Sample records for observed surface brightness

  1. Suzaku observations of low surface brightness cluster Abell 1631

    NASA Astrophysics Data System (ADS)

    Babazaki, Yasunori; Mitsuishi, Ikuyuki; Ota, Naomi; Sasaki, Shin; Böhringer, Hans; Chon, Gayoung; Pratt, Gabriel W.; Matsumoto, Hironori

    2018-04-01

    We present analysis results for a nearby galaxy cluster Abell 1631 at z = 0.046 using the X-ray observatory Suzaku. This cluster is categorized as a low X-ray surface brightness cluster. To study the dynamical state of the cluster, we conduct four-pointed Suzaku observations and investigate physical properties of the Mpc-scale hot gas associated with the A 1631 cluster for the first time. Unlike relaxed clusters, the X-ray image shows no strong peak at the center and an irregular morphology. We perform spectral analysis and investigate the radial profiles of the gas temperature, density, and entropy out to approximately 1.5 Mpc in the east, north, west, and south directions by combining with the XMM-Newton data archive. The measured gas density in the central region is relatively low (a few ×10-4 cm-3) at the given temperature (˜2.9 keV) compared with X-ray-selected clusters. The entropy profile and value within the central region (r < 0.1 r200) are found to be flatter and higher (≳400 keV cm2). The observed bolometric luminosity is approximately three times lower than that expected from the luminosity-temperature relation in previous studies of relaxed clusters. These features are also observed in another low surface brightness cluster, Abell 76. The spatial distributions of galaxies and the hot gas appear to be different. The X-ray luminosity is relatively lower than that expected from the velocity dispersion. A post-merger scenario may explain the observed results.

  2. Suzaku observations of low surface brightness cluster Abell 1631

    NASA Astrophysics Data System (ADS)

    Babazaki, Yasunori; Mitsuishi, Ikuyuki; Ota, Naomi; Sasaki, Shin; Böhringer, Hans; Chon, Gayoung; Pratt, Gabriel W.; Matsumoto, Hironori

    2018-06-01

    We present analysis results for a nearby galaxy cluster Abell 1631 at z = 0.046 using the X-ray observatory Suzaku. This cluster is categorized as a low X-ray surface brightness cluster. To study the dynamical state of the cluster, we conduct four-pointed Suzaku observations and investigate physical properties of the Mpc-scale hot gas associated with the A 1631 cluster for the first time. Unlike relaxed clusters, the X-ray image shows no strong peak at the center and an irregular morphology. We perform spectral analysis and investigate the radial profiles of the gas temperature, density, and entropy out to approximately 1.5 Mpc in the east, north, west, and south directions by combining with the XMM-Newton data archive. The measured gas density in the central region is relatively low (a few ×10-4 cm-3) at the given temperature (˜2.9 keV) compared with X-ray-selected clusters. The entropy profile and value within the central region (r < 0.1 r200) are found to be flatter and higher (≳400 keV cm2). The observed bolometric luminosity is approximately three times lower than that expected from the luminosity-temperature relation in previous studies of relaxed clusters. These features are also observed in another low surface brightness cluster, Abell 76. The spatial distributions of galaxies and the hot gas appear to be different. The X-ray luminosity is relatively lower than that expected from the velocity dispersion. A post-merger scenario may explain the observed results.

  3. K-band observations of boxy bulges - I. Morphology and surface brightness profiles

    NASA Astrophysics Data System (ADS)

    Bureau, M.; Aronica, G.; Athanassoula, E.; Dettmar, R.-J.; Bosma, A.; Freeman, K. C.

    2006-08-01

    In this first paper of a series on the structure of boxy and peanut-shaped (B/PS) bulges, Kn-band observations of a sample of 30 edge-on spiral galaxies are described and discussed. Kn-band observations best trace the dominant luminous galactic mass and are minimally affected by dust. Images, unsharp-masked images, as well as major-axis and vertically summed surface brightness profiles are presented and discussed. Galaxies with a B/PS bulge tend to have a more complex morphology than galaxies with other bulge types, more often showing centred or off-centred X structures, secondary maxima along the major-axis and spiral-like structures. While probably not uniquely related to bars, those features are observed in three-dimensional N-body simulations of barred discs and may trace the main bar orbit families. The surface brightness profiles of galaxies with a B/PS bulge are also more complex, typically containing three or more clearly separated regions, including a shallow or flat intermediate region (Freeman Type II profiles). The breaks in the profiles offer evidence for bar-driven transfer of angular momentum and radial redistribution of material. The profiles further suggest a rapid variation of the scaleheight of the disc material, contrary to conventional wisdom but again as expected from the vertical resonances and instabilities present in barred discs. Interestingly, the steep inner region of the surface brightness profiles is often shorter than the isophotally thick part of the galaxies, itself always shorter than the flat intermediate region of the profiles. The steep inner region is also much more prominent along the major-axis than in the vertically summed profiles. Similarly to other recent work but contrary to the standard `bulge + disc' model (where the bulge is both thick and steep), we thus propose that galaxies with a B/PS bulge are composed of a thin concentrated disc (a disc-like bulge) contained within a partially thick bar (the B/PS bulge), itself

  4. Giant Low Surface Brightness Galaxies

    NASA Astrophysics Data System (ADS)

    Mishra, Alka; Kantharia, Nimisha G.; Das, Mousumi

    2018-04-01

    In this paper, we present radio observations of the giant low surface brightness (LSB) galaxies made using the Giant Metrewave Radio Telescope (GMRT). LSB galaxies are generally large, dark matter dominated spirals that have low star formation efficiencies and large HI gas disks. Their properties suggest that they are less evolved compared to high surface brightness galaxies. We present GMRT emission maps of LSB galaxies with an optically-identified active nucleus. Using our radio data and archival near-infrared (2MASS) and near-ultraviolet (GALEX) data, we studied morphology and star formation efficiencies in these galaxies. All the galaxies show radio continuum emission mostly associated with the centre of the galaxy.

  5. Do Low Surface Brightness Galaxies Host Stellar Bars?

    NASA Astrophysics Data System (ADS)

    Cervantes Sodi, Bernardo; Sánchez García, Osbaldo

    2017-09-01

    With the aim of assessing if low surface brightness galaxies host stellar bars and by studying the dependence of the occurrence of bars as a function of surface brightness, we use the Galaxy Zoo 2 data set to construct a large volume-limited sample of galaxies and then segregate these galaxies as having low or high surface brightness in terms of their central surface brightness. We find that the fraction of low surface brightness galaxies hosting strong bars is systematically lower than that found for high surface brightness galaxies. The dependence of the bar fraction on the central surface brightness is mostly driven by a correlation of the surface brightness with the spin and the gas richness of the galaxies, showing only a minor dependence on the surface brightness. We also find that the length of the bars is strongly dependent on the surface brightness, and although some of this dependence is attributed to the gas content, even at a fixed gas-to-stellar mass ratio, high surface brightness galaxies host longer bars than their low surface brightness counterparts, which we attribute to an anticorrelation of the surface brightness with the spin.

  6. Do Low Surface Brightness Galaxies Host Stellar Bars?

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Cervantes Sodi, Bernardo; Sánchez García, Osbaldo, E-mail: b.cervantes@irya.unam.mx, E-mail: o.sanchez@irya.unam.mx

    With the aim of assessing if low surface brightness galaxies host stellar bars and by studying the dependence of the occurrence of bars as a function of surface brightness, we use the Galaxy Zoo 2 data set to construct a large volume-limited sample of galaxies and then segregate these galaxies as having low or high surface brightness in terms of their central surface brightness. We find that the fraction of low surface brightness galaxies hosting strong bars is systematically lower than that found for high surface brightness galaxies. The dependence of the bar fraction on the central surface brightness ismore » mostly driven by a correlation of the surface brightness with the spin and the gas richness of the galaxies, showing only a minor dependence on the surface brightness. We also find that the length of the bars is strongly dependent on the surface brightness, and although some of this dependence is attributed to the gas content, even at a fixed gas-to-stellar mass ratio, high surface brightness galaxies host longer bars than their low surface brightness counterparts, which we attribute to an anticorrelation of the surface brightness with the spin.« less

  7. Galaxy Selection and the Surface Brightness Distribution

    NASA Astrophysics Data System (ADS)

    McGaugh, Stacy S.; Bothun, Gregory D.; Schombert, James M.

    1995-08-01

    Optical surveys for galaxies are biased against the inclusion of low surface brightness (LSB) galaxies. Disney [Nature, 263,573(1976)] suggested that the constancy of disk central surface brightness noticed by Freeman [ApJ, 160,811(1970)] was not a physical result, but instead was an artifact of sample selection. Since LSB galaxies do exist, the pertinent and still controversial issue is if these newly discovered galaxies constitute a significant percentage of the general galaxy population. In this paper, we address this issue by determining the space density of galaxies as a function of disk central surface brightness. Using the physically reasonable assumption (which is motivated by the data) that central surface brightness is independent of disk scale length, we arrive at a distribution which is roughly flat (i.e., approximately equal numbers of galaxies at each surface brightness) faintwards of the Freeman (1970) value. Brightwards of this, we find a sharp decline in the distribution which is analogous to the turn down in the luminosity function at L^*^. An intrinsically sharply peaked "Freeman law" distribution can be completely ruled out, and no Gaussian distribution can fit the data. Low surface brightness galaxies (those with central surface brightness fainter than 22 B mag arcsec^-2^) comprise >~ 1/2 the general galaxy population, so a representative sample of galaxies at z = 0 does not really exist at present since past surveys have been insensitive to this component of the general galaxy population.

  8. Inferring Land Surface Model Parameters for the Assimilation of Satellite-Based L-Band Brightness Temperature Observations into a Soil Moisture Analysis System

    NASA Technical Reports Server (NTRS)

    Reichle, Rolf H.; De Lannoy, Gabrielle J. M.

    2012-01-01

    The Soil Moisture and Ocean Salinity (SMOS) satellite mission provides global measurements of L-band brightness temperatures at horizontal and vertical polarization and a variety of incidence angles that are sensitive to moisture and temperature conditions in the top few centimeters of the soil. These L-band observations can therefore be assimilated into a land surface model to obtain surface and root zone soil moisture estimates. As part of the observation operator, such an assimilation system requires a radiative transfer model (RTM) that converts geophysical fields (including soil moisture and soil temperature) into modeled L-band brightness temperatures. At the global scale, the RTM parameters and the climatological soil moisture conditions are still poorly known. Using look-up tables from the literature to estimate the RTM parameters usually results in modeled L-band brightness temperatures that are strongly biased against the SMOS observations, with biases varying regionally and seasonally. Such biases must be addressed within the land data assimilation system. In this presentation, the estimation of the RTM parameters is discussed for the NASA GEOS-5 land data assimilation system, which is based on the ensemble Kalman filter (EnKF) and the Catchment land surface model. In the GEOS-5 land data assimilation system, soil moisture and brightness temperature biases are addressed in three stages. First, the global soil properties and soil hydraulic parameters that are used in the Catchment model were revised to minimize the bias in the modeled soil moisture, as verified against available in situ soil moisture measurements. Second, key parameters of the "tau-omega" RTM were calibrated prior to data assimilation using an objective function that minimizes the climatological differences between the modeled L-band brightness temperatures and the corresponding SMOS observations. Calibrated parameters include soil roughness parameters, vegetation structure parameters

  9. A 5-micron-bright spot on Titan: evidence for surface diversity.

    PubMed

    Barnes, Jason W; Brown, Robert H; Turtle, Elizabeth P; McEwen, Alfred S; Lorenz, Ralph D; Janssen, Michael; Schaller, Emily L; Brown, Michael E; Buratti, Bonnie J; Sotin, Christophe; Griffith, Caitlin; Clark, Roger; Perry, Jason; Fussner, Stephanie; Barbara, John; West, Richard; Elachi, Charles; Bouchez, Antonin H; Roe, Henry G; Baines, Kevin H; Bellucci, Giancarlo; Bibring, Jean-Pierre; Capaccioni, Fabrizio; Cerroni, Priscilla; Combes, Michel; Coradini, Angioletta; Cruikshank, Dale P; Drossart, Pierre; Formisano, Vittorio; Jaumann, Ralf; Langevin, Yves; Matson, Dennis L; McCord, Thomas B; Nicholson, Phillip D; Sicardy, Bruno

    2005-10-07

    Observations from the Cassini Visual and Infrared Mapping Spectrometer show an anomalously bright spot on Titan located at 80 degrees W and 20 degrees S. This area is bright in reflected light at all observed wavelengths, but is most noticeable at 5 microns. The spot is associated with a surface albedo feature identified in images taken by the Cassini Imaging Science Subsystem. We discuss various hypotheses about the source of the spot, reaching the conclusion that the spot is probably due to variation in surface composition, perhaps associated with recent geophysical phenomena.

  10. A catalog of low surface brightness galaxies - List II

    NASA Technical Reports Server (NTRS)

    Schombert, James M.; Bothun, Gregory D.; Schneider, Stephen E.; Mcgaugh, Stacy S.

    1992-01-01

    A list of galaxies characterized by low surface brightness (LSB) is presented which facilitates the recognition of galaxies with brightnesses close to that of the sky. A total of 198 objects and 140 objects are listed in the primary and secondary catalogs respectively, and LSB galaxies are examined by means of H I redshift distributions. LSB disk galaxies are shown to have similar sizes and masses as the high-surface-brightness counterparts, and ellipticals and SOs are rarely encountered. Many LSB spirals have stellarlike nuclei, and most of the galaxies in the present catalog are late-type galaxies in the Sc, Sm, and Im classes. The LSB region of observational parameter space is shown to encompass a spectrum of types as full as that of the Hubble sequence. It is suggested that studies of LSB galaxies can provide important data regarding the formation and star-formation history of all galaxies.

  11. Planetary science: A 5-micron-bright spot on Titan: Evidence for surface diversity

    USGS Publications Warehouse

    Barnes, J.W.; Brown, R.H.; Turtle, E.P.; McEwen, A.S.; Lorenz, R.D.; Janssen, M.; Schaller, E.L.; Brown, M.E.; Buratti, B.J.; Sotin, Christophe; Griffith, C.; Clark, R.; Perry, J.; Fussner, S.; Barbara, J.; West, R.; Elachi, C.; Bouchez, A.H.; Roe, H.G.; Baines, K.H.; Bellucci, G.; Bibring, J.-P.; Capaccioni, F.; Cerroni, P.; Combes, M.; Coradini, A.; Cruikshank, D.P.; Drossart, P.; Formisano, V.; Jaumann, R.; Langevin, Y.; Matson, D.L.; McCord, T.B.; Nicholson, P.D.; Sicardy, B.

    2005-01-01

    Observations from the Cassini Visual and Infrared Mapping Spectrometer show an anomalously bright spot on Titan located at 80??W and 20??S. This area is bright in reflected tight at all observed wavelengths, but is most noticeable at 5 microns. The spot is associated with a surface albedo feature identified in images taken by the Cassini Imaging Science Subsystem. We discuss various hypotheses about the source of the spot, reaching the conclusion that the spot is probably due to variation in surface composition, perhaps associated with recent geophysical phenomena.

  12. The effect of monomolecular surface films on the microwave brightness temperature of the sea surface

    NASA Technical Reports Server (NTRS)

    Alpers, W.; Blume, H.-J. C.; Garrett, W. D.; Huehnerfuss, H.

    1982-01-01

    It is pointed out that monomolecular surface films of biological origin are often encountered on the ocean surface, especially in coastal regions. The thicknesses of the monomolecular films are of the order of 3 x 10 to the -9th m. Huehnerfuss et al. (1978, 1981) have shown that monomolecular surface films damp surface waves quite strongly in the centimeter to decimeter wavelength regime. Other effects caused by films are related to the reduction of the gas exchange at the air-sea interface and the decrease of the wind stress. The present investigation is concerned with experiments which reveal an unexpectedly large response of the microwave brightness temperature to a monomolecular oleyl alcohol slick at 1.43 GHz. Brightness temperature is a function of the complex dielectric constant of thy upper layer of the ocean. During six overflights over an ocean area covered with an artificial monomolecular alcohol film, a large decrease of the brightness temperature at the L-band was measured, while at the S-band almost no decrease was observed.

  13. Surface and Atmospheric Contributions to Passive Microwave Brightness Temperatures

    NASA Technical Reports Server (NTRS)

    Jackson, Gail Skofronick; Johnson, Benjamin T.

    2010-01-01

    Physically-based passive microwave precipitation retrieval algorithms require a set of relationships between satellite observed brightness temperatures (TB) and the physical state of the underlying atmosphere and surface. These relationships are typically non-linear, such that inversions are ill-posed especially over variable land surfaces. In order to better understand these relationships, this work presents a theoretical analysis using brightness temperature weighting functions to quantify the percentage of the TB resulting from absorption/emission/reflection from the surface, absorption/emission/scattering by liquid and frozen hydrometeors in the cloud, the emission from atmospheric water vapor, and other contributors. The results are presented for frequencies from 10 to 874 GHz and for several individual precipitation profiles as well as for three cloud resolving model simulations of falling snow. As expected, low frequency channels (<89 GHz) respond to liquid hydrometeors and the surface, while the higher frequency channels become increasingly sensitive to ice hydrometeors and the water vapor sounding channels react to water vapor in the atmosphere. Low emissivity surfaces (water and snow-covered land) permit energy downwelling from clouds to be reflected at the surface thereby increasing the percentage of the TB resulting from the hydrometeors. The slant path at a 53deg viewing angle increases the hydrometeor contributions relative to nadir viewing channels and show sensitivity to surface polarization effects. The TB percentage information presented in this paper answers questions about the relative contributions to the brightness temperatures and provides a key piece of information required to develop and improve precipitation retrievals over land surfaces.

  14. The local metallicity-surface brightness relationship in galactic disks

    NASA Technical Reports Server (NTRS)

    Ryder, Stuart D.

    1995-01-01

    We present the results of a first attempt to employ multiaperture masks to obtain spectrophotometry of H II regions in nearby galaxies. A total of 97 H II regions in six southern spiral galaxies were observed using a combination of multiaperture masks and conventional long-slit spectrophotometry. The oxygen abundances derived from the multiaperture mask observations using the empirical abundance diagnostic R(sub 23) are shown to be consistent with those from long-slit spectra and generally show better reproducibility and object definition. Although the number of objects that can be observed simultaneously with this particular system is still quite limited compared with either imaging spectrophotometry or fiber-fed spectrographs, the spectral resolution offered and high throughput in the blue help make multiaperture spectrophotometry a competitive technique for increasing the sampling of H II regions in both radial distance and luminosity. There is still no clear trend of abundance gradient with either the galaxy's luminosity or its Hubble type, although the extrapolated central abundance does appear to correlate with galaxy luminosity/mass. In order to avoid difficulty in choosing an appropriate normalizing radius, we instead plot the oxygen abundance against the underlying I-band surface brightness at the radial distance of the H II region and confirm the existence of a local metallicity-surface brightness reltaionship within the disks of spiral galaxies. Although the simple closed-boc model of galaxy evolution predicts almost the right form of this relationship, a more realistic multizone model employing expnentially decreasing gas infall provides a more satisfactory fit to the observational data, provided the expected enriched gas return from dying low-mass stars shedding their envelopes at late epochs is properly taken into account. This same model, with a star formation law based upon self-regulating star formation in a three-dimensional disk (Dopita & Ryder

  15. Thermal measurements of dark and bright surface features on Vesta as derived from Dawn/VIR

    USGS Publications Warehouse

    Tosi, Federico; Capria, Maria Teresa; De Sanctis, M.C.; Combe, J.-Ph.; Zambon, F.; Nathues, A.; Schröder, S.E.; Li, J.-Y.; Palomba, E.; Longobardo, A.; Blewett, D.T.; Denevi, B.W.; Palmer, E.; Capaccioni, F.; Ammannito, E.; Titus, Timothy N.; Mittlefehldt, D.W.; Sunshine, J.M.; Russell, C.T.; Raymond, C.A.; Dawn/VIR Team,

    2014-01-01

    Remote sensing data acquired during Dawn’s orbital mission at Vesta showed several local concentrations of high-albedo (bright) and low-albedo (dark) material units, in addition to spectrally distinct meteorite impact ejecta. The thermal behavior of such areas seen at local scale (1-10 km) is related to physical properties that can provide information about the origin of those materials. We use Dawn’s Visible and InfraRed (VIR) mapping spectrometer hyperspectral data to retrieve surface temperatures and emissivities, with high accuracy as long as temperatures are greater than 220 K. Some of the dark and bright features were observed multiple times by VIR in the various mission phases at variable spatial resolution, illumination and observation angles, local solar time, and heliocentric distance. This work presents the first temperature maps and spectral emissivities of several kilometer-scale dark and bright material units on Vesta. Results retrieved from the infrared data acquired by VIR show that bright regions generally correspond to regions with lower temperature, while dark regions correspond to areas with higher temperature. During maximum daily insolation and in the range of heliocentric distances explored by Dawn, i.e. 2.23-2.54 AU, the warmest dark unit found on Vesta rises to a temperature of 273 K, while bright units observed under comparable conditions do not exceed 266 K. Similarly, dark units appear to have higher emissivity on average compared to bright units. Dark-material units show a weak anticorrelation between temperature and albedo, whereas the relation is stronger for bright material units observed under the same conditions. Individual features may show either evanescent or distinct margins in the thermal images, as a consequence of the cohesion of the surface material. Finally, for the two categories of dark and bright materials, we were able to highlight the influence of heliocentric distance on surface temperatures, and estimate an

  16. Surface brightness profiles of 10 comets

    NASA Astrophysics Data System (ADS)

    Jewitt, D. C.; Meech, K. J.

