fields that have spatial scales larger than the outer scale of the turbulence and that fluctuate on very long time scales compared to the turbulent turnover time. Thus, the shear dynamo effect appears to be quite general and robust. As the combination of a shear flow and turbulence is a very common feature in astrophysical systems, the shear dynamo potentially represents a generic mechanism for making large-scale fields. While much needs to be understood about its properties before its relevance to real astrophysical systems can be more than an appealing speculation, the simplicity of the idea of the shear dynamo certainly makes it a worthwhile object of study. It is also important to determine how generic the shear dynamo is and how it combines with other large-scale features present in real astrophysical systems: various differential rotation laws, temperature and density gradients, linear instabilities, etc.. Studies in this vein are already being undertaken. For example, recent numerical experiments by \\citet{kapyla08} have shown that large-scale dynamo action is also possible in local simulations of magnetoconvection with imposed horizontal shear. They also report a growth rate $\\gamma \\sim S$ and find large-scale dynamo action for magnetic Reynolds numbers above the critical threshold for the fluctuation dynamo. Another topical recent study is by \\citet{gressel08}, who simulated the supernova-driven galactic turbulence and found that they needed to impose a linear velocity shear to obtain the amplification of a large-scale field. These studies clearly demonstrate the key role of shear in producing a mean-field dynamo. However, in comparing their results to ours, one has to keep in mind that their simulations had rotation and vertical stratification, so the turbulence in these simulations is likely to be helical and may also host an $\\alpha$ effect. In motivating our choice of Keplerian rotation law, we mentioned the possible relevance to accretion-disk turbulence, which is believed to be driven by the magnetorotational instability (MRI) \\citep{balbus03}. Given a (weak) large-scale field, the MRI will generate velocity and magnetic-field fluctuations at small scales. These fluctuations, in conjunction with Keplerian rotation and shear, must then amplify the large-scale field to close the loop. The mechanisms for such an ``MRI dynamo'' have been discussed and simulated for some time \\citep[e.g.,][]{brand95,hawley96,fromang07,rincon07b,rincon08,lesur08}. It is tempting to observe in the context of the results reported above that a combination of small-scale turbulence (magnetohydrodynamic turbulence in the case of the MRI) and large-scale shear does indeed appear to work as a dynamo giving rise to a large-scale azimuthal magnetic field ($B_y$ in the shearing sheet model). We note, however, that \\citet{lesur08}, who have analysed this process in detail, find some important differences between what happens in MRI-driven shearing sheet simulations and the forced case studied by us. To conclude, we believe that the discovery of shear dynamo has opened a number of new and exciting avenues of research and produced some promising leads towards unravelling the ways in which cosmic magnetic fields emerge. Further investigations will help assess the range of applicability and relevance of the shear dynamo effect and the physical mechanisms that are responsible for it." }, "0807/0807.1913_arXiv.txt": { "abstract": "Rotation plays a major role in the evolution of massive stars. A revised grid of stellar evolutionary tracks accounting for rotation has recently been released by the Geneva group and implemented into the Starburst99 evolutionary synthesis code. Massive stars are predicted to be hotter and more luminous than previously thought, and the spectral energy distributions of young populations mirror this trend. The hydrogen ionizing continuum in particular increases by a factor of up to 3 in the presence of rotating massive stars. The effects of rotation generally increase towards shorter wavelengths and with decreasing metallicity. Revised relations between star-formation rates and monochromatic luminosities for the new stellar models are presented. ", "introduction": "The star-formation rates of galaxies are commonly determined from the integrated light emitted at a certain wavelengths, such as the $V$ band. Comparison with theoretically predicted mass-to-light ($M/L$) ratios can in principle provide the total stellar mass and, in combination with an appropriate timescale, the star-formation rate. This fundamental methodology goes back to Tinsley's (1980) pioneering work. Although the reasoning is immediately intuitive, two major challenges need to be tackled. The first challenge is the a priori unknown stellar initial mass function (IMF). Observed values of $M/L_{\\rm V}$ in galaxy centers are around 5~--~100. At the same time we know that the main contributors to the galaxy light are upper main-sequence (MS) and evolved low-mass stars. These stars have $M/L_{\\rm V} \\approx 1$. To put these numbers into perspective, the average $M/L_{\\rm V}$ of all known stars within 20~pc of the Sun is about 1~--~2 (Faber \\& Gallagher 1979), and an early-type MS star has $M/L_{\\rm V} \\approx 10^{-2}$. Since the mass-to-light ratio in galaxy centers is not too sensitive to dark matter (at least for disk galaxies; cf. E. Brinks' talk at this conference), the apparent discrepancy suggests that most of the stellar mass is hidden from view because most stars have lower luminosity, and therefore have lower mass than indicated by the spectrum. For the purpose of this paper I will assume we can correct for this effect by assuming a known, universal IMF (Kroupa 2007 and this conference). The second challenge involves the $M/L$ of individual stars of all masses. At the high-mass end, this quantity is not accessible to direct measurements and can only be predicted by stellar evolution models: masses are poorly known because of the scarcity of very massive binaries with mass determinations, and luminosities are elusive because most of the stellar light is emitted in the ionizing ultraviolet. The purpose of this paper is to discuss how the latest generation of stellar evolution models including rotation differs from its predecessor, and how these new models affect the predictions of the evolutionary synthesis code Starburst99 (Leitherer et al. 1999; V\\'azquez \\& Leitherer 2005). Some of the results presented here can be found in V\\'azquez et al. (2007). ", "conclusions": "" }, "0807/0807.4477_arXiv.txt": { "abstract": "Within both dynamical and thermodynamical approaches using the equation of state for neutron-rich nuclear matter constrained by the recent isospin diffusion data from heavy-ion reactions in the same sub-saturation density range as the neutron star crust, the density and pressure at the inner edge separating the liquid core from the solid crust of neutron stars are determined to be $0.040$ fm$^{-3}$ $\\leq \\rho _{t}\\leq 0.065$ fm$^{-3}$ and $0.01$ MeV/fm$^{3}$ $\\leq P_{t}\\leq 0.26$ MeV/fm$^{3}$, respectively. These together with the observed minimum crustal fraction of the total moment of inertia allow us to set a new limit for the radius of the Vela pulsar significantly different from the previous estimate. It is further shown that the widely used parabolic approximation to the equation of state of asymmetric nuclear matter leads systematically to significantly higher core-crust transition densities and pressures, especially with stiffer symmetry energy functionals. ", "introduction": "Having been the major testing grounds of our knowledge on the nature of matter under extreme conditions, neutron stars are among the most mysterious objects in the universe. To understand their structures and properties has long been a very challenging task for both the astrophysics and the nuclear physics community~\\cite{Lat04}. Theoretically, neutron stars are expected to have a solid inner crust surrounding a liquid core. Knowledge on properties of the crust plays an important role in understanding many astrophysical observations~\\cite% {BPS71,BBP71,Pet95a,Pet95b,Lat00,Lat07,Ste05,Lin99,Hor04,Bur06,Owe05}. The inner crust spans the region from the neutron drip-out point to the inner edge separating the solid crust from the homogeneous liquid core. While the neutron drip-out density $\\rho _{out}$ is relatively well determined to be about $4\\times 10^{11}$ g/cm$^{3}$ \\cite{Rus06}, the transition density $% \\rho _{t}$ at the inner edge is still largely uncertain mainly because of our very limited knowledge on the equation of state (EOS), especially the density dependence of the symmetry energy, of neutron-rich nucleonic matter~\\cite{Lat00,Lat07}. These uncertainties have hampered our accurate understanding of many important properties of neutron stars~\\cite{Lat04,Lat00,Lat07}. Recently, significant progress has been made in constraining the EOS of neutron-rich nuclear matter using terrestrial laboratory experiments (See Ref.~\\cite{LCK08} for the most recent review). In particular, the analysis of isospin-diffusion data \\cite{Tsa04,Che05a,LiBA05c} in heavy-ion collisions has constrained tightly the density dependence of the symmetry energy in exactly the same sub-saturation density region around the expected inner edge of neutron star crust. Moreover, the obtained constraint on the symmetry energy was found to agree with isoscaling analyses in heavy-ion collisions \\cite{She07}, the isotopic dependence of the giant monopole resonance in even-A Sn isotopes \\cite{LiT07}, and the neutron-skin thickness of $^{208}$Pb \\cite{Ste05b,LiBA05c,Che05b}. In this paper, using the equation of state for neutron-rich nuclear matter constrained by the recent isospin diffusion data from heavy-ion reactions in the same sub-saturation density range as the neutron star crust, we determine the inner edge of neutron star crusts. Consequently, the limit on the radius of the Vela pulsar is significantly different from the previous estimate. In addition, we find that the widely used parabolic approximation (PA) to the EOS of asymmetric nuclear matter enhances significantly the transition densities and pressures, especially with stiffer symmetry energy functionals. ", "conclusions": "In summary, the density and pressure at the inner edge separating the liquid core from the solid crust of neutron stars are determined to be $0.040$ fm$^{-3}$ $\\leq \\rho _{t}\\leq 0.065$ fm$^{-3}$ and $0.01$ MeV/fm$^{3}$ $\\leq P_{t}\\leq 0.26$ MeV/fm$^{3}$, respectively, using the MDI EOS of neutron-rich nuclear matter constrained by the recent isospin diffusion data from heavy-ion reactions in the same sub-saturation density range as the neutron star crust. These constraints allow us to set a new limit on the radius of the Vela pulsar. It is significantly different from the previous estimate and thus puts a new constraint for the mass-radius relation of neutron stars. Furthermore, we find that the widely used parabolic approximation to the EOS of asymmetric nuclear matter leads systematically to significantly higher core-crust transition densities and pressures, especially for the energy density functional with stiffer symmetry energies. Our results thus indicate that one may introduce a huge error by assuming {\\it a priori} that the EOS is parabolic with respect to isospin asymmetry for a given interaction in locating the inner edge of neutron star crust." }, "0807/0807.3844_arXiv.txt": { "abstract": "Cosmic rays with energies exceeding $10^{17}$~eV are frequently registered by measurements of the fluorescence light emitted by extensive air showers. The main uncertainty for the absolute energy scale of the measured air showers is coming from the fluorescence light yield of electrons in air. The fluorescence light yield has been studied in laboratory experiments. Pioneering measurements between 1954 and 2000 are reviewed. ", "introduction": "\\Label{intro} Cosmic rays with energies exceeding $10^{17}$~eV are frequently observed by measurements of the fluorescence light emitted in air showers. The latter are induced by high energy cosmic rays interacting in the Earth's atmosphere initiating a cascade of secondary particles. Relativistic electrons (and positrons) are the most numerous charged particles in air showers. On their way through the atmosphere they excite nitrogen molecules. The nitrogen molecules release their excitation energy partly through isotropic emission of fluorescence light. The fluorescence light is measured with imaging telescopes in air shower experiments, allowing for a three-dimensional reconstruction of the shower profile in the atmosphere. The main uncertainty of the absolute energy scale of the measured showers is coming from the fluorescence light yield of electrons in air. The bulk of the fluorescence light is induced by electrons with MeV energies. Thus, the emission mechanism can be studied in laboratory experiments using particles, mainly electrons, from radioactive sources or particle accelerators. Particles are injected into volumes of nitrogen and air under well controlled circumstances and the fluorescence light yield is measured as a function of various parameters, like electron energy, gas pressure, temperature, and humidity. Recent developments have been discussed during the 5th Fluorescence Workshop in El Escorial, Spain from September 16th to 20th, 2007. The results are summarized in an accompanying article \\Cite{summary}. The present article summarizes measurements of the fluorescence light yield conducted between 1954 and 2000. The objective is to compile information relevant for air shower observations from the (sometimes hardly accessible) historical papers. Their implications on the contemporary understanding of the subject are discussed in \\Cite{summary}. \\Label{sec:fl_history} ", "conclusions": "" }, "0807/0807.4180_arXiv.txt": { "abstract": "Extrasolar terrestrial planets with the potential to host life might have large obliquities or be subject to strong obliquity variations. We revisit the habitability of oblique planets with an energy balance climate model (EBM) allowing for dynamical transitions to ice-covered snowball states as a result of ice-albedo feedback. Despite the great simplicity of our EBM, it captures reasonably well the seasonal cycle of global energetic fluxes at Earth's surface. It also performs satisfactorily against a full-physics climate model of a highly oblique Earth--like planet, in an unusual regime of circulation dominated by heat transport from the poles to the equator. Climates on oblique terrestrial planets can violate global radiative balance through much of their seasonal cycle, which limits the usefulness of simple radiative equilibrium arguments. High obliquity planets have severe climates, with large amplitude seasonal variations, but they are not necessarily more prone to global snowball transitions than low obliquity planets. We find that terrestrial planets with massive $\\rm CO_2$ atmospheres, typically expected in the outer regions of habitable zones, can also be subject to such dynamical snowball transitions. Some of the snowball climates investigated for $\\rm CO_2$--rich atmospheres experience partial atmospheric collapse. Since long-term $\\rm CO_2$ atmospheric build-up acts as a climatic thermostat for habitable planets, partial $\\rm CO_2$ collapse could limit the habitability of such planets. A terrestrial planet's habitability may thus depend sensitively on its short-term climatic stability. ", "introduction": "\\label{obl_sec:intro} The Earth's obliquity is remarkably stable: the angle between the spin--axis and the normal to the orbital plane varies by no more than a few degrees from its present value of $\\sim 23.5\\degr$. This stability is maintained by torque from the Moon \\citep{laskar_et_al1993,nerondesurgy+laskar1997}. Even within our own Solar System, though, the obliquity of other terrestrial planets has varied significantly more; the analysis of \\citet{laskar+robutel1993} indicates that Mars' obliquity exhibits chaotic variations between $\\sim 0\\degr$ and $\\sim 60\\degr$. How does climate depend on obliquity and its possible variations in time? How does the range of orbital radii around a star at which a planet could support water-based life depend on the planet's obliquity? Has the stability of Earth's obliquity made it a more climatically hospitable home? The answers to these questions will be important to evaluate the fraction of stars that have potentially habitable planets. There are now more than 300 extrasolar planets known,\\footnote{See http://exoplanet.eu/.} several of which are close to the terrestrial regime with masses less than 10 times that of the Earth (e.g., \\citealt{beaulieu_et_al2006}, \\citealt{udry_et_al2007}, \\citealt{bennett_et_al2008}, \\citealt{mayor_et_al2008}). {\\it COROT}, which has already launched, and {\\it Kepler}, scheduled to launch in less than one year, are dedicated space-based transit-detecting observatories that will monitor a large number of stars to detect the small decreases in stellar flux that occur when terrestrial planets cross in front of their host stars \\citep{baglin2003,borucki_et_al2003,borucki_et_al2007}. These missions are expected to multiply by perhaps several hundredfold or more the number of known terrestrial planets, depending on the distribution of such planets around solar-type stars (\\citealt{borucki_et_al2007, borucki_et_al2003, basri_et_al2005}; although see revised predictions in \\citealt{beatty+gaudi2008}). NASA and ESA have plans for ambitious future missions to obtain spectra of nearby Earth-like planets in the hope that they would reveal the first unambiguous signatures of life on a remote world: NASA's {\\it Terrestrial Planet Finder} and ESA's {\\it Darwin} \\citep{leger+herbst2007}. The design of such observatories, and the urgency with which they will be built and deployed, will depend on the habitability potential of terrestrial planets that will be found in the next 5-10 years. Over the last 50 years, various authors have addressed how to predict the way in which terrestrial planet habitability depends on star-planet distance (see \\citealt{kasting+catling2003} for a recent review). Several of the important initial calculations predated the first discoveries of extrasolar planets, including \\citet{dole1964}, \\citet{hart1979}, and the seminal work of \\citet{kasting_et_al1993}. \\citet{selsis_et_al2007}, \\citet{vonbloh_et_al2008}, and \\citet{barnes_et_al2008} have reconsidered habitability in light of recent exoplanetary detections. \\citet{williams_et_al1996}, \\citet[hereafter WK97]{williams+kasting1997} and \\citet{williams+pollard2003} have tackled precisely the questions relating to obliquity posed above, and have concluded that variations in obliquity do not necessarily render a planet non-habitable (see also \\citealt{hunt1982}, \\citealt{williams1988b,williams1988c}, \\citealt{oglesby+ogg1998}, \\citealt{chandler+sohl2000} and \\citealt{jenkins2000,jenkins2001,jenkins2003} in the context of Earth's paleoclimate studies). Here we seek to generalize these analyses to model planets that are less close analogs to Earth than have been considered previously. In \\citet[hereafter SMS08]{spiegel_et_al2008}, we examined how regionally and temporally habitable climates are affected by variations in the efficiency of latitudinal heat transport on a planet, and by variations in the ocean fraction. Importantly, we found that otherwise habitable Earth-like terrestrial planets can be subject to dynamical climate transitions into globally-frozen snowball states. Since it is not trivial to escape a snowball state (e.g., \\citealt{pierrehumbert2005}) and such globally-frozen climates may have profound influences on the development or existence of life (e.g., \\citealt{hoffman+schrag2002}), identifying the likelihood of such transitions on terrestrial exoplanets should be central to any habitability assessment. Following in the footsteps of our first analysis, here we focus on obliquity and consider the influence on habitability of several planetary attributes a priori unknown for exoplanets, such as the efficiency of latitudinal heat transport and the land-ocean distribution. The remainder of this paper is structured as follows: In \\S~\\ref{obl_sec:model} we describe the energy