    1987-06-01

    CCD photometric observations of the comae of 10 comets, obtained at the 4-m and 2.1-m telescopes at KPNO during 1985-1986 using filters centered at 700.5, 650.0, or 546.0 nm, are reported. The data are presented in extensive tables and graphs and characterized in detail. The radial surface brightness profiles are shown to be steeper than predicted by an idealized spherically symmetric steady-state comet model, the steepness increasing with the projected distance from the nucleus. These profiles are attributed, on the basis of Monte Carlo simulations, to imperfect coupling between the sublimated gas and the optically dominant grains of the coma.

  17. Skylab experiment SO73: Gegenschein/zodiacal light. [electrophotometry of surface brightness and polarization

    NASA Technical Reports Server (NTRS)

    Weinberg, J. L.

    1976-01-01

    A 10 color photoelectric polarimeter was used to measure the surface brightness and polarization associated with zodiacal light, background starlight, and spacecraft corona during each of the Skylab missions. Fixed position and sky scanning observations were obtained during Skylab missions SL-2 and SL-3 at 10 wavelenghts between 4000A and 8200A. Initial results from the fixed-position data are presented on the spacecraft corona and on the polarized brightness of the zodiacal light. Included among the fixed position regions that were observed are the north celestial pole, south ecliptic pole, two regions near the north galactic pole, and 90 deg from the sun in the ecliptic. The polarized brightness of the zodiacal light was found to have the color of the sun at each of these positions. Because previous observations found the total brightness to have the color of the sun from the near ultraviolet out to 2.4 micrometers, the degree of polarization of the zodiacal light is independent of wavelength from 4000A to 8200A.

  18. Erratum - the Lowest Surface Brightness Disc Galaxy Known

    NASA Astrophysics Data System (ADS)

    Davies, J. I.; Phillipps, S.; Disney, M. J.

    1988-11-01

    The paper "The lowest surface brightness disc galaxy known' by J.I. Davies, S. Phillipps and M.J. Disney was published in Mon. Not. R. astr. Soc. (1988), 231, 69p. The declination of the object given in section 2 of the paper is incorrect and should be changed to +19^deg^48'23". Thus the object cannot be identified with GP 1444 as in the original paper. To minimize confusion we propose to refer to the low surface brightness galaxy as GP 1444A.

  19. Low-Surface-Brightness Galaxies: Hidden Galaxies Revealed

    NASA Astrophysics Data System (ADS)

    Bothun, G.; Impey, C.; McGaugh, S.

    1997-07-01

    In twenty years, low surface brightness (LSB) galaxies have evolved from being an idiosyncratic notion to being one of the major baryonic repositories in the Universe. The story of their discovery and the characterization of their properties is told here. Their recovery from the noise of the night sky background is a strong testament to the severity of surface brightness selection effects. LSB galaxies have a number of remarkable properties which distinguish them from the more familiar Hubble Sequence of spirals. The two most important are 1) they evolve at a significantly slower rate and may well experience star formation outside of the molecular cloud environment, 2) they are embedded in dark matter halos which are of lower density and more extended than the halos around high surface brightness (HSB) disk galaxies. Compared to HSB disks, LSB disks are strongly dark matter dominated at all radii and show a systematic increase in $M/L$ with decreasing central surface brightness. In addition, the recognition that large numbers of LSB galaxies actually exist has changed the form of the galaxy luminosity function and has clearly increased the space density of galaxies at z =0. Recent CCD surveys have uncovered a population of red LSB disks that may be related to the excess of faint blue galaxies detected at moderate redshifts. LSB galaxies offer us a new window into galaxy evolution and formation which is every bit as important as those processes which have produced easy to detect galaxies. Indeed, the apparent youth of some LSB galaxies suggest that galaxy formation is a greatly extended process. While the discovery of LSB galaxies have lead to new insights, it remains unwise to presume that we now have a representative sample which encompasses all galaxy types and forms. (SECTION: Invited Review Paper)

  20. The X-ray surface brightness distribution and spectral properties of six early-type galaxies

    NASA Technical Reports Server (NTRS)

    Trinchieri, G.; Fabbiano, G.; Canizares, C. R.

    1986-01-01

    Detailed analysis is presented of the Einstein X-ray observations of six early-type galaxies. The results show that effective cooling is probably present in these systems, at least in the innermost regions. Interaction with the surrounding medium has a major effect on the X-ray surface brightness distribution at large radii, at least for galaxies in clusters. The data do not warrant the general assumptions of isothermality and gravitational hydrostatic equilibrium at large radii. Comparison of the X-ray surface brightness profiles with model predictions indicate that 1/r-squared halos with masses of the order of 10 times the stellar masses are required to match the data. The physical model of White and Chevalier (1984) for steady cooling flows in a King law potential with no heavy halo gives a surface brightness distribution that resembles the data if supernovae heating is present.

  1. IMPACT OF SUPERNOVA AND COSMIC-RAY DRIVING ON THE SURFACE BRIGHTNESS OF THE GALACTIC HALO IN SOFT X-RAYS

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Peters, Thomas; Girichidis, Philipp; Gatto, Andrea

    2015-11-10

    The halo of the Milky Way contains a hot plasma with a surface brightness in soft X-rays of the order 10{sup −12} erg cm{sup −2} s{sup −1} deg{sup −2}. The origin of this gas is unclear, but so far numerical models of galactic star formation have failed to reproduce such a large surface brightness by several orders of magnitude. In this paper, we analyze simulations of the turbulent, magnetized, multi-phase interstellar medium including thermal feedback by supernova explosions as well as cosmic-ray feedback. We include a time-dependent chemical network, self-shielding by gas and dust, and self-gravity. Pure thermal feedback alonemore » is sufficient to produce the observed surface brightness, although it is very sensitive to the supernova rate. Cosmic rays suppress this sensitivity and reduce the surface brightness because they drive cooler outflows. Self-gravity has by far the largest effect because it accumulates the diffuse gas in the disk in dense clumps and filaments, so that supernovae exploding in voids can eject a large amount of hot gas into the halo. This can boost the surface brightness by several orders of magnitude. Although our simulations do not reach a steady state, all simulations produce surface brightness values of the same order of magnitude as the observations, with the exact value depending sensitively on the simulation parameters. We conclude that star formation feedback alone is sufficient to explain the origin of the hot halo gas, but measurements of the surface brightness alone do not provide useful diagnostics for the study of galactic star formation.« less

  2. Extracting Galaxy Cluster Gas Inhomogeneity from X-Ray Surface Brightness: A Statistical Approach and Application to Abell 3667

    NASA Astrophysics Data System (ADS)

    Kawahara, Hajime; Reese, Erik D.; Kitayama, Tetsu; Sasaki, Shin; Suto, Yasushi

    2008-11-01

    Our previous analysis indicates that small-scale fluctuations in the intracluster medium (ICM) from cosmological hydrodynamic simulations follow the lognormal probability density function. In order to test the lognormal nature of the ICM directly against X-ray observations of galaxy clusters, we develop a method of extracting statistical information about the three-dimensional properties of the fluctuations from the two-dimensional X-ray surface brightness. We first create a set of synthetic clusters with lognormal fluctuations around their mean profile given by spherical isothermal β-models, later considering polytropic temperature profiles as well. Performing mock observations of these synthetic clusters, we find that the resulting X-ray surface brightness fluctuations also follow the lognormal distribution fairly well. Systematic analysis of the synthetic clusters provides an empirical relation between the three-dimensional density fluctuations and the two-dimensional X-ray surface brightness. We analyze Chandra observations of the galaxy cluster Abell 3667, and find that its X-ray surface brightness fluctuations follow the lognormal distribution. While the lognormal model was originally motivated by cosmological hydrodynamic simulations, this is the first observational confirmation of the lognormal signature in a real cluster. Finally we check the synthetic cluster results against clusters from cosmological hydrodynamic simulations. As a result of the complex structure exhibited by simulated clusters, the empirical relation between the two- and three-dimensional fluctuation properties calibrated with synthetic clusters when applied to simulated clusters shows large scatter. Nevertheless we are able to reproduce the true value of the fluctuation amplitude of simulated clusters within a factor of 2 from their two-dimensional X-ray surface brightness alone. Our current methodology combined with existing observational data is useful in describing and inferring the

  3. Analysis of Mass Profiles and Cooling Flows of Bright, Early-Type Galaxies AO2, AO3 and Surface Brightness Profiles and Energetics of Intracluster Gas in Cool Galaxy Clusters AO3

    NASA Technical Reports Server (NTRS)

    White, Raymond E., III

    1998-01-01

    This final report uses ROSAT observations to analyze two different studies. These studies are: Analysis of Mass Profiles and Cooling Flows of Bright, Early-Type Galaxies; and Surface Brightness Profiles and Energetics of Intracluster Gas in Cool Galaxy Clusters.

  4. Stellar Surface Brightness Profiles of Dwarf Galaxies

    NASA Astrophysics Data System (ADS)

    Herrmann, Kimberly A.; LITTLE THINGS Team

    2012-01-01

    Radial stellar surface brightness profiles of spiral galaxies can be classified into three types: (I) single exponential, (II) truncated: the light falls off with one exponential out to a break radius and then falls off more steeply, and (III) anti-truncated: the light falls off with one exponential out to a break radius and then falls off less steeply. Stellar surface brightness profile breaks are also found in dwarf disk galaxies, but with an additional category: (FI) flat-inside: the light is roughly constant or increasing and then falls off beyond a break. We have been re-examining the multi-wavelength stellar disk profiles of 141 dwarf galaxies, primarily from Hunter & Elmegreen (2006, 2004). Each dwarf has data in up to 11 wavelength bands: FUV and NUV from GALEX, UBVJHK and H-alpha from ground-based observations, and 3.6 and 4.5 microns from Spitzer. In this talk, I will highlight results from a semi-automatic fitting of this data set, including: (1) statistics of break locations and other properties as a function of wavelength and profile type, (2) color trends and radial mass distribution as a function of profile type, and (3) the relationship of the break radius to the kinematics and density profiles of atomic hydrogen gas in the 41 dwarfs of the LITTLE THINGS subsample. We gratefully acknowledge funding for this research from the National Science Foundation (AST-0707563).

  5. The Barnes-Evans color-surface brightness relation: A preliminary theoretical interpretation

    NASA Technical Reports Server (NTRS)

    Shipman, H. L.

    1980-01-01

    Model atmosphere calculations are used to assess whether an empirically derived relation between V-R and surface brightness is independent of a variety of stellar paramters, including surface gravity. This relationship is used in a variety of applications, including the determination of the distances of Cepheid variables using a method based on the Beade-Wesselink method. It is concluded that the use of a main sequence relation between V-R color and surface brightness in determining radii of giant stars is subject to systematic errors that are smaller than 10% in the determination of a radius or distance for temperature cooler than 12,000 K. The error in white dwarf radii determined from a main sequence color surface brightness relation is roughly 10%.

  6. Titan's Surface Brightness Temperatures and H2 Mole Fraction from Cassini CIRS

    NASA Technical Reports Server (NTRS)

    Jennings, Donald E.; Flasar, F. M.; Kunde, V. G.; Samuelson, R. E.; Pearl, J. C.; Nixon, C. A.; Carlson, R. C.; Mamoutkine, A. A.; Brasunas, J. C.; Guandique, E.; hide

    2008-01-01

    The atmosphere of Titan has a spectral window of low opacity around 530/cm in the thermal infrared where radiation from the surface can be detected from space. The Composite Infrared spectrometer1 (CIRS) uses this window to measure the surface brightness temperature of Titan. By combining all observations from the Cassini tour it is possible to go beyond previous Voyager IRIS studies in latitude mapping of surface temperature. CIRS finds an average equatorial surface brightness temperature of 93.7+/-0.6 K, which is close to the 93.65+/-0.25 K value measured at the surface by Huygens HASi. The temperature decreases toward the poles, reaching 91.6+/-0.7 K at 90 S and 90.0+/-1.0 K at 87 N. The temperature distribution is centered in latitude at approximately 12 S, consistent with Titan's season of late northern winter. Near the equator the temperature varies with longitude and is higher in the trailing hemisphere, where the lower albedo may lead to relatively greater surface heating5. Modeling of radiances at 590/cm constrains the atmospheric H2 mole fraction to 0.12+/-0.06 %, in agreement with results from Voyager iris.

  7. The Fundamental Plane and the Surface Brightness Test for the Expansion of the Universe

    NASA Astrophysics Data System (ADS)

    Kjaergaard, Per; Jorgensen, Inger; Moles, Mariano

    1993-12-01

    We have determined the Petrosian radius, rη , and the enclosed mean surface brightness within the Petrosian radius, <μ>η, for 33 elliptical and S0 galaxies in the Coma cluster from new accurate CCD surface photometry. For the Petrosian parameter η = 1.39, rη and <μ>η are compared with the effective radius, re, and the effective mean surface brightness, <μ>e derived from fitting a de Vaucouleurs law. The fundamental plane (FP) expressed using rη and <μ>η is the same as the FP found by Jørgensen, Franx, & Kjaergaard (1993) using re and <μ>e. The FP can be used to predict the mean surface brightness within the effective radius or the corresponding Petrosian radius (η = 1.39) with an uncertainty of ±0.14 mag for Coma cluster ellipticals. Thus the FP, applied to clusters, appears to be a suitable tool for performing the surface brightness test (SBT) for the expansion of the universe. We suggest that instead of correcting individual galaxies to some standard conditions, e.g., the same metric radius, the fundamental plane itself should be considered the standard. It is argued that the metric size enclosing around 75% of the total light represents a reasonable compromise between resolution and faint level detection when performing the SBT. This radius could be derived as the Petrosian radius corresponding to η = 2.0 or from a global fit to that part of the observed profile which encompasses 75% of the total light. In case both small and large galaxies are well described by a de Vaucouleurs law the global fit can be performed on a smaller central part of the brightness profile. The use of the FP involves the time consuming determinations of velocity dispersions. We find that <μ>η (η = 1.39) can be predicted from the log rη alone with an accuracy of 0.3 mag for the Coma cluster ellipticals. Our discussion of the various error contributions to the predicted mean surface brightness for faint cluster ellipticals at redshifts z < 0.5 shows that the final

  8. Temporal observations of bright soil exposures at Gusev crater, Mars

    USGS Publications Warehouse

    Rice, M.S.; Bell, J.F.; Cloutis, E.A.; Wray, J.J.; Herkenhoff, K. E.; Sullivan, R.; Johnson, J. R.; Anderson, R.B.

    2011-01-01

    The Mars Exploration Rover Spirit has discovered bright soil deposits in its wheel tracks that previously have been confirmed to contain ferric sulfates and/or opaline silica. Repeated Pancam multispectral observations have been acquired at four of these deposits to monitor spectral and textural changes over time during exposure to Martian surface conditions. Previous studies suggested that temporal spectral changes occur because of mineralogic changes (e.g., phase transitions accompanying dehydration). In this study, we present a multispectral and temporal analysis of eight Pancam image sequences at the Tyrone exposure, three at the Gertrude Weise exposure, two at the Kit Carson exposure, and ten at the Ulysses exposure that have been acquired as of sol 2132 (1 January 2010). We compare observed variations in Pancam data to spectral changes predicted by laboratory experiments for the dehydration of ferric sulfates. We also present a spectral analysis of repeated Mars Reconnaissance Orbiter HiRISE observations spanning 32 sols and a textural analysis of Spirit Microscopic Imager observations of Ulysses spanning 102 sols. At all bright soil exposures, we observe no statistically significant spectral changes with time that are uniquely diagnostic of dehydration and/or mineralogic phase changes. However, at Kit Carson and Ulysses, we observe significant textural changes, including slumping within the wheel trench, movement of individual grains, disappearance of fines, and dispersal of soil clods. All observed textural changes are consistent with aeolian sorting and/or minor amounts of air fall dust deposition.

  9. GLOBAL PROPERTIES OF M31'S STELLAR HALO FROM THE SPLASH SURVEY. I. SURFACE BRIGHTNESS PROFILE

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Gilbert, Karoline M.; Guhathakurta, Puragra; Beaton, Rachael L.

    2012-11-20

    We present the surface brightness profile of M31's stellar halo out to a projected radius of 175 kpc. The surface brightness estimates are based on confirmed samples of M31 red giant branch stars derived from Keck/DEIMOS spectroscopic observations. A set of empirical spectroscopic and photometric M31 membership diagnostics is used to identify and reject foreground and background contaminants. This enables us to trace the stellar halo of M31 to larger projected distances and fainter surface brightnesses than previous photometric studies. The surface brightness profile of M31's halo follows a power law with index -2.2 {+-} 0.2 and extends to amore » projected distance of at least {approx}175 kpc ({approx}2/3 of M31's virial radius), with no evidence of a downward break at large radii. The best-fit elliptical isophotes have b/a = 0.94 with the major axis of the halo aligned along the minor axis of M31's disk, consistent with a prolate halo, although the data are also consistent with M31's halo having spherical symmetry. The fact that tidal debris features are kinematically cold is used to identify substructure in the spectroscopic fields out to projected radii of 90 kpc and investigate the effect of this substructure on the surface brightness profile. The scatter in the surface brightness profile is reduced when kinematically identified tidal debris features in M31 are statistically subtracted; the remaining profile indicates that a comparatively diffuse stellar component to M31's stellar halo exists to large distances. Beyond 90 kpc, kinematically cold tidal debris features cannot be identified due to small number statistics; nevertheless, the significant field-to-field variation in surface brightness beyond 90 kpc suggests that the outermost region of M31's halo is also comprised to a significant degree of stars stripped from accreted objects.« less

  10. New Observations of Subarcsecond Photospheric Bright Points

    NASA Technical Reports Server (NTRS)

    Berger, T. E.; Schrijver, C. J.; Shine, R. A.; Tarbell, T. D.; Title, A. M.; Scharmer, G.

    1995-01-01

    We have used an interference filter centered at 4305 A within the bandhead of the CH radical (the 'G band') and real-time image selection at the Swedish Vacuum Solar Telescope on La Palma to produce very high contrast images of subarcsecond photospheric bright points at all locations on the solar disk. During the 6 day period of 1993 September 15-20 we observed active region NOAA 7581 from its appearance on the East limb to a near-disk-center position on September 20. A total of 1804 bright points were selected for analysis from the disk center image using feature extraction image processing techniques. The measured Full Width at Half Maximum (FWHM) distribution of the bright points in the image is lognormal with a modal value of 220 km (0 sec .30) and an average value of 250 km (0 sec .35). The smallest measured bright point diameter is 120 km (0 sec .17) and the largest is 600 km (O sec .69). Approximately 60% of the measured bright points are circular (eccentricity approx. 1.0), the average eccentricity is 1.5, and the maximum eccentricity corresponding to filigree in the image is 6.5. The peak contrast of the measured bright points is normally distributed. The contrast distribution variance is much greater than the measurement accuracy, indicating a large spread in intrinsic bright-point contrast. When referenced to an averaged 'quiet-Sun' area in the image, the modal contrast is 29% and the maximum value is 75%; when referenced to an average intergranular lane brightness in the image, the distribution has a modal value of 61% and a maximum of 119%. The bin-averaged contrast of G-band bright points is constant across the entire measured size range. The measured area of the bright points, corrected for pixelation and selection effects, covers about 1.8% of the total image area. Large pores and micropores occupy an additional 2% of the image area, implying a total area fraction of magnetic proxy features in the image of 3.8%. We discuss the implications of this

  11. New Observations of Subarcsecond Photospheric Bright Points

    NASA Technical Reports Server (NTRS)

    Berger, T. E.; Schrijver, C. J.; Shine, R. A.; Tarbell, T. D.; Title, A. M.; Scharmer, G.