balance climate model we use. In \\S~\\ref{obl_sec:valid} discuss several validation tests in which our model performs well enough to give us some confidence in its behavior for conditions that differ from those found on Earth. In \\S~\\ref{obl_sec:results} we examine the influence on regional and seasonal habitability of various excursions from Earth-like conditions. Finally, we conclude in \\S~\\ref{obl_sec:disc+conc}. ", "conclusions": "\\label{obl_sec:disc+conc} We have presented a series of energy balance models to address the variety of climatic conditions that might exist on oblique terrestrial planets with circular orbits. We considered dynamic climate forcings and responses determined by several planetary attributes a priori unknown for extrasolar planets, including obliquity, rotation rate, distribution of land/ocean coverage, and the detailed nature of the radiative cooling and heating functions. We find that planets with small ocean fractions or polar continents can experience very severe seasonal climatic variations, but that these planets also might maintain seasonally and regionally habitable conditions over a larger range of orbital radii than more Earth-like planets. Climates on high obliquity planets with nonuniform distributions of land and ocean can be far from global radiative balance, as compared to the Earth. Our results provide indications that the modeled climates are somewhat less prone to dynamical snowball transitions at high obliquity. Fast rotating Earth-like planets may fall victim to global glaciation events at closer orbital radii than slower rotating planets. This is also the case for planets with massive CO$_2$ atmospheres, which are expected to be found in the outer orbital range of habitable zones. Snowball transitions could be particularly significant for such planets since partial collapse of their CO$_2$-rich atmospheres may occur and possibly interfere with the thermostatic effect of their carbonate-silicate weathering cycle, thus affecting their long-term habitability." }, "0807/0807.3769_arXiv.txt": { "abstract": "We investigate the effects of weakly-interacting massive particle (WIMP) dark matter annihilation on the formation of Population III.1 stars, which are theorized to form from the collapse of gas cores at the centers of dark matter minihalos. We consider the relative importance of cooling due to baryonic radiative processes and heating due to WIMP annihilation. We analyze the dark matter and gas profiles of several halos formed in cosmological-scale numerical simulations. The heating rate depends sensitively on the dark matter density profile, which we approximate with a power law $\\rho_{\\chi}\\propto r^{-\\alpha_\\chi}$, in the numerically unresolved inner regions of the halo. If we assume a self-similar structure so that $\\alpha_{\\chi}\\simeq 1.5$ as measured on the resolved scales $\\sim 1$~pc, then for a fiducial WIMP mass of 100~GeV, the heating rate is typically much smaller ($<10^{-3}$) than the cooling rate for densities up to $n_{\\rm H}=10^{17}\\:{\\rm cm^{-3}}$. In one case, where $\\alpha_{\\chi}=1.65$, the heating rate becomes similar to the cooling rate by a density of $n_{\\rm H}=10^{15}\\:{\\rm cm^{-3}}$. The dark matter density profile is expected to steepen in the central baryon-dominated region $\\lesssim 1$~pc due to adiabatic contraction, and we observe this effect (though with relatively low resolution) in our numerical models. From these we estimate $\\alpha_{\\chi}\\simeq 2.0$. The heating now dominates cooling above $n_{\\rm H}\\simeq 10^{14}\\:{\\rm cm^{-3}}$, in agreement with the previous study of Spolyar, Freese \\& Gondolo. We expect this leads to the formation of an equilibrium structure with a baryonic and dark matter density distribution exhibiting a flattened central core. Examining such equilibria, we find total luminosities due to WIMP annihilation are relatively constant and $\\sim 10^3\\:L_\\odot$, set by the radiative luminosity of the baryonic core. We discuss the implications for Pop III.1 star formation, particularly the subsequent growth and evolution of the protostar. Even if the initial protostar fails to accumulate any additional dark matter, its contraction to the main sequence could be significantly delayed by WIMP annihilation heating, potentially raising the mass scale of Pop III.1 stars to masses $\\gg100\\:M_\\odot$. ", "introduction": "Population III stars are defined to be those whose formation and evolution are independent of metallicity (McKee \\& Tan 2008; O'Shea et al. 2008), since their metallicity is extremely low: close to or equal to that arising from big bang nucleosynthesis. Population III.1 stars are defined as having their formation be independent of other stars or other astrophysical objects, so that their initial conditions are determined solely by cosmology. These stars will be the first objects to form in a given region of the universe and they are likely to play an important role in cosmic reionization and in laying the foundations for the formation and build-up of galaxies. It is possible that they are the direct or indirect progenitors of supermassive black holes. Within the commonly accepted $\\Lambda$CDM framework, Pop III.1 stars form within dark matter halos. Indeed, in those halos that do form stars, only one star appears to form in the center of each halo (Abel et al. 2002; Bromm, Coppi, \\& Larson 2002). One of the most theoretically well-motivated cold dark matter candidates is a Weakly-Interacting Massive Particle (WIMP). Supersymmetric theories with R-parity naturally provide a stable dark matter candidate which could compose all or part of the dark matter in the Universe. It has been pointed out that, if dark matter consists of a WIMP such as the supersymmetric neutralino, the energy released by the annihilation of these particles could influence early structure formation, star formation and protostellar evolution (Ripamonti, Mapelli \\& Ferrara 2007; Ascasibar 2007; Spolyar, Freese, \\& Gondolo 2008; Iocco 2008; Freese, Spolyar \\& Aguirre 2008; Freese et al. 2008b,c). The effects on stellar evolution at fixed mass have also been investigated (Taoso et al. 2008; Yoon, Iocco \\& Akiyama 2008). Spolyar et al. (2008) show that, for the adiabatically-contracted Navarro, Frenk \\& White (1996) (NFW) dark matter density profiles they considered, dark matter heating can overwhelm gas cooling in the innermost region of a star-forming minihalo, and they propose that this can then lead to a dark matter powered star. In this paper we revisit the scenario investigated by Spolyar et al. (2008). In \\S\\ref{S:analytic} we derive an analytic expression for the dark matter heating rate, including a simplified treatment of radiative transport. In \\S\\ref{S:results} we present our results of the assessment of the importance of WIMP annihilation heating for several halos formed in numerical simulations of cosmic structure formation (O'Shea \\& Norman 2007). We describe the dark matter density structure in \\S\\ref{S:dm} and the properties of the baryons in \\S\\ref{S:baryon}. We compare the WIMP annihilation heating rates and the baryonic cooling rates and discuss the equilibrium structure of dark matter powered protostars in \\S\\ref{S:equilibrium}. We discuss the implications for subsequent protostellar evolution in \\S\\ref{S:protostar}. We conclude in \\S\\ref{S:conclusions}. ", "conclusions": "\\label{S:conclusions} We have investigated the effects of WIMP dark matter annihilation on the formation of Population III.1 stars by analyzing the results of cosmological simulations that follow the gravitational collapse of baryons and dark matter. While these simulations (O'Shea \\& Norman 2007; Yoshida et al. 2006) have followed the baryons to very high densities at scales $\\lesssim 1$~AU, the dark matter is only well-resolved down to scales $\\sim 1$~pc. Thus we have considered various power law extrapolations of the dark matter density profile towards the center. If one assumes the dark matter profile is self-similar, extending inwards from the dark matter dominated regime with $\\rho_{\\chi}\\propto r^{-\\alpha_{\\chi}}$ and $\\alpha_{\\chi}\\simeq 1.5$, then, for a fiducial WIMP mass of 100~GeV, the dark matter annihilation heating is typically negligible ($\\sim10^{-4}$ of the cooling rate). This conclusion would be unchanged for a reduction in the WIMP mass by a factor of 10 or more. One of the simulated minihalos (C) exhibits a slightly steeper density profile in the well-resolved region ($\\alpha_\\chi=1.65$, and in this case WIMP annihilation heating does become important at $n_{\\rm H}\\gtrsim 10^{15}\\:{\\rm cm^{-3}}$. However, there are theoretical reasons to expect that the dark matter density profile will steepen because of adiabatic contraction in the baryon-dominated core. Indeed, this process appears to be occurring in the simulations of O'Shea \\& Norman (2007), although it is not well-resolved. A value of $\\alpha_{\\chi}\\simeq2.0$ appears to be a better description of the dark matter density profile in this region. For such a profile, and again for a 100~GeV particle, the dark matter annihilation heating now exceeds baryonic cooling for densities $n_{\\rm H}>10^{14}\\:{\\rm cm^{-3}}$, in agreement with the previous study of Spolyar et al. (2008). We considered the properties of equilibrium halos in which the density distributions of the baryons and dark matter exhibit a constant density central core. The luminosity that is generated is $\\sim 10^3\\:L_\\odot$ and is relatively invariant, being set by baryonic cooling processes. The sizes of the central cores range from $\\sim 1$ to 40~AU. The detailed effects of this extra heating on the protostellar structure remain to be determined. We expect that subsequent baryonic growth of the protostar will occur more rapidly than its accumulation of dark matter, because the baryons are undergoing rapid free fall collapse followed by disk accretion. However, even if the initial protostar does not gain any additional dark matter, its initial dark matter content in the subsolar mass core could be sufficient to prevent contraction to the zero age main sequence for masses of 100~$M_\\odot$ or greater. Such circumstances could have dramatic implications for the masses of Pop III.1 stars, conceivably raising the mass scale to a regime important for the formation of supermassive black holes. These conclusions depend sensitively on the initial protostellar core, which now needs to be studied with self-consistent cosmological simulations that include the influence of WIMP annihilation heating on the baryons." }, "0807/0807.3133_arXiv.txt": { "abstract": "In this paper, we study the photon-photon pair production optical depth for 10 GeV--1 TeV gamma rays from 3C 279 due to the diffuse radiation of broad-line region (BLR). Assuming a power-law spectrum of $E_{\\gamma}^{-a_2}$ for the photon intensity of very high energy (VHE) gamma rays, $a_1 \\gtrsim 405$ and $a_2\\gtrsim 6.4$ are inferred by the integrated photon fluxes measured by MAGIC and HESS. Based on this power-law spectrum, the pre-absorbed spectra are inferred by correcting the photon-photon absorption on the diffuse photons of the BLR (internal absorption) and the extragalactic background light (external absorption). Position of gamma-ray emitting region $R_{\\rm{\\gamma}}$ determines the relative contributions of this two diffuse radiation to the total absorption for 10 GeV--1 TeV gamma rays. The internal absorption could make spectral shape of gamma rays more complex than only corrected for the external absorption, and could lead to the formation of arbitrary softening and hardening gamma-ray spectra. It should be necessary for the internal absorption to be considered in studying 10 GeV--1 TeV gamma rays from powerful blazars. The energies of annihilated gamma-ray photons due to the internal absorption are likely to be mainly reradiated around GeV. Our results indicate that $R_{\\rm{\\gamma}}$ may be between the inner and outer radii of the BLR for 3C 279. This implies for powerful blazars that $R_{\\rm{\\gamma}}$ might be neither inside the BLR cavity nor outside the BLR, but be within the BLR shell. Observations of $\\it GLAST$, MAGIC, HESS, and VERITAS in the near future could give more constraints on the position of the gamma-ray emitting region relative to the BLR. ", "introduction": "The classical flat spectrum radio quasar (FSRQ) 3C 279 is one of the brightest extragalactic objects in the gamma-ray sky. It was detected by the EGRET, and its spectrum does not show any signature of gamma-ray absorption by pair production up to $\\sim$ 10 GeV (Fichtel et al. 1994; von Montigny et al. 1995). With the detection of high energy gamma rays in 66 blazars, containing 51 FSRQs and 15 BL Lac objects, in the GeV energy range by the EGRET experiment aboard the Compton Gamma Ray Observatory (Catanese et al. 1997; Fichtel et al. 1994; Lin et al. 1997; Mukherjee et al. 1997; Thompson et al. 1995, 1996; Villata et al. 1997; Hartman et al. 1999; Nolan et al. 2003), an exceptional opportunity is presented for the understanding of the central engine operating in blazars. Some of blazars have been also firmly detected by atmospheric Cerenkov telescopes at energies above 1 TeV, such as Mrk 421 (Punch et al. 1992), and Mrk 501 (Quinn et al. 1996). The High Energy Stereoscopic System (HESS) is an imaging atmospheric Cerenkov detector with the energy threshold above 100 GeV (Funk et al. 2004; Hinton 2004; Hofmann 2003). The Very Energetic Radiation Imaging Telescope Array System (VERITAS) provides unprecedented sensitivity to photon energies between 50 GeV and 50 TeV (see e.g., Holder et al. 2006). The Major Atmospheric Gamma Imaging Cerenkov telescope (MAGIC) is currently the largest single-dish Imaging Air Cerenkov Telescope in operation with the lowest energy threshold, $\\sim$ 30 GeV, among the new Cerenkov telescopes (see e.g., Baixeras et al. 2004). At present, 23 active galactic nuclei (AGNs) have been detected in very high energy (VHE) gamma rays, containing 21 BL Lac objects, one radio source M87, and the first FSRQ 3C 279 with the highest redshift $z=0.536$ in these VHE AGNs\\footnote{http://www.mppmu.mpg.de/$\\sim$rwagner/sources/}. The MAGIC telescope detected VHE gamma rays from 3C 279 (Teshima et al. 2007). HESS observations measured an upper limit of integrated photon flux (Aharonian et al. 2008). The Large Area Telescope instrument on the Gamma-Ray Large Area Space Telescope ($\\it GLAST$), the new-generation high energy gamma-ray telescope with sufficient angular resolution to allow identification of a large fraction of their optical counterparts, will observe gamma rays with energies from 20 MeV to greater than 300 GeV, and have the unique capability to detect thousands of gamma-ray blazars to redshifts of at least $z=4$ (see e.g., Chen et al. 2004). $\\it GLAST$ was launched on June 11 2008. $\\it GLAST$, combined with new-generation TeV instruments such as MAGIC, HESS, and VERITAS, will tremendously improve blazar spectra studies, filling in the band from 20 MeV to 10 TeV with high significance data for hundreds of AGNs (see Gehrels \\& Michelson 1999). Future measurements of the gamma-ray spectrum shape and its variability of blazars would tremendously improve our understanding to blazars. These gamma rays from blazars are generally believed to be attributed by emission from a relativistic jet oriented at a small angle to the line of sight (Blandford \\& Rees 1978). These gamma-ray components are contributed by inverse Compton emissions, including synchrotron self-Compton (SSC) scattering synchrotron seed photons, and external Compton (EC) scattering seed photons from sources outside the jet (see e.g., B\\\"ottcher 1999). The diffuse radiation fields of broad-line region (BLR) could have a strong impact on the expected EC spectra of powerful blazars, FSRQs (Liu \\& Bai 2006; Reimer 2007; Tavecchio \\& Ghisellini 2008). The external soft photon fields not only provide target photons for the EC processes to produce these gamma-ray components, but also absorb gamma rays from the EC processes, because gamma rays between 10 GeV and 1 TeV interact with infrared-ultraviolet photons to be attenuated by photon-photon pair production. Many efforts to study the absorption of gamma rays focus on the photon-photon annihilation by the diffuse extragalactic background radiation at the infrared (IR), optical, and ultraviolet (UV) bands (see e.g., Stecker et al. 1992, 2006, 2007; Stecker \\& de Jager 1998; Oh 2001; Renault et al. 2001; Chen et al. 2004; Dwek \\& Krennrich 2005; Schroedter 2005). This external absorption of gamma rays on the diffuse extragalactic background light (EBL) is also proposed and used to probe the EBL (Renault et al. 2001; Chen et al. 2004; Dwek \\& Krennrich 2005; Schroedter 2005). Indeed, the internal absorption of gamma rays inside FSRQs could result in serious problem for the possibility to use the external absorption of gamma rays to probe the IR--optical--UV extragalactic background, because the intrinsic spectra of gamma rays are masked by the internal absorption inside blazars (Donea \\& Protheroe 2003; Liu \\& Bai 2006; Reimer 2007). The intrinsic spectra of gamma rays are complicated by the complex spectra of the diffuse radiation fields of the BLRs in FSRQs (Tavecchio \\& Ghisellini 2008). The mean of intrinsic spectral index is around $2.3$ for 17 BL Lac objects detected in the VHE regime (Wagner 2008). The positions of gamma-ray emitting regions are still an open and controversial issue in the researches on blazars. It is suggested that gamma rays are produced inside the BLRs and the gamma-ray emitting radius $R_{\\gamma}$ ranges roughly between 0.03 and 0.3 pc (Ghisellini \\& Madau 1996). It is argued by Georganopoulos et al. (2001) that the radiative plasma in relativistic jets of powerful blazars are within cavities formed by the BLRs. However, other researchers argued that the gamma-ray emitting regions are outside the BLRs (Lindfors et al. 2005; Sokolov \\& Marscher 2005). In our previous research (Liu \\& Bai 2006, hereafter Paper I), the position of gamma-ray emitting region is a key parameter to determine whether high energy gamma rays could escape the diffuse radiation fields of the BLRs for FSRQs. It is unknown whether these gamma rays could be detected by $\\it GLAST$, MAGIC, HESS, and VERITAS even if blazars intrinsically produce 10 GeV--1 TeV gamma rays, because the gamma-ray emitting radii are unknown. These gamma rays around 200 GeV is optically thick for 3C 279 if the emitting region is within the BLR cavity (see Paper I). However, the MAGIC observations on 2006 February 23 have shown a clear gamma-ray signal in the VHE regime (Teshima et al. 2007). This indicates that the emitting region of these VHE gamma rays should not be inside the BLR cavity, otherwise, the intrinsic flux of these VHE gamma rays is likely to be extremely high. In Paper I, we addressed an important topic in gamma-ray astrophysics, namely the absorption of high energy gamma rays inside of FSRQs by photons of the BLR. In this paper, we attempt to address the particular topic of absorption in the gamma-ray quasar 3C 279 using the available observational data, and its potential effect on the spectra of gamma rays. In order to constrain the position of gamma-ray emitting region in 3C 279, we study the internal absorption of gamma rays by the diffuse radiation from the BLR and its potential effect on the spectra of gamma rays from 10 GeV to 1 TeV. The structure of this paper is as follows. $\\S$ 2 presents intensity of VHE gamma rays. $\\S$ 3 presents theoretical calculation, and consists of two subsections. $\\S$ 3.1 presents calculations of temperature profiles, and $\\S$ 3.2 photon-photon optical depth for 3C 279. $\\S$ 4 presents external absorption on IR--optical--UV extragalactic background. $\\S$ 5 is spectral shape of VHE gamma rays. $\\S$ 6 presents pair spectrum due to photon-photon annihilation and their radiation. $\\S$ 7 is for discussions and conclusions. Throughout this paper, we use a flat cosmology with a deceleration factor $q_0=0.5$ and a Hubble constant $H_0=75 \\/\\ \\rm{km \\/\\ s^{-1} \\/\\ Mpc^{-1}}$. ", "conclusions": "HESS observations in 2007 January measured an upper limit of integrated photon flux $F(E_{\\gamma}>300\\/\\ \\rm{GeV})<3.98\\times 10^{-12}\\/\\ \\rm{cm^{-2}s^{-1}}$ (Aharonian et al. 2008). In calculations, we assumed $F(E_{\\gamma}>300\\/\\ \\rm{GeV})=3.9\\times 10^{-12}\\/\\ \\rm{cm^{-2}s^{-1}}$, allowed by HESS observations. Combining this flux with integrated photon flux $F(E_{\\gamma}>200\\/\\ \\rm{GeV})=3.5\\times 10^{-11}\\/\\ \\rm{cm^{-2}s^{-1}}$ measured by MAGIC (Teshima et al. 2007), a power-law spectrum of $\\frac{dI}{dE_{\\gamma}}=a_1 E_{\\gamma}^{-a_2}$ is inferred with $a_1 =505$ and $a_2 =6.4$. If $F(E_{\\gamma}>300\\/\\ \\rm{GeV})=3.6\\times 10^{-12}\\/\\ \\rm{cm^{-2}s^{-1}}$ is adopted for HESS observations, $a_1 =1510$ and $a_2 =6.6$. If $F(E_{\\gamma}>300\\/\\ \\rm{GeV})=3.98\\times 10^{-12}\\/\\ \\rm{cm^{-2}s^{-1}}$ is adopted, $a_1 =405$ and $a_2 =6.36$. Thus $a_1 \\gtrsim 405$ and $a_2 \\gtrsim 6.4$ are likely to be limited by MAGIC and HESS observations for a VHE gamma-ray spectrum of $\\frac{dI}{dE_{\\gamma}}=a_1 E_{\\gamma}^{-a_2}$. Based on the gamma-ray spectrum of $\\frac{dI}{dE_{\\gamma}}=505 E_{\\gamma}^{-6.4}$ but corrected for the internal and external absorption (see Figs. 2$b$, 3$b$, and 4$b$), photon indices of these corrected VHE gamma rays are around 5 as $R_{\\gamma}=0.1 \\/\\ \\rm{pc}$, 2.0 as $R_{\\gamma}=0.25 \\/\\ \\rm{pc}$, and 1.7 as $R_{\\gamma}=0.4 \\/\\ \\rm{pc}$. These inferred photon indices as $R_{\\gamma}=$ 0.25 and 0.4 $\\rm{pc}$ are inside of those known intrinsic photon indices. If $a_2 =6.6$, photon indices of $1.9\\pm0.03$ are inferred for $R_{\\gamma}=0.4 \\/\\ \\rm{pc}$, $2.2\\pm0.01$ ($a_{\\ast}=0.5$) and $2.1\\pm 0.003$ ($a_{\\ast}=0.998$) for $R_{\\gamma}=0.25 \\/\\ \\rm{pc}$, and $6.0\\pm0.2$ ($a_{\\ast}=0.5$) and $5.1\\pm0.2$ ($a_{\\ast}=0.998$) for $R_{\\gamma}=0.1 \\/\\ \\rm{pc}$. As $R_{\\gamma}$ varies between 0.1 and 0.4 $\\rm{pc}$, photon indices of pre-absorbed VHE gamma-ray spectra are likely to be inside of intrinsic photon index range from 1.3 to 3.6 (Wagner 2008). If one smaller $F(E_{\\gamma}>300\\/\\ \\rm{GeV})=2.0\\times 10^{-12}\\/\\ \\rm{cm^{-2}s^{-1}}$ is adopted, one larger $a_2=8.0$ is obtained. As $a_2 =8.0$, photon indices of $3.3\\pm0.03$ are inferred for $R_{\\gamma}=0.4 \\/\\ \\rm{pc}$, $3.6\\pm0.01$ ($a_{\\ast}=0.5$) and $3.5\\pm 0.003$ ($a_{\\ast}=0.998$) for $R_{\\gamma}=0.25 \\/\\ \\rm{pc}$, and $7.3\\pm0.2$ ($a_{\\ast}=0.5$) and $6.5\\pm0.2$ ($a_{\\ast}=0.998$) for $R_{\\gamma}=0.1 \\/\\ \\rm{pc}$. Thus, larger photon indices of these pre-absorbed spectra are obtained for a fixed $R_{\\gamma}$ as larger $a_2$ is adopted. As $R_{\\gamma}=r_{\\rm{BLR,in}}$, photon indices of these pre-absorbed gamma-ray spectra are always larger than the typical value $\\Gamma_{\\rm{in}}=2.3$ and the maximum $\\Gamma_{\\rm{in}}=3.6$ of those known intrinsic photon indices. Thus, it is unlikely for 3C 279 that $R_{\\gamma}$ be inside the BLR cavity, i.e. it is likely $R_{\\gamma}> r_{\\rm{BLR,in}}$. If $F(E_{\\gamma}>300\\/\\ \\rm{GeV})$ adopted is closer to the upper limit of HESS observations, $a_2$ is not too large. Too large $a_2$, such as $a_2\\simeq 10$ corresponding to $F(E_{\\gamma}>300\\/\\ \\rm{GeV})\\simeq 10^{-12}\\/\\ \\rm{cm^{-2}s^{-1}}$, seems to be impossible for VHE gamma-ray spectra because these spectra with $a_2 \\gtrsim 10$ are much steeper (softer) than the steepest (softest) spectrum measured for PG 1553+113 (Albert et al. 2007). When gamma-ray emitting region is already beyond the BLR, the EC mechanisms, where external photons originate from the BLR, are made inefficient to produce the observed gamma rays. For 3C 279, soft photon energy density around $r_{\\rm{BLR,out}}$ is lower by a factor of 6 to 7 than that around $r_{\\rm{BLR,in}}$ for the BLR diffuse radiation produced by the spherical shell of clouds (see Fig. 1 in Paper I). Thus it is unlikely $R_{\\gamma}>r_{\\rm{BLR,out}}$, i.e. it is likely $R_{\\gamma}\\lesssim r_{\\rm{BLR,out}}$. The external absorption softens the observed spectra relative to the emission ones in the interval from 10 GeV to 1 TeV. Whether the internal absorption softens or hardens the observed spectra relies on gamma-ray emitting radius $R_{\\gamma}$ and energy $E_{\\gamma}$. Photon index variations of gamma-ray spectra relative to $a_2=6.4$ are calculated for the internal absorption, and the internal and external absorption (see Figure 7). As $R_{\\gamma}=r_{\\rm{BLR,in}}$, the local photon indices of gamma-ray spectra decrease with $E_{\\gamma}$ from 10 to 110 GeV, and increase with $E_{\\gamma}$ from 110 GeV to 1 TeV (see Figure 7$a$). The local photon indices of gamma-ray spectra corrected for the internal absorption recover around 270 GeV ($\\Delta \\Gamma(\\rm{in})=0$). The internal absorption make gamma-ray spectra softer and softer from 10 to 110 GeV and from 270 down to 110 GeV, and make gamma-ray spectra harder and harder from 270 GeV to 1 TeV. After corrected for the internal and external absorption, the local photon indices of gamma rays recover around 500 GeV ($\\Delta \\Gamma(\\rm{in+ext})=0$). The external absorption can not change the trend of the local photon indices of gamma-ray spectra corrected for the internal absorption. As $R_{\\gamma}=(r_{\\rm{BLR,in}}+r_{\\rm{BLR,out}})/2$, the local photon indices of gamma-ray spectra corrected for the internal absorption behave as in the case of $R_{\\gamma}=r_{\\rm{BLR,in}}$ (see Figure 7$b$). These local photon indices recover around 400 GeV ($\\Delta \\Gamma(\\rm{in})=0$). The internal absorption make gamma-ray spectra softer and softer from 10 to 110 GeV and from 400 down to 110 GeV, and make gamma-ray spectra harder and harder from 400 GeV to 1 TeV. After corrected for the internal and external absorption, the local photon indices of pre-absorbed gamma-ray spectra basically decrease with $E_{\\gamma}$. The external absorption can not change the trend of the local photon indices of gamma-ray spectra corrected for the internal absorption below 100 GeV, but change that above 100 GeV. Gamma-ray spectra become softer and softer relative to pre-absorbed ones from 10 GeV to 1 TeV ($\\Delta \\Gamma(\\rm{in+ext})<0$). As $R_{\\gamma}=r_{\\rm{BLR,out}}$, the local photon indices have similar trend to that in the case of $R_{\\gamma}=(r_{\\rm{BLR,in}}+r_{\\rm{BLR,out}})/2$ (see Figure 7$c$). The internal absorption make gamma-ray spectra softer and softer from 10 to $\\sim$ 420 GeV and from 1 TeV down to $\\sim$ 420 GeV. After corrected for the internal and external absorption, the local photon indices monotonically decrease with $E_{\\gamma}$, and gamma-ray spectra become softer and softer relative to pre-absorbed ones from 10 GeV to 1 TeV ($\\Delta \\Gamma(\\rm{in+ext})<0$). The external absorption change the trend of the local photon indices of gamma-ray spectra corrected for the internal absorption. The calculated results presented in Figure 7 basically are independent on the particular value of $a_2$. For example, the trends of the local photon indices for $a_2=5.4$ are basically identical to those in the case of $a_2=6.4$. In summary, the internal absorption could make spectral shape more complex than only considering the external absorption, and could lead to the formation of arbitrary softening and hardening gamma-ray spectra (see Figs. 2--4 and 7). Thus, it should be necessary for the internal absorption to be considered in studying gamma rays of 10 GeV--1 TeV from FSRQs. Assuming $\\Gamma_{\\rm{jet}}=15$ for 3C 279, most of gamma rays are contained within a radiation cone with a half open angle of $\\Delta \\varphi\\sim 1/ \\Gamma_{\\rm{jet}} \\sim 3.8^{\\circ}$, because of the relativistic beaming effect. If the central IR--optical--UV photons coming directly from accretion disks travel through the radiation cone, the IR--optical--UV photons can have photon-photon pair creation processes with gamma rays within the radiation cone. The gamma-ray emitting region $R_{\\rm{\\gamma}}\\sim 0.1 \\/\\ \\rm{pc}$ is assumed, and the radii of the UV radiation regions are $R_{D}\\lesssim 30 r_{\\rm{g}}\\sim 0.0004 \\/\\ \\rm{pc}$ (see Fig. 1), the angle between the jet direction and the travelling direction of UV photons at $R_{\\rm{\\gamma}}$ is $\\sim \\arcsin R_{\\rm{D}}/R_{\\rm{\\gamma}}\\lesssim 0.2^{\\circ}$. Then the photons within the radiation cone have the collision angles of $\\theta \\lesssim 4.0^{\\circ}$. For the UV photons at the frequency $\\nu \\simeq10^{16.5} \\/\\ \\rm{Hz}$ with energies of $\\varepsilon_2\\simeq 2.56\\times 10^{-4}$ and the gamma rays of $\\varepsilon_1\\simeq 2.0\\times 10^{6}$ corresponding to energies around 1 TeV, the left of threshold condition (eq. [3] in Paper I) has a upper limit of $\\lesssim 0.6$, which is less than unity, and thus the two kinds of photons cannot be absorbed by the photon-photon pair creation processes. Therefore, the central UV radiation have negligible contributions to the absorption for gamma rays relative to the diffuse radiation from the BLR. For optical photons, the radii of the optical radiation regions are $R_{D}< 200 r_{\\rm{g}}\\sim 0.003 \\/\\ \\rm{pc}$, the angle between the jet direction and the travelling direction of optical photons at $R_{\\rm{\\gamma}}$ is $\\sim \\arcsin R_{\\rm{D}}/R_{\\rm{\\gamma}}\\lesssim 1.7^{\\circ}$. Then the photons within the radiation cone have the collision angles of $\\theta \\lesssim 5.5^{\\circ}$. For the optical photons at the frequency $\\nu \\simeq10^{15} \\/\\ \\rm{Hz}$ with energies of $\\varepsilon_2\\simeq 8.1\\times 10^{-6}$ and the gamma rays of $\\varepsilon_1\\simeq 2.0\\times 10^{6}$, the left of threshold condition has a upper limit of $\\lesssim 0.04$, which is less than unity, and thus the two kinds of photons cannot be absorbed by the photon-photon pair creation processes. Therefore, the central optical radiation have negligible contributions to the absorption for gamma rays relative to the diffuse radiation from the BLR. For IR photons, the central radiation also have negligible contributions to the absorption for gamma rays. Thus, the central IR--optical--UV radiation have negligible contributions to the absorption for gamma rays relative to the diffuse radiation from the BLR. Absorption for gamma rays by photon-photon annihilation and where the energies carried by the annihilated gamma rays reradiate are important to gamma-ray research. The intense creation of pairs would produce a strong radiation at low energy X-rays or at GeV energy (Protheroe \\& Stanev 1993; Saug\\'e \\& Henri 2004; Zdziarski \\& Coppi 1991). The electron-positron pair cascade could cause the soft X-ray excesses (Zdziarski \\& Coppi 1991). The produced electron-positron pairs could make a difference around 1 GeV, where the pileup of photons below the absorption threshold results in a significant flattening in the observed spectrum relative to the emission spectrum (Protheroe \\& Stanev 1993). In $\\S$ 6, we studied the pair spectrum due to photon-photon absorption, and the synchrotron and EC spectra emitted by the equilibrium pair distribution of steady state. The synchrotron radiation peaks around keV X-rays, and the EC radiation peaks around GeV gamma rays (see Fig. 6$b$). Thus, pairs due to annihilation absorption for gamma rays by the diffuse radiation fields of the BLR are likely to make a difference around 1 GeV for 3C 279. In this paper, in order to limit the gamma-ray emitting radius $R_{\\gamma}$, we used a BLR model to study the photon-photon absorption by the diffuse radiation of the BLR in 3C 279 for gamma rays of 10 GeV to 1 TeV in the observed spectrum. We calculated the internal absorption of gamma rays from 10 GeV to 1 TeV for $R_{\\gamma}=r_{\\rm{BLR,in}}$, $r_{\\rm{BLR,out}}$, and $(r_{\\rm{BLR,in}}+r_{\\rm{BLR,out}})/2$ (see Figs. 2$a$, 3$a$, and 4$a$). For a fixed $R_{\\gamma}$, dependence of photon-photon absorption optical depth $\\tau_{\\gamma\\gamma}$ on energies of gamma rays $E_{\\gamma}$ relies on $R_{\\gamma}$. Dependence of $\\tau_{\\gamma\\gamma}$ on $R_{\\gamma}$ was also studied for a fixed $E_{\\gamma}$ (see Fig. 5). For a fixed $E_{\\gamma}$, $\\tau_{\\gamma\\gamma}$ decreases with increasing $R_{\\gamma}$. The external absorption on the IR--optical--UV EBL was also estimated for gamma rays of 10 GeV--1 TeV, and it monotonically increases as $E_{\\gamma}$ increases. Comparing the internal absorption with the external one shows that $R_{\\gamma}$ determines the relative contributions of the internal and external absorption to the total photon-photon annihilation absorption of 10 GeV--1 TeV gamma rays (see Figs. 2$a$, 3$a$, and 4$a$). Based on MAGIC and HESS observations, a power-law spectrum as in equation (1) was adopted for the photon intensity of VHE gamma rays. The pre-absorbed gamma-ray spectra are inferred by this power-law corrected for the internal and external absorption. The internal absorption could make spectral shape of gamma rays more complex than that only corrected for the external absorption, and could lead to the formation of arbitrary softening and hardening gamma-ray spectra (see Figs. 2$a$, 3$a$, 4$a$, and 7). Thus, it should be necessary for the internal absorption to be considered in studying 10 GeV--1 TeV gamma rays from FSRQs. $R_{\\gamma}$ significantly influences the variations of spectral shape due to the internal absorption. Calculations imply that the energies of annihilated gamma rays due to the internal absorption are mainly reradiated around GeV regime (see Fig. 6$b$). Considering the possible variations of photon index $a_2$, the photon indices of the pre-absorbed VHE gamma-ray spectra were compared with those known intrinsic photon indices. As $R_{\\gamma}=r_{\\rm{BLR,in}}$ and $a_2\\gtrsim 6.4$, the photon indices of the pre-absorbed gamma-ray spectra are always larger than the typical value $\\Gamma_{\\rm{in}}=2.3$ and the maximum $\\Gamma_{\\rm{in}}=3.6$ of those known intrinsic photon indices, and $\\tau_{\\gamma\\gamma}$ is larger than unity. Thus, it is likely $R_{\\gamma}> r_{\\rm{BLR,in}}$ for 3C 279. As $R_{\\gamma}=r_{\\rm{BLR,out}}$ and $6.4 \\lesssim a_2 \\lesssim 8.3$, photon indices of pre-absorbed gamma-ray spectra are not larger than $\\Gamma_{\\rm{in}}=3.6$. For a fixed $a_2$, photon indices of pre-absorbed gamma-ray spectra decrease as $R_{\\gamma}$ increases. Too large $a_2$ seems to be impossible for VHE gamma-ray spectra. In addition, the EC processes may be inefficient to produce the observed gamma rays as gamma-ray emitting region is already beyond the BLR. Thus, it is likely $R_{\\gamma}\\lesssim r_{\\rm{BLR,out}}$ for 3C 279. Our results suggest that $R_{\\gamma}$ for powerful blazars might be neither inside the BLR cavity nor outside the BLR, but be within the BLR shell. This is neither consistent with suggestions of Ghisellini \\& Madau (1996) and Georganopoulos et al. (2001) nor consistent with suggestions of Lindfors et al. (2005) and Sokolov \\& Marscher (2005). Our results are model dependent, especially dependent on the assumed power-law spectrum for the VHE gamma rays. If Teshima et al. published the spectral indices of gamma-ray spectra measured by MAGIC (Teshima et al. 2007), the power-law spectrum assumed in this paper could be tested. Tavecchio \\& Ghisellini (2008) used the photoionization code CLOUDY, described by Ferland et al. (1998), to calculate the detailed spectra from the BLRs for powerful blazars, and then used these spectra to calculate the EC spectra. Approximate spectra of the BLRs are used in this paper and Paper I, and another approximate spectra are used by Reimer (2007). Difference between the detailed and approximate spectra should exist. It should be useful in the future researches to study the effects of this difference on the results of previous researches using approximate spectra (e.g., Liu \\& Bai 2006; Reimer 2007). Observations of $\\it GLAST$, MAGIC, HESS, and VERITAS in the near future could give more observational constraints on the gamma-ray emitting regions and the BLRs for the powerful blazars. Publications of intrinsic photon indices predicted by theoretical researches and photon indices measured by observations in the VHE regime could give stronger constraints on $R_{\\gamma}$." }, "0807/0807.4844_arXiv.txt": { "abstract": "The Tautenburg Exoplanet Search Telescope (TEST) is a robotic telescope system. The telescope uses a folded Schmidt Camera with a 300mm main mirror. The focal length is 940mm and it gives a $2.2^{\\circ}$ x $2.2^{\\circ}$ field of view. Dome, mount, and CCD cameras are controlled by a software bundle made by Software Bisque. The automation of the telescope includes selection of the night observing program from a given framework, taking darks and skyflats, field identification, guiding, data taking, and archiving. For the search for transiting exoplanets and variable stars an automated psf photometry based on IRAF and a lightcurve analysis based on ESO-Midas are conducted. The images and the results are managed using a PostgreSQL database. ", "introduction": "In 1999 the first transiting extrasolar planet was discovered (\\cite[Charbonneau et al. 2000]{Charbonneau_etal00}). As the knowledge of transit parameters allows an unique insight into the nature of the planet, many efforts have been made to increase the number of known transiting planets. Until now about 50 transiting extrasolar planets are discovered. Inspired by the Berlin Exoplanet Search Telescope (BEST) (\\cite[Rauer et al. 2004]{Rauer_etal04}), which was operated by the Deutsches Zentrum f{\\\"u}r Luft- und Raumfahrt at the Th{\\\"u}ringer Landessternwarte Tautenburg in the years 2001~-~2003, a new small aperture telescope has now been installed at this site to search for transiting exoplanets. ", "conclusions": "" }, "0807/0807.1185_arXiv.txt": { "abstract": "The Extreme ultraviolet Imaging Spectrometer (EIS) on board \\hinode\\ is the first solar telescope to obtain wide slit spectral images that can be used for detecting Doppler flows in transition region and coronal lines on the Sun and to relate them to their surrounding small scale dynamics. We select EIS lines covering the temperature range $6\\times10^4$~K to $2\\times10^6$~K that give spectrally pure images of the Sun with the 40\\arcsec\\ slit. In these images Doppler shifts are seen as horizontal brightenings. Inside the image it is difficult to distinguish shifts from horizontal structures but emission beyond the image edge can be unambiguously identified as a line shift in several lines separated from others on their blue or red side by more than the width of the spectrometer slit (40 pixels). In the blue wing of \\HeII, we find a large number of events with properties (size and lifetime) similar to the well-studied explosive events seen in the ultraviolet spectral range. Comparison with X-Ray Telescope (XRT) images shows many Doppler shift events at the footpoints of small X-ray loops. The most spectacular event observed showed a strong blue shift in transition region and lower corona lines from a small X-ray spot that lasted less than 7~min. The emission appears to be near a cool coronal loop connecting an X-ray bright point to an adjacent region of quiet Sun. The width of the emission implies a line-of-sight velocity of 220~\\kms. In addition, we show an example of an \\FeXV\\ shift with a velocity about 120~\\kms, coming from what looks like a narrow loop leg connecting a small X-ray brightening to a larger region of X-ray emission. ", "introduction": "The Extreme ultraviolet Imaging Spectrometer \\citep[EIS;][]{Culhane07} on \\hinode\\ obtains images and spectra of many transition region and coronal lines in the wavelength ranges $170-211$~\\AA\\ and $246-292$~\\AA. The EIS wide slits are an interesting compromise between narrow slit spectra where the time to raster across a particular structure is often longer than the lifetime of the structure and filter images which give no direct measurement of flow velocities. The 40\\arcsec\\ wide slit provides overlapping spectra from the observed 40\\arcsec\\ wide region of the Sun. Images in different lines overlap if the lines are separated by less than the 40 pixel width of the slit image times the wavelength plate scale, 0.0223 \\AA/pixel. Figure~\\ref{slots}, shows a 40\\arcsec\\ wide slit spectrum of a small active region for the wavelength region $268-292$~\\AA. Isolated lines produce well-defined 40\\arcsec\\ wide images. Within the EIS spectrum there are several relatively strong isolated lines that can be used to study the dynamics of the outer atmosphere. \\citet{Hansteen07} report on rapid temporal variations in quiet Sun \\HeII\\ 256~\\AA\\ and \\FeXII\\ 195~\\AA\\ wide slit emission features observed with a cadence of 30~s. They also mention the detection of many \\HeII\\ blue shifts in quiet Sun narrow slit rasters observed after the wide slit sequence. In this paper we show how the detection of \\HeII\\ blue shifts and large-scale temporal variations, can be made simultaneously with wide slit observations. The key is that the emission beyond the edge of the 40\\arcsec\\ main line image is either line broadening/Doppler shifts from the main line or emission in a neighbouring line. Figure~\\ref{slots} demonstrates both effects. \\FeXV\\ and \\FeXIV\\ (labelled) are both well separated lines and the images are basically straight along both edges. On the edge that cuts the active region, the emission from both lines bulges slightly because the bright active region lines are broader. A closer inspection of the emission shows that both lines have the same structure along the edge. On the other hand, the strong unmarked line on the left is mainly \\FeXIV\\ 270.5~\\AA, but there is clearly a blend producing extensions on the left hand side. Positions of the projections coincide with bright \\MgVII\\ and \\SiVII\\ because the blend is the lower coronal line \\MgVI\\ 270.4~\\AA. \\begin{figure} \\centering \\includegraphics[width=12 cm]{slots.eps} \\caption{A 40\\arcsec\\ wide slit spectrum of a small active region, showing well separated and blended lines. Several of the lines discussed in this paper are marked.