    1995-01-01

    We have used an interference filter centered at 4305 A within the bandhead of the CH radical (the 'G band') and real-time image selection at the Swedish Vacuum Solar Telescope on La Palma to produce very high contrast images of subarcsecond photospheric bright points at all locations on the solar disk. During the 6 day period of 15-20 Sept. 1993 we observed active region NOAA 7581 from its appearance on the East limb to a near-disk-center position on 20 Sept. A total of 1804 bright points were selected for analysis from the disk center image using feature extraction image processing techniques. The measured FWHM distribution of the bright points in the image is lognormal with a modal value of 220 km (0.30 sec) and an average value of 250 km (0.35 sec). The smallest measured bright point diameter is 120 km (0.17 sec) and the largest is 600 km (O.69 sec). Approximately 60% of the measured bright points are circular (eccentricity approx. 1.0), the average eccentricity is 1.5, and the maximum eccentricity corresponding to filigree in the image is 6.5. The peak contrast of the measured bright points is normally distributed. The contrast distribution variance is much greater than the measurement accuracy, indicating a large spread in intrinsic bright-point contrast. When referenced to an averaged 'quiet-Sun' area in the image, the modal contrast is 29% and the maximum value is 75%; when referenced to an average intergranular lane brightness in the image, the distribution has a modal value of 61% and a maximum of 119%. The bin-averaged contrast of G-band bright points is constant across the entire measured size range. The measured area of the bright points, corrected for pixelation and selection effects, covers about 1.8% of the total image area. Large pores and micropores occupy an additional 2% of the image area, implying a total area fraction of magnetic proxy features in the image of 3.8%. We discuss the implications of this area fraction measurement in the context of

  12. Metre-size bright spots at the surface of comet 67P/Churyumov-Gerasimenko: Interpretation of OSIRIS data using laboratory experiments

    NASA Astrophysics Data System (ADS)

    Pommerol, Antoine; Thomas, Nicolas; Antonella Barucci, M.; Bertaux, Jean-Loup; Davidsson, Björn; Ramy El-Maarry, Mohamed; La Forgia, Fiorengela; Fornasier, Sonia; Gracia, Antonio; Groussin, Olivier; Jost, Bernhard; Keller, Horst Uwe; Kuehrt, Ekkehard; Marschall, Raphael; Massironi, Matteo; Motolla, Stefano; Naletto, Giampiero; Oklay, Nilda; Pajola, Maurizio; Poch, Olivier

    2015-04-01

    Since the beginning of Rosetta's orbital observations, over a hundred small bright spots have been identified in images returned by its OSIRIS NAC camera, in all types of morphological regions on the nucleus. Bright spots are found as clusters of several tens of individuals in the vicinity of cliffs, or isolated without clear structural relation to the surrounding terrain. They are however mostly observed in the areas of the nucleus currently receiving the lowest amount of insolation and some of the best examples appear completely surrounded by shadows. Their typical sizes are of the order of a few metres and they are often observed at the surfaces of boulders of larger dimension. The brightness of these spots is up to ten times the average brightness of the surrounding terrain and multi-spectral analyses show a significantly bluer spectrum over the 0.3-1µm range. Comparisons of images taken in September and November 2014 under similar illumination conditions do not show any significant change of these features. Analysis of the results of past and present laboratory experiments with H2O-ice/dust mixtures provide interesting insights about the nature and origin of the bright spots. In particular, recent sublimation experiments conducted at the University of Bern reproduce the spectro-photometric variability observed at the surface of the nucleus by sequences of formation and ejection of a mantle of refractory organic-rich dust at the surface of the icy material. The formation of hardened layers of ice by sintering/re-condensation below the uppermost dust layer can also have strong implications for both the photometric and mechanical properties of the subsurface layer. Based on the comparison between OSIRIS observations and laboratory results, our favoured interpretation of the observed features is that the bright spots are exposures of water ice, resulting from the removal of the uppermost layer of refractory dust that covers the rest of the nucleus. Some of the

  13. Using SMOS brightness temperature and derived surface-soil moisture to characterize surface conditions and validate land surface models.

    NASA Astrophysics Data System (ADS)

    Polcher, Jan; Barella-Ortiz, Anaïs; Piles, Maria; Gelati, Emiliano; de Rosnay, Patricia

    2017-04-01

    The SMOS satellite, operated by ESA, observes the surface in the L-band. On continental surface these observations are sensitive to moisture and in particular surface-soil moisture (SSM). In this presentation we will explore how the observations of this satellite can be exploited over the Iberian Peninsula by comparing its results with two land surface models : ORCHIDEE and HTESSEL. Measured and modelled brightness temperatures show a good agreement in their temporal evolution, but their spatial structures are not consistent. An empirical orthogonal function analysis of the brightness temperature's error identifies a dominant structure over the south-west of the Iberian Peninsula which evolves during the year and is maximum in autumn and winter. Hypotheses concerning forcing-induced biases and assumptions made in the radiative transfer model are analysed to explain this inconsistency, but no candidate is found to be responsible for the weak spatial correlations. The analysis of spatial inconsistencies between modelled and measured TBs is important, as these can affect the estimation of geophysical variables and TB assimilation in operational models, as well as result in misleading validation studies. When comparing the surface-soil moisture of the models with the product derived operationally by ESA from SMOS observations similar results are found. The spatial correlation over the IP between SMOS and ORCHIDEE SSM estimates is poor (ρ 0.3). A single value decomposition (SVD) analysis of rainfall and SSM shows that the co-varying patterns of these variables are in reasonable agreement between both products. Moreover the first three SVD soil moisture patterns explain over 80% of the SSM variance simulated by the model while the explained fraction is only 52% of the remotely sensed values. These results suggest that the rainfall-driven soil moisture variability may not account for the poor spatial correlation between SMOS and ORCHIDEE products. Other reasons have to

  14. Directional Emissivity Effects on Martian Surface Brightness Temperatures

    NASA Astrophysics Data System (ADS)

    Pitman, K. M.; Wolff, M. J.; Bandfield, J. L.; Clancy, R. T.; Clayton, G. C.

    2001-11-01

    The angular dependence of thermal emission from the surface of Mars has not been well characterized. Although nadir sequences constitute most of the MGS/TES Martian surface observations [1,2], a significant number scans of Martian surfaces at multiple emission angles (emission phase function (EPF) sequences) also exist. Such data can provide insight into surface structures, thermal inertias, and non-isotropic corrections to thermal emission measurements [3]. The availability of abundant EPF data as well as the added utility of such observations for atmospheric characterization provide the impetus for examining the phenomenon of directional emissivity. We present examples of directional emissivity effects on brightness temperature spectra for a variety of typical Martian surfaces. We examine the theoretical development by Hapke (1993, 1996) [4,5] and compare his algorithm to that of Mishchenko et al. (1999) [6]. These results are then compared to relevant TES EPF data. This work is supported through NASA grant NAGS-9820 (MJW) and JPL contract no. 961471 (RTC). [1] Smith et al. (1998), AAS-DPS meeting # 30, # 11.P07. [2] Kieffer, Mullins, & Titus (1998), EOS, 79, 533. [3] Jakosky, Finiol, & Henderson (1990), JGR, 17, 985--988. [4] Hapke, B. (1993), Theory of Reflectance & Emittance Spectroscopy, Cambridge Univ. Press, NY. [5] Hapke, B. (1996), JGR, 101, E7, 16817--16831. [6] Mishchenko et al. (1999), JQSRT, 63, 409--432.

  15. Temporal observations of bright soil exposures at Gusev crater, Mars

    USGS Publications Warehouse

    Rice, M.S.; Bell, J.F.; Cloutis, E.A.; Wray, J.J.; Herkenhoff, K. E.; Sullivan, R.; Johnson, J. R.; Anderson, R.B.

    2011-01-01

    The Mars Exploration Rover Spirit has discovered bright soil deposits in its wheel tracks that previously have been confirmed to contain ferric sulfates and/or opaline silica. Repeated Pancam multispectral observations have been acquired at four of these deposits to monitor spectral and textural changes over time during exposure to Martian surface conditions. Previous studies suggested that temporal spectral changes occur because of mineralogic changes (e.g., phase transitions accompanying dehydration). In this study, we present a multispectral and temporal analysis of eight Pancam image sequences at the Tyrone exposure, three at the Gertrude Weise exposure, two at the Kit Carson exposure, and ten at the Ulysses exposure that have been acquired as of sol 2132 (1 January 2010). We compare observed variations in Pancam data to spectral changes predicted by laboratory experiments for the dehydration of ferric sulfates. We also present a spectral analysis of repeated Mars Reconnaissance Orbiter HiRISE observations spanning 32 sols and a textural analysis of Spirit Microscopic Imager observations of Ulysses spanning 102 sols. At all bright soil exposures, we observe no statistically significant spectral changes with time that are uniquely diagnostic of dehydration and/or mineralogic phase changes. However, at Kit Carson and Ulysses, we observe significant textural changes, including slumping within the wheel trench, movement of individual grains, disappearance of fines, and dispersal of soil clods. All observed textural changes are consistent with aeolian sorting and/or minor amounts of air fall dust deposition. Copyright 2011 by the American Geophysical Union.

  16. Global Properties of M31's Stellar Halo from the SPLASH Survey. I. Surface Brightness Profile

    NASA Astrophysics Data System (ADS)

    Gilbert, Karoline M.; Guhathakurta, Puragra; Beaton, Rachael L.; Bullock, James; Geha, Marla C.; Kalirai, Jason S.; Kirby, Evan N.; Majewski, Steven R.; Ostheimer, James C.; Patterson, Richard J.; Tollerud, Erik J.; Tanaka, Mikito; Chiba, Masashi

    2012-11-01

    We present the surface brightness profile of M31's stellar halo out to a projected radius of 175 kpc. The surface brightness estimates are based on confirmed samples of M31 red giant branch stars derived from Keck/DEIMOS spectroscopic observations. A set of empirical spectroscopic and photometric M31 membership diagnostics is used to identify and reject foreground and background contaminants. This enables us to trace the stellar halo of M31 to larger projected distances and fainter surface brightnesses than previous photometric studies. The surface brightness profile of M31's halo follows a power law with index -2.2 ± 0.2 and extends to a projected distance of at least ~175 kpc (~2/3 of M31's virial radius), with no evidence of a downward break at large radii. The best-fit elliptical isophotes have b/a = 0.94 with the major axis of the halo aligned along the minor axis of M31's disk, consistent with a prolate halo, although the data are also consistent with M31's halo having spherical symmetry. The fact that tidal debris features are kinematically cold is used to identify substructure in the spectroscopic fields out to projected radii of 90 kpc and investigate the effect of this substructure on the surface brightness profile. The scatter in the surface brightness profile is reduced when kinematically identified tidal debris features in M31 are statistically subtracted; the remaining profile indicates that a comparatively diffuse stellar component to M31's stellar halo exists to large distances. Beyond 90 kpc, kinematically cold tidal debris features cannot be identified due to small number statistics; nevertheless, the significant field-to-field variation in surface brightness beyond 90 kpc suggests that the outermost region of M31's halo is also comprised to a significant degree of stars stripped from accreted objects. The data presented herein were obtained at the W. M. Keck Observatory, which is operated as a scientific partnership among the California

  17. Simulated X-ray galaxy clusters at the virial radius: Slopes of the gas density, temperature and surface brightness profiles

    NASA Astrophysics Data System (ADS)

    Roncarelli, M.; Ettori, S.; Dolag, K.; Moscardini, L.; Borgani, S.; Murante, G.

    2006-12-01

    Using a set of hydrodynamical simulations of nine galaxy clusters with masses in the range 1.5 × 1014 < Mvir < 3.4 × 1015Msolar, we have studied the density, temperature and X-ray surface brightness profiles of the intracluster medium in the regions around the virial radius. We have analysed the profiles in the radial range well above the cluster core, the physics of which are still unclear and matter of tension between simulated and observed properties, and up to the virial radius and beyond, where present observations are unable to provide any constraints. We have modelled the radial profiles between 0.3R200 and 3R200 with power laws with one index, two indexes and a rolling index. The simulated temperature and [0.5-2] keV surface brightness profiles well reproduce the observed behaviours outside the core. The shape of all these profiles in the radial range considered depends mainly on the activity of the gravitational collapse, with no significant difference among models including extraphysics. The profiles steepen in the outskirts, with the slope of the power-law fit that changes from -2.5 to -3.4 in the gas density, from -0.5 to -1.8 in the gas temperature and from -3.5 to -5.0 in the X-ray soft surface brightness. We predict that the gas density, temperature and [0.5-2] keV surface brightness values at R200 are, on average, 0.05, 0.60, 0.008 times the measured values at 0.3R200. At 2R200, these values decrease by an order of magnitude in the gas density and surface brightness, by a factor of 2 in the temperature, putting stringent limits on the detectable properties of the intracluster-medium (ICM) in the virial regions.

  18. Chemical abundances in low surface brightness galaxies: Implications for their evolution

    NASA Technical Reports Server (NTRS)

    Mcgaugh, S. S.; Bothun, G. D.

    1993-01-01

    Low Surface Brightness (LSB) galaxies are an important but often neglected part of the galaxy content of the universe. Their importance stems both from the selection effects which cause them to be under-represented in galaxy catalogs, and from what they can tell us about the physical processes of galaxy evolution that has resulted in something other than the traditional Hubble sequence of spirals. An important constraint for any evolutionary model is the present day chemical abundances of LSB disks. Towards this end, spectra for a sample of 75 H 2 regions distributed in 20 LSB disks galaxies were obtained. Structurally, this sample is defined as having B(0) fainter than 23.0 mag arcsec(sup -2) and scale lengths that cluster either around 3 kpc or 10 kpc. In fact, structurally, these galaxies are very similar to the high surface brightness spirals which define the Hubble sequence. Thus, our sample galaxies are not dwarf galaxies but instead have masses comparable to or in excess of the Milky Way. The basic results from these observations are summarized.

  19. The distribution of star formation and metals in the low surface brightness galaxy UGC 628

    NASA Astrophysics Data System (ADS)

    Young, J. E.; Kuzio de Naray, Rachel; Wang, Sharon X.

    2015-09-01

    We introduce the MUSCEL Programme (MUltiwavelength observations of the Structure, Chemistry and Evolution of LSB galaxies), a project aimed at determining the star-formation histories of low surface brightness galaxies. MUSCEL utilizes ground-based optical spectra and space-based UV and IR photometry to fully constrain the star-formation histories of our targets with the aim of shedding light on the processes that led low surface brightness galaxies down a different evolutionary path from that followed by high surface brightness galaxies, such as our Milky Way. Here we present the spatially resolved optical spectra of UGC 628, observed with the VIRUS-P IFU at the 2.7-m Harlen J. Smith Telescope at the McDonald Observatory, and utilize emission-line diagnostics to determine the rate and distribution of star formation as well as the gas-phase metallicity and metallicity gradient. We find highly clustered star formation throughout UGC 628, excluding the core regions, and a log(O/H) metallicity around -4.2, with more metal-rich regions near the edges of the galactic disc. Based on the emission-line diagnostics alone, the current mode of star formation, slow and concentrated in the outer disc, appears to have dominated for quite some time, although there are clear signs of a much older stellar population formed in a more standard inside-out fashion.

  20. Surface Brightness Test and Plasma Redshift

    NASA Astrophysics Data System (ADS)

    Brynjolfsson, Ari

    2006-03-01

    The plasma redshift of photons in a hot sparse plasma follows from basic axioms of physics. It has no adjustable parameters (arXiv:astro-ph/0406437). Both the distance-redshift relation and the magnitude-redshift relation for supernovae and galaxies are well-defined functions of the average electron densities in intergalactic space. We have previously shown that the predictions of the magnitude-redshift relation in plasma- redshift cosmology match well the observed relations for the type Ia supernovae (SNe). No adjustable parameters such as the time variable ``dark energy'' and ``dark matter'' are needed. We have also shown that plasma redshift cosmology predicts well the intensity and black body spectrum of the cosmic microwave background (CMB). Plasma redshift explains also the spectrum below and above the 2.73 K black body CMB, and the X-ray background. In the following, we will show that the good observations and analyses of the relation between surface brightness and redshift for galaxies, as determined by Allan Sandage and Lori M. Lubin in 2001, are well predicted by the plasma redshift. All these relations are inconsistent with cosmic time dilation and the contemporary big-bang cosmology.

  1. Truncated disc surface brightness profiles produced by flares

    NASA Astrophysics Data System (ADS)

    Borlaff, Alejandro; Eliche-Moral, M. Carmen; Beckman, John; Font, Joan

    2017-03-01

    Previous studies have discarded that flares in galactic discs may explain the truncation that are frequently observed in highly-inclined galaxies (Kregel et al. 2002). However, no study has systematically analysed this hypothesis using realistic models for the disc, the flare and the bulge. We derive edge-on and face-on surface brightness profiles for a series of realistic galaxy models with flared discs that sample a wide range of structural and photometric parameters across the Hubble Sequence, accordingly to observations. The surface brightness profile for each galaxy model has been simulated for edge-on and face-on views to find out whether the flared disc produces a significant truncation in the disc in the edge-on view compared to the face-on view or not. In order to simulate realistic images of disc galaxies, we have considered the observational distribution of the photometric parameters as a function of the morphological type for three mass bins (10 < log10(M/M ⊙) < 10.7, 10.7 < log10(M/M ⊙) < 11 and log10(M/M ⊙) > 11), and four morphological type bins (S0-Sa, Sb-Sbc, Sc-Scd and Sd-Sdm). For each mass bin, we have restricted the photometric and structural parameters of each modelled galaxy to their characteristic observational ranges (μ0, disc, μeff, bulge, B/T, M abs, r eff, n bulge, h R, disc) and the flare in the disc (h z, disc/h R, disc, ∂h z, disc/∂R, see de Grijs & Peletier 1997, Graham 2001, López-Corredoira et al. 2002, Yoachim & Dalcanton 2006, Bizyaev et al. 2014, Mosenkov et al. 2015). Contrary to previous claims, the simulations show that realistic flared disks can be responsible for the truncations observed in many edge-on systems, preserving the profile of the non-flared analogous model in face-on view. These breaks reproduce the properties of the weak-to-intermediate breaks observed in many real Type-II galaxies in the diagram relating the radial location of the break (R brkII) in units of the inner disk scale-length with the

  2. Observations During GRIP from HIRAD: Images of C-Band Brightness Temperatures and Ocean Surface Wind Speed and Rain Rate

    NASA Technical Reports Server (NTRS)

    Miller, Timothy L.; James, M. W.; Jones, W. L.; Ruf, C. S.; Uhlhorn, E. W.; Biswas, S.; May, C.; Shah, G.; Black, P.; Buckley, C. D.

    2012-01-01

    HIRAD (Hurricane Imaging Radiometer) flew on the WB-57 during NASA s GRIP (Genesis and Rapid Intensification Processes) campaign in August - September of 2010. HIRAD is a new C-band radiometer using a synthetic thinned array radiometer (STAR) technology to obtain cross-track resolution of approximately 3 degrees, out to approximately 60 degrees to each side of nadir. By obtaining measurements of emissions at 4, 5, 6, and 6.6 GHz, observations of ocean surface wind speed and rain rate can be inferred. This technique has been used for many years by precursor instruments, including the Stepped Frequency Microwave Radiometer (SFMR), which has been flying on the NOAA and USAF hurricane reconnaissance aircraft for several years. The advantage of HIRAD over SFMR is that HIRAD can observe a +/- 60-degree swath, rather than a single footprint at nadir angle. Results from the flights during the GRIP campaign will be shown, including images of brightness temperatures, wind speed, and rain rate. To the extent possible, comparisons will be made with observations from other instruments on the GRIP campaign, for which HIRAD observations are either directly comparable or are complementary. Features such as storm eye and eyewall, location of vortex wind and rain maxima, and indications of dynamical features such as the merging of a weaker outer wind/rain maximum with the main vortex may be seen in the data. Potential impacts on operational ocean surface wind analyses and on numerical weather forecasts will also be discussed.

  3. Stellar populations of bulges in galaxies with a low surface-brightness disc

    NASA Astrophysics Data System (ADS)

    Morelli, L.; Corsini, E. M.; Pizzella, A.; Dalla Bontà, E.; Coccato, L.; Méndez-Abreu, J.

    2015-03-01

    The radial profiles of the Hβ, Mg, and Fe line-strength indices are presented for a sample of eight spiral galaxies with a low surface-brightness stellar disc and a bulge. The correlations between the central values of the line-strength indices and velocity dispersion are consistent to those known for early-type galaxies and bulges of high surface-brightness galaxies. The age, metallicity, and α/Fe enhancement of the stellar populations in the bulge-dominated region are obtained using stellar population models with variable element abundance ratios. Almost all the sample bulges are characterized by a young stellar population, on-going star formation, and a solar α/Fe enhancement. Their metallicity spans from high to sub-solar values. No significant gradient in age and α/Fe enhancement is measured, whereas only in a few cases a negative metallicity gradient is found. These properties suggest that a pure dissipative collapse is not able to explain formation of all the sample bulges and that other phenomena, like mergers or acquisition events, need to be invoked. Such a picture is also supported by the lack of a correlation between the central value and gradient of the metallicity in bulges with very low metallicity. The stellar populations of the bulges hosted by low surface-brightness discs share many properties with those of high surface-brightness galaxies. Therefore, they are likely to have common formation scenarios and evolution histories. A strong interplay between bulges and discs is ruled out by the fact that in spite of being hosted by discs with extremely different properties, the bulges of low and high surface-brightness discs are remarkably similar.

  4. Effects of cloud size and cloud particles on satellite-observed reflected brightness

    NASA Technical Reports Server (NTRS)

    Reynolds, D. W.; Mckee, T. B.; Danielson, K. S.

    1978-01-01

    Satellite observations allowed obtaining data on the visible brightness of cumulus clouds over South Park, Colorado, while aircraft observations were made in cloud to obtain the drop size distributions and liquid water content of the cloud. Attention is focused on evaluating the relationship between cloud brightness, horizontal dimension, and internal microphysical structure. A Monte Carlo cloud model for finite clouds was run using different distributions of drop sizes and numbers, while varying the cloud depth and width to determine how theory would predict what the satellite would view from its given location in space. Comparison of these results to the satellite observed reflectances is presented. Theoretical results are found to be in good agreement with observations. For clouds of optical thickness between 20 and 60, monitoring cloud brightness changes in clouds of uniform depth and variable width gives adequate information about a cloud's liquid water content. A cloud having a 10:1 width to depth ratio is almost reaching its maximum brightness for a specified optical thickness.

  5. Titan's Surface Temperatures Maps from Cassini - CIRS Observations

    NASA Astrophysics Data System (ADS)

    Cottini, Valeria; Nixon, C. A.; Jennings, D. E.; Anderson, C. M.; Samuelson, R. E.; Irwin, P. G. J.; Flasar, F. M.

    2009-09-01

    The Cassini Composite Infrared Spectrometer (CIRS) observations of Saturn's largest moon, Titan, are providing us with the ability to detect the surface temperature of the planet by studying its outgoing radiance through a spectral window in the thermal infrared at 19 μm (530 cm-1) characterized by low opacity. Since the first acquisitions of CIRS Titan data the instrument has gathered a large amount of spectra covering a wide range of latitudes, longitudes and local times. We retrieve the surface temperature and the atmospheric temperature profile by modeling proper zonally averaged spectra of nadir observations with radiative transfer computations. Our forward model uses the correlated-k approximation for spectral opacity to calculate the emitted radiance, including contributions from collision induced pairs of CH4, N2 and H2, haze, and gaseous emission lines (Irwin et al. 2008). The retrieval method uses a non-linear least-squares optimal estimation technique to iteratively adjust the model parameters to achieve a spectral fit (Rodgers 2000). We show an accurate selection of the wide amount of data available in terms of footprint diameter on the planet and observational conditions, together with the retrieved results. Our results represent formal retrievals of surface brightness temperatures from the Cassini CIRS dataset using a full radiative transfer treatment, and we compare to the earlier findings of Jennings et al. (2009). In future, application of our methodology over wide areas should greatly increase the planet coverage and accuracy of our knowledge of Titan's surface brightness temperature. References: Irwin, P.G.J., et al.: "The NEMESIS planetary atmosphere radiative transfer and retrieval tool" (2008). JQSRT, Vol. 109, pp. 1136-1150, 2008. Rodgers, C. D.: "Inverse Methods For Atmospheric Sounding: Theory and Practice". World Scientific, Singapore, 2000. Jennings, D.E., et al.: "Titan's Surface Brightness Temperatures." Ap. J. L., Vol. 691, pp. L103-L

  6. The GALEX/S4G Surface Brightness and Color Profiles Catalog. I. Surface Photometry and Color Gradients of Galaxies

    NASA Astrophysics Data System (ADS)

    Bouquin, Alexandre Y. K.; Gil de Paz, Armando; Muñoz-Mateos, Juan Carlos; Boissier, Samuel; Sheth, Kartik; Zaritsky, Dennis; Peletier, Reynier F.; Knapen, Johan H.; Gallego, Jesús

    2018-02-01

    We present new spatially resolved surface photometry in the far-ultraviolet (FUV) and near-ultraviolet (NUV) from images obtained by the Galaxy Evolution Explorer (GALEX) and IRAC1 (3.6 μm) photometry from the Spitzer Survey of Stellar Structure in Galaxies (S4G). We analyze the radial surface brightness profiles μ FUV, μ NUV, and μ [3.6], as well as the radial profiles of (FUV ‑ NUV), (NUV ‑ [3.6]), and (FUV ‑ [3.6]) colors in 1931 nearby galaxies (z < 0.01). The analysis of the 3.6 μm surface brightness profiles also allows us to separate the bulge and disk components in a quasi-automatic way and to compare their light and color distribution with those predicted by the chemo-spectrophotometric models for the evolution of galaxy disks of Boissier & Prantzos. The exponential disk component is best isolated by setting an inner radial cutoff and an upper surface brightness limit in stellar mass surface density. The best-fitting models to the measured scale length and central surface brightness values yield distributions of spin and circular velocity within a factor of two of those obtained via direct kinematic measurements. We find that at a surface brightness fainter than μ [3.6] = 20.89 mag arcsec‑2, or below 3 × 108 M ⊙ kpc‑2 in stellar mass surface density, the average specific star formation rate (sSFR) for star-forming and quiescent galaxies remains relatively flat with radius. However, a large fraction of GALEX Green Valley galaxies show a radial decrease in sSFR. This behavior suggests that an outside-in damping mechanism, possibly related to environmental effects, could be testimony of an early evolution of galaxies from the blue sequence of star-forming galaxies toward the red sequence of quiescent galaxies.

  7. Calculations of microwave brightness temperature of rough soil surfaces: Bare field

    NASA Technical Reports Server (NTRS)

    Mo, T.; Schmugge, T. J.; Wang, J. R.

    1985-01-01

    A model for simulating the brightness temperatures of soils with rough surfaces is developed. The surface emissivity of the soil media is obtained by the integration of the bistatic scattering coefficients for rough surfaces. The roughness of a soil surface is characterized by two parameters, the surface height standard deviation sigma and its horizontal correlation length l. The model calculations are compared to the measured angular variations of the polarized brightness temperatures at both 1.4 GHz and 5 GHz frequences. A nonlinear least-squares fitting method is used to obtain the values of delta and l that best characterize the surface roughness. The effect of shadowing is incorporated by introducing a function S(theta), which represents the probability that a point on a rough surface is not shadowed by other parts of the surface. The model results for the horizontal polarization are in excellent agreement with the data. However, for the vertical polarization, some discrepancies exist between the calculations and data, particularly at the 1.4 GHz frequency. Possible causes of the discrepancy are discussed.

  8. New Observations of C-band Brightness Temperatures and Ocean Surface Wind Speed and Rain Rate From the Hurricane Imaging Radiometer (HIRAD)

    NASA Technical Reports Server (NTRS)

    Miller, Timothy L.; James, M. W.; Roberts, J. B.; Buckley, C. D.; Biswas, S.; May, C.; Ruf, C. S.; Uhlhorn, E. W.; Atlas, R.; Black, P.; hide

    2012-01-01

    HIRAD flew on the WB-57 during NASA's GRIP (Genesis and Rapid Intensification Processes) campaign in August September of 2010. HIRAD is a new C-band radiometer using a synthetic thinned array radiometer (STAR) technology to obtain cross-track resolution of approximately 3 degrees, out to approximately 60 degrees to each side of nadir. By obtaining measurements of emissions at 4, 5, 6, and 6.6 GHz, observations of ocean surface wind speed and rain rate can be retrieved. This technique has been used for many years by precursor instruments, including the Stepped Frequency Microwave Radiometer (SFMR), which has been flying on the NOAA and USAF hurricane reconnaissance aircraft for several years to obtain observations within a single footprint at nadir angle. Results from the flights during the GRIP campaign will be shown, including images of brightness temperatures, wind speed, and rain rate. Comparisons will be made with observations from other instruments on the GRIP campaign, for which HIRAD observations are either directly comparable or are complementary. Features such as storm eye and eyewall, location of storm wind and rain maxima, and indications of dynamical features such as the merging of a weaker outer wind/rain maximum with the main vortex may be seen in the data. Potential impacts on operational ocean surface wind analyses and on numerical weather forecasts will also be discussed.

  9. First Observation of Bright Solitons in Bulk Superfluid ^{4}He.

    PubMed

    Ancilotto, Francesco; Levy, David; Pimentel, Jessica; Eloranta, Jussi

    2018-01-19

    The existence of bright solitons in bulk superfluid ^{4}He is demonstrated by time-resolved shadowgraph imaging experiments and density functional theory (DFT) calculations. The initial liquid compression that leads to the creation of nonlinear waves is produced by rapidly expanding plasma from laser ablation. After the leading dissipative period, these waves transform into bright solitons, which exhibit three characteristic features: dispersionless propagation, negligible interaction in a two-wave collision, and direct dependence between soliton amplitude and the propagation velocity. The experimental observations are supported by DFT calculations, which show rapid evolution of the initially compressed liquid into bright solitons. At high amplitudes, solitons become unstable and break down into dispersive shock waves.

  10. Unveiling the Low Surface Brightness Stellar Peripheries of Galaxies

    NASA Astrophysics Data System (ADS)

    Ferguson, Annette M. N.

    2018-01-01

    The low surface brightness peripheral regions of galaxies contain a gold mine of information about how minor mergers and accretions have influenced their evolution over cosmic time. Enormous stellar envelopes and copious amounts of faint tidal debris are natural outcomes of the hierarchical assembly process and the search for and study of these features, albeit highly challenging, offers the potential for unrivalled insight into the mechanisms of galaxy growth. Over the last two decades, there has been burgeoning interest in probing galaxy outskirts using resolved stellar populations. Wide-field surveys have uncovered vast tidal debris features and new populations of very remote globular clusters, while deep Hubble Space Telescope photometry has provided exquisite star formation histories back to the earliest epochs. I will highlight some recent results from studies within and beyond the Local Group and conclude by briefly discussing the great potential of future facilities, such as JWST, Euclid, LSST and WFIRST, for major breakthroughs in low surface brightness galaxy periphery science.

  11. Hi-C Observations of Penumbral Bright Dots

    NASA Astrophysics Data System (ADS)

    Alpert, S.; Tiwari, S. K.; Moore, R. L.; Savage, S. L.; Winebarger, A. R.

    2014-12-01

    We use high-quality data obtained by the High Resolution Coronal Imager (Hi-C) to examine bright dots (BDs) in a sunspot's penumbra. The sizes of these BDs are on the order of 1 arcsecond (1") and are therefore hard to identify using the Atmospheric Imaging Assembly's (AIA) 0.6" pixel-1 resolution. These BDs become readily apparent with Hi-C's 0.1" pixel-1 resolution. Tian et al. (2014) found penumbral BDs in the transition region (TR) by using the Interface Region Imaging Spectrograph (IRIS). However, only a few of their dots could be associated with any enhanced brightness in AIA channels. In this work, we examine the characteristics of the penumbral BDs observed by Hi-C in a sunspot penumbra, including their sizes, lifetimes, speeds, and intensity. We also attempt to relate these BDs to the IRIS BDs. There are fewer Hi-C BDs in the penumbra than seen by IRIS, though different sunspots were studied. We use 193Å Hi-C data from July 11, 2012 which observed from ~18:52:00 UT--18:56:00 UT and supplement it with data from AIA's 193Å passband to see the complete lifetime of the dots that were born before and/or lasted longer than Hi-C's 5-minute observation period. We use additional AIA passbands and compare the light curves of the BDs at different temperatures to test whether the Hi-C BDs are TR BDs. We find that most Hi-C BDs show clear movement, and of those that do, they move in a radial direction, toward or away from the sunspot umbra. Single BDs interact with other BDs, combining to fade away or brighten. The BDs that do not interact with other BDs tend to move less. Our BDs are similar to the exceptional IRIS BDs: they move slower on average and their sizes and lifetimes are on the high end of the distribution of IRIS BDs. We infer that our penumbral BDs are some of the larger BDs observed by IRIS, those that are bright enough in TR emission to be seen in the 193Å band of Hi-C.

  12. Illuminating Low Surface Brightness Galaxies with the Hyper Suprime-Cam Survey

    NASA Astrophysics Data System (ADS)

    Greco, Johnny P.; Greene, Jenny E.; Strauss, Michael A.; Macarthur, Lauren A.; Flowers, Xzavier; Goulding, Andy D.; Huang, Song; Kim, Ji Hoon; Komiyama, Yutaka; Leauthaud, Alexie; Leisman, Lukas; Lupton, Robert H.; Sifón, Cristóbal; Wang, Shiang-Yu

    2018-04-01

    We present a catalog of extended low surface brightness galaxies (LSBGs) identified in the Wide layer of the Hyper Suprime-Cam Subaru Strategic Program (HSC-SSP). Using the first ∼200 deg2 of the survey, we have uncovered 781 LSBGs, spanning red (g ‑ i ≥ 0.64) and blue (g ‑ i < 0.64) colors and a wide range of morphologies. Since we focus on extended galaxies (r eff = 2.″5–14″), our sample is likely dominated by low-redshift objects. We define LSBGs to have mean surface brightnesses {\\bar{μ }}eff}(g)> 24.3 mag arcsec‑2, which allows nucleated galaxies into our sample. As a result, the central surface brightness distribution spans a wide range of μ 0(g) = 18–27.4 mag arcsec‑2, with 50% and 95% of galaxies fainter than 24.3 and 22 mag arcsec‑2, respectively. Furthermore, the surface brightness distribution is a strong function of color, with the red distribution being much broader and generally fainter than that of the blue LSBGs, and this trend shows a clear correlation with galaxy morphology. Red LSBGs typically have smooth light profiles that are well characterized by single-component Sérsic functions. In contrast, blue LSBGs tend to have irregular morphologies and show evidence for ongoing star formation. We cross-match our sample with existing optical, H I, and ultraviolet catalogs to gain insight into the physical nature of the LSBGs. We find that our sample is diverse, ranging from dwarf spheroidals and ultradiffuse galaxies in nearby groups to gas-rich irregulars to giant LSB spirals, demonstrating the potential of the HSC-SSP to provide a truly unprecedented view of the LSBG population.

  13. The surface brightness of reflection nebulae. Ph.D. Thesis, Dec. 1972

    NASA Technical Reports Server (NTRS)

    Rush, W. F.

    1974-01-01

    Hubble's equation relating the maximum apparent angular extent of a reflection nebula to the apparent magnitude of the illuminating star has been reconsidered under a set of less restrictive assumptions. A computational technique is developed which permits the use of fits to observed m, log a values to determine the albedo of the particles composing reflection nebulae, providing only that one assumes a particular phase function. Despite the fact that all orders of scattering, anisotropic phase functions, and illumination by the general stellar field are considered, the albedo which is determined for reflection nebulae by this method appears larger than that for interstellar particles in general. The possibility that the higher surface brightness might be due to a continuous fluorescence mechanism is considered both theoretically and observationally.

  14. Exploring the extremely low surface brightness sky: distances to 23 newly discovered objects in Dragonfly fields

    NASA Astrophysics Data System (ADS)

    van Dokkum, Pieter

    2016-10-01

    We are obtaining deep, wide field images of nearby galaxies with the Dragonfly Telephoto Array. This telescope is optimized for low surface brightness imaging, and we are finding many low surface brightness objects in the Dragonfly fields. In Cycle 22 we obtained ACS imaging for 7 galaxies that we had discovered in a Dragonfly image of the galaxy M101. Unexpectedly, the ACS data show that only 3 of the galaxies are members of the M101 group, and the other 4 are very large Ultra Diffuse Galaxies (UDGs) at much greater distance. Building on our Cycle 22 program, here we request ACS imaging for 23 newly discovered low surface brightness objects in four Dragonfly fields centered on the galaxies NGC 1052, NGC 1084, NGC 3384, and NGC 4258. The immediate goals are to construct the satellite luminosity functions in these four fields and to constrain the number density of UDGs that are not in rich clusters. More generally, this complete sample of extremely low surface brightness objects provides the first systematic insight into galaxies whose brightness peaks at >25 mag/arcsec^2.

  15. The nucleus of Comet Borrelly: A study of morphology and surface brightness

    USGS Publications Warehouse

    Oberst, J.; Howington-Kraus, E.; Kirk, R.; Soderblom, L.; Buratti, B.; Hicks, M.; Nelson, R.; Britt, D.

    2004-01-01

    Stereo images obtained during the DS1 flyby were analyzed to derive a topographic model for the nucleus of Comet 19P/Borrelly for morphologic and photometric studies. The elongated nucleus has an overall concave shape, resembling a peanut, with the lower end tilted towards the camera. The bimodal character of surface-slopes and curvatures support the idea that the nucleus is a gravitational aggregate, consisting of two fragments in contact. Our photometric modeling suggests that topographic shading effects on Borrelly's surface are very minor (<10%) at the given resolution of the terrain model. Instead, albedo effects are thought to dominate Borrelly's large variations in surface brightness. With 90% of the visible surface having single scattering albedos between 0.008 and 0.024, Borrelly is confirmed to be among the darkest of the known Solar System objects. Photometrically corrected images emphasize that the nucleus has distinct, contiguous terrains covered with either bright or dark, smooth or mottled materials. Also, mapping of the changes in surface brightness with phase angle suggests that terrain roughness at subpixel scale is not uniform over the nucleus. High surface roughness is noted in particular near the transition between the upper and lower end of the nucleus, as well as near the presumed source region of Borrelly's main jets. Borrelly's surface is complex and characterized by distinct types of materials that have different compositional and/or physical properties. ?? 2003 Elsevier Inc. All rights reserved.

  16. Observations of C-band Brightness Temperatures and Ocean Surface Wind Speed and Rain Rate from the Hurricane Imaging Radiometer (HIRAD)

    NASA Technical Reports Server (NTRS)

    Miller, Timothy L.; James, M. W.; Roberts, J. B.; Jones, W. L.; May, C.; Ruf, C. S.; Uhlhorn, E. W.; Atlas, R.; Black, P.

    2012-01-01

    HIRAD flew on the WB-57 over Earl and Karl during NASA s GRIP (Genesis and Rapid Intensification Processes) campaign in August - September of 2010. HIRAD is a new Cband radiometer using a synthetic thinned array radiometer (STAR) technology to obtain cross-track resolution of approximately 3 degrees, out to approximately 60 degrees to each side of nadir. (The resulting swath width for a platform at 60,000 feet is roughly 60 km, and resolution for most of the swath is around 2 km.) By obtaining measurements of emissions at 4, 5, 6, and 6.6 GHz, observations of ocean surface wind speed and rain rate can be retrieved. This technique has been used for many years by precursor instruments, including the Stepped Frequency Microwave Radiometer (SFMR), which has been flying on the NOAA and USAF hurricane reconnaissance aircraft for several years to obtain observations within a single footprint at nadir angle. Results from the flights during the GRIP campaign will be shown, including images of brightness temperatures, wind speed, and rain rate. Comparisons will be made with observations from other instruments on the GRIP campaign, for which HIRAD observations are either directly comparable or are complementary. Features such as storm eye and eyewall, location of storm wind and rain maxima, and indications of dynamical features such as the merging of a weaker outer wind/rain maximum with the main vortex may be seen in the data. Potential impacts on operational ocean surface wind analyses and on numerical weather forecasts will also be discussed.

  17. Posterior uncertainty of GEOS-5 L-band radiative transfer model parameters and brightness temperatures after calibration with SMOS observations

    NASA Astrophysics Data System (ADS)

    De Lannoy, G. J.; Reichle, R. H.; Vrugt, J. A.

    2012-12-01

    Simulated L-band (1.4 GHz) brightness temperatures are very sensitive to the values of the parameters in the radiative transfer model (RTM). We assess the optimum RTM parameter values and their (posterior) uncertainty in the Goddard Earth Observing System (GEOS-5) land surface model using observations of multi-angular brightness temperature over North America from the Soil Moisture Ocean Salinity (SMOS) mission. Two different parameter estimation methods are being compared: (i) a particle swarm optimization (PSO) approach, and (ii) an MCMC simulation procedure using the differential evolution adaptive Metropolis (DREAM) algorithm. Our results demonstrate that both methods provide similar "optimal" parameter values. Yet, DREAM exhibits better convergence properties, resulting in a reduced spread of the posterior ensemble. The posterior parameter distributions derived with both methods are used for predictive uncertainty estimation of brightness temperature. This presentation will highlight our model-data synthesis framework and summarize our initial findings.

  18. The Tolman Surface Brightness Test for the Reality of the Expansion. III. Hubble Space Telescope Profile and Surface Brightness Data for Early-Type Galaxies in Three High-Redshift Clusters

    NASA Astrophysics Data System (ADS)

    Lubin, Lori M.; Sandage, Allan

    2001-09-01

    Photometric data for 34 early-type galaxies in the three high-redshift clusters Cl 1324+3011 (z=0.76), Cl 1604+4304 (z=0.90), and Cl 1604+4321 (z=0.92), observed with the Hubble Space Telescope (HST) and with the Keck 10 m telescopes by Oke, Postman, & Lubin, are analyzed to obtain the photometric parameters of mean surface brightness, magnitudes for the growth curves, and angular radii at various Petrosian η radii. The angular radii at η=1.3 mag for the program galaxies are all larger than 0.24". All the galaxies are well resolved at this angular size using HST, whose point-spread function is 0.05", half-width at half-maximum. The data for each of the program galaxies are listed at η=1.0, 1.3, 1.5, 1.7, and 2.0 mag. They are corrected by color equations and K-terms for the effects of redshift to the rest-frame Cape/Cousins I for Cl 1324+3011 and Cl 1604+4304 and R for Cl 1604+4321. The K-corrections are calculated from synthetic spectral energy distributions derived from evolving stellar population models of Bruzual & Charlot, that have been fitted to the observed broadband (BVRI) AB magnitudes of each program galaxy. The listed photometric data are independent of all cosmological parameters. They are the source data for the Tolman surface brightness test made in Paper IV.

  19. Surface brightness profiles and structural parameters for 53 rich stellar clusters in the Large Magellanic Cloud

    NASA Astrophysics Data System (ADS)

    Mackey, A. D.; Gilmore, G. F.