} \\label{slots} \\end{figure} When the Sun's structure is known from images in isolated lines, it is possible to work out from the spectrum whether neighbouring lines are expected to produce emission. If there is no spectral line that can cause the emission, it is very likely due to Doppler shifts/broadening in the image line. Thus the EIS 40\\arcsec\\ wide slit can be used to obtain both images of the dynamics over a wide range of temperatures, and Doppler shifts from structures on the edge of the images. Inside the image, Doppler broadened lines appear as horizontal streaks which are difficult to distinguish from a jet-like structure. Comparison with filter images is very important in order to positively identify shifts inside the image. Simultaneous images obtained by other instruments on \\hinode\\ can, in principle, be used to help identify shifts but there will always be some uncertainty because the filters to not detect exactly the same plasma. In this letter we show examples of Doppler shift events seen in the quiet Sun with the EIS 40\\arcsec\\ wide slit. Comparisons are made with X-ray images obtained by the X-Ray Telescope \\citep[XRT;][]{Golub07} in order to show the relationship to hot loop structures. Further work will report on detailed analyses with data from the Solar Optical Telescope \\citep[SOT;][]{Tsuneta08} on \\hinode\\ and the Solar Ultraviolet Measurements of Emitted Radiation \\citep[SUMER;][]{Wetal95} spectrograph on SoHO. ", "conclusions": "There is a wealth of information in the EIS 40\\arcsec\\ wide slit images. The primary use of wide slit images is to see rapid time variations of structures in specific spectral lines \\citep{Hansteen07}, typically with a cadence 20 times (the slit width) faster than rastering with the 2~\\arcsec\\ slit. After passing through the slit, the light is dispersed producing rows of overlapping spectral images. If a Doppler shift occurs in the center of the image, this is indistinguishable from a horizontal structure. It is therefore beneficial to have simultaneous filtergrams to show the underlying structures. Here we show how the wide slit can be used to detect Doppler shifts at the edge of the image. We have highlighted three types of Doppler shift event seen in EIS wide slit spectra: explosive events in \\HeII\\ at the image edge; a fast jet related to an X-ray spot brightening; and \\FeXV\\ flow connecting an X-ray spot brightening to a larger X-ray region. The latter observations were done in sit-and-stare at two positions spaced 40\\arcsec\\ apart in order to pick up first the blue and then the red wings of the lines. At present our understanding of the line profile at the edge of the slit is not sufficiently well understood to determine accurate line shifts but flows producing emission more than two pixels away from the average line profile at the edge are clearly detectable. This suggests new ways of investigating, for example, explosive event dynamics because one can see the loop motion and heating beyond the position of the explosive event. It could also be useful when investigating Doppler shift fluctuations from hot loops because narrow slit observations in sit-and-stare are not able to distinguish real intensity fluctuations at a point from movement of loops into and out of the slit field-of-view \\citep[cf][]{Wetal03b, Mariska07}. The aim of this short paper is to show the possibilities for future studies, not to present event details and their scientific implications. Discussions on the events themselves, including detailed analyses from SOT on \\hinode\\, and SUMER on SOHO will be given in more extensive papers." }, "0807/0807.3239_arXiv.txt": { "abstract": "We discuss a new criterion to estimate the mass in the outer, non-equilibrium region of galaxy clusters, where the galaxy dynamics is dominated by an overall infall motion towards the cluster centre. In the framework of the spherical infall model the local mean velocity of the infalling galaxies at every radius provides information about the integrated matter overdensity $\\delta$. Thus, a well-defined value of the overdensity $\\delta_t$ is expected at the turnaround radius $r_t$, i.e. the radius where the Hubble flow balances the infall motion. Within this scenario, we analysed the kinematical properties of a large catalogue of simulated clusters, using both dark matter particles and member galaxies as tracer of the infall motion. We also compared the simulation with analytical calculation performed in the spherical infall approximation, to analyze the dependence of the results on cosmology in spatially flat universe. If we normalize cluster mass profiles by means of the turnaround mass $M_t$ (i.e. the mass within $r_t$), they are consistent with an exponential profile in the whole non-equilibrium region ($0.5\\la r/r_t\\la 2$). Turnaround radii are proportional to virialization radii ($r_t\\simeq 3.5 r_v$), while turnaround masses are proportional to virialization masses, i.e. $M_t\\simeq 1.7 M_v$, where $M_v$ is the mass within $r_v$. Actually, the mass evaluated within the turnaround radius is a more exhaustive evaluation of the total mass of the cluster. These results can be applied to the analysis of observed clusters. ", "introduction": "The gravitational collapse of galaxies towards the centre of clusters is usually described within the framework of the spherical infall model, as the motion of a set of concentric, spherically symmetrical mass shells \\citep{GG,Si,Sc}. Actually the spherical infall model is widely accepted in literature, since it describes fairly well the dynamics of the non-equilibrium region of galaxy clusters, defined as the region where the effects of virialization and the crossing of the above-mentioned shells are negligible and some overall infall motion of member galaxies is recognizable. Under the spherical symmetry assumption, the infall motion produces a pattern of caustic surfaces in the galaxy redshift-space distribution (which is obtained representing the line-of-sight velocities of galaxies \\emph{versus} their projected position on the sky plane). These caustics envelop all galaxies whose infall motion overwhelms the Hubble flow \\citep{K}. Caustics with a characteristical ``trumpet'' shape were actually observed in the redshift-space distribution of clusters \\citep{Oal}. Diaferio \\& Geller \\citep{DG} and Diaferio \\citep{D} showed that the caustic amplitude provides a direct measure of the escape velocity of galaxies, and therefore allows to estimate the mass profile of the cluster in the innermost part of the non-equilibrium region, up to the turnaround radius $r_t$ (i.e. the radial distance where the velocity of the infall motion is equal to the Hubble flow velocity.). The caustic technique was applied to the observation of many local clusters \\citep{GDK,Ral1,Ral2,Ral3}. These mass estimates are consistent with those based on virial theorem \\citep{Giral,BivGir} and weak lensing observations (\\citealt{DGR}, and references therein). In fact, up to now the sampled volumes were always restricted within the turnaround radius, and this is due to the definition of caustics surfaces \\citep{RG}. In this paper, we discuss an approach to the issue of mass estimation, which can be applied to larger sampled volumes, well beyond the turnaround radius. We use the radial velocity of galaxies as the key quantity, instead of the escape velocity, as in the caustic technique. According to Silk \\citep{Si}, Peebles \\citep{P1,P2}, and Gunn \\citep{G}, within the spherical symmetry hypothesis, the velocity of the matter infall motion at a certain distance from the centre depends on the encompassed mass. Our purpose is to use this dependence to constrain the value of the overdensity at the turnaround radius. In fact, we see that the turnaround radius is far outside the virialization core of clusters, and is therefore a suitable normalization scale for the cluster mass profile in the non-equilibrium region \\citep{VH}. To test our assumptions and to verify the results, we will analys a large galaxy population extracted from a simulated cluster catalogue \\citep{Borgal,Bivial}. We will study all clusters both as a whole and one by one. We will prove that the actual turnaround overdensity of clusters is in good agreement with the predictions of the spherical infall model. Moreover, we will show that the normalized mass profiles are generally consistent with a power-law profile, which extends the standard Navarro--Frenk--White profile \\citep[hereafter NFW]{NFW1,NFW2,NFW3} to the non-equilibrium region. In Section \\ref{sec:model} we present the details of our model, concerning the theoretical framework (\\ref{sec:theory}) and the simulated data sample (\\ref{sec:data}). In Section \\ref{sec:results} we discuss the results of our analysis, focusing on the mass estimation at the turnaround radius (\\ref{sec:turn}) and in the whole non-equilibrium region, up to 8 virialization radii (\\ref{sec:prof}). Finally, in Section \\ref{sec:concl} we draw the conclusions of our work. ", "conclusions": "\\label{sec:concl} We analized a large sample of simulated galaxy clusters in order to reconstruct the mass profile in the non-equilibrium region, where the galaxy dynamics is dominated by an overall infall motion towards the cluster centre. Within the assumptions of the spherical infall model, the turnaround overdensity $\\delta_t$ can be theoretically computed as a function of only the matter density parameter $\\Omega_0$, assuming a spatially flat universe. We obtained the overdensity $\\delta_t\\simeq 6-15$, depending on the infall velocity profile we adopted. We interpolated the infall velocity profile of member galaxies extracted from the simulated clusters of our catalogue, and we showed that: \\begin{enumerate} \\p The turnaround radius $r_t$ can be quite well approximated by a multiple of the virialization radius $r_v$: $r_t\\simeq 3.5r_v$; \\p The turnaround overdensity $\\delta_t$ is consistent with the prediction of the spherical infall model, as long as the infall velocity profile is described by the Meiksin approximation \\citep{VD}. \\end{enumerate} Points (i) and (ii) are in agreement with \\citet{VH} and \\citet{RG} and imply a proportionality between the turnaround mass $M_t$ and the virialization mass $M_v$: $M_t\\simeq 1.7 M_v$. Moreover, $M_t$ turns out to depend on the 3-d DM velocity dispersion within the virialization core $\\sigma_{v,\\rmn{DM}}$ approximately in the form of a cubic relation. The turnaround values can be assumed as a suitable normalization scale for the mass profiles in the non-equilibrium region of clusters. We showed that the normalized mass profiles are generally consistent with a cosmic profile, which can be described (for $0.5\\la r/r_t\\la 2$) by: \\be\\label{eq:M_concl} M(r)\\simeq M_t\\exp\\left[\\f{0.6}{\\Omega_0^{1/4}}\\left(\\f{r}{r_t}-1\\right)\\right]. \\ee While in the inner, relaxed or almost-relaxed regions the mass can be considered independent on cosmological parameters, in the outer regions a dependence on $\\Omega_0$ (even if small) has to be taken, at least in principle, into account. We used a synthetic cluster, obtained by summing all catalogue clusters, to determine a robust estimate of $r_t$ and $\\delta_t$. If we assume this values, our model is able to predict the mass profile in the non-equilibrium region for about $80\\%$ of clusters. So, it is possible to speak about a mass profile even in the region where mass accretion takes places along isolated radial filaments rather than in a spherically symmetric way. Our model may be useful in observational analysis in order to estimate the total mass of clusters using the redshift-space distribution of galaxies. The method is the following: \\begin{enumerate} \\item one estimates the virialization radius and the virialization mass from the galaxy velocity dispersion in the cluster core; \\item using equation (\\ref{eq:st}) and equation (\\ref{eq:Mt}) one computes the turnaround radius $r_t$ and the turnaround mass of the cluster $M_t$; \\item once the turnaround radius and the turnaround mass are known, one can estimate the mass profile in the non-equilibrium region using the exponential law in equation (\\ref{eq:M_concl}). \\end{enumerate} The advantage of this approach lies in the possibility to estimate the mass profiles up to the far outskirts of clusters, where the caustic pattern is not generally recognizable \\citep{DG,D}; up to now, these cluster outer volumes have been usually neglected in the evaluation of cluster total masses. Actually, the turnaround mass is a more exhaustive evaluation of the total mass of the cluster. The steps leading to it consist in the abovementioned points (i), (ii), and (iii), and are expected to be applied to observed clusters." }, "0807/0807.2523_arXiv.txt": { "abstract": "A dynamical analysis of an effective homogeneous and irrotational Weyssenhoff fluid in general relativity is performed using the $1+3$ covariant approach that enables the dynamics of the fluid to be determined without assuming any particular form for the space-time metric. The spin contributions to the field equations produce a bounce that averts an initial singularity, provided that the spin density exceeds the rate of shear. At later times, when the spin contribution can be neglected, a Weyssenhoff fluid reduces to a standard cosmological fluid in general relativity. Numerical solutions for the time evolution of the generalised scale factor $R(t)$ in spatially-curved models are presented, some of which exhibit eternal oscillatory behaviour without any singularities. In spatially-flat models, analytical solutions for particular values of the equation-of-state parameter are derived. Although the scale factor of a Weyssenhoff fluid generically has a positive temporal curvature near a bounce (i.e. $\\ddot{R}(t)>0$), it requires unreasonable fine tuning of the equation-of-state parameter to produce a sufficiently extended period of inflation to fit the current observational data. ", "introduction": "The Einstein-Cartan (EC) theory of gravity is an extension of Einstein's theory of general relativity (GR) that includes the spin properties of matter and their influence on the geometrical structure of space-time ($\\cite{Cartan:1922}$; see also $\\cite{Kleinert:1989}$, $\\cite{Kleinert:2008}$). In GR, the energy-momentum of the matter content is assumed to be the source of curvature of a Riemannian space-time manifold $V_4$. In the EC theory, the spin of the matter has been postulated, in addition, to be the source of torsion of a Riemann-Cartan space-time manifold $U_4$ $\\cite{Hehl:1973}$. Weyssenhoff and Raabe $\\cite{Weyssenhoff:1947}$ were the first to study the behaviour of perfect fluids with spin. Obukhov and Korotky extended their work in order to build cosmological models based on the EC theory $\\cite{Obukhov:1987}$ and showed that by assuming the Frenkel condition$\\footnote{Note that the Frenkel condition arises naturally when performing a rigorous variation of the action. It simply means that the spin pseudovector is spacelike in the fluid rest frame.}$ the theory may be described by an effective fluid in GR where the effective stress-energy momentum tensor contains some additional spin terms. The aim of this publication is two-fold. First, we wish to investigate the possibility that the spin contributions for a Weyssenhoff fluid may avert an initial singularity, as first suggested by Trautman $\\cite{Trautman:1973}$. Second, since any realistic cosmological model has to include an inflation phase to fit the current observational data, it is also of particular interest to see if the spin contributions are able to generate a dynamical model endowed with an early inflationary era, as first suggested by Gasperini $\\cite{Gasperini:1986}$. Scalars fields can generate inflation, but they have not yet been observed. Therefore, it is of interest to examine possible alternatives, such as a Weyssenhoff fluid. In contrast to the approaches of Trautman $\\cite{Trautman:1973}$ and Gasperini $\\cite{Gasperini:1986}$, our use of the $1+3$ covariant formalism enables us to determine the dynamics of a Weyssenhoff fluid without assuming any particular form for the space-time metric. The study of the dynamics of a Weyssenhoff fluid in a $1+3$ covariant approach was initiated by Palle $\\cite{Palle:1998}$. His work has been revised and extended in our previous publication $\\cite{Brechet:2007a}$. The present paper builds on $\\cite{Brechet:2007a}$ to extend the work carried out first by Trautman $\\cite{Trautman:1973}$ in an isotropic space-time, and Kopczynski $\\cite{Kopczynski:1973}$ and Stewart $\\cite{Stewart:1973}$ in an anisotropic space-time. It also generalises the analysis of the inflationary behaviour of Weyssenhoff fluid models made by Gasperini $\\cite{Gasperini:1986}$ to anisotropic space-times. In our dynamical analysis, we choose to restrict our study to a spatially homogeneous and irrotational Weyssenhoff fluid. This particular choice, which implies a vanishing vorticity and peculiar acceleration, has been motivated by underlying fundamental physical reasons. For a vanishing vorticity, the fluid flow is hypersurface-orthogonal, which means that the instantaneous rest spaces defined at each space-time point should mesh together to form a set of 3-surfaces in space-time $\\cite{Ellis:1971}$. These hypersurfaces, which are surfaces of simultaneity for all the fluid observers, define a global cosmic time coordinate determined by the fluid flow. Moreover, by assuming that any peculiar acceleration vanishes, the cosmic time is then uniquely defined. It is worth mentioning that the absence of vorticity is an involutive property, which means that if it is true initially then it will remain so at later times as shown by Ellis et al $\\cite{Ellis:1998}$. Finally, the assumption that there is no vorticity on all scales implies that the fluid has no global rotation. This is in line with recent Bayesian MCMC analysis of WMAP data performed by Bridges et {\\it al.