    2003-01-01

    We have compiled a pseudo-snapshot data set of two-colour observations from the Hubble Space Telescope archive for a sample of 53 rich LMC clusters with ages of 106-1010 yr. We present surface brightness profiles for the entire sample, and derive structural parameters for each cluster, including core radii, and luminosity and mass estimates. Because we expect the results presented here to form the basis for several further projects, we describe in detail the data reduction and surface brightness profile construction processes, and compare our results with those of previous ground-based studies. The surface brightness profiles show a large amount of detail, including irregularities in the profiles of young clusters (such as bumps, dips and sharp shoulders), and evidence for both double clusters and post-core-collapse (PCC) clusters. In particular, we find power-law profiles in the inner regions of several candidate PCC clusters, with slopes of approximately -0.7, but showing considerable variation. We estimate that 20 +/- 7 per cent of the old cluster population of the Large Magellanic Cloud (LMC) has entered PCC evolution, a similar fraction to that for the Galactic globular cluster system. In addition, we examine the profile of R136 in detail and show that it is probably not a PCC cluster. We also observe a trend in core radius with age that has been discovered and discussed in several previous publications by different authors. Our diagram has better resolution, however, and appears to show a bifurcation at several hundred Myr. We argue that this observed relationship reflects true physical evolution in LMC clusters, with some experiencing small-scale core expansion owing to mass loss, and others large-scale expansion owing to some unidentified characteristic or physical process.

  20. Synoptic maps constructed from brightness observations of Thomson scattering by heliospheric electrons

    NASA Technical Reports Server (NTRS)

    Hick, P.; Jackson, B.; Schwenn, R.

    1991-01-01

    Observations of the Thomson scattering brightness by electrons in the inner heliosphere provide a means of probing the heliospheric electron distributions. An extensive data base of Thomson scattering observations, stretching over many years, is available from the zodiacal light photometers on board the two Helios spacecraft. A survey of these data is in progress, presenting these scattering intensities in the form of synoptic maps for successive Carrington rotations. The Thomson scattering maps reflect conditions at typically several tenths of an astronomical unit from the sun. Some representative examples from the survey in comparison with other solar/heliospheric data, such as in situ observations of the Helios plasma experiment and synoptic maps constructed from magnetic field, H alpha and K-coronameter data are presented. The comparison will provide some information about the extension of solar surface features into the inner heliosphere.

  1. Calibrating the Type Ia Supernova Distance Scale Using Surface Brightness Fluctuations

    NASA Astrophysics Data System (ADS)

    Potter, Cicely; Jensen, Joseph B.; Blakeslee, John; Milne, Peter; Garnavich, Peter M.; Brown, Peter

    2018-06-01

    We have observed 20 supernova host galaxies with HST WFC3/IR in the F110W filter, and prepared the data for Surface Brightness Fluctuation (SBF) distance measurements. The purpose of this study is to determine if there are any discrepancies between the SBF distance scale and the type-Ia SN distance scale, for which local calibrators are scarce. We have now measured SBF magnitudes to all early-type galaxies that have hosted SN Ia within 80 Mpc for which SBF measurements are possible. SBF is the only distance measurement technique with statistical uncertainties comparable to SN Ia that can be applied to galaxies out to 80 Mpc.

  2. Observation of Bright Ring Phenomenon for Red Blood Cells by Lattice Boltzmann Method

    NASA Astrophysics Data System (ADS)

    Kim, Young Woo; Moon, Ji Young; Lee, Joon Sang

    2017-11-01

    RBC (Red Blood Cell) aggregation is one of interests for various biomechanical fields such as cell chip or visualization. The unique phenomenon called ``bright ring'' is due to RBC aggregation in pulsatile flow of blood. Shear rate and flow acceleration on RBC causes them to repeat aggregating and scattering from center of the channel. The reason that this phenomenon is called bright ring is because that when observed by ultrasound imaging, the bright ring occurs periodically. Many studies tried to observe this bright ring phenomenon experimentally. However, there are yet not many studies trying to make use of this phenomenon for practical purposes. Bright ring phenomenon has high potential when used for cell separation or other microchip devices. In this paper, the Lattice Boltzmann method is used to control this bright ring phenomenon. The purpose of this paper is to find conditions when bright ring phenomenon occurs, and to control the aggregating-scattering frequency and degree. Deformability of RBC is calculated following the work of Moon JY et al. (2016). The result of this paper could be further extended to the optimization of cell-separating microchips. This work was also supported by the National Research Foundation of Korea (NRF) Grant funded by the Korean Government (MSIP) (No. 2015R1A5A1037668) and Brain Korea 21 Plus.

  3. Observation of spatial and temporal variations in X-ray bright point emergence patterns. [at solar surface

    NASA Technical Reports Server (NTRS)

    Golub, L.; Krieger, A. S.; Vaiana, G. S.

    1976-01-01

    Observations of X-ray bright points (XBP) over a six-month interval in 1973 show significant variations in both the number density of XBP as a function of heliographic longitude and in the full-sun average number of XBP from one rotation to the next. The observed increases in XBP emergence are estimated to be equivalent to several large active regions emerging per day for several months. The number of XBP emerging at high latitudes varies in phase with the low-latitude variation and reaches a maximum approximately simultaneous with a major outbreak of active regions. The quantity of magnetic flux emerging in the form of XBP at high latitudes alone is estimated to be as large as the contribution from all active regions.

  4. The MESSIER surveyor: unveiling the ultra-low surface brightness universe

    NASA Astrophysics Data System (ADS)

    Valls-Gabaud, David; MESSIER Collaboration

    2017-03-01

    The MESSIER surveyor is a small mission designed at exploring the very low surface brightness universe. The satellite will drift-scan the entire sky in 6 filters covering the 200-1000 nm range, reaching unprecedented surface brightness levels of 34 and 37 mag arcsec-2 in the optical and UV, respectively. These levels are required to achieve the two main science goals of the mission: to critically test the ΛCDM paradigm of structure formation through (1) the detection and characterisation of ultra-faint dwarf galaxies, which are predicted to be extremely abundant around normal galaxies, but which remain elusive; and (2) tracing the cosmic web, which feeds dark matter and baryons into galactic haloes, and which may contain the reservoir of missing baryons at low redshifts. A large number of science cases, ranging from stellar mass loss episodes to intracluster light through fluctuations in the cosmological UV-optical background radiation are free by-products of the full-sky maps produced.

  5. The response of the SSM/I to the marine environment. I - An analytic model for the atmospheric component of observed brightness temperatures

    NASA Technical Reports Server (NTRS)

    Petty, Grant W.; Katsaros, Kristina B.

    1992-01-01

    A detailed parameterization is developed for the contribution of the nonprecipitating atmosphere to the microwave brightness temperatures observed by the Special Sensor Microwave/Imager (SSM/I). The atmospheric variables considered include the viewing angle, the integrated water vapor amount and scale height, the effective tropospheric lapse rate and near-surface temperature, the total cloud liquid water, the effective cloud height, and the surface pressure. The dependence of the radiative variables on meteorological variables is determined for each of the SSM/I frequencies 19.35, 22.235, 37.0, and 85.5 GHz, based on the values computed from 16,893 maritime temperature and humidity profiles representing all latitude belts and all seasons. A comparison of the predicted brightness temperatures with brightness temperatures obtained by direct numerical integration of the radiative transfer equation for the radiosonde-profile dataset yielded rms differences well below 1 K for all four SSM/I frequencies.

  6. Soil moisture estimation by assimilating L-band microwave brightness temperature with geostatistics and observation localization.

    PubMed

    Han, Xujun; Li, Xin; Rigon, Riccardo; Jin, Rui; Endrizzi, Stefano

    2015-01-01

    The observation could be used to reduce the model uncertainties with data assimilation. If the observation cannot cover the whole model area due to spatial availability or instrument ability, how to do data assimilation at locations not covered by observation? Two commonly used strategies were firstly described: One is covariance localization (CL); the other is observation localization (OL). Compared with CL, OL is easy to parallelize and more efficient for large-scale analysis. This paper evaluated OL in soil moisture profile characterizations, in which the geostatistical semivariogram was used to fit the spatial correlated characteristics of synthetic L-Band microwave brightness temperature measurement. The fitted semivariogram model and the local ensemble transform Kalman filter algorithm are combined together to weight and assimilate the observations within a local region surrounding the grid cell of land surface model to be analyzed. Six scenarios were compared: 1_Obs with one nearest observation assimilated, 5_Obs with no more than five nearest local observations assimilated, and 9_Obs with no more than nine nearest local observations assimilated. The scenarios with no more than 16, 25, and 36 local observations were also compared. From the results we can conclude that more local observations involved in assimilation will improve estimations with an upper bound of 9 observations in this case. This study demonstrates the potentials of geostatistical correlation representation in OL to improve data assimilation of catchment scale soil moisture using synthetic L-band microwave brightness temperature, which cannot cover the study area fully in space due to vegetation effects.

  7. Soil Moisture Estimation by Assimilating L-Band Microwave Brightness Temperature with Geostatistics and Observation Localization

    PubMed Central

    Han, Xujun; Li, Xin; Rigon, Riccardo; Jin, Rui; Endrizzi, Stefano

    2015-01-01

    The observation could be used to reduce the model uncertainties with data assimilation. If the observation cannot cover the whole model area due to spatial availability or instrument ability, how to do data assimilation at locations not covered by observation? Two commonly used strategies were firstly described: One is covariance localization (CL); the other is observation localization (OL). Compared with CL, OL is easy to parallelize and more efficient for large-scale analysis. This paper evaluated OL in soil moisture profile characterizations, in which the geostatistical semivariogram was used to fit the spatial correlated characteristics of synthetic L-Band microwave brightness temperature measurement. The fitted semivariogram model and the local ensemble transform Kalman filter algorithm are combined together to weight and assimilate the observations within a local region surrounding the grid cell of land surface model to be analyzed. Six scenarios were compared: 1_Obs with one nearest observation assimilated, 5_Obs with no more than five nearest local observations assimilated, and 9_Obs with no more than nine nearest local observations assimilated. The scenarios with no more than 16, 25, and 36 local observations were also compared. From the results we can conclude that more local observations involved in assimilation will improve estimations with an upper bound of 9 observations in this case. This study demonstrates the potentials of geostatistical correlation representation in OL to improve data assimilation of catchment scale soil moisture using synthetic L-band microwave brightness temperature, which cannot cover the study area fully in space due to vegetation effects. PMID:25635771

  8. Surface-plasmon resonance-enhanced multiphoton emission of high-brightness electron beams from a nanostructured copper cathode.

    PubMed

    Li, R K; To, H; Andonian, G; Feng, J; Polyakov, A; Scoby, C M; Thompson, K; Wan, W; Padmore, H A; Musumeci, P

    2013-02-15

    We experimentally investigate surface-plasmon assisted photoemission to enhance the efficiency of metallic photocathodes for high-brightness electron sources. A nanohole array-based copper surface was designed to exhibit a plasmonic response at 800 nm, fabricated using the focused ion beam milling technique, optically characterized and tested as a photocathode in a high power radio frequency photoinjector. Because of the larger absorption and localization of the optical field intensity, the charge yield observed under ultrashort laser pulse illumination is increased by more than 100 times compared to a flat surface. We also present the first beam characterization results (intrinsic emittance and bunch length) from a nanostructured photocathode.

  9. The night sky brightness at McDonald Observatory

    NASA Technical Reports Server (NTRS)

    Kalinowski, J. K.; Roosen, R. G.; Brandt, J. C.

    1975-01-01

    Baseline observations of the night sky brightness in B and V are presented for McDonald Observatory. In agreement with earlier work by Elvey and Rudnick (1937) and Elvey (1943), significant night-to-night and same-night variations in sky brightness are found. Possible causes for these variations are discussed. The largest variation in sky brightness found during a single night is approximately a factor of two, a value which corresponds to a factor-of-four variation in airglow brightness. The data are used to comment on the accuracy of previously published surface photometry of M 81.

  10. Global Monitoring of Martian Surface Albedo Changes from Orbital Observations

    NASA Astrophysics Data System (ADS)

    Geissler, P.; Enga, M.; Mukherjee, P.

    2013-12-01

    Martian surface changes were first observed from orbit during the Mariner 9 and Viking Orbiter missions. They were found to be caused by eolian processes, produced by deposition of dust during regional and global dust storms and subsequent darkening of the surface through erosion and transportation of dust and sand. The albedo changes accumulated in the 20 years between Viking and Mars Global Surveyor were sufficient to alter the global circulation of winds and the climate of Mars according to model calculations (Fenton et al., Nature 2007), but little was known about the timing or frequency of the changes. Since 1999, we have had the benefit of continuous monitoring by a series of orbiting spacecraft that continues today with Mars Reconnaissance Orbiter, Mars Odyssey, and Mars Express. Daily synoptic observations enable us to determine whether the surface albedo changes are gradual or episodic in nature and to record the seasons that the changes take place. High resolution images of surface morphology and atmospheric phenomena help identify the physical mechanisms responsible for the changes. From these data, we hope to learn the combinations of atmospheric conditions and sediment properties that produce surface changes on Mars and possibly predict when they will take place in the future. Martian surface changes are particularly conspicuous in low albedo terrain, where even a thin layer of bright dust brightens the surface drastically. Equatorial dark areas are repeatedly coated and recoated by dust, which is later shed from the surface by a variety of mechanisms. An example is Syrtis Major, suddenly buried in bright dust by the global dust storm of 2001. Persistent easterly winds blew much of the dust cover away over the course of the next Martian year, but episodic changes continue today, particularly during southern summer when regional dust storms are rife. Another such region is Solis Planum, south of the Valles Marineris, where changes take place

  11. From 20 cm - 1 micron: Measuring the Gas and Dust in Massive Low Surface Brightness Galaxies

    NASA Astrophysics Data System (ADS)

    Kearsley, E.; O'Neil, K.

    2005-12-01

    Archival data from the IRAS, 2MASS, NVSS, and FIRST catalogs, supplemented with new measurements of HI, are used to analyze the relationship between the relative mass of the various components of galaxies (stars, atomic hydrogen, dust, and molecular gas) using a small sample of nearby (z<0.1), massive low surface brightness galaxies. The sample is compared to three sets of published data: a large collection of radio sources from the UGC having a radio continuum intensity >2.5 mJy (Condon, Cotton, & Broderick 2002 AJ 124, 675) ; a smaller sample of low surface brightness galaxies (Galaz, et al 2002 2002 AJ 124, 1360); and a collection of NIR low surface brightness galaxies (Monnier-Ragaigne, et al 2002 Ap&SS 281, 145). Overall, our sample properties are similar to the comparison samples in regard to NIR color, gas, stellar, and dynamic mass ratios, etc. Based off the galaxies' q-value (determined from the FIR/1.4 GHz ratio), it appears likely that at least two of the 28 galaxies studied harbor AGN. Notably, we also find that if we naively assume the ratio of the dust and molecular gas mass relative to the mass of HI is a constant we are unable to predict the observed ratio of stellar mass to HI mass, indicating that the HI mass ratio is a poor indicator of the total baryonic mass in the studied galaxies. HI measurements obtained during this study using the Green Bank Telescope also provide a correction to the velocity of UGC 11068.

  12. Simultaneous Multi-band Detection of Low Surface Brightness Galaxies with Markovian Modeling

    NASA Astrophysics Data System (ADS)

    Vollmer, B.; Perret, B.; Petremand, M.; Lavigne, F.; Collet, Ch.; van Driel, W.; Bonnarel, F.; Louys, M.; Sabatini, S.; MacArthur, L. A.

    2013-02-01

    We present to the astronomical community an algorithm for the detection of low surface brightness (LSB) galaxies in images, called MARSIAA (MARkovian Software for Image Analysis in Astronomy), which is based on multi-scale Markovian modeling. MARSIAA can be applied simultaneously to different bands. It segments an image into a user-defined number of classes, according to their surface brightness and surroundings—typically, one or two classes contain the LSB structures. We have developed an algorithm, called DetectLSB, which allows the efficient identification of LSB galaxies from among the candidate sources selected by MARSIAA. The application of the method to two and three bands simultaneously was tested on simulated images. Based on our tests, we are confident that we can detect LSB galaxies down to a central surface brightness level of only 1.5 times the standard deviation from the mean pixel value in the image background. To assess the robustness of our method, the method was applied to a set of 18 B- and I-band images (covering 1.3 deg2 in total) of the Virgo Cluster to which Sabatini et al. previously applied a matched-filter dwarf LSB galaxy search algorithm. We have detected all 20 objects from the Sabatini et al. catalog which we could classify by eye as bona fide LSB galaxies. Our method has also detected four additional Virgo Cluster LSB galaxy candidates undetected by Sabatini et al. To further assess the completeness of the results of our method, both MARSIAA, SExtractor, and DetectLSB were applied to search for (1) mock Virgo LSB galaxies inserted into a set of deep Next Generation Virgo Survey (NGVS) gri-band subimages and (2) Virgo LSB galaxies identified by eye in a full set of NGVS square degree gri images. MARSIAA/DetectLSB recovered ~20% more mock LSB galaxies and ~40% more LSB galaxies identified by eye than SExtractor/DetectLSB. With a 90% fraction of false positives from an entirely unsupervised pipeline, a completeness of 90% is reached

  13. Luminosity and surface brightness distribution of K-band galaxies from the UKIDSS Large Area Survey

    NASA Astrophysics Data System (ADS)

    Smith, Anthony J.; Loveday, Jon; Cross, Nicholas J. G.

    2009-08-01

    We present luminosity and surface-brightness distributions of 40111 galaxies with K-band photometry from the United Kingdom Infrared Telescope (UKIRT) Infrared Deep Sky Survey (UKIDSS) Large Area Survey (LAS), Data Release 3 and optical photometry from Data Release 5 of the Sloan Digital Sky Survey (SDSS). Various features and limitations of the new UKIDSS data are examined, such as a problem affecting Petrosian magnitudes of extended sources. Selection limits in K- and r-band magnitude, K-band surface brightness and K-band radius are included explicitly in the 1/Vmax estimate of the space density and luminosity function. The bivariate brightness distribution in K-band absolute magnitude and surface brightness is presented and found to display a clear luminosity-surface brightness correlation that flattens at high luminosity and broadens at low luminosity, consistent with similar analyses at optical wavelengths. Best-fitting Schechter function parameters for the K-band luminosity function are found to be M* - 5 logh = -23.19 +/- 0.04,α = -0.81 +/- 0.04 and φ* = (0.0166 +/- 0.0008)h3Mpc-3, although the Schechter function provides a poor fit to the data at high and low luminosity, while the luminosity density in the K band is found to be j = (6.305 +/- 0.067) × 108LsolarhMpc-3. However, we caution that there are various known sources of incompleteness and uncertainty in our results. Using mass-to-light ratios determined from the optical colours, we estimate the stellar mass function, finding good agreement with previous results. Possible improvements are discussed that could be implemented when extending this analysis to the full LAS.

  14. Diurnal Variations of Titan's Surface Temperatures From Cassini -CIRS Observations

    NASA Astrophysics Data System (ADS)

    Cottini, Valeria; Nixon, Conor; Jennings, Don; Anderson, Carrie; Samuelson, Robert; Irwin, Patrick; Flasar, F. Michael

    The Cassini Composite Infrared Spectrometer (CIRS) observations of Saturn's largest moon, Titan, are providing us with the ability to detect the surface temperature of the planet by studying its outgoing radiance through a spectral window in the thermal infrared at 19 m (530 cm-1) characterized by low opacity. Since the first acquisitions of CIRS Titan data the in-strument has gathered a large amount of spectra covering a wide range of latitudes, longitudes and local times. We retrieve the surface temperature and the atmospheric temperature pro-file by modeling proper zonally averaged spectra of nadir observations with radiative transfer computations. Our forward model uses the correlated-k approximation for spectral opacity to calculate the emitted radiance, including contributions from collision induced pairs of CH4, N2 and H2, haze, and gaseous emission lines (Irwin et al. 2008). The retrieval method uses a non-linear least-squares optimal estimation technique to iteratively adjust the model parameters to achieve a spectral fit (Rodgers 2000). We show an accurate selection of the wide amount of data available in terms of footprint diameter on the planet and observational conditions, together with the retrieved results. Our results represent formal retrievals of surface brightness temperatures from the Cassini CIRS dataset using a full radiative transfer treatment, and we compare to the earlier findings of Jennings et al. (2009). The application of our methodology over wide areas has increased the planet coverage and accuracy of our knowledge of Titan's surface brightness temperature. In particular we had the chance to look for diurnal variations in surface temperature around the equator: a trend with slowly increasing temperature toward the late afternoon reveals that diurnal temperature changes are present on Titan surface. References: Irwin, P.G.J., et al.: "The NEMESIS planetary atmosphere radiative transfer and retrieval tool" (2008). JQSRT, Vol. 109, pp

  15. THE DISCOVERY OF SEVEN EXTREMELY LOW SURFACE BRIGHTNESS GALAXIES IN THE FIELD OF THE NEARBY SPIRAL GALAXY M101

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Merritt, Allison; Van Dokkum, Pieter; Abraham, Roberto, E-mail: allison.merritt@yale.edu

    2014-06-01

    Dwarf satellite galaxies are a key probe of dark matter and of galaxy formation on small scales and of the dark matter halo masses of their central galaxies. They have very low surface brightness, which makes it difficult to identify and study them outside of the Local Group. We used a low surface brightness-optimized telescope, the Dragonfly Telephoto Array, to search for dwarf galaxies in the field of the massive spiral galaxy M101. We identify seven large, low surface brightness objects in this field, with effective radii of 10-30 arcseconds and central surface brightnesses of μ {sub g} ∼ 25.5-27.5 magmore » arcsec{sup –2}. Given their large apparent sizes and low surface brightnesses, these objects would likely be missed by standard galaxy searches in deep fields. Assuming the galaxies are dwarf satellites of M101, their absolute magnitudes are in the range –11.6 ≲ M{sub V} ≲ –9.3 and their effective radii are 350 pc-1.3 kpc. Their radial surface brightness profiles are well fit by Sersic profiles with a very low Sersic index (n ∼ 0.3-0.7). The properties of the sample are similar to those of well-studied dwarf galaxies in the Local Group, such as Sextans I and Phoenix. Distance measurements are required to determine whether these galaxies are in fact associated with M101 or are in its foreground or background.« less

  16. Synchronized observations of bright points from the solar photosphere to the corona

    NASA Astrophysics Data System (ADS)

    Tavabi, Ehsan

    2018-05-01

    One of the most important features in the solar atmosphere is the magnetic network and its relationship to the transition region (TR) and coronal brightness. It is important to understand how energy is transported into the corona and how it travels along the magnetic field lines between the deep photosphere and chromosphere through the TR and corona. An excellent proxy for transportation is the Interface Region Imaging Spectrograph (IRIS) raster scans and imaging observations in near-ultraviolet (NUV) and far-ultraviolet (FUV) emission channels, which have high time, spectral and spatial resolutions. In this study, we focus on the quiet Sun as observed with IRIS. The data with a high signal-to-noise ratio in the Si IV, C II and Mg II k lines and with strong emission intensities show a high correlation with TR bright network points. The results of the IRIS intensity maps and dopplergrams are compared with those of the Atmospheric Imaging Assembly (AIA) and Helioseismic and Magnetic Imager (HMI) instruments onboard the Solar Dynamical Observatory (SDO). The average network intensity profiles show a strong correlation with AIA coronal channels. Furthermore, we applied simultaneous observations of the magnetic network from HMI and found a strong relationship between the network bright points in all levels of the solar atmosphere. These features in the network elements exhibited regions of high Doppler velocity and strong magnetic signatures. Plenty of corona bright points emission, accompanied by the magnetic origins in the photosphere, suggest that magnetic field concentrations in the network rosettes could help to couple the inner and outer solar atmosphere.

  17. Simulating a slow bar in the low surface brightness galaxy UGC 628

    NASA Astrophysics Data System (ADS)

    Chequers, Matthew H.; Spekkens, Kristine; Widrow, Lawrence M.; Gilhuly, Colleen

    2016-12-01

    We present a disc-halo N-body model of the low surface brightness galaxy UGC 628, one of the few systems that harbours a `slow' bar with a ratio of corotation radius to bar length of R ≡ R_c/a_b ˜ 2. We select our initial conditions using SDSS DR10 photometry, a physically motivated radially variable mass-to-light ratio profile, and rotation curve data from the literature. A global bar instability grows in our submaximal disc model, and the disc morphology and dynamics agree broadly with the photometry and kinematics of UGC 628 at times between peak bar strength and the onset of buckling. Prior to bar formation, the disc and halo contribute roughly equally to the potential in the galaxy's inner region, giving the disc enough self-gravity for bar modes to grow. After bar formation, there is significant mass redistribution, creating a baryon-dominated inner and dark matter-dominated outer disc. This implies that, unlike most other low surface brightness galaxies, UGC 628 is not dark matter dominated everywhere. Our model nonetheless implies that UGC 628 falls on the same relationship between dark matter fraction and rotation velocity found for high surface brightness galaxies, and lends credence to the argument that the disc mass fraction measured at the location where its contribution to the potential peaks is not a reliable indicator of its dynamical importance at all radii.

  18. Prospects for Near Ultraviolet Astronomical Observations from the Lunar Surface — LUCI

    NASA Astrophysics Data System (ADS)

    Mathew, J.; Kumar, B.; Sarpotdar, M.; Suresh, A.; Nirmal, K.; Sreejith, A. G.; Safonova, M.; Murthy, J.; Brosch, N.

    2018-04-01

    We have explored the prospects for UV observations from the lunar surface and developed a UV telescope (LUCI-Lunar Ultraviolet Cosmic Imager) to put on the Moon, with the aim to detect bright UV transients such as SNe, novae, TDE, etc.

  19. Retrieval of Ocean Surface Windspeed and Rainrate from the Hurricane Imaging Radiometer (HIRAD) Brightness Temperature Observations

    NASA Technical Reports Server (NTRS)

    Biswas, Sayak K.; Jones, Linwood; Roberts, Jason; Ruf, Christopher; Ulhorn, Eric; Miller, Timothy

    2012-01-01

    The Hurricane Imaging Radiometer (HIRAD) is a new airborne synthetic aperture passive microwave radiometer capable of wide swath imaging of the ocean surface wind speed under heavy precipitation e.g. in tropical cyclones. It uses interferometric signal processing to produce upwelling brightness temperature (Tb) images at its four operating frequencies 4, 5, 6 and 6.6 GHz [1,2]. HIRAD participated in NASA s Genesis and Rapid Intensification Processes (GRIP) mission during 2010 as its first science field campaign. It produced Tb images with 70 km swath width and 3 km resolution from a 20 km altitude. From this, ocean surface wind speed and column averaged atmospheric liquid water content can be retrieved across the swath. The column averaged liquid water then could be related to an average rain rate. The retrieval algorithm (and the HIRAD instrument itself) is a direct descendant of the nadir-only Stepped Frequency Microwave Radiometer that is used operationally by the NOAA Hurricane Research Division to monitor tropical cyclones [3,4]. However, due to HIRAD s slant viewing geometry (compared to nadir viewing SFMR) a major modification is required in the algorithm. Results based on the modified algorithm from the GRIP campaign will be presented in the paper.

  20. Observation of a rapid decrease in the brightness of the coma of 2060 Chiron in 1990 January

    NASA Technical Reports Server (NTRS)

    Buratti, Bonnie J.; Dunbar, R. Scott

    1991-01-01

    Photometric observations of 2060 Chiron in the V and R filters were obtained with the 1.5-m telescope on Palomar Mountain during a 7-hr period on January 20, 1990 (UT). A general decrease of about 10 percent in integrated brightness occurred in both filters. No color dependence to the decrease was observed. A small (about 0.02 mag) rotational light curve, far smaller than the 0.09 mag (peak-to-peak) one observed by Bus et al. (1989) is superposed on the general decrease. On January 29, 1990, Luu and Jewitt (1990) observed an impulsive brightening of Chiron of approximately the same magnitude and time scale as the presently observed decrease in brightness. The combined results provide evidence that Chiron is currently exhibiting short-term fluctuations in the brightness of its coma, in addition to its well-established general decrease in brightness.

  1. Observation of a rapid decrease in the brightness of the coma of 2060 Chiron in 1990 January

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Buratti, B.J.; Dunbar, R.S.

    Photometric observations of 2060 Chiron in the V and R filters were obtained with the 1.5-m telescope on Palomar Mountain during a 7-hr period on January 20, 1990 (UT). A general decrease of about 10 percent in integrated brightness occurred in both filters. No color dependence to the decrease was observed. A small (about 0.02 mag) rotational light curve, far smaller than the 0.09 mag (peak-to-peak) one observed by Bus et al. (1989) is superposed on the general decrease. On January 29, 1990, Luu and Jewitt (1990) observed an impulsive brightening of Chiron of approximately the same magnitude and timemore » scale as the presently observed decrease in brightness. The combined results provide evidence that Chiron is currently exhibiting short-term fluctuations in the brightness of its coma, in addition to its well-established general decrease in brightness. 14 refs.« less

  2. Flattening and surface-brightness of the fast-rotating star δ Persei with the visible VEGA/CHARA interferometer

    NASA Astrophysics Data System (ADS)

    Challouf, M.; Nardetto, N.; Domiciano de Souza, A.; Mourard, D.; Tallon-Bosc, I.; Aroui, H.; Farrington, C.; Ligi, R.; Meilland, A.; Mouelhi, M.

    2017-08-01

    Context. Rapid rotation is a common feature for massive stars, with important consequences on their physical structure, flux distribution and evolution. Fast-rotating stars are flattened and show gravity darkening (non-uniform surface intensity distribution). Another important and less studied impact of fast-rotation in early-type stars is its influence on the surface brightness colour relation (hereafter SBCR), which could be used to derive the distance of eclipsing binaries. Aims: The purpose of this paper is to determine the flattening of the fast-rotating B-type star δ Per using visible long-baseline interferometry. A second goal is to evaluate the impact of rotation and gravity darkening on the V - K colour and surface brightness of the star. Methods: The B-type star δ Per was observed with the VEGA/CHARA interferometer, which can measure spatial resolutions down to 0.3 mas and spectral resolving power of 5000 in the visible. We first used a toy model to derive the position angle of the rotation axis of the star in the plane of the sky. Then we used a code of stellar rotation, CHARRON, in order to derive the physical parameters of the star. Finally, by considering two cases, a static reference star and our best model of δ Per, we can quantify the impact of fast rotation on the surface brightness colour relation (SBCR). Results: We find a position angle of 23 ± 6 degrees. The polar axis angular diameter of δ Per is θp = 0.544 ± 0.007 mas, and the derived flatness is r = 1.121 ± 0.013. We derive an inclination angle for the star of I = 85+ 5-20 degrees and a projected rotation velocity Vsini = 175+ 8-11 km s-1 (or 57% of the critical velocity). We find also that the rotation and inclination angle of δ Per keeps the V - K colour unchanged while it decreasing its surface-brightness by about 0.05 mag. Conclusions: Correcting the impact of rotation on the SBCR of early-type stars appears feasible using visible interferometry and dedicated models.

  3. Hurricane Imaging Radiometer (HIRAD) Observations of Brightness Temperatures and Ocean Surface Wind Speed and Rain Rate During NASA's GRIP and HS3 Campaigns

    NASA Technical Reports Server (NTRS)

    Miller, Timothy L.; James, M. W.; Roberts, J. B.; Jones, W. L.; Biswas, S.; Ruf, C. S.; Uhlhorn, E. W.; Atlas, R.; Black, P.; Albers, C.

    2012-01-01

    HIRAD flew on high-altitude aircraft over Earl and Karl during NASA s GRIP (Genesis and Rapid Intensification Processes) campaign in August - September of 2010, and plans to fly over Atlantic tropical cyclones in September of 2012 as part of the Hurricane and Severe Storm Sentinel (HS3) mission. HIRAD is a new C-band radiometer using a synthetic thinned array radiometer (STAR) technology to obtain spatial resolution of approximately 2 km, out to roughly 30 km each side of nadir. By obtaining measurements of emissions at 4, 5, 6, and 6.6 GHz, observations of ocean surface wind speed and rain rate can be retrieved. The physical retrieval technique has been used for many years by precursor instruments, including the Stepped Frequency Microwave Radiometer (SFMR), which has been flying on the NOAA and USAF hurricane reconnaissance aircraft for several years to obtain observations within a single footprint at nadir angle. Results from the flights during the GRIP and HS3 campaigns will be shown, including images of brightness temperatures, wind speed, and rain rate. Comparisons will be made with observations from other instruments on the campaigns, for which HIRAD observations are either directly comparable or are complementary. Features such as storm eye and eye-wall, location of storm wind and rain maxima, and indications of dynamical features such as the merging of a weaker outer wind/rain maximum with the main vortex may be seen in the data. Potential impacts on operational ocean surface wind analyses and on numerical weather forecasts will also be discussed.

  4. Little Bright Spot

    NASA Image and Video Library

    2015-01-12

    A bright spot can be seen on the left side of Rhea in this image. The spot is the crater Inktomi, named for a Lakota spider spirit. Inktomi is believed to be the youngest feature on Rhea (949 miles or 1527 kilometers across). The relative youth of the feature is evident by its brightness. Material that is newly excavated from below the moon's surface and tossed across the surface by a cratering event, appears bright. But as the newly exposed surface is subjected to the harsh space environment, it darkens. This is one technique scientists use to date features on surfaces. This view looks toward the trailing hemisphere of Rhea. North on Rhea is up and rotated 21 degrees to the left. The image was taken in visible light with the Cassini spacecraft narrow-angle camera on July 29, 2013. The view was obtained at a distance of approximately 1.0 million miles (1.6 million kilometers) fro http://photojournal.jpl.nasa.gov/catalog/PIA18300

  5. The formation of giant low surface brightness galaxies

    NASA Technical Reports Server (NTRS)

    Hoffman, Yehuda; Silk, Joseph; Wyse, Rosemary F. G.

    1992-01-01

    It is demonstrated that the initial structure of galaxies can be strongly affected by their large-scale environments. In particular, rare (about 3 sigma) massive galaxies in voids will have normal bulges, but unevolved, extended disks; it is proposed that the low surface brightness objects Malin I and Malin II are prototypes of this class of object. The model predicts that searches for more examples of 'crouching giants' should be fruitful, but that such galaxies do not provide a substantial fraction of mass in the universe. The identification of dwarf galaxies is relatively unaffected by their environment.

  6. Hi-C Observations of Penumbral Bright Dots

    NASA Technical Reports Server (NTRS)

    Alpert, S. E.; Tiwari, S. K.; Moore, R. L.; Savage, S. L.; Winebarger, A. R.

    2014-01-01

    We use high-quality data obtained by the High Resolution Coronal Imager (Hi-C) to examine bright dots (BDs) in a sunspot's penumbra. The sizes of these BDs are on the order of 1 arcsecond (1") and are therefore hard to identify using the Atmospheric Imaging Assembly's (AIA) 0.6" pixel(exp -1) resolution. These BD become readily apparent with Hi-C's 0.1" pixel(exp -1) resolution. Tian et al. (2014) found penumbral BDs in the transition region (TR) by using the Interface Region Imaging Spectrograph (IRIS). However, only a few of their dots could be associated with any enhanced brightness in AIA channels. In this work, we examine the characteristics of the penumbral BDs observed by Hi-C in a sunspot penumbra, including their sizes, lifetimes, speeds, and intensity. We also attempt to find any association of these BDs to the IRIS BDs. There are fewer Hi-C BDs in the penumbra than seen by IRIS, though different sunspots were studied. We use 193 Angstroms Hi-C data from July 11, 2012 which observed from approximately 18:52:00 UT- 18:56:00 UT and supplement it with data from AIA's 193 Angstrom passband to see the complete lifetime of the dots that were born before and/or lasted longer than Hi- C's 5-minute observation period. We use additional AIA passbands and compare the light curves of the BDs at different temperatures to test whether the Hi-C BDs are TR BDs. We find that most Hi-C BDs show clear movement, and of those that do, they move in a radial direction, toward or away from the sunspot umbra. Single BDs interact with other BDs, combining to fade away or brighten. The BDs that do not interact with other BDs tend to move less. Many of the properties of our BDs are similar to the extreme values of the IRIS BDs, e.g., they move slower on average and their sizes and lifetimes are on the higher end of the IRIS BDs. We infer that our penumbral BDs are the large-scale end of the distribution of BDs observed by IRIS.

  7. Stellar Surface Brightness Profiles of Dwarf Galaxies

    NASA Astrophysics Data System (ADS)

    Herrmann, K. A.

    2014-03-01

    Radial stellar surface brightness profiles of spiral galaxies can be classified into three types: (I) single exponential, or the light falls off with one exponential out to a break radius and then falls off (II) more steeply (“truncated”), or (III) less steeply (“anti-truncated”). Why there are three different radial profile types is still a mystery, including why light falls off as an exponential at all. Profile breaks are also found in dwarf disks, but some dwarf Type IIs are flat or increasing (FI) out to a break before falling off. I have been re-examining the multi-wavelength stellar disk profiles of 141 dwarf galaxies, primarily from Hunter & Elmegreen (2004, 2006). Each dwarf has data in up to 11 wavelength bands: FUV and NUV from GALEX, UBVJHK and Hα from ground-based observations, and 3.6 and 4.5μm from Spitzer. Here I highlight some results from a semi-automatic fitting of this data set including: (1) statistics of break locations and other properties as a function of wavelength and profile type, (2) color trends and radial mass distribution as a function of profile type, and (3) the relationship of the break radius to the kinematics and density profiles of atomic hydrogen gas in the 40 dwarfs of the LITTLE THINGS subsample.

  8. Lunar and Venusian radar bright rings

    NASA Technical Reports Server (NTRS)

    Thompson, T. W.; Saunders, R. S.; Weissman, D. E.

    1986-01-01

    Twenty-one lunar craters have radar bright ring appearances which are analogous to eleven complete ring features in the earth-based 12.5 cm observations of Venus. Radar ring diameters and widths for the lunar and Venusian features overlap for sizes from 45 to 100 km. Radar bright areas for the lunar craters are associated with the slopes of the inner and outer rim walls, while level crater floors and level ejecta fields beyond the raised portion of the rim have average radar backscatter. It is proposed that the radar bright areas of the Venusian rings are also associated with the slopes on the rims of craters. The lunar craters have evolved to radar bright rings via mass wasting of crater rim walls and via post-impact flooding of crater floors. Aeolian deposits of fine-grained material on Venusian crater floors may produce radar scattering effects similar to lunar crater floor flooding. These Venusian aeolian deposits may preferentially cover blocky crater floors producing a radar bright ring appearance. It is proposed that the Venusian features with complete bright ring appearances and sizes less than 100 km are impact craters. They have the same sizes as lunar craters and could have evolved to radar bright rings via analogous surface processes.

  9. Characterizing bars in low surface brightness disc galaxies

    NASA Astrophysics Data System (ADS)

    Peters, Wesley; Kuzio de Naray, Rachel

    2018-05-01

    In this paper, we use B-band, I-band, and 3.6 μm azimuthal light profiles of four low surface brightness galaxies (LSBs; UGC 628, F568-1, F568-3, F563-V2) to characterize three bar parameters: length, strength, and corotation radius. We employ three techniques to measure the radius of the bars, including a new method using the azimuthal light profiles. We find comparable bar radii between the I-band and 3.6 μm for all four galaxies when using our azimuthal light profile method, and that our bar lengths are comparable to those in high surface brightness galaxies (HSBs). In addition, we find the bar strengths for our galaxies to be smaller than those for HSBs. Finally, we use Fourier transforms of the B-band, I-band, and 3.6 μm images to characterize the bars as either `fast' or `slow' by measuring the corotation radius via phase profiles. When using the B- and I-band phase crossings, we find three of our galaxies have faster than expected relative bar pattern speeds for galaxies expected to be embedded in centrally dense cold dark matter haloes. When using the B-band and 3.6 μm phase crossings, we find more ambiguous results, although the relative bar pattern speeds are still faster than expected. Since we find a very slow bar in F563-V2, we are confident that we are able to differentiate between fast and slow bars. Finally, we find no relation between bar strength and relative bar pattern speed when comparing our LSBs to HSBs.

  10. A list of some bright objects which S-052 can observe

    NASA Technical Reports Server (NTRS)

    Mcquire, J. P.

    1972-01-01

    In order to find out the precise orientation of the photographs obtained by the High Altitude Observatory's ATM white light coronagraph, celestial objects must appear on each roll of film. A list of such bright objects and the times during which they can be observed is presented.

  11. Bright points and ejections observed on the sun by the KORONAS-FOTON instrument TESIS

    NASA Astrophysics Data System (ADS)

    Ulyanov, A. S.; Bogachev, S. A.; Kuzin, S. V.

    2010-10-01

    Five-second observations of the solar corona carried out in the FeIX 171 Å line by the KORONAS-FOTON instrument TESIS are used to study the dynamics of small-scale coronal structures emitting in and around coronal bright points. The small-scale structures of the lower corona display complex dynamics similar to those of magnetic loops located at higher levels of the solar corona. Numerous detected oscillating structures with sizes below 10 000 km display oscillation periods from 50 to 350 s. The period distributions of these structures are different for P < 150 s and P > 150 s, which implies that different oscillation modes are excited at different periods. The small-scale structures generate numerous flare-like events with energies 1024-1026 erg (nanoflares) and with a spatial density of one event per arcsecond or more observed over an area of 4 × 1011 km2. Nanoflares are not associated with coronal bright points, and almost uniformly cover the solar disk in the observation region. The ejections of solar material from the coronal bright points demonstrate velocities of 80-110 km/s.

  12. Synoptic maps of heliospheric Thomson scattering brightness from 1974-1985 as observed by the Helios photometers

    NASA Technical Reports Server (NTRS)

    Hick, P.; Jackson, B. V.; Schwenn, R.

    1992-01-01

    We display the electron Thomson scattering intensity of the inner heliosphere as observed by the zodiacal light photometers on board the Helios spacecraft in the form of synoptic maps. The technique extrapolates the brightness information from each photometer sector near the Sun and constructs a latitude/longitude map at a given solar height. These data are unique in that they give a determination of heliospheric structures out of the ecliptic above the primary region of solar wind acceleration. The spatial extent of bright, co-rotating heliospheric structures is readily observed in the data north and south of the ecliptic plane where the Helios photometer coverage is most complete. Because the technique has been used on the complete Helios data set from 1974 to 1985, we observe the change in our synoptic maps with solar cycle. Bright structures are concentrated near the heliospheric equator at solar minimum, while at solar maximum bright structures are found at far higher heliographic latitudes. A comparison of these maps with other forms of synoptic data are shown for two available intervals.