} $\\cite{Bridges:2006}$. Their work confirms that a physical Bianchi $\\mathrm{VII_h}$ model, which has a non-vanishing vorticity, is statistically disfavored by the data. It is worth pointing out that Szydlowski and Krawiec $\\cite{Szydlowski:2004}$ have considered an isotropic and homogeneous cosmological model in which a Weyssenhoff fluid is proposed as a potential candidate to describe dark energy at late times. In a subsequent publication $\\cite{Krawiec:2005}$, the authors showed that it is not disfavoured by SNIa data, but it may be in conflict with CMB and BBN observational constraints. By contrast, in this paper, we consider the full evolutionary history of an, in general, anisotropic universe with a Weyssenhoff fluid as its matter source, concentrating in particular on the `early universe' behaviour when the spin terms are significant. Indeed, at late times, when the spin contributions can be neglected, the Weyssenhoff fluid reduces to a standard cosmological fluid. We thus allow for the presence of a non-zero cosmological constant, in accord with current observational constraints. In $\\Sref{Weyssenhoff fluid description}$, we give a concise description of a Weyssenhoff fluid using a 1+3 covariant approach outlined in {\\it Appendix} A. The spatial symmetries and macroscopic spin averaging procedure are discussed in $\\Sref{Spatial symmetries and macroscopic spin averaging}$. In $\\Sref{Dymamics of a homogeneous and irrotational Weyssenhoff fluid}$, we establish the relevant dynamical relations for a homogeneous and irrotational Weyssenhoff fluid. In $\\Sref{Geodesic singularity analysis}$, we perform a geodesic singularity analysis for such a fluid. In $\\Sref{Dynamical evolution: general considerations}$, we analyse the fluid dynamics. The behaviour of the generalised scale factor $R(t)$ of such a fluid in a spatially-curved models is discussed in $\\Sref{Quantitative dynamical evolution of spatially-curved models}$ and explicit analytical solutions in spatially-flat models are given in $\\Sref{homogeneous solutions}$ . For the reader's convenience, certain main results obtained in our earlier work $\\cite{Brechet:2007a}$ will be repeated in the case of a homogeneous and irrotational Weyssenhoff fluid in $\\Sref{Weyssenhoff fluid description}$ and $\\Sref{Dymamics of a homogeneous and irrotational Weyssenhoff fluid}$. In this paper, we use the $(+,-,-,-)$ signature. To express our results in the opposite signature used by Ellis $\\cite{Ellis:1998}$, the correspondence between physical variables can be found in $\\cite{Brechet:2007a}$. ", "conclusions": "We have used the $1+3$ covariant approach to perform a dynamical analysis of an effective homogeneous and irrotational Weyssenhoff fluid. Contrary to the case of a perfect fluid in GR, the effective spin contributions to the fluid dynamics act like centrifugal forces preventing the formation of singularities for isotropic and anisotropic models satisfying the spin-shear constraint $\\eref{cond spin shear}$. The temporal evolution of the models is symmetric with respect to $t=0$. In a cosmological context, the energy density at the bounce state $\\rho_0$ has to be sufficiently dense in order to seed large scale structures from primordial quantum fluctuations. For cosmological parameters which are consistent with current cosmological data $\\eref{lambda condition}$ $\\eref{curvature condition}$, the temporal curvature of scale factor of a Weyssenhoff fluid is positively defined near the bounce $\\eref{a}$. However such a fluid is not a suitable candidate for inflation given that the only way to include an inflation phase of about $50-70$ e-folds, is by considering a fluid with a very fine-tuned equation-of-state $\\eref{State tunning}$, which does not reduce to the standard cosmological fluid at later times. It is worth emphasizing that the time evolution of the scale factor of a homogeneous and irrotational Weyssenhoff fluid exhibits eternal oscillations, without any singularities. By contrast, the corresponding solutions obtained for a perfect fluid in GR are cycloids, which do exhibit singularities. Hence, the absence of singularities for a specific range of parameters is a genuinely new feature of cosmological models based on a Weyssenhoff fluid. \\ack S~D~B thanks the Isaac Newton Studentship and the Sunburst Fund for their support. The authors also thank Reece Heineke for giving an insightful talk entitled ``Inflation via Einstein-Cartan theory\", and John M. Stewart for useful discussions. \\appendix" }, "0807/0807.2653_arXiv.txt": { "abstract": "{ We employ a long \\xmmn observation of the core of the Perseus cluster to validate claims of a non-thermal component discovered with \\chandran. From a meticulous analysis of our dataset, which includes a detailed treatment of systematic errors, we find the 2-10 keV surface brightness of the non-thermal component to be smaller than about 5$\\times10^{-16}$erg$~$cm$^{-2}$s$^{-1}$arcsec$^{-2}$. The most likely explanation for the discrepancy between the \\xmmn and \\chandra estimates is a problem in the effective area calibration of the latter. Our \\epic based magnetic field lower limits are not in disagreement with Faraday rotation measure estimates on a few cool cores and with a minimum energy estimate on Perseus. In the not too distant future \\emph{Simbol-X} may allow detection of non-thermal components with intensities more than 10 times smaller than those that can be measured with \\epicn; nonetheless even the exquisite sensitivity within reach for \\emph{Simbol-X} might be insufficient to detect the IC emission from Perseus. ", "introduction": "\\label{sec: intro} Although the bulk of the energy radiated from clusters is of thermal nature, non-thermal mechanisms play an important role. Indeed the characterization of non-thermal components provides much needed clues on the physical process presiding over the formation and evolution of clusters. In some of the more disturbed objects, evidence of non-thermal processes has been known for quite some time. Radio observations indicate that merging clusters are often the site of cluster-wide synchrotron emission, the so called radio halos and radio relics \\citep[and references therein]{cassano07}. Radio data, polarimetric or not, has been used to provide estimates of magnetic fields in clusters. Unfortunately estimates based on Faraday rotation measures (hereafter RM) or minimum-energy arguments are affected by large uncertainties. In the case of RM estimates the unknown field topology \\citep{ensslin03} and the accessability of only a few, not necessarily representative, lines of sight \\citep{rudnick03}, are a major source of concern, while for minimum energy arguments the proton to electron ratio, and the applicability of the argument itself play an equally critical role. It has been recognized for quite some time that detection of inverse Compton (hereafter IC) emission at X-ray wavelengths can provide an alternative method to estimate cluster magnetic fields \\citep{rephaeli87}. Detection and characterization of the so called hard tails is by no means a trivial task: the signal, if there is one, is caught between the hammer of the thermal emission and the anvil of the instrumental background. So far, only detections at a few sigma level have been reported \\citep[for a recent review see][]{rephaeli08}, and at least in one case, Coma \\citep{fusco99,fusco04,fusco07}, they have been challenged \\citep{rossetti04,rossetti07}. Aim of the present work is to employ a long \\xmmn observation of the Perseus core to validate claims about a non-thermal emission component discovered with \\chandra \\citep{sand04,sand05,sand07}. The major advantage of the \\xmmn observation is the better sensitivity at high energies ($\\sim$ 6 keV) of the \\epic detectors with respect to the \\stn ; this is illustrated in Fig.~\\ref{fig: cxo_epc} where we show the ratio of background to source intensity in a representative region, a ring with bounding radii 1$^{\\prime}$-2$^{\\prime}$ centered on the emission peak. The \\chandra ratio rises very rapidly from 1\\% at 5keV to unity at 8 keV; at 10 keV the \\st background is about 20 times larger than the source. \\mos and \\pn ratios rise above 1\\% around 7 keV and remain below 30\\% and 10\\% respectively at all energies. \\begin{figure} \\centering \\resizebox{75mm}{!}{\\includegraphics[angle=-90]{fig_cxo_pn_m2.ps}} \\\\ \\caption{Ratio of background to source spectra for the \\st (blue), \\md (green) and \\pn (red) instruments from the 1$^{\\prime}$-2$^{\\prime}$ ring. The \\st ratio rises rapidly from 1\\% at 5keV to unity at 8 keV; at 10 keV the \\st background is about 20 times larger than the source. \\mos and \\pn ratios rise above 1\\% around 7 keV and remain below 30\\% and 10\\% respectively at all energies.} \\label{fig: cxo_epc} \\end{figure} The outline of the paper is as follows: in Sect.~\\ref{sec: ev_prep} we provide details on our data preparation; in Sect.~\\ref{sec: spec_anal} we describe results from the spectra analysis; in Sect.~\\ref{sec: chandra} we compare results from the analysis of \\epic data with those from \\chandran; in Sect.~\\ref{sec: discussion} we discuss our results and in Sect.~\\ref{sec: summary} we summarize them. We assume $H_0 = 70$~km s$^{-1}$Mpc$^{-1}$ so that 1 arcsec$=$0.35 kpc at the redshift of NGC1275 \\citep[0.0176,][]{huchra99}; all errors are 1$\\sigma$ unless otherwise stated. ", "conclusions": "\\label{sec: discussion} From a general perspective our analysis may be viewed as an attempt to characterize the spectrum of an X-ray source beyond a simple one/two component model. Given the modest intensity of the component we are specifically interested in and the lack of spectral signatures, such as lines or edges, the task is by no means a trivial one. Clearly it can only be attempted on sources where adequate statistics is available. It is often the case that, when the statistics is abundant, systematics become the dominant source of error. At the present time X-ray astronomers do not have a standard method of analyzing data under these predicaments and mostly resort, has we have done, to a trial and error approach. The development of a standard strategy, possibly descending from first principles, is highly desirable particularly if we consider that in the not too distant future X-ray missions such as \\emph{XEUS} or \\emph{Con-X} will have, as one of their primary goals, the detailed characterization of relatively bright X-ray spectra. In some instances the advent of high resolution spectrometry will provide a valid solution. In others, where features that need to be characterized are broad, either because they are of non-thermal nature, such as the putative hard tails in clusters, or because they are smeared by other processes (i.e. the broad iron line observed in nearby AGN), solutions will have to be found elsewhere. Recent work by \\citet{drake07} provides an interesting starting point. \\subsection{\\epic estimate} \\label{sec: dis_epic} In this paper we have characterized the surface brightness of the non-thermal component in Perseus by resorting to a heuristic approach which can be divided into 2 steps: 1) inclusion of 2\\% systematic errors on spectra; 2) modification of spectral models. As far as the first step is concerned, we find a significant reduction of the surface brightness measured from the \\pn spectra when the 2\\% systematic error is included, no significant changes are found for the \\mos measurements ( note however that a reduction is observed starting from a 3\\% systematic error, see Sect.~\\ref{sec: ring_1_2_mos} ). As far as the second step is concerned, the stability of our results with respect to the inclusion of corrections for some imperfections in redistribution matrix, effective area, energy scale and choice of astrophysical model indicate that our measures are relatively robust with respect to the residual systematic effects we have been able to identify. Comparison of \\mos with \\pn results shows that the former favour a somewhat larger value for the surface brightness of the non-thermal component than the latter, roughly speaking 2 against 1. The difference between the two measures is most likely caused by minor cross-calibration issues, possibly in the high energy response of either \\mos or \\pnn. Since we do not have firm evidence as to which of the two experiments is better calibrated, rather than privileging measures from one over the other, we take the common envelope of \\mos and \\pn measurements (i.e. 0- 5$\\times10^{-16}$erg$~$cm$^{-2}$s$^{-1}$arcsec$^{-2}$) as our best estimate for the interval constraining the surface brightness of the non-thermal component. Some of our readers might find this range rather broad, particularly in light of the high statistical quality of our data, we reiterate that the major source of indetermination here as elsewhere are systematic and not statistical errors. \\subsection{\\epic vs. \\acis } \\label{sec: dis_acis} Results from the analysis of \\epic spectra, either \\pn or \\mosn, cannot be reconciled with those obtained with \\stn. As discussed in Sect.~\\ref{sec: chandra} the difference is to be ascribed to a cross-calibration issue between \\epic and \\acisn. Similar problems have been identified by a group of calibrators who have compared \\chandra and \\xmmn observations of hot clusters \\citep{david07}. Recent efforts by \\chandra calibrators \\citep{david07,marshall08} have shown that the above difference follows from problems in the calibration of the \\chandra high energy effective areas. \\subsection{Magnetic field estimates} \\label{sec: dis_b} Under the assumption that the non-thermal emission originates from inverse Compton (IC) scattering of seed microwave and infrared photons by relativistic electrons, responsible of the synchrotron emission in the mini radio-halo, the Perseus core magnetic field can be estimated from the energy density of the seed photon field, $U_{rad}$, the synchrotron and inverse-Compton luminosities, $L_R$ and $L_X$, i.e. $ U_B = U_{rad} \\cdot L_R / L_X $, where $ U_B$ is the energy density of the magnetic field. \\citet{sand05}, from their measures of the non-thermal component, find magnetic field intensities ranging from several $\\mu$G at the very center, to $\\sim$ 1.0~$\\mu$G at 8 kpc (0.4$^\\prime$) and $\\sim$ 0.1~$\\mu$G at 40 kpc (2$^\\prime$) \\citep[see Fig.~9 of][]{sand05}. Our measurements can be used to provide a revised estimate of the B field. As discussed above we estimate the surface brightness of the non-thermal component to be somewhere between 0 and 5$\\times 10^{-16}$erg$~$cm$^{-2}$s$^{-1}$arcsec$^{-2}$. We therefore start by assuming an upper limit of 5$\\times 10^{-16}$erg$~$cm$^{-2}$s$^{-1}$arcsec$^{-2}$, from this we subtract a value of 2.4$\\times 10^{-16}$erg$~$cm$^{-2}$s$^{-1}$arcsec$^{-2}$, as in \\citet{sand05}, to account for projection effects. The resulting surface brightness is used to derive lower limits for the magnetic field using essentially the same method and radio measures described in \\citet{sand05}. In the 0.5$^{\\prime}$-1$^{\\prime}$ and 1$^{\\prime}$-2$^{\\prime}$ annuli our X-ray estimate convert into lower limits of roughly 0.4~$\\mu$G and 0.3~$\\mu$G respectively. We note that these numbers have been derived through a series of rather drastic approximations, (i.e. we have a rather poor determination of the X-ray upper limit, the radio emission is measured at a frequency which is substantially larger than that at which non-thermal electrons responsible for the inverse Compton emission emit via synchrotron, the B field is assumed to be constant, etc.) what really matter is that they are lower limits and not detections. Faraday rotation measures (RM) in cool cores on scales of tens of kpc are in the order of several hundreds to thousands \\citep{taylor02}. Estimates of magnetic fields from rotation measures have undergone some revision in the last few years with more recent estimates typically in the order of a few $\\mu$G \\citep{clarke04,ensslin06}. In the case of the Perseus mini radio-halo Faraday rotation measures are available only on very small scales \\citep{taylor06}, i.e. few tens of pc. RM estimates are in the order of $\\sim$7000 rad m$^{-2}$ leading to B field values of $\\sim 25~\\mu$G under the assumption the screen is localized in the ICM. This, however, appears to be unlikely as variations of 10\\% in the RM are observed on scales of $\\sim$ 1 pc \\citep{taylor06}, while ICM magnetic fields are expected to be ordered on significantly larger scales \\citep[few kpc:][]{taylor02,vogt05,ensslin06}. Application of the classical minimum-energy argument to the Perseus mini radio-halo data, leads to estimates for the central (i.e. $r=0$) magnetic field strength of $\\sim 7~\\mu$G \\citep{pfrommer04}. IC estimates of the magnetic field based on \\chandra measurements \\citep{sand05} are about an order of magnitude below the RM and minimum-energy estimates detailed above. A similar discrepancy has been found when comparing magnetic field estimates on cluster wide scales. There IC measures are in the 0.2-1~$\\mu$G range \\citep[and references therein]{rephaeli08}, while RM estimates are about an order of magnitude larger \\citep[e.g.][]{carilli02,rephaeli08}. While the intricate nature of the astrophysical scenario lends itself to various possible explanations for the discrepancy \\citep[e.g.][]{carilli02,govoni04}, a more trivial alternative, namely that the IC estimates might be incorrect, should also be considered. There is at least one object, Coma, where IC magnetic field measures \\citep{fusco99,fusco04,fusco07} have been challenged \\citep{rossetti04,rossetti07}. Moreover first measures from the \\emph{Suzaku} mission on a few objects, (i.e. A3667, Coma) are turning out to be lower limits on IC measures \\citep[and references therein]{fuka07}. \\xmmn measures presented in this paper seem to play a similar role in the determination of the Perseus core IC magnetic field estimates. \\subsection{Future prospects} \\label{sec: dis_future} It is unlikely that either \\chandra or \\xmmn will provide significantly better estimates of the intensity of the non-thermal component in Perseus than those reported here. Clearly experiments with sensitivities extending into the hard X-ray band are the most appropriate for these kind of studies. Recently the \\emph{BAT} experiment on board the \\emph{Swift} satellite has been used to determine that an extrapolation of the non-thermal flux measured with \\chandra to the 50-100 keV band overshoots the flux measured with \\emph{BAT} by a factor of about 4 \\citep{ajello08}. The \\emph{Suzaku} high energy experiment \\citep{taka07}, limited as it is by modest spatial resolution and by a non optimal background treatment \\citep{kokubun07}, may provide some useful indications but will most likely not allow substantial advancement. Amongst missions that are currently under development \\emph{Simbol-X} \\citep{ferrando05} is arguably one of the most promising. The combination of large throughput in the 1-60 keV range, low instrumental background and \\epicn-like spatial resolution, will allow a sensitive measurement of non-thermal components extending beyond the thermal cutoff. We have performed simulations of \\emph{Simbol-X} spectra based on our best fits for the 1$^{\\prime}$-2$^{\\prime}$ annulus trying out different values of the normalization of the non-thermal component to determine how far down we might go. We started by taking a value of the surface brightness of 1$\\times10^{-16}$erg$~$cm$^{-2}$s$^{-1}$arcsec$^{-2}$. We find that the non-thermal component dominates thermal emission above $\\simeq$ 20 keV and remains above background emission up to $\\simeq$ 30 keV. Thus, if the non-thermal emission has an intensity of this magnitude or larger it will be detected rather easily. Repeating the same exercise with a surface brightness that is ten times smaller, i.e. 1$\\times10^{-17}$erg$~$cm$^{-2}$s$^{-1}$arcsec$^{-2}$, we find that formal fitting of the simulated spectrum still allows a detection of the non-thermal component at the $\\sim 3-4 \\sigma$ level, provided the observation is longer than $3\\times 10^5$s and current estimates of the \\emph{Simbol-X} instrumental background are within 10-20\\% of the real value. We note that in this case the relative intensity of the non-thermal component in the 20-30 keV band is about 10\\%. As we have learned from the analysis of our \\epic data, at these intensity levels, systematics, which are at this time unknown for \\emph{Simbol-X}, will play an important, possibly dominant role. In the specific case of \\emph{Simbol-X} measurements, the critical elements which need to be kept under control at the few percent level are: the effective area calibration; the cross-calibration between the high and low energy detectors and the background. These requirements, albeit challenging, are not beyond reach, particularly if we consider that the formation flight strategy adopted by \\emph{Simbol-X} will entail some important advantages over previous missions. Extensive calibration