  13. Structural properties of faint low surface brightness galaxies

    NASA Astrophysics Data System (ADS)

    Pahwa, Isha; Saha, Kanak

    2018-05-01

    We study the structural properties of Low Surface Brightness galaxies (LSB) using a sample of 263 galaxies observed by the Green Bank Telescope (Schneider et al. 1992). We perform 2D decompositions of these galaxies in the SDSS g, r and i bands using the GALFIT software. Our decomposition reveals that about 60% of these galaxies are bulgeless i.e., their light distributions are well modelled by pure exponential disks. The rest of the galaxies were fitted with two components: a Sersic bulge and an exponential disk. Most of these galaxies have bulge-to-total (B/T) ratio less than 0.1. However, of these 104 galaxies, 20% have B/T > 0.1 i.e., hosting significant bulge component and they are more prominent amongst the fainter LSBs. According to g - r colour criteria, most of the LSB galaxies in our sample are blue, with only 7 classified as red LSBs. About 15% of the LSB galaxies (including both blue and red) in our sample host stellar bars. The incidence of bars is more prominent in relatively massive blue LSB galaxies with very high gas fraction. These findings may provide important clues to the formation and evolution of LSB galaxies - in particular on the bar/bulge formation in faint LSB disks.

  14. IRAS surface brightness maps of reflection nebulae in the Pleiades

    NASA Technical Reports Server (NTRS)

    Castelaz, Michael W.; Werner, M. W.; Sellgren, K.

    1987-01-01

    Surface brightness maps at 12, 25, 60, and 100 microns were made of a 2.5 deg x 2.5 deg area of the reflection nebulae in the Pleiades by coadding IRAS scans of this region. Emission is seen surrounding 17 Tau, 20 Tau, 23 Tau, and 25 Tau in all four bands, coextensive with the visible reflection nebulosity, and extending as far as 30 arcminutes from the illuminating stars. The infrared energy distributions of the nebulae peak in the 100 micron band, but up to 40 percent of the total infrared power lies in the 12 and 25 micron bands. The brightness of the 12 and 25 micron emission and the absence of temperature gradients at these wavelengths are inconsistent with the predictions of equilibrium thermal emission models. The emission at these wavelengths appears to be the result of micron nonequilibrium emission from very small grains, or from molecules consisting of 10-100 carbon atoms, which have been excited by ultraviolet radiation from the illuminating stars.

  15. Assimilation of SMOS Brightness Temperatures or Soil Moisture Retrievals into a Land Surface Model

    NASA Technical Reports Server (NTRS)

    De Lannoy, Gabrielle J. M.; Reichle, Rolf H.

    2016-01-01

    Three different data products from the Soil Moisture Ocean Salinity (SMOS) mission are assimilated separately into the Goddard Earth Observing System Model, version 5 (GEOS-5) to improve estimates of surface and root-zone soil moisture. The first product consists of multi-angle, dual-polarization brightness temperature (Tb) observations at the bottom of the atmosphere extracted from Level 1 data. The second product is a derived SMOS Tb product that mimics the data at a 40 degree incidence angle from the Soil Moisture Active Passive (SMAP) mission. The third product is the operational SMOS Level 2 surface soil moisture (SM) retrieval product. The assimilation system uses a spatially distributed ensemble Kalman filter (EnKF) with seasonally varying climatological bias mitigation for Tb assimilation, whereas a time-invariant cumulative density function matching is used for SM retrieval assimilation. All assimilation experiments improve the soil moisture estimates compared to model-only simulations in terms of unbiased root-mean-square differences and anomaly correlations during the period from 1 July 2010 to 1 May 2015 and for 187 sites across the US. Especially in areas where the satellite data are most sensitive to surface soil moisture, large skill improvements (e.g., an increase in the anomaly correlation by 0.1) are found in the surface soil moisture. The domain-average surface and root-zone skill metrics are similar among the various assimilation experiments, but large differences in skill are found locally. The observation-minus-forecast residuals and analysis increments reveal large differences in how the observations add value in the Tb and SM retrieval assimilation systems. The distinct patterns of these diagnostics in the two systems reflect observation and model errors patterns that are not well captured in the assigned EnKF error parameters. Consequently, a localized optimization of the EnKF error parameters is needed to further improve Tb or SM retrieval

  16. HST and ground-based observations of bright storms on Uranus during 2014-2015.

    NASA Astrophysics Data System (ADS)

    Sayanagi, K. M.; Sromovsky, L. A.; Fry, P. M.; De Pater, I.; Hammel, H. B.; Rages, K. A.; Baranec, C.; Delcroix, M.; Wesley, A.; Hueso, R.; Sanchez-Lavega, A.; Simon, A. A.; Wong, M. H.; Orton, G. S.; Irwin, P. G.

    2015-12-01

    We report the temporal evolution of bright, long-lived cloud features on Uranus. We observed and tracked the features between August 2014 and January 2015 with the Hubble Space Telescope, the Keck 2 10-m telescope, VLT, Gran Telescopio Canarias, Gemini, William Herschel Telescope, Robo-AO, Pic du Midi 1-m telescope, and multiple smaller telescopes operated by amateur astronomers. Surprisingly bright features were first revealed in the Keck adaptive-optics images in August; this initial set of observations motivated follow-up observations around the world. One of the storms (identified as "Feature F" in Sromovsky et al. 2015, and Feature 2 in de Pater et al. 2015), which was the deepest in that dataset, was bright enough that it was detected by multiple amateur observers, permitting us to trigger a Hubble Target of Opportunity (ToO) observation on October 14th, 2014. A complex of features at this latitude was also observed by Hubble as part of the Outer Planet Atmospheres Legacy (OPAL) program on November 8-9, 2014. We will present the temporal evolution of the cloud activities from August 2014 through January 2015, and analyze the vertical structure of the cloud features in the Hubble datasets. The Hubble images used in our study were collected with support of HST grants GO13712 to KMS and GO13937 to AAS. Sromovsky et al. 2015, "High S/N Keck and Gemini AO imaging of Uranus during 2012-2014: New cloud patterns, increasing activity, and improved wind measurements." Icarus 258, 192-223. de Pater et al. 2014, "Record-breaking storm activity on Uranus in 2014." Icarus 252, 121-128

  17. Observations of C-Band Brightness Temperatures and Ocean Surface Wind Speed and Rain Rate from the Hurricane Imaging Radiometer (HIRAD) during GRIP and HS3

    NASA Technical Reports Server (NTRS)

    Miller, Timothy L.; James, M. W.; Roberts, J. B.; Jones, W. L.; Biswas, S.; Ruf, C. S.; Uhlhorn, E. W.; Atlas, R.; Black, P.; Albers, C.

    2013-01-01

    HIRAD flew on high-altitude aircraft over Earl and Karl during NASA s GRIP (Genesis and Rapid Intensification Processes) campaign in August - September of 2010, and at the time of this writing plans to fly over Atlantic tropical cyclones in September of 2012 as part of the Hurricane and Severe Storm Sentinel (HS3) mission. HIRAD is a new C-band radiometer using a synthetic thinned array radiometer (STAR) technology to obtain cross-track resolution of approximately 3 degrees, out to approximately 60 degrees to each side of nadir. By obtaining measurements of emissions at 4, 5, 6, and 6.6 GHz, observations of ocean surface wind speed and rain rate can be retrieved. This technique has been used for many years by precursor instruments, including the Stepped Frequency Microwave Radiometer (SFMR), which has been flying on the NOAA and USAF hurricane reconnaissance aircraft for several years to obtain observations within a single footprint at nadir angle. Results from the flights during the GRIP and HS3 campaigns will be shown, including images of brightness temperatures, wind speed, and rain rate. Comparisons will be made with observations from other instruments on the campaigns, for which HIRAD observations are either directly comparable or are complementary. Features such as storm eye and eye-wall, location of storm wind and rain maxima, and indications of dynamical features such as the merging of a weaker outer wind/rain maximum with the main vortex may be seen in the data. Potential impacts on operational ocean surface wind analyses and on numerical weather forecasts will also be discussed.

  18. Energy-exchange collisions of dark-bright-bright vector solitons.

    PubMed

    Radhakrishnan, R; Manikandan, N; Aravinthan, K

    2015-12-01

    We find a dark component guiding the practically interesting bright-bright vector one-soliton to two different parametric domains giving rise to different physical situations by constructing a more general form of three-component dark-bright-bright mixed vector one-soliton solution of the generalized Manakov model with nine free real parameters. Moreover our main investigation of the collision dynamics of such mixed vector solitons by constructing the multisoliton solution of the generalized Manakov model with the help of Hirota technique reveals that the dark-bright-bright vector two-soliton supports energy-exchange collision dynamics. In particular the dark component preserves its initial form and the energy-exchange collision property of the bright-bright vector two-soliton solution of the Manakov model during collision. In addition the interactions between bound state dark-bright-bright vector solitons reveal oscillations in their amplitudes. A similar kind of breathing effect was also experimentally observed in the Bose-Einstein condensates. Some possible ways are theoretically suggested not only to control this breathing effect but also to manage the beating, bouncing, jumping, and attraction effects in the collision dynamics of dark-bright-bright vector solitons. The role of multiple free parameters in our solution is examined to define polarization vector, envelope speed, envelope width, envelope amplitude, grayness, and complex modulation of our solution. It is interesting to note that the polarization vector of our mixed vector one-soliton evolves in sphere or hyperboloid depending upon the initial parametric choices.

  19. Using Simplistic Shape/Surface Models to Predict Brightness in Estimation Filters

    NASA Astrophysics Data System (ADS)

    Wetterer, C.; Sheppard, D.; Hunt, B.

    The prerequisite for using brightness (radiometric flux intensity) measurements in an estimation filter is to have a measurement function that accurately predicts a space objects brightness for variations in the parameters of interest. These parameters include changes in attitude and articulations of particular components (e.g. solar panel east-west offsets to direct sun-tracking). Typically, shape models and bidirectional reflectance distribution functions are combined to provide this forward light curve modeling capability. To achieve precise orbit predictions with the inclusion of shape/surface dependent forces such as radiation pressure, relatively complex and sophisticated modeling is required. Unfortunately, increasing the complexity of the models makes it difficult to estimate all those parameters simultaneously because changes in light curve features can now be explained by variations in a number of different properties. The classic example of this is the connection between the albedo and the area of a surface. If, however, the desire is to extract information about a single and specific parameter or feature from the light curve, a simple shape/surface model could be used. This paper details an example of this where a complex model is used to create simulated light curves, and then a simple model is used in an estimation filter to extract out a particular feature of interest. In order for this to be successful, however, the simple model must be first constructed using training data where the feature of interest is known or at least known to be constant.

  20. Assimilation of Global Radar Backscatter and Radiometer Brightness Temperature Observations to Improve Soil Moisture and Land Evaporation Estimates

    NASA Technical Reports Server (NTRS)

    Lievens, H.; Martens, B.; Verhoest, N. E. C.; Hahn, S.; Reichle, R. H.; Miralles, D. G.

    2017-01-01

    Active radar backscatter (s?) observations from the Advanced Scatterometer (ASCAT) and passive radiometer brightness temperature (TB) observations from the Soil Moisture Ocean Salinity (SMOS) mission are assimilated either individually or jointly into the Global Land Evaporation Amsterdam Model (GLEAM) to improve its simulations of soil moisture and land evaporation. To enable s? and TB assimilation, GLEAM is coupled to the Water Cloud Model and the L-band Microwave Emission from the Biosphere (L-MEB) model. The innovations, i.e. differences between observations and simulations, are mapped onto the model soil moisture states through an Ensemble Kalman Filter. The validation of surface (0-10 cm) soil moisture simulations over the period 2010-2014 against in situ measurements from the International Soil Moisture Network (ISMN) shows that assimilating s? or TB alone improves the average correlation of seasonal anomalies (Ran) from 0.514 to 0.547 and 0.548, respectively. The joint assimilation further improves Ran to 0.559. Associated enhancements in daily evaporative flux simulations by GLEAM are validated based on measurements from 22 FLUXNET stations. Again, the singular assimilation improves Ran from 0.502 to 0.536 and 0.533, respectively for s? and TB, whereas the best performance is observed for the joint assimilation (Ran = 0.546). These results demonstrate the complementary value of assimilating radar backscatter observations together with brightness temperatures for improving estimates of hydrological variables, as their joint assimilation outperforms the assimilation of each observation type separately.

  1. Joint Assimilation of SMOS Brightness Temperature and GRACE Terrestrial Water Storage Observations for Improved Soil Moisture Estimation

    NASA Technical Reports Server (NTRS)

    Girotto, Manuela; Reichle, Rolf H.; De Lannoy, Gabrielle J. M.; Rodell, Matthew

    2017-01-01

    Observations from recent soil moisture missions (e.g. SMOS) have been used in innovative data assimilation studies to provide global high spatial (i.e. 40 km) and temporal resolution (i.e. 3-days) soil moisture profile estimates from microwave brightness temperature observations. In contrast with microwave-based satellite missions that are only sensitive to near-surface soil moisture (0 - 5 cm), the Gravity Recovery and Climate Experiment (GRACE) mission provides accurate measurements of the entire vertically integrated terrestrial water storage column but, it is characterized by low spatial (i.e. 150,000 km2) and temporal (i.e. monthly) resolutions. Data assimilation studies have shown that GRACE-TWS primarily affects (in absolute terms) deeper moisture storages (i.e., groundwater). This work hypothesizes that unprecedented soil water profile accuracy can be obtained through the joint assimilation of GRACE terrestrial water storage and SMOS brightness temperature observations. A particular challenge of the joint assimilation is the use of the two different types of measurements that are relevant for hydrologic processes representing different temporal and spatial scales. The performance of the joint assimilation strongly depends on the chosen assimilation methods, measurement and model error spatial structures. The optimization of the assimilation technique constitutes a fundamental step toward a multi-variate multi-resolution integrative assimilation system aiming to improve our understanding of the global terrestrial water cycle.

  2. Joint assimilation of SMOS brightness temperature and GRACE terrestrial water storage observations for improved soil moisture estimation

    NASA Astrophysics Data System (ADS)

    Girotto, M.; Reichle, R. H.; De Lannoy, G.; Rodell, M.

    2017-12-01

    Observations from recent soil moisture missions (e.g. SMOS) have been used in innovative data assimilation studies to provide global high spatial (i.e. 40 km) and temporal resolution (i.e. 3-days) soil moisture profile estimates from microwave brightness temperature observations. In contrast with microwave-based satellite missions that are only sensitive to near-surface soil moisture (0-5 cm), the Gravity Recovery and Climate Experiment (GRACE) mission provides accurate measurements of the entire vertically integrated terrestrial water storage column but, it is characterized by low spatial (i.e. 150,000 km2) and temporal (i.e. monthly) resolutions. Data assimilation studies have shown that GRACE-TWS primarily affects (in absolute terms) deeper moisture storages (i.e., groundwater). This work hypothesizes that unprecedented soil water profile accuracy can be obtained through the joint assimilation of GRACE terrestrial water storage and SMOS brightness temperature observations. A particular challenge of the joint assimilation is the use of the two different types of measurements that are relevant for hydrologic processes representing different temporal and spatial scales. The performance of the joint assimilation strongly depends on the chosen assimilation methods, measurement and model error spatial structures. The optimization of the assimilation technique constitutes a fundamental step toward a multi-variate multi-resolution integrative assimilation system aiming to improve our understanding of the global terrestrial water cycle.

  3. Double Bright Band Observations with High-Resolution Vertically Pointing Radar, Lidar, and Profiles

    NASA Technical Reports Server (NTRS)

    Emory, Amber E.; Demoz, Belay; Vermeesch, Kevin; Hicks, Michael

    2014-01-01

    On 11 May 2010, an elevated temperature inversion associated with an approaching warm front produced two melting layers simultaneously, which resulted in two distinct bright bands as viewed from the ER-2 Doppler radar system, a vertically pointing, coherent X band radar located in Greenbelt, MD. Due to the high temporal resolution of this radar system, an increase in altitude of the melting layer of approximately 1.2 km in the time span of 4 min was captured. The double bright band feature remained evident for approximately 17 min, until the lower atmosphere warmed enough to dissipate the lower melting layer. This case shows the relatively rapid evolution of freezing levels in response to an advancing warm front over a 2 h time period and the descent of an elevated warm air mass with time. Although observations of double bright bands are somewhat rare, the ability to identify this phenomenon is important for rainfall estimation from spaceborne sensors because algorithms employing the restriction of a radar bright band to a constant height, especially when sampling across frontal systems, will limit the ability to accurately estimate rainfall.

  4. Double bright band observations with high-resolution vertically pointing radar, lidar, and profilers

    NASA Astrophysics Data System (ADS)

    Emory, Amber E.; Demoz, Belay; Vermeesch, Kevin; Hicks, Micheal

    2014-07-01

    On 11 May 2010, an elevated temperature inversion associated with an approaching warm front produced two melting layers simultaneously, which resulted in two distinct bright bands as viewed from the ER-2 Doppler radar system, a vertically pointing, coherent X band radar located in Greenbelt, MD. Due to the high temporal resolution of this radar system, an increase in altitude of the melting layer of approximately 1.2 km in the time span of 4 min was captured. The double bright band feature remained evident for approximately 17 min, until the lower atmosphere warmed enough to dissipate the lower melting layer. This case shows the relatively rapid evolution of freezing levels in response to an advancing warm front over a 2 h time period and the descent of an elevated warm air mass with time. Although observations of double bright bands are somewhat rare, the ability to identify this phenomenon is important for rainfall estimation from spaceborne sensors because algorithms employing the restriction of a radar bright band to a constant height, especially when sampling across frontal systems, will limit the ability to accurately estimate rainfall.

  5. Bright Stuff on Ceres = Sulfates and Carbonates on CI Chondrites

    NASA Technical Reports Server (NTRS)

    Zolensky, Michael; Chan, Queenie H. S.; Gounelle, Matthieu; Fries, Marc

    2016-01-01

    Recent reports of the DAWN spacecraft's observations of the surface of Ceres indicate that there are bright areas, which can be explained by large amounts of the Mg sulfate hexahydrate (MgSO4•6(H2O)), although the identification appears tenuous. There are preliminary indications that water is being evolved from these bright areas, and some have inferred that these might be sites of contemporary hydro-volcanism. A heat source for such modern activity is not obvious, given the small size of Ceres, lack of any tidal forces from nearby giant planets, probable age and presumed bulk composition. We contend that observations of chondritic materials in the lab shed light on the nature of the bright spots on Ceres

  6. L-Band Brightness Temperature Variations at Dome C and Snow Metamorphism at the Surface

    NASA Technical Reports Server (NTRS)

    Brucker, Ludovic; Dinnat, Emmanuel; Picard, Ghislain; Champollion, Nicolas

    2014-01-01

    The Antarctic Plateau is a promising site to monitor microwave radiometers' drift, and to inter-calibrate microwave radiometers, especially 1.4 GigaHertz (L-band) radiometers on board the Soil Moisture and Ocean Salinity (SMOS), and AquariusSAC-D missions. The Plateau is a thick ice cover, thermally stable in depth, with large dimensions, and relatively low heterogeneities. In addition, its high latitude location in the Southern Hemisphere enables frequent observations by polar-orbiting satellites, and no contaminations by radio frequency interference. At Dome C (75S, 123E), on the Antarctic Plateau, the substantial amount of in-situ snow measurements available allows us to interpret variations in space-borne microwave brightness temperature (TB) (e.g. Macelloni et al., 2007, 2013, Brucker et al., 2011, Champollion et al., 2013). However, to analyze the observations from the Aquarius radiometers, whose sensitivity is 0.15 K, the stability of the snow layers near the surface that are most susceptible to rapidly change needs to be precisely assessed. This study focuses on the spatial and temporal variations of the Aquarius TB over the Antarctic Plateau, and at Dome C in particular, to highlight the impact of snow surface metamorphism on the TB observations at L-band.

  7. Aquarius Brightness Temperature Variations at Dome C and Snow Metamorphism at the Surface. [29

    NASA Technical Reports Server (NTRS)

    Brucker, Ludovic; Dinnat, Emmanuel Phillippe; Picard, Ghislain; Champollion, Nicolas

    2014-01-01

    The Antarctic Plateau is a promising site to monitor microwave radiometers' drift, and to inter-calibrate microwave radiometers, especially 1.4 GHz (L-band) radiometers on board the Soil Moisture and Ocean Salinity (SMOS), and AquariusSAC-D missions. The Plateau is a thick ice cover, thermally stable in depth, with large dimensions, and relatively low heterogeneities. In addition, its high latitude location in the Southern Hemisphere enables frequent observations by polar-orbiting satellites, and no contaminations by radio frequency interference. At Dome C (75S, 123E), on the Antarctic Plateau, the substantial amount of in-situ snow measurements available allows us to interpret variations in space-borne microwave brightness temperature (TB) (e.g. Macelloni et al., 2007, 2013, Brucker et al., 2011, Champollion et al., 2013). However, to analyze the observations from the Aquarius radiometers, whose sensitivity is 0.15 K, the stability of the snow layers near the surface that are most susceptible to rapidly change needs to be precisely assessed. This study focuses on the spatial and temporal variations of the Aquarius TB over the Antarctic Plateau, and at Dome C in particular, to highlight the impact of snow surface metamorphism on the TB observations at L-band.