of the telescope effective areas can be performed in flight by comparing observations of calibration sources performed with and without the telescope in the optical path. Moreover direct illumination of the \\emph{Simbol-X} focal plane, with the telescope removed, will allow an in-flight verification of the detector quantum efficiency. Assuming that \\emph{Simbol-X} can indeed reach a sensitivity of 1$\\times10^{-17}$erg$~$cm$^{-2}$s$^{-1}$arcsec$^{-2}$ this will allow to detect IC components with associated B fields of roughly 1~$\\mu$G, thereby providing important constraints on the magnetic field in Perseus. Given the current magnetic field estimates from Faraday rotation measures on a few cool cores \\citep[few $\\mu$G:][]{ensslin06} and the classical minimum-energy argument estimate on Perseus \\citep[$\\sim~7\\mu$G:][]{pfrommer04}, even the exquisite sensitivity within reach for \\emph{Simbol-X} might be insufficient to detect the IC emission. We have carried out a detailed analysis of a long \\epic observation of the Perseus core in an attempt to detect and characterize a non-thermal component, our main findings may be summarized as follows: \\begin{itemize} \\item systematic uncertainties play an important role in the characterization of the non-thermal component; in the absence of a strategy descending from first principles we have developed a heuristic approach to include them in our analysis; \\item at variance with our preliminary estimates, we find that the non-thermal component is not detected; the surface brightness is determined to be smaller than $\\sim$ 5$\\times10^{-16}$erg$~$cm$^{-2}$s$^{-1}$arcsec$^{-2}$; \\item our \\xmmn estimates are at variance with \\chandra estimates from \\citet{sand05,sand07} and from our own analysis; the most likely explanation for the discrepancy between \\chandra and \\xmmn is a problem in the \\chandra effective area calibration; \\item our \\epic based upper-limit on the surface brightness converts into IC magnetic field lower limits of $\\sim$0.4~$\\mu$G for the 0.5$^{\\prime}$-1$^{\\prime}$ annulus and $\\sim$0.3~$\\mu$G for the 1$^{\\prime}$-2$^{\\prime}$ annulus; these measures are not in disagreement with RM estimates on a few cool cores \\citep[few $\\mu$G:][]{ensslin06} and the minimum energy estimate on Perseus \\citep[10~$\\mu$G:][]{pfrommer04}; \\item in the not too distant future \\emph{Simbol-X} may allow detection of non-thermal components with intensities more than 10 times smaller than those that can be measured with \\epicn; nonetheless even the exquisite sensitivity within reach for \\emph{Simbol-X} might be insufficient to detect the IC emission from Perseus. \\end{itemize}" }, "0807/0807.1424.txt": { "abstract": "We have conducted aperture polarimetry of $\\sim$500 stars of the Orion Nebula Cluster (ONC) in M42 based on our wide-field ({$\\sim$}8$'${$\\times$}8$'$) $JHKs$ band polarimetry. Most of the near-infrared (NIR) polarizations are dichroic, with position angles of polarization agreeing, both globally and locally, with previous far-infrared (FIR) and submillimeter observations, having taken into account the 90$^\\circ $ difference in angles between dichroic absorption and emission. % thermal emission polarizations, %\\cup except for the 90$^{\\circ}$ difference of the position angles. This is consistent with the idea that both NIR dichroic polarizations and FIR/submillimeter thermal polarizations trace the magnetic fields in the OMC-1 region. The magnetic fields inferred from these observations show a pinch at scales less than 0.5 pc with a centroid near IRc2. The hourglass-shaped magnetic field pattern is explained by the models in which the magnetic field lines are dragged along with the contracting gas and then wound up by rotation in a disk. The highly polarized region to the northwest of IRc2 and the low-polarized region near the bright bar are also common among NIR and FIR/submillimeter data, although a few regions of discrepancy exist. We have also discerned {$\\sim$}50 possible highly polarized sources whose polarizations are more likely to be intrinsic rather than dichroic. Their polarization efficiencies ($P(H)/A(H)$) are too large to be explained by the interstellar polarization. These include 10 young brown dwarfs that suggest a higher polarization efficiency, which may present geometrical evidence for (unresolved) circumstellar structures around young brown dwarfs. ", "introduction": "Magnetic fields are believed to play an important role in the evolution of molecular clouds, from their large-scale structures to dense cores, protostar envelopes, and protoplanetary disks. Magnetic fields can be measured by observing polarizations of electromagnetic waves. Dichroic polarizations are believed to be caused by spinning non-spherical dust grains becoming aligned with their short axis precessing around the direction of the local magnetic field. In absorption this produces polarization parallel to the magnetic field and perpendicular to the magnetic field in emission (e.g., Weintraub et al. 2000). %Dichroic polarizations are believed to be caused by magnetically aligned, nonspherical grains, %which thus follow the magnetic fields (e.g., Weintraub et al. 2000). Optical stellar polarimetry is useful for revealing magnetic field structures in the periphery of molecular clouds (e.g., Appenzeller 1974; Vrba et al. 1976; Moneti et al. 1984; Goodman et al. 1990). However, because of heavy dust extinction, optical polarimetry is not always a good tracer of the magnetic fields inside molecular clouds. Since the pioneering work by Vrba et al. (1976) and Wilking et al. (1979), near-infrared (NIR) stellar polarimetry in star-forming regions has been a powerful tool for tracing magnetic field structures in molecular clouds (Tamura et al. 1987; Sato et al. 1988; Tamura et al. 1988; Klebe \\& Jones 1990). The stars could be either background stars or even embedded sources if they are not associated with circumstellar structures (i.e., have no intrinsic polarization). Although arguments have been made against the usefulness of NIR polarimetry in tracing magnetic fields inside molecular clouds and despite a suggestion that the dust grains are not well aligned in the dense regions (Goodman et al. 1995), recent submillimeter polarimetry has clearly shown aligned dust grains, even in cold regions such as starless globules and cores (Kirk, Ward-Thompson, \\& Crutcher 2006). With the advent of wide-field NIR polarimetric capability (Tamura et al. 1996), reevaluating the utility of NIR stellar polarimetry is warranted because the wide field allows simultaneous polarimetry of many stars in the field of view. The Orion Nebula Cluster (ONC), or Trapezium cluster, is an ideal site for such a study. The ONC is one of the nearest (450 pc), massive, star-forming regions to the Sun and the most populous young cluster within 2 kpc, composed of some 3500 young, low-mass stars (O'Dell 2001).\\\\ % The magnetic field structure of the Orion region was most extensively studied by observing linearly polarized thermal emission from aligned dust grains. Houde et al. (2004) and Schleuning (1998) showed at 350 $\\mu$m and 100$\\mu$m, respectively, that the magnetic field in OMC-1 is generally oriented northwest--southeast, with the field pinched along the northeast--southwest direction on a scale of several arcmin. The polarization percentage is low at the location of the Becklin--Neugebauer object and Kleinmann--Low nebula (BN/KL) compared to elsewhere, but that may be due to a small-scale variation that is undetectable in their 12$''$ - 35$''$ resolution maps. See Cudlip et al. 1982, Hildebrand et al. 1984, Dragovan 1986, Barvainis et al. 1988, Gonatas et al. 1990, Leach et al. 1991, and Rao et al. 1998 for earlier or other millimeter and submillimeter studies in the Orion region. Imaging polarimetry of M42 has been conducted either only at optical wavelengths (Pallister et al. 1977) or toward a small region near IRc2 or BN at near-infrared wavelengths (Minchin et al. 1991; Jiang et al. 2005; Simpson et al. 2006). % Hough et al. (1986), Burton et al. (1991), and Chrysostomou et al. (1994) measured the polarization of the H$_2$~v~=~1-0~S(1) line at $2.12\\mu$m around IRc2. They attributed the polarization in the vicinity of IRc2 to dichroic absorption, and in particular, the twist in the polarization vectors 5$''$ east of IRc2 to a twist in the magnetic field direction. % At NIR wavelengths, the other IRc sources (IRc2 is not readily visible at 2~$\\mu$m, probably because it is behind the edge of the hot core) are all strongly polarized, with centrosymmetric polarization vectors indicating scattered light; that is, the IRc sources are illuminated by either BN (IRc1) or some star in the vicinity of IRc2 (Werner et al. 1983; Minchin et al. 1991; Chrysostomou et al. 1994). The high degree of polarization of BN, which is proportional to the extinction with the FIR-inferred magnetic field directions, has been used to deduce that BN is polarized by dichroic absorption at all NIR and MIR wavelengths where it has been measured (e.g., Lee \\& Draine 1985; Hough et al. 1996; Aitken et al. 1997; Smith et al. 2000). However, Jiang et al. (2005) used NIR adaptive optics-aided, very high-resolution imaging polarimetry to successfully reveal an outflow/disk system around BN. % % In this paper, we present the aperture polarimetry in the $J, H,$ and $Ks$ bands of $\\sim 500$ stars of the ONC with the polarimetry mode of the IRSF/SIRIUS instrument (SIRPOL). We also discuss possible intrinsically highly polarized sources and the polarization of young brown dwarfs. ", "conclusions": "We conducted deep $JHKs$ imaging polarimetry of a 7$\\farcm$7 $\\times$ 7$\\farcm$7 area of M42 and here present the results of the aperture polarimetry. Our main conclusions are summarized as follows: \\begin{enumerate} \\item Most of the NIR polarizations are dichroic. Their global and local vector patterns are in good agreement with previous FIR and submillimeter polarization patterns, except for the 90$^{\\circ}$ difference of position angles. \\item A positive correlation exists between $P(H)$ and $H-Ks$; most of $P(H)$ is below $P_{max}$, except for possible highly polarized sources. This supports an intracloud origin of the NIR polarizations, excluding (a) a non-intracloud magnetic origin of polarization, and (b) a weaker alignment of dust grains in the cloud. \\item We argue that both NIR dichroic polarizations and FIR/submillimeter thermal polarizations trace the magnetic fields in the OMC-1 region. \\item The magnetic fields are pinched at scales less than 0.5 pc with a centroid near IRc2. The hourglass-shaped magnetic field pattern is explained by models in which the magnetic field lines are dragged along with the contracting gas and then wound up by rotation in a disk. \\item The highly polarized region to the northwest of IRc2 and the low-polarized region near the bright bar, commonly seen in both NIR and FIR/submillimeter data, are explained by the field geometry; the latter is the field along and the former perpendicular to the line of sight. \\item We also discriminated $\\sim$50 possible highly polarized sources whose polarizations are more likely to be intrinsic rather than dichroic. Their polarization efficiencies ($P(H)/A(H)$) are too large to be explained by normal interstellar polarization. \\item For 9 young brown dwarf candidates, we also suggest the existence of higher polarization efficiency, which may present geometrical evidence for (unresolved) circumstellar structures around young brown dwarfs. \\end{enumerate} We are grateful to Shuji Sato for helpful suggestions. Thanks are due to the staff in SAAO for their kind help during the observations. We thank D. C. Lisfor kindly providing the 350 \u0083\u00cam map in FITS format. We also thank Noboru Ebizuka Tetsuo Nishino, and Toshihide Kawai for their technical support in the development of SIRPOL. The IRSF/SIRIUS project was initiated and supported by Nagoya University, National Astronomical Observatory of Japan, and The University of Tokyo in collaboration with South African Astronomical Observatory under a financial support of Grants-in-Aid for Scientific Research on Priority Area (A) No. 10147207 and No. 10147214, and Grants-in-Aid No. 13573001 and No. 16340061 of the Ministry of Education, Culture, Sports, Science, and Technology of Japan. This study was partly supported by a MEXT Grant-in-Aid for Scientific Research in Priority Areas, \u0081gDevelopment of Extrasolar Planetary Science,\u0081h and by grants-in-aid fromMEXT(Nos. 16077204, 16077171, and 19204018)." }, "0807/0807.2605_arXiv.txt": { "abstract": "The mass reinserted by young stars of an emerging massive compact cluster shows a bimodal hydrodynamic behaviour. In the inner part of the cluster, it is thermally unstable, while in its outer parts it forms an out-blowing wind. The chemical homogeneity/inhomogeneity of low/high mass clusters demonstrates the relevance of this solution to the presence of single/multiple stellar populations. We show the consequences that the thermal instability of the reinserted mass has to the galactic super-winds. ", "introduction": "The open star clusters and stellar moving groups have internally homogeneous chemical composition. Clusters like the Hyades, Collinder 261, the Herculis stream or the moving group HR 1614 are chemically unique, distinguishable one from the other, showing no pollution from secondary star formation \\cite[(De Silva et al., 2008)]{Silva3}. The chemical homogeneity of open star clusters like the Hyades \\cite[(De Silva et al., 2006)]{Silva1} and Collinder 261 \\cite[(De Silva et al., 2007)]{Silva2} proves that they have been formed out of a well-mixed cloud and that any self-enrichment of stars did not take place there. Young and massive stellar clusters, frequently called super star clusters, are preferentially observed in interacting galaxies. Their stellar mass amounts to several million M$_{\\odot }$ within a region less than a few parsecs in diameter. They represent the dominant mode of star formation in starburst galaxies. Their high stellar densities resemble those of globular clusters, where several stellar populations have been observed \\cite[(Piotto, 2008)]{GP}. To explain the presence of multiple stellar generations in globular clusters, the slow wind emerging from a first generation of fast rotating massive stars is invoked by \\cite[Decressin et al. (2007)]{TDetal}. (See also the review by \\cite[Meynet (2008)]{Meynet} in this volume.) The authors argue that the fast rotating massive stars function as a filter separating the H-burning products from later products of He-burning. However, it is not clear why all the massive stars rotate fast, or why the slow wind produced by stellar rotation is just retained inside the potential well of the stellar cluster. An alternative solution, how to form the second generation of stars in massive star clusters, is proposed in models of star cluster winds described by \\cite[Tenorio-Tagle et al. (2007)]{GTTetal}, \\cite[W\\\" unsch et al. (2007)]{RWetal1}, and \\cite[W\\\" unsch et al. (2008)]{RWetal2}. There, we argue that a critical mass of a cluster exists, below which the single-mode hydrodynamical solution to the cluster winds applies. Such clusters should have one stellar generation only, and show strong winds corresponding to the momentum and energy feedback of all their stars. The clusters above the critical mass should follow the bi-modal solution to their winds, where only the outer skin of the cluster participates in the wind. Their inner parts are thermally unstable, and hence being the potential places of secondary star formation. ", "conclusions": "We propose a bimodal solution, where in the central part of a massive star cluster, a thermally unstable region forms (see Fig. 2). In this region, the thermal instability creates cold regions surrounded by hot medium imploding into them. The high-velocity wing of broad spectral lines \\cite[Gilbert, \\& Graham (2007)]{AMG&JRG} observed in SSC may be created by imploding shock in the vicinity of thermally unstable parcels of gas. The second generation of stars may be formed out of cold clumps produced by thermal instability. During the early evolution of a massive cluster, the first Myrs, the mechanical energy input is dominated by stellar winds \\cite[(Leitherer et al. 1999)]{Leithereretal}. The efficiency of thermalization $\\eta $ may be low in this case, and a massive cluster may be above $M_{SC, crit}$, since its value is low. Later, the importance of winds fades out, and the mechanical input is dominated by supernovae. This may increase the thermalization efficiency $\\eta $, increasing at the same time the value $M_{SC, crit}$. Thus the same massive cluster, which was initially in the bi-modal situation moves to single-mode situation. The cold parcels of gas form, in the cluster central part during the early bi-modal situation, from the winds that are enriched by products of H-burning. The later He-burning products are inserted into the cluster volume when the mechanical energy input is dominated by supernovae, which may mean that the cluster is in the single-mode situation, and the wind clears its volume from He burning products. Thus, the thermal instability, which operates during a few initial Myr in the central part of the cluster, may produce a second generation of stars enriched by H-burning products. Later, the cluster moves into the single-mode situation, which means that it is able to expel the He-burning products. The feedback of massive stars in super star clusters creates galactic winds, or super winds, reaching to large distances from the parent galaxies, transporting the products of stellar burning into intergalactic space. The bimodal solution, providing a possible explanation of multiple stellar populations in globular clusters, limits the super winds. During the initial period of cluster evolution, when the stellar winds dominate the mechanical energy input, the super wind is restricted only to the outer skin of the cluster, which means that it is rather weak. Only later, when supernovae become dominant in mechanical energy input, strong super winds blowing out from all the cluster volume may reach large distances from their parent galaxies. How effective the super winds of super star clusters can be in transporting the products of stellar evolution into intergalactic spaces should be discussed in future." }, "0807/0807.2433_arXiv.txt": { "abstract": "Much progress has been made in measuring black hole (BH) masses in (non-active) galactic nuclei using the tight correlation between stellar velocity dispersions $\\sigma$ in galaxies and the mass of their central BH. The use of this correlation in quasars, however, is hampered by the difficulty in measuring sigma in host galaxies that tend to be overpowered by their very bright nuclei. We discuss results from a project that focuses on $z\\sim0.3$ quasars suffering from heavy extinction at shorter wavelengths. This makes it possible to obtain clean spectra of the hosts in the spectral regions of interest, while broad lines (like H$\\alpha$) are still visible at longer wavelengths. We compare BH masses obtained from velocity dispersions to those obtained from the BLR and thus probe the evolution of this relation and BH growth with redshift and luminosity. Our preliminary results show an offset between the position of our objects and the local relation, in the sense that red quasars have, on average, lower velocity dispersions than local galaxies. We discuss possible biases and systematic errors that may affect our results. ", "introduction": "Black hole (BH) mass is believed to be one of the fundamental parameters that characterize quasar activity and much effort has been devoted to obtaining accurate BH masses for quasars and other AGN \\citep[\\eg][]{ho99}. In recent years, much progress has been made in measuring BH masses in galactic nuclei, particularly with the remarkable discovery by \\citet{geb00a} and \\citet{fer00} of a tight correlation between stellar velocity dispersion in galaxies and the mass of their central BH (M$_{\\rm BH}\\propto\\sigma^{n}$). The use of this correlation to derive BH masses in AGN, however, is hampered by the difficulty in measuring velocity dispersions in host galaxies that tend to be overpowered by their very bright nuclei. Nevertheless, the correlation has been shown to be present at low redshift ($z<0.1$) in low luminosity AGN (\\eg\\ BL Lac objects: \\cite{barth03}; or Seyfert galaxies: \\cite{geb00b}). Seyfert galaxies at higher redshift ($z\\sim0.36$ and $z\\sim0.57$), however, appear to show an offset from the local relation \\citep[][and references therein]{Woo2008}. It is not yet known whether the M$_{\\rm BH}-\\sigma$ correlation holds for the highest luminosity AGN. A loose correlation has been found by using the width of [O\\,III] lines in active nuclei \\citep{nel00,shi03}, but the width of these lines is dependent upon other parameters (outflows, radio luminosity, etc.) and therefore lead to a correlation with a large scatter. BH masses derived from [O\\,III] emission line widths can only be accurate to within a factor of five at best \\citep{bor03}. More accurate determinations are necessary if we hope to use them to disentangle some of the other fundamental relationships among quasar parameters. ", "conclusions": "" }, "0807/0807.0436_arXiv.txt": { "abstract": " ", "introduction": "Cataclysmic variable stars (CVs) consist of a Roche-lobe filling main sequence star, called the secondary star, orbiting a white dwarf - the primary star. Mass flows from the secondary star through the inner Lagrangian point towards the primary star, ultimately leading to the formation of an accretion disc (AD). After the AD's formation, a bright spot (BS) is formed at the contact point where the gas stream hits the AD. Dwarf novae (DNe) are a subclass of cataclysmic variables with two different brightness states, quiescence and outburst. In the latter, the system usually brightens by 2--5\\,mag with a time-scale that can vary from tens to hundreds of days. DNe can also show a third brightness state, the superoutburst, where the system brightens by typically an additional magnitude or two compared to normal outburst. IP Peg is the brightest eclipsing ($i \\approx81^{\\circ}$) DN above the period gap, and was discovered by \\citet{lip81}. It has a 3.8-h orbital period and a $V$ magnitude in the range of 14--12\\,mag, the extreme values corresponding to quiescence and outburst, respectively. From the first photometric light curves during quiescence and return to quiescence \\citep{woo86,woo89}, it was evident that the BS produces a prominent orbital hump, when it rotates into view from behind the disc with a single-stepped ingress and double-stepped egress superimposed on a more gradual disc-eclipse. Infrared light curves showed 0.2\\,mag ellipsoidal variations from the secondary star superposed on a deep eclipse of the WD and the BS \\citep{szk86}. \\citet{woo89}, based on the analysis of 39 eclipses, found out that IP Peg has a rapidly increasing orbital period. However, a new study performed by \\citet{wol93} and combining 49 eclipses along with previously reported ones, reached the conclusion that the orbital period varies sinusoidally with a period of 4.7 years. This variation they attributed to a third body, a late M dwarf star with a most probable mass of 0.10\\,$\\rm M_\\odot$. They also showed that the rate of shrinkage in disc radius during decline from outburst is constant during different outburst cycles, following an exponential decay. A study of the flickering in IP Peg \\citep{bru00} showed that the BS is the dominant source of flickering, with the AD and WD contributing practically nothing. Spectroscopy of the red star \\citep{mar89} showed TiO bands and the NaI doublet (red dwarf features), several broad double-peaked emission lines (disc features) and a narrow third peak in the emission lines moving in anti-phase with the double peaks (a chromospheric emission feature). \\citet{bee00} performed a TiO study of IP Peg and concluded that TiO instead of the NaI doublet should be used for radial velocity studies of the secondary star. In this way they found that the contamination from disc emission, which in other studies returned an elliptical orbit, could be avoided and higher signal-to-noise ratio could be achieved. They gave a final projected orbital velocity of the companion star $K_2=331.3\\,\\rm km\\,s^{-1}$ and refined values of the primary and secondary star masses $M_1=1.05\\,\\rm M_\\odot$ and $M_2=0.33\\,\\rm M_\\odot$, respectively. Covering a complete cycle from quiescence to outburst and quiescence again, the spectral characteristics of IP Peg will now be described. {\\it Quiescence spectra} \\citep{mar88} show broad double-peaked emission lines on a blue continuum, a strong orbital hump in the continuum (but much weaker in the lines) and a flat Balmer decrement indicative of optically thick emission. The trailed spectra show the signature of a prograde rotation of the AD, where the blue-shifted emission is eclipsed prior to the red-shifted one, a weak S-wave component in the red-shifted wing and a 10\\% blue/red asymmetry before eclipse, most probably arising from line emission from the BS. {\\it Early rise-to-outburst spectra} \\citep{wol98} show Balmer emission, TiO absorption bands, KI and NaI absorption doublets, a weak S-wave and two bright spots in both Doppler and eclipse maps. The study of \\citet{wol98}, as well as a re-analysis of their results using an extension of the classical eclipse mapping method \\citep{bob99}, concluded that the first spot should be identified as the BS while two possibilities existed for the second spot. It either was the beginning of formation of a spiral arm or a second inner BS due to gas-stream overflow as predicted by, e.g., simulations performed by \\citet{arm96}. {\\it Rise-to-outburst spectra} \\citep{har94} show a decrease in the equivalent width of H$\\alpha$ emission line, down to a level similar to that observed immediately after outburst and, simultaneously, weak H$\\alpha$ emission from the secondary due to irradiation of the secondary star by the boundary layer. {\\it Near-outburst spectra} \\citep{pic89} show unusually strong high-excitation lines, Balmer lines in emission, indications of a very disturbed AD with a non-uniform He emission distribution, an outflowing wind from the WD vicinity and chromospheric emission from the secondary. {\\it Outburst spectra} as presented by \\citet{mar90} show: i) strong Balmer and HeII $\\lambda$4686 double-peaked emission lines rising from the AD, ii) non-Keplerian emission filling in the centre of HeII $\\lambda$4686, probably due to a compact outflowing wind or inflowing magnetic accretion columns from close to the WD, iii) Balmer but not HeII emission from the secondary, located in the polar region on the side facing the AD, iv) no evidence for an enhanced gas-stream compared to quiescence, v) large azimuthal asymmetry in the outer disc, with the blue-shifted side of the disc being brighter and possibly more extended at phase 0 and vi) a sharp transient component in the HeII line possibly generated from the interaction of the gas stream and disc. \\citet{mar90} suggested that He emission can be explained as photoionisation of the disc by the boundary layer and found that Balmer emission in outburst can be attributed by 3/4 to the disc, 1/4 to irradiation of the red star and also to boundary layer irradiation. Another study performed by \\citet{ste96} showed, from trailed spectra and Doppler maps, asymmetric disc emission, strong secondary star emission and a low velocity emission in the H$\\alpha$ line only. This stationary emission, appearing as an emission blob in the centre of the Doppler maps, was interpreted as slingshot prominences, material trapped in magnetic loops and co-rotating with the secondary star. During another outburst maximum, \\citet{har99a} found metal line emission such as MgII. Doppler tomography of the He and metal lines locates those lines' emission on the inner Roche lobe of the secondary star. {\\it Early decline-from-outburst spectra} \\citep{hes89} reveal for the first time in such strength a blend of high-excitation lines of HeII, CIII and NIII along with a very strong chromospheric emission-line component from the secondary. {\\it Decline-from-outburst spectra} \\citep{mor00} show irradiation-induced emission from the companion star in Balmer, He, MgII and CII lines. This emission, located only near the poles of the secondary, suggests shielding of the equatorial regions of the secondary star by the vertically extended AD. {\\it Early-quiescence spectra} in the infrared \\citep{lit01} reveal the existence of mirror eclipses in IP Peg. A mirror eclipse is an eclipse of the secondary star by an optically thin AD and appears in the trailed spectra as a reduction in the equivalent widths of the lines, at phase 0.5. A mirror eclipse also has a mirror-symmetry with the classical rotational disturbance, i.e. the red-shifted part of the line is eclipsed prior to the blue-shifted one. Even though IP Peg had shown many peculiarities and novelties in the studies cited previously, it had to reveal another breakthrough. \\citet{ste97} were the first to find convincing evidence for the existence of spiral structure in the AD of a close binary. Spiral waves had long been predicted in studies of tidal interactions and numerical simulations (see \\citet{bof01} and references therein) and may play a crucial role in our understanding of the angular momentum transport. \\citet{ste97} obtained spectrophotometric observations of IP Peg during the rise to an outburst, covering H$\\alpha$ and HeI $\\lambda$6678. Disc emission was centred on the WD and had a strong azimuthal asymmetry in the forms of two spiral arms. The repeatability of the two-armed spiral structure was confirmed during another outburst maximum. \\citet{har99a} applied Doppler tomography of the HeII $\\lambda$4686 line and clarified the shock nature of the spiral arms which showed a jump of a factor of more than 2 in intensity. Additionally they calculated an azimuthal extent of 90$^{\\circ}$, a 30$^{\\circ}$ upper limit in opening angle and a 15\\% contribution of the shocks to the total disc emission. \\citet{mor00} showed that the spiral pattern persists even 5 and 6 days after the outburst maximum, without showing any diminution in strength. Time-resolved optical spectra obtained during late outburst by \\citet{bap00} reveal, through the eclipse mapping technique, that spirals are present even 8 days after the onset of an outburst. Their study showed a similar contribution to the continuum emission, as that measured during the peak of the outburst but also a retrograde azimuthal rotation of 58$^{\\circ}$ compared to outburst maximum. \\citet{bap05} performed a re-analysis of previous spectroscopic data of IP Peg during outburst \\citep{mor00}, based on the eclipse mapping technique. The difference in their new study was based on the fact that they separated the inner and outer regions of the emission line light curves, obtaining in this way the ``core'' and ``spiral arms'' lights curves, respectively. In this way they improved the previously noisy view of the spiral arms and were able to trace them in detail. The two-armed spiral structures were clearly visible in all their eclipse maps, albeit the ``blue'' arm appearing farther out from the disc's centre than the ``red'' one. However, their crucial findings were the sub-Keplerian velocities along the spiral structures, as well as the clear correlation between the opening angle of the spirals and the outburst stage (the later in the outburst, the smaller the opening angle of the spirals). Both these findings greatly favour the interpretation of the spiral asymmetric structures as tidally-induced spiral shocks instead of mere radiation patterns on an otherwise unperturbed AD. A study performed by \\citet{ste01} presents a review of the spiral structure that far, and plots in detail the azimuthal angles and extent of the spirals for the HeII $\\lambda$4686 emission line during an outburst presented by \\citet{har99}. In more detail, it is shown that the two spirals can be traced for almost 180$^{\\circ}$, while their velocity varies from 495 to 780\\,km\\,s$^{-1}$. Under the assumption of Keplerian velocities, they show that the arms cover a substantial part of the disc, 0.3--0.9 times the distance to the inner Langragian point $L_1$. It is pointed out that tracing the spiral arms and plotting their position and intensity as a function of azimuth throughout an outburst, or even better applying this procedure to a range of emission lines simultaneously, can yield important results on the evolution of the spirals. A future paper, as stated in the Discussion section, is intended to address this issue. A first approach on tracing the spiral arms of two different emission lines (H$\\alpha$, HeI $\\lambda$6678), as well as plotting their emissivity as a function of azimuth, is presented in \\citet{ste01}. It is shown that, even though there is a similar position for the spirals in terms of velocity as a function of azimuth, the intensity modulation is significantly different. \\begin{figure} \\centering \\includegraphics[width=13cm]{fig01.eps} \\caption{AAVSO long-term light curve of visual validated data. The arrow shows the timing of our observations, just 1 day after the peak of the outburst.} \\label{f_aavso} \\end{figure} Our work is part of a project of studying CVs through high resolution time-resolved echelle spectra and applying the indirect imaging technique of Doppler tomography. The simultaneous study of several emission lines can yield important results on the AD structure, as presented for another CV system in \\citet{pap08}. Following the introduction on IP Peg one can conclude on the benefits of applying such an analysis to this object, especially since it was observed during outburst. In this way we can study the behaviour of the spiral arms and map the structure of the AD in detail. The application of Modulation Doppler tomography, an extension of the classic Doppler tomography method \\citep{ste03}, gives us the opportunity to map emission sources which can vary harmonically with the orbital period. Such an analysis is a first for IP Peg. This paper is organised as follows: in Sect.~2 we present our observations, in Sec.~3 the data reduction, in Sect.~4 we perform the spectral analysis and Sect.~5 deals with the application and results of Modulation Doppler tomography. Last, Sect.~6 is devoted to the discussion and summary of our results. ", "conclusions": "" }, "0807/0807.1430_arXiv.txt": { "abstract": "Resonant relaxation (RR) is a rapid relaxation process that operates in the nearly-Keplerian potential near a massive black hole (MBH). RR dominates the dynamics of compact remnants that inspiral into a MBH and emit gravitational waves (extreme mass ratio inspiral events, EMRIs). RR can either increase the EMRI rate, or strongly suppress it, depending on its still poorly-determined efficiency. We use small-scale Newtonian $N$-body simulations to measure the RR efficiency and to explore its possible dependence on the stellar number density profile around the MBH, and the mass-ratio between the MBH and a star (a single-mass stellar population is assumed). We develop an efficient and robust procedure for detecting and measuring RR in $N$-body simulations. We present a suite of simulations with a range of stellar density profiles and mass-ratios, and measure the mean RR efficiency in the near-Keplerian limit. We do not find a statistically significant dependence on the density profile or the mass-ratio. Our numerical determination of the RR efficiency in the Newtonian, single-mass population approximations, suggests that RR will likely \\emph{enhance} the EMRI rate by a factor of a few over the rates predicted assuming only slow stochastic two-body relaxation. ", "introduction": "Dynamical relaxation processes near massive black holes (MBH) in galactic centers affect the rates of strong stellar interactions with the MBH, such as tidal disruption, tidal dissipation, or gravitational wave (GW) emission \\citep[e.g.][]{ale05}. These relaxation processes may also be reflected by the dynamical properties of the different stellar populations there \\citep{hop+06a}, as observed in the Galactic Center \\citep{gen+00,pau+06}. Of particular importance, in anticipation of the planned Laser Interferometer Space Antenna (LISA) GW detector, is to understand the role of relaxation in regulating the rate of GW emission events from compact remnants undergoing quasi-periodic extreme mass ratio inspiral (EMRI) into MBHs. Two-body relaxation, or non-coherent relaxation (NR), is inherent to any discrete large-N system, due to the cumulative effect of uncorrelated two-body encounters. These cause the orbital energy $E$ and the angular momentum $J$ to change in a random-walk fashion ($\\propto\\!\\sqrt{t}$) on the typically long NR timescale $T_{\\mathrm{NR}}$. In contrast, when the gravitational potential has approximate symmetries that restrict orbital evolution (e.g. fixed ellipses in a Keplerian potential; fixed orbital planes in a spherical potential), the perturbations on a test star are no longer random, but correlated, leading to coherent ($\\propto\\! t$) torquing of $J$ on short timescales, while the symmetries hold. Over longer times, this results in resonant relaxation (RR) (\\citealt{rau+96,rau+98}; \\S \\ref{ss:RR}), a rapid random walk of $J$ on the typically short RR timescale $T_{\\mathrm{RR}}\\!\\ll\\! T_{\\mathrm{NR}}$. RR in a near-Keplerian potential can change both the direction and magnitude of $\\mathbf{J}$ ({}``scalar RR''), thereby driving stars to near-radial orbits that interact strongly with the MBH. RR in a near-spherical potential can only change the direction of $\\mathbf{J}$ ({}``vector RR''). RR is particularly relevant in the potential near a MBH, where compact EMRI candidates originate. \\citet{hop+06a} show that RR dominates EMRI source dynamics. Depending on its still poorly-determined efficiency, RR can either increase the EMRI rate over that predicted assuming NR only, or if too efficient, it can strongly suppress the EMRI rate by throwing the compact remnants into infall (plunge) orbits (cf Fig. \\ref{f:EMRI} below) that emit a single, non-periodic and hard to detect GW burst. A prime motivation for the systematic numerical investigation of RR efficiency presented here, are the still open questions about the implications of RR for EMRI rates and orbital properties. This paper is organized as follows. In \\S \\ref{s:theory} we briefly review the theory of NR and RR relaxation and derive a new relation between scalar and vector RR. In \\S \\ref{s:ACF} we describe our method of analyzing and quantifying the effects of RR in $N$-body simulations, which are described in \\S \\ref{s:simulations}. We present our results in \\S \\ref{s:results} and discuss and summarize them in \\S \\ref{s:conclusions}. ", "conclusions": "\\label{s:conclusions} We characterized and measured the mean efficiency coefficients of NR ($\\alpha_{\\Lambda},\\eta_{s,\\Lambda},\\eta_{v,\\Lambda}$) and RR ($\\beta_{s},\\beta_{v}$) in Newtonian $N$-body simulations of isotropic, thermal, near-Keplerian stellar cusps around a MBH. We derived a simple analytical form for the rms RR auto-correlation curves of $\\mathbf{J}$ and $J$, and showed that the two are simply proportional to each other. We then measured these coefficients in a large suite of small scale $N$-body simulations with different stellar density distributions and MBH/star mass ratios. We found no statistically significant trends in the values of these coefficients as function of the system properties. This may require better statistics. Our measured RR efficiency suggests that RR increases the EMRI rate by a factor of $\\sim\\!5$ above what is predicted for NR only. This estimate of RR efficiency is consistent with that suggested by the analysis of the dynamical properties of the different stellar populations in the Galactic Center \\citep{hop+06a}. However, this conclusion is still preliminary, since several important open issues remain, which should be addressed by larger scale simulations. These include (1) The dependence of the RR coefficients on the orbit of the test star, for example its eccentricity \\citep{gur+07}. We find that $N\\!\\sim\\!200$ is not enough for reliable statistics on sub-samples within given \\emph{E} or \\emph{J}-bins. The eccentricity dependence of RR is particularly relevant for the supply rate of stars to the MBH from $J\\!\\rightarrow\\!0$ orbits (the loss-cone refilling problem). (2) The effects of a stellar mass spectrum. This will likely affect RR by decreasing the RR timescale, and changing the stellar density distribution through strong mass segregation (\\citealt{ale07}; Alexander \\& Hopman 2008, in prep.). (3) The robustness of RR against perturbations from the larger non-Keplerian stellar system in which the inner near-Keplerian region of interest is embedded. (4) The role of post-Newtonian effects in RR, such as General Relativistic precession and GW emission. These are expected to play a key role in enabling inspiral by quenching RR just as the compact remnant enters the EMRI phase, and in regulating the GW inspiral rate \\citep{hop+06a}." }, "0807/0807.3329_arXiv.txt": { "abstract": "{ We review the results of the first multi-scale, hydrodynamical simulations of mergers between galaxies with central supermassive black holes (SMBHs) to investigate the formation of SMBH binaries in galactic nuclei. We demonstrate that strong gas inflows due to tidal torques produce nuclear disks at the centers of merger remnants whose properties depend sensitively on the details of gas thermodynamics. In numerical simulations with parsec-scale spatial resolution in the gas component and an effective equation of state appropriate for a starburst galaxy, we show that a SMBH binary forms very rapidly, less than a million years after the merger of the two galaxies, owing to the drag exerted by the surrounding gaseous nuclear disk. Binary formation is significantly suppressed in the presence of a strong heating source such as radiative feedback by the accreting SMBHs. We also present preliminary results of numerical simulations with ultra-high spatial resolution of $0.1$~pc in the gas component. These simulations resolve the internal structure of the resulting nuclear disk down to parsec scales and demonstrate the formation of a central massive object ($\\sim 10^8 \\Mo$) by efficient angular momentum transport due to the disk's extended spiral arms. This is the first time that a radial gas inflow is shown to extend to parsec scales as a result of the dynamics and hydrodynamics involved in a galaxy merger, and has important implications for the fueling of SMBHs. Due to the rapid formation of the central clump, the density of the nuclear disk decreases significantly in its outer region, reducing dramatically the effect of dynamical friction and leading to the stalling of the two SMBHs at a separation of $\\sim 1$~pc. We discuss how the orbital decay of the black holes might continue in a more realistic model which incorporates star formation and the multi-phase nature of the ISM. ", "introduction": "In recent years, compelling dynamical evidence has indicated that supermassive black holes (SMBHs) are ubiquitous in galactic nuclei (e.g., Ferrarese \\& Ford 2005). According to the standard modern theory of cosmological structure formation, the Cold Dark Matter (CDM) paradigm (e.g., Blumenthal et al. 1984), galaxies in the Universe grow through a complex process of continuous mergers and agglomeration of smaller systems. Thus, if more than one of the protogalactic fragments contained a SMBH, the formation of SMBH binaries during galaxy assembly will be almost inevitable (e.g., Begelman et al. 1980). In a purely stellar background, as the binary separation decays, the effectiveness of dynamical friction slowly declines, and the pair can become tightly bound via three-body interactions, namely by capturing stars that pass close to the black holes and ejecting them at much higher velocities (e.g., Milosavljevi{\\' c} \\& Merritt 2001). If the hardening continues to sufficiently small relative distances, gravitational wave emission becomes the dominant source of orbital energy loss and the two SMBHs may coalesce in less than a Hubble time. However, the binary orbit may stop shrinking before gravitational radiation becomes relevant as there is a finite supply of stars on intersecting orbits (e.g., Berczik et al. 2005). During the assembly of galaxies, especially at high $z$, their SMBHs likely evolve within gas-rich environments. Merging systems such as the Ultraluminous Infrared Galaxies (ULIRGs) NGC 6240 and Arp 220 harbor large concentrations of gas, in excess of $10^9 \\Mo$, at their center, in the form of either a turbulent irregular structure or of a kinematically coherent, rotating disk (e.g., Downes \\& Solomon 1998). Massive rotating nuclear disks of molecular gas are also ubiquitous in galaxies that appear to have just undergone a major merger, such as Markarian 231 (Davies et al. 2004). Gas dynamics may thus profoundly affect the pairing of SMBHs both during and after their host galaxies merge (e.g., Escala et al. 2004; Kazantzidis et al. 2005). Recent simulations of the orbital evolution of SMBHs within an equilibrium, rotationally-supported, gaseous disk have shown that dynamical friction against the gaseous background leads to the formation of a tightly bound SMBH binary with a final separation of $<1$~pc in about $10^7$~yr (Escala et al. 2005; Dotti et al. 2006; Dotti et al., these proceedings). Here we review the results of high-resolution $N$-body + smoothed particle hydrodynamics (SPH) simulations of mergers between galaxies with central SMBHs having enough dynamic range to follow the black holes from hundreds of kiloparsecs down to sub-parsec scales, bridging more than ten orders of magnitude in density. ", "conclusions": "" }, "0807/0807.3745_arXiv.txt": { "abstract": "We revisit arguments that simple models of inflation with a small red tilt in the scalar power spectrum generically yield an observable tensor spectrum. We show that criteria for fine-tuning based upon the algebraic simplicity of the potential depend strongly upon the explicit assumptions they incorporate, particularly regarding the end of inflation. In addition, some models with algebraically simple potentials require carefully tuned initial field configurations, and not all types of fine-tuning are identifiable via the algebraic simplicity of the potential. Conversely, in the absence of a strong prior on the mechanism that ends inflation, we demonstrate the existence of potentials with vanishingly small tensor amplitudes which are natural in terms of both their algebraic form and initial conditions. We thus argue that proposed experiments (CMBPol or BBO) which make highly sensitive measurements of the tensor amplitude cannot definitively rule out the inflationary paradigm. ", "introduction": "In the simplest scenarios, inflation is driven by a single scalar field, the ``inflaton'', rolling in a slowly varying potential \\cite{Guth:1981,Sato:1981,Mukhanov:1981,Linde:1982,Albrecht:1982}. Sufficient inflation ($\\approx 60$ $e$-folds) smooths the universe and solves the flatness, horizon, and monopole problems. Initial inhomogeneities originate as quantum fluctuations in the inflaton field and the metric, sourcing scalar \\cite{Hawking:1982,Guth:1982, Starobinsky:1982, Bardeen:1983} and tensor \\cite{Grishchuk:1975, Starobinsky:1979} perturbations which are almost Gaussian and scale-invariant \\cite{Bardeen:1983}. These broad-brush predictions are consistent with observations (e.g. \\cite{Komatsu:2008hk,Peiris:2006ug,Peiris:2006sj,Peiris:2008be}). The shape of the inflaton potential is not predicted by theory; it is only required to satisfy the slow-roll conditions. However, the shape of the potential is constrained by the scalar spectral index, $n_s$, which measures the deviation of the scalar power spectrum from scale invariance, and the tensor-scalar ratio, $r$, which measures the amplitude of the tensor modes relative to that of the scalar modes. Data from the {\\sl Wilkinson Microwave Anisotropy Probe} (WMAP) combined with a compilation of supernovae and baryon acoustic oscillations data gives $n_s =0.960^{+0.014}_{-0.013}$ at 68\\% CL if the tensor-scalar ratio $r=0$, and $n_s=0.968 \\pm 0.015$ when we marginalize over $r$ \\cite{Dunkley:2008ie, Komatsu:2008hk}. There is currently no evidence for the existence of a primordial tensor background: the 95\\% CL upper limit from the same data compilation is $r<0.20$. A detection of primordial tensors would have tremendous implications for cosmology as their existence is a key prediction of inflation \\cite{Starobinsky:1979}. Moreover, measuring a non-zero value of $r$ would eliminate the ekpyrotic/cyclic scenarios, which predict an unobservably small primordial tensor amplitude \\cite{Khoury:2001bz,Boyle:2003km}. Finally, the observed amplitude of the tensor spectrum fixes the energy scale of inflation, removing a major uncertainty in inflationary model building. On the other hand, an observable tensor spectrum would imply the need for super-Planckian field variations during inflation in simple single field models \\cite{Lyth:1996im}, which raises specific theoretical challenges. Cosmic variance, polarized foregrounds, and weak lensing of the $E$-mode may ensure that $r \\sim 10^{-3}$ is the effective lower limit for a detection of tensor modes in the cosmic microwave background (CMB) (e.g. \\cite{Knox:2002pe,Kesden:2002,Verde:2005ff,Amarie:2005in}). Consequently, it is important to know whether inflation makes a {\\em generic\\/} prediction for the value of $r$ and, if so, whether it falls within the detectable range. A plethora of explicit inflationary models with varying degrees of physical motivation have been proposed. Many models do have observable values of $r$ (e.g. chaotic inflation \\cite{Linde:1983}); conversely, a detection of tensors would rule out many \\cite{Baumann:2006cd}, but not all \\cite{Dimopoulos:2005ac,Easther:2005zr,Silverstein:2008}, explicit string-theory constructions. Consequently, to show that inflation has a ``natural'' range of $r$ values, we must establish a ``weight'' that will give preference to certain classes of models. One weighting proposal is that of \\BST\\ (BST), who offer a fine-tuning statistic based on the algebraic simplicity of the potential, along with five conditions which a successful inflationary potential should satisfy. These criteria are then applied to quartic polynomial potentials, with the conclusion that a high degree of fine-tuning is required for $n_s>0.98$, or for $r<10^{-2}$ with $n_s>0.95$. If future observations confirm that $0.95 < n_s< 0.98$, the BST criterion suggests that $r$ is within the observable range. In addition to tuning the {\\em parameters\\/} of the potential, some potentials need a very special initial field configuration if inflation is to begin. This tuning can involve both the suppression of the kinetic energy relative to the potential energy (to ensure that the inflaton does not ``overshoot'' the inflationary portion of the potential) or the absence of inhomogeneities on the initial inflationary patch. This topic has been investigated for many of the common models of inflation \\cite{Belinsky:1985, Goldwirth:1990, Mendes:2000,Underwood:2008dh, Chiba:2008}. In particle physics, models are regarded as fine-tuned if they require correlations between parameters beyond those imposed by the symmetries of the underlying theory \\cite{tHooft:1979bh}, or when certain parameters are arbitrarily set to zero; models which are not fine-tuned in this sense are known as ``technically natural'' \\cite{Hotchkiss:2008sa}. There are other definitions of naturalness based on the sensitivity of observables to small changes in their parameters \\cite{Barbieri:1988, Anderson:1994dz}. Ref. \\cite{Hotchkiss:2008sa} examined technically natural potentials which did not obey all of the BST viability conditions, and found red-tilted examples with $r<0.01$ which were not fine-tuned under their definition. Other studies came to similar conclusions \\cite{Efstathiou:2006, Smith:2008pf}. Inflationary potentials are often written down without reference to an underlying theory, so we do not necessarily know the relevant symmetries, nor the couplings between the inflaton and other fields which contribute loop corrections to the inflaton potential. Consequently, while technical naturalness is a powerful tool, it is less useful when applied to generic inflationary potentials. Moreover, inflationary potentials typically require some tuning in order that inflation happens at all. We tackle two questions. Firstly, we show that even a mild relaxation of the rules imposed by BST significantly weakens their conclusions for the likely range of $r$. In addition, we ask whether the inflaton's initial conditions must be tuned in order for viable inflation to occur. We consider only the ratio of the kinetic to potential energy, treating the universe as a homogeneous system and ignoring the ability of primordial inhomogeneities to suppress the onset of inflation. However, this requirement places a strong constraint on the inflationary model space, which does not overlap with the BST proposal, demonstrating the interplay between the {\\em definition\\/} of tuning, and the resulting conclusions for the likely values of $r$ and $n_s$. Our aim is not to assess the intrinsic merit of different naturalness criteria. Rather, our purpose is to investigate whether conclusions about the natural value of $r$ are sensitive to assumptions encoded in the tuning criteria. This question is of considerable importance when assessing and justifying CMB $B$-mode or gravitational wave experiments sensitive to $r$. Likewise, any conclusions we reach regarding the ``natural'' value of $r$ in inflation will determine the significance of a null result from these experiments. ", "conclusions": "\\label{sec:conclusions} We have investigated whether, given the observed limits on the scalar tilt $n_s$, it is possible to claim that certain ranges of values for the tensor-scalar ratio $r$ are ``fine-tuned'' in a way that is independent of the definition of fine-tuning. We find that, on the contrary, different criteria are sensitive to different types of fine-tuning, and hence can give differing answers about the naturalness of a given potential. The BST criterion, based on the number of unnecessary ``features\" in the potential, tells us that models with $r<10^{-2}$ are fine-tuned for $n_s > 0.95$, while an alternative criterion based on sensitivity to initial conditions shows that models with $n_s<0.94$ and $r<10^{-3}$ are fine-tuned, but with $n_s>0.94$ there are ``natural'' models with $r \\sim 10^{-6}$. Moreover, BST's conclusions are not robust against relatively minor modifications to their criteria. In particular, the claim that models with $n_s>0.98$ are fine-tuned is softened by a small technical change in the fine-tuning criterion $Z_\\eta$: our upper limit on natural values of the spectral index is $n_s <0.99$. This may seem like a fine distinction, but recall that the breaking of scale invariance is measured by $n_s-1$, and thus differs by a factor of 2. Moreover, Planck is expected to measure $n_s$ with a 1-$\\sigma$ error of $0.0045$ \\cite{Planck} and $n_s = 0.99$ cannot be easily distinguished from $n_s = 1$, but $n_s=0.98$ could be detected with considerable confidence. Likewise, BST's conclusion that $n_s>0.95$ implies $r>10^{-2}$ in the absence of fine-tuning only holds if the potential evolves smoothly to an analytic minimum with $V(0) \\approx 0$. This is equivalent to requiring that the parameters fixing the shape of the potential as cosmological perturbations are laid down also govern the shape of the potential as inflation ends. While it is not inherently unreasonable, this restriction immediately excludes both hybrid inflation and many stringy constructions, even though their underlying potentials may be technically natural from a particle physics perspective. Pending a better understanding of physics at the energy scale of inflation, this condition therefore seems overly restrictive. Our alternative criterion is based on the sensitivity of a particular potential to the inflaton's initial conditions. It shows that potentials with $n_s<0.94$ and $r<10^{-3}$ are very sensitive to their initial conditions, as are all models with $r<10^{-6}$. Since inflation was originally designed to solve the initial condition problems of a hot big bang, one should certainly check that inflation itself does not require the universe to be in a special initial state. For our ``scoring'' of the initial conditions problem, there is a small preference for models with higher $r$, but the {\\em specific\\/} results we find depend on the explicit construction of our fine-tuning criterion, so we do not see this as a robust conclusion. Rather, we emphasize that this criterion singles out a different set of potentials from the BST criterion, and neither set of potentials is a strict subset of the other. The above conclusions apply to potentials that are well-described by a quartic potential. In Appendix \\ref{potentialparams}, we explore the regions of parameter space which yield untuned potentials. \\begin{figure*}[!htp] \\includegraphics[scale=0.175]{ri-05_10-zn-1.png} \\hfill \\includegraphics[scale=0.175]{cubic-both-05.png} \\\\ \\caption{(Top) The set of models which are deemed natural by both the BST and initial conditions criteria, i.e. $Z_\\eta<2$ and $R_i>0.5$. The color bar shows $R_i$. (Left) Imposing all five BST conditions. (Right) Neglecting selection conditions (\\ref{cond:min}) and (\\ref{cond:meta}). Potentials which have low $r \\sim 10^{-4}$ and $n_s \\sim 0.95$ generically contain a flat plateau terminating in a drop, similar to Case 2 in Figure~\\ref{fig:cases}.} \\label{combinedft} \\end{figure*} We have shown that conclusions drawn from BST-like criteria depend on the explicit assumptions they encode. We also discussed a second type of fine-tuning -- sensitivity to the initial field configuration -- which is distinct from that described of BST. One may restrict attention to models deemed acceptable by {\\em both\\/} criteria. The left panel of Figure~\\ref{combinedft} indicates that, at face value, this would lead one to conclude that the ``natural'' lower limit on the tensor amplitude is $r \\gtrsim 0.04$, and thus well within the range that is experimentally accessible. However, the right panel shows that this conclusion only holds if condition (\\ref{cond:min}) is imposed. Moreover, there are types of fine-tuning missed by our initial conditions criterion (e.g. initial inhomogeneity or spatial curvature); including these would further reduce the number of models for which inflation can be regarded as generic. The BST criterion aims to characterize the fine-tuning of the {\\it shape} of the potential. The initial conditions criterion might, at face value, appear to measure only the sensitivity to an initially large kinetic term. However, it is important to realize that it also encodes information about the functional form of the potential: for instance, a potential with a steep hill, a small plateau, and another steep downhill stretch will have difficulty supporting inflation without carefully tuned initial conditions. Thus, $R_i$ penalizes the functional form of the potential, but in a different way to $Z_\\eta$. Hence, Occam's razor cannot characterize fine-tuning (even of a single property, such as the shape of the potential) in an unequivocal way. These ambiguities will remain after the arrival of data with sufficient precision to select a small region of the $(r, n_s)$ plane. If this region is disfavored according to a specific fine-tuning criterion, should we interpret this as evidence against inflation, or even (in a more limited sense) as evidence against a class of inflationary models? Absent a detailed understanding of the physics of inflation, the answer is unclear, since we would not know whether the fine-tuning identified by that criterion is relevant to the underlying theory. As is well known, simple single field models of inflation have a direct correlation between the total excursion of the inflaton field and the value of $r$. If values of $r>10^{-3}$ are ruled out by experiment, this would also eliminate all single field models with $\\Delta \\phi>M_P$, with dramatic consequences for inflationary model building, as well as for alternative theories of the early universe \\cite{Khoury:2001bz,Boyle:2003km}. Many proposed inflationary models predict signals in this range but only a handful of models live in the lower part, the so-called ``tensor desert'' at $10^{-3}