  8. A new, bright and hard aluminum surface produced by anodization

    NASA Astrophysics Data System (ADS)

    Hou, Fengyan; Hu, Bo; Tay, See Leng; Wang, Yuxin; Xiong, Chao; Gao, Wei

    2017-07-01

    Anodized aluminum (Al) and Al alloys have a wide range of applications. However, certain anodized finishings have relatively low hardness, dull appearance and/or poor corrosion resistance, which limited their applications. In this research, Al was first electropolished in a phosphoric acid-based solution, then anodized in a sulfuric acid-based solution under controlled processing parameters. The anodized specimen was then sealed by two-step sealing method. A systematic study including microstructure, surface morphology, hardness and corrosion resistance of these anodized films has been conducted. Results show that the hardness of this new anodized film was increased by a factor of 10 compared with the pure Al metal. Salt spray corrosion testing also demonstrated the greatly improved corrosion resistance. Unlike the traditional hard anodized Al which presents a dull-colored surface, this newly developed anodized Al alloy possesses a very bright and shiny surface with good hardness and corrosion resistance.

  9. RadioAstron Observations of the Quasar 3C273: A Challenge to the Brightness Temperature Limit

    NASA Astrophysics Data System (ADS)

    Kovalev, Y. Y.; Kardashev, N. S.; Kellermann, K. I.; Lobanov, A. P.; Johnson, M. D.; Gurvits, L. I.; Voitsik, P. A.; Zensus, J. A.; Anderson, J. M.; Bach, U.; Jauncey, D. L.; Ghigo, F.; Ghosh, T.; Kraus, A.; Kovalev, Yu. A.; Lisakov, M. M.; Petrov, L. Yu.; Romney, J. D.; Salter, C. J.; Sokolovsky, K. V.

    2016-03-01

    Inverse Compton cooling limits the brightness temperature of the radiating plasma to a maximum of 1011.5 K. Relativistic boosting can increase its observed value, but apparent brightness temperatures much in excess of 1013 K are inaccessible using ground-based very long baseline interferometry (VLBI) at any wavelength. We present observations of the quasar 3C 273, made with the space VLBI mission RadioAstron on baselines up to 171,000 km, which directly reveal the presence of angular structure as small as 26 μas (2.7 light months) and brightness temperature in excess of 1013 K. These measurements challenge our understanding of the non-thermal continuum emission in the vicinity of supermassive black holes and require a much higher Doppler factor than what is determined from jet apparent kinematics.

  10. Anticorrelation of X-ray bright points with sunspot number, 1970-1978

    NASA Technical Reports Server (NTRS)

    Golub, L.; Davis, J. M.; Krieger, A. S.

    1979-01-01

    Soft X-ray observations of the solar corona over the period 1970-1978 show that the number of small short-lived bipolar magnetic features (X-ray bright points) varies inversely with the sunspot index. During the entire period from 1973 to 1978 most of the magnetic flux emerging at the solar surface appeared in the form of bright points. In 1970, near the peak of solar cycle 20, the contributions from bright points and from active regions appear to be approximately equal. These observations strongly support an earlier suggestion that the solar cycle may be characterized as an oscillator in wave-number space with relatively little variation in the average total rate of flux emergence.

  11. Brightness and transparency in the early visual cortex.

    PubMed

    Salmela, Viljami R; Vanni, Simo

    2013-06-24

    Several psychophysical studies have shown that transparency can have drastic effects on brightness and lightness. However, the neural processes generating these effects have remained unresolved. Several lines of evidence suggest that the early visual cortex is important for brightness perception. While single cell recordings suggest that surface brightness is represented in the primary visual cortex, the results of functional magnetic resonance imaging (fMRI) studies have been discrepant. In addition, the location of the neural representation of transparency is not yet known. We investigated whether the fMRI responses in areas V1, V2, and V3 correlate with brightness and transparency. To dissociate the blood oxygen level-dependent (BOLD) response to brightness from the response to local border contrast and mean luminance, we used variants of White's brightness illusion, both opaque and transparent, in which luminance increments and decrements cancel each other out. The stimuli consisted of a target surface and a surround. The surround luminance was always sinusoidally modulated at 0.5 Hz to induce brightness modulation to the target. The target luminance was constant or modulated in counterphase to null brightness modulation. The mean signal changes were calculated from the voxels in V1, V2, and V3 corresponding to the retinotopic location of the target surface. The BOLD responses were significantly stronger for modulating brightness than for stimuli with constant brightness. In addition, the responses were stronger for transparent than for opaque stimuli, but there was more individual variation. No interaction between brightness and transparency was found. The results show that the early visual areas V1-V3 are sensitive to surface brightness and transparency and suggest that brightness and transparency are represented separately.

  12. The nature of solar brightness variations

    NASA Astrophysics Data System (ADS)

    Shapiro, A. I.; Solanki, S. K.; Krivova, N. A.; Cameron, R. H.; Yeo, K. L.; Schmutz, W. K.

    2017-09-01

    Determining the sources of solar brightness variations1,2, often referred to as solar noise3, is important because solar noise limits the detection of solar oscillations3, is one of the drivers of the Earth's climate system4,5 and is a prototype of stellar variability6,7—an important limiting factor for the detection of extrasolar planets. Here, we model the magnetic contribution to solar brightness variability using high-cadence8,9 observations from the Solar Dynamics Observatory (SDO) and the Spectral And Total Irradiance REconstruction (SATIRE)10,11 model. The brightness variations caused by the constantly evolving cellular granulation pattern on the solar surface were computed with the Max Planck Institute for Solar System Research (MPS)/University of Chicago Radiative Magnetohydrodynamics (MURaM)12 code. We found that the surface magnetic field and granulation can together precisely explain solar noise (that is, solar variability excluding oscillations) on timescales from minutes to decades, accounting for all timescales that have so far been resolved or covered by irradiance measurements. We demonstrate that no other sources of variability are required to explain the data. Recent measurements of Sun-like stars by the COnvection ROtation and planetary Transits (CoRoT)13 and Kepler14 missions uncovered brightness variations similar to that of the Sun, but with a much wider variety of patterns15. Our finding that solar brightness variations can be replicated in detail with just two well-known sources will greatly simplify future modelling of existing CoRoT and Kepler as well as anticipated Transiting Exoplanet Survey Satellite16 and PLAnetary Transits and Oscillations of stars (PLATO)17 data.

  13. Plasmonic EIT-like switching in bright-dark-bright plasmon resonators.

    PubMed

    Chen, Junxue; Wang, Pei; Chen, Chuncong; Lu, Yonghua; Ming, Hai; Zhan, Qiwen

    2011-03-28

    In this paper we report the study of the electromagnetically induced transparency (EIT)-like transmission in the bright-dark-bright plasmon resonators. It is demonstrated that the interferences between the dark plasmons excited by two bright plasmon resonators can be controlled by the incident light polarization. The constructive interference strengthens the coupling between the bright and dark resonators, leading to a more prominent EIT-like transparency window of the metamaterial. In contrary, destructive interference suppresses the coupling between the bright and dark resonators, destroying the interference pathway that forms the EIT-like transmission. Based on this observation, the plasmonic EIT switching can be realized by changing the polarization of incident light. This phenomenon may find applications in optical switching and plasmon-based information processing.

  14. Liquid Hydrocarbons on Titan's Surface? How Cassini ISS Observations Fit into the Story (So Far)

    NASA Technical Reports Server (NTRS)

    Turtle, E. P.; Dawson, D. D.; Fussner, S.; Hardegree-Ullman, E.; Ewen, A. S.; Perry, J.; Porco, C. C.; West, R. A.

    2005-01-01

    Titan is the only satellite in our Solar System with a substantial atmosphere, the origins and evolution of which are still not well understood. Its primary (greater than 90%) component is nitrogen, with a few percent methane and lesser amounts of other species. Methane and ethane are stable in the liquid state under the temperature and pressure conditions in Titan s lower atmosphere and at the surface; indeed, clouds, likely composed of methane, have been detected. Photochemical processes acting in the atmosphere convert methane into more complex hydrocarbons, creating Titan s haze and destroying methane over relatively short timescales. Therefore, it has been hypothesized that Titan s surface has reservoirs of liquid methane which serve to resupply the atmosphere. Early observations of Titan s surface revealed albedo patterns which have been interpreted as dark hydrocarbon liquids occupying topographically low regions between higher-standing exposures of bright, water-ice bedrock, although this is far from being the only explanation for the observed albedo contrast. Observations made by the Imaging Science Subsystem during Cassini's approach to Saturn and its first encounters with Titan show the bright and dark regions in greater detail but have yet to resolve the question of whether there are liquids on the surface.

  15. Percentage Contributions from Atmospheric and Surface Features to Computed Brightness Temperatures

    NASA Technical Reports Server (NTRS)

    Jackson, Gail Skofronick

    2006-01-01

    Over the past few years, there has become an increasing interest in the use of millimeter-wave (mm-wave) and sub-millimeter-wave (submm-wave) radiometer observations to investigate the properties of ice particles in clouds. Passive radiometric channels respond to both the integrated particle mass throughout the volume and field of view, and to the amount, location, and size distribution of the frozen (and liquid) particles with the sensitivity varying for different frequencies and hydrometeor types. One methodology used since the 1960's to discern the relationship between the physical state observed and the brightness temperature (TB) is through the temperature weighting function profile. In this research, the temperature weighting function concept is exploited to analyze the sensitivity of various characteristics of the cloud profile, such as relative humidity, ice water path, liquid water path, and surface emissivity. In our numerical analysis, we compute the contribution (in Kelvin) from each of these cloud and surface characteristics, so that the sum of these various parts equals the computed TB. Furthermore, the percentage contribution from each of these characteristics is assessed. There is some intermingling/contamination of the contributions from various components due to the integrated nature of passive observations and the absorption and scattering between the vertical layers, but all in all the knowledge gained is useful. This investigation probes the sensitivity over several cloud classifications, such as cirrus, blizzards, light snow, anvil clouds, and heavy rain. The focus is on mm-wave and submm-wave frequencies, however discussions of the effects of cloud variations to frequencies as low as 10 GHz and up to 874 GHz will also be presented. The results show that nearly 60% of the TB value at 89 GHz comes from the earth's surface for even the heaviest blizzard snow rates. On the other hand, a significant percentage of the TB value comes from the snow

  16. Are solar brightness variations faculae- or spot-dominated?

    NASA Astrophysics Data System (ADS)

    Shapiro, A. I.; Solanki, S. K.; Krivova, N. A.; Yeo, K. L.; Schmutz, W. K.

    2016-05-01

    Context. Regular spaceborne measurements have revealed that solar brightness varies on multiple timescales, variations on timescales greater than a day being attributed to a surface magnetic field. Independently, ground-based and spaceborne measurements suggest that Sun-like stars show a similar, but significantly broader pattern of photometric variability. Aims: To understand whether the broader pattern of stellar variations is consistent with the solar paradigm, we assess relative contributions of faculae and spots to solar magnetically-driven brightness variability. We investigate how the solar brightness variability and its facular and spot contributions depend on the wavelength, timescale of variability, and position of the observer relative to the ecliptic plane. Methods: We performed calculations with the SATIRE model, which returns solar brightness with daily cadence from solar disc area coverages of various magnetic features. We took coverages as seen by an Earth-based observer from full-disc SoHO/MDI and SDO/HMI data and projected them to mimic out-of-ecliptic viewing by an appropriate transformation. Results: Moving the observer away from the ecliptic plane increases the amplitude of 11-year variability as it would be seen in Strömgren (b + y)/2 photometry, but decreases the amplitude of the rotational brightness variations as it would appear in Kepler and CoRoT passbands. The spot and facular contributions to the 11-year solar variability in the Strömgren (b + y)/2 photometry almost fully compensate each other so that the Sun appears anomalously quiet with respect to its stellar cohort. Such a compensation does not occur on the rotational timescale. Conclusions: The rotational solar brightness variability as it would appear in the Kepler and CoRoT passbands from the ecliptic plane is spot-dominated, but the relative contribution of faculae increases for out-of-ecliptic viewing so that the apparent brightness variations are faculae-dominated for

  17. The abundance properties of nearby late-type galaxies. II. The relation between abundance distributions and surface brightness profiles

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Pilyugin, L. S.; Grebel, E. K.; Zinchenko, I. A.

    2014-12-01

    The relations between oxygen abundance and disk surface brightness (OH–SB relation) in the infrared W1 band are examined for nearby late-type galaxies. The oxygen abundances were presented in Paper I. The photometric characteristics of the disks are inferred here using photometric maps from the literature through bulge-disk decomposition. We find evidence that the OH–SB relation is not unique but depends on the galactocentric distance r (taken as a fraction of the optical radius R{sub 25}) and on the properties of a galaxy: the disk scale length h and the morphological T-type. We suggest a general, four-dimensional OH–SB relation with themore » values r, h, and T as parameters. The parametric OH–SB relation reproduces the observed data better than a simple, one-parameter relation; the deviations resulting when using our parametric relation are smaller by a factor of ∼1.4 than that of the simple relation. The influence of the parameters on the OH–SB relation varies with galactocentric distance. The influence of the T-type on the OH–SB relation is negligible at the centers of galaxies and increases with galactocentric distance. In contrast, the influence of the disk scale length on the OH–SB relation is at a maximum at the centers of galaxies and decreases with galactocentric distance, disappearing at the optical edges of galaxies. Two-dimensional relations can be used to reproduce the observed data at the optical edges of the disks and at the centers of the disks. The disk scale length should be used as a second parameter in the OH–SB relation at the center of the disk while the morphological T-type should be used as a second parameter in the relation at optical edge of the disk. The relations between oxygen abundance and disk surface brightness in the optical B and infrared K bands at the center of the disk and at optical edge of the disk are also considered. The general properties of the abundance–surface brightness relations are similar for

  18. Automated detection of very Low Surface Brightness galaxies in the Virgo Cluster

    NASA Astrophysics Data System (ADS)

    Prole, D. J.; Davies, J. I.; Keenan, O. C.; Davies, L. J. M.

    2018-04-01

    We report the automatic detection of a new sample of very low surface brightness (LSB) galaxies, likely members of the Virgo cluster. We introduce our new software, DeepScan, that has been designed specifically to detect extended LSB features automatically using the DBSCAN algorithm. We demonstrate the technique by applying it over a 5 degree2 portion of the Next-Generation Virgo Survey (NGVS) data to reveal 53 low surface brightness galaxies that are candidate cluster members based on their sizes and colours. 30 of these sources are new detections despite the region being searched specifically for LSB galaxies previously. Our final sample contains galaxies with 26.0 ≤ ⟨μe⟩ ≤ 28.5 and 19 ≤ mg ≤ 21, making them some of the faintest known in Virgo. The majority of them have colours consistent with the red sequence, and have a mean stellar mass of 106.3 ± 0.5M⊙ assuming cluster membership. After using ProFit to fit Sérsic profiles to our detections, none of the new sources have effective radii larger than 1.5 Kpc and do not meet the criteria for ultra-diffuse galaxy (UDG) classification, so we classify them as ultra-faint dwarfs.

  19. Using SMOS observations in the development of the SMAP level 4 surface and root-zone soil moisture project

    USDA-ARS?s Scientific Manuscript database

    The Soil Moisture and Ocean Salinity (SMOS; [1]) mission was launched by ESA in November 2009 and has since been observing L-band (1.4 GHz) upwelling passive microwaves. Along with these brightness temperature observations, ESA also disseminates retrievals of surface soil moisture that are derived ...

  20. WINDII airglow observations of wave superposition and the possible association with historical "bright nights"

    NASA Astrophysics Data System (ADS)

    Shepherd, G. G.; Cho, Y.-M.

    2017-07-01

    Longitudinal variations of airglow emission rate are prominent in all midlatitude nighttime O(1S) lower thermospheric data obtained with the Wind Imaging Interferometer (WINDII) on the Upper Atmosphere Research Satellite (UARS). The pattern generally appears as a combination of zonal waves 1, 2, 3, and 4 whose phases propagate at different rates. Sudden localized enhancements of 2 to 4 days duration are sometimes evident, reaching vertically integrated emission rates of 400 R, a factor of 10 higher than minimum values for the same day. These are found to occur when the four wave components come into the same phase at one longitude. It is shown that these highly localized longitudinal maxima are consistent with the historical phenomena known as "bright nights" in which the surroundings of human dark night observers were seen to be illuminated by this enhanced airglow.Plain Language SummaryFor centuries, going back to the Roman era, people have recorded experiences of brightened skies during the night, called "<span class="hlt">bright</span> nights." Currently, scientists study airglow, an emission of light from the high atmosphere, 100 km above us. Satellite <span class="hlt">observations</span> of a green airglow have shown that it consists of waves 1, 2, 3, and 4 around the earth. It happens that when the peaks of the different waves coincide there is an airglow brightening, and this article demonstrates that this event produces a <span class="hlt">bright</span> night. The modern data are shown to be entirely consistent with the historical <span class="hlt">observations</span>.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_6 --> <div id="page_7" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="121"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990116038&hterms=EIT&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DEIT','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990116038&hterms=EIT&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DEIT"><span>Micro Coronal <span class="hlt">Bright</span> Points <span class="hlt">Observed</span> in the Quiet Magnetic Network by SOHO/EIT</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Falconer, D. A.; Moore, R. L.; Porter, J. G.</p> <p>1997-01-01</p> <p>When one looks at SOHO/EIT Fe XII images of quiet regions, one can see the conventional coronal <span class="hlt">bright</span> points (> 10 arcsec in diameter), but one will also notice many smaller faint enhancements in <span class="hlt">brightness</span> (Figure 1). Do these micro coronal <span class="hlt">bright</span> points belong to the same family as the conventional <span class="hlt">bright</span> points? To investigate this question we compared SOHO/EIT Fe XII images with Kitt Peak magnetograms to determine whether the micro <span class="hlt">bright</span> points are in the magnetic network and mark magnetic bipoles within the network. To identify the coronal <span class="hlt">bright</span> points, we applied a picture frame filter to the Fe XII images; this brings out the Fe XII network and <span class="hlt">bright</span> points (Figure 2) and allows us to study the <span class="hlt">bright</span> points down to the resolution limit of the SOHO/EIT instrument. This picture frame filter is a square smoothing function (hlargelyalf a network cell wide) with a central square (quarter of a network cell wide) removed so that a <span class="hlt">bright</span> point's intensity does not effect its own background. This smoothing function is applied to the full disk image. Then we divide the original image by the smoothed image to obtain our filtered image. A <span class="hlt">bright</span> point is defined as any contiguous set of pixels (including diagonally) which have enhancements of 30% or more above the background; a micro <span class="hlt">bright</span> point is any <span class="hlt">bright</span> point 16 pixels or smaller in size. We then analyzed the <span class="hlt">bright</span> points that were fully within quiet regions (0.6 x 0.6 solar radius) centered on disk center on six different days.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...856..170G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...856..170G"><span>Tracers of Stellar Mass-loss. II. Mid-IR Colors and <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Fluctuations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>González-Lópezlira, Rosa A.</p> <p>2018-04-01</p> <p>I present integrated colors and <span class="hlt">surface</span> <span class="hlt">brightness</span> fluctuation magnitudes in the mid-infrared (IR), derived from stellar population synthesis models that include the effects of the dusty envelopes around thermally pulsing asymptotic giant branch (TP-AGB) stars. The models are based on the Bruzual & Charlot CB* isochrones; they are single-burst, range in age from a few Myr to 14 Gyr, and comprise metallicities between Z = 0.0001 and Z = 0.04. I compare these models to mid-IR data of AGB stars and star clusters in the Magellanic Clouds, and study the effects of varying self-consistently the mass-loss rate, the stellar parameters, and the output spectra of the stars plus their dusty envelopes. I find that models with a higher than fiducial mass-loss rate are needed to fit the mid-IR colors of “extreme” single AGB stars in the Large Magellanic Cloud. <span class="hlt">Surface</span> <span class="hlt">brightness</span> fluctuation magnitudes are quite sensitive to metallicity for 4.5 μm and longer wavelengths at all stellar population ages, and powerful diagnostics of mass-loss rate in the TP-AGB for intermediate-age populations, between 100 Myr and 2–3 Gyr.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016MNRAS.463.2746W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016MNRAS.463.2746W"><span>Galaxy And Mass Assembly (GAMA): detection of low-<span class="hlt">surface-brightness</span> galaxies from SDSS data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Williams, Richard P.; Baldry, I. K.; Kelvin, L. S.; James, P. A.; Driver, S. P.; Prescott, M.; Brough, S.; Brown, M. J. I.; Davies, L. J. M.; Holwerda, B. W.; Liske, J.; Norberg, P.; Moffett, A. J.; Wright, A. H.</p> <p>2016-12-01</p> <p>We report on a search for new low-<span class="hlt">surface-brightness</span> galaxies (LSBGs) using Sloan Digital Sky Survey (SDSS) data within the Galaxy And Mass Assembly (GAMA) equatorial fields. The search method consisted of masking objects detected with SDSS PHOTO, combining gri images weighted to maximize the expected signal-to-noise ratio, and smoothing the images. The processed images were then run through a detection algorithm that finds all pixels above a set threshold and groups them based on their proximity to one another. The list of detections was cleaned of contaminants such as diffraction spikes and the faint wings of masked objects. From these, selecting potentially the brightest in terms of total flux, a list of 343 LSBGs was produced having been confirmed using VISTA Kilo-degree Infrared Galaxy Survey (VIKING) imaging. The photometry of this sample was refined using the deeper VIKING Z band as the aperture-defining band. Measuring their g - I and J - K colours shows that most are consistent with being at redshifts less than 0.2. The photometry is carried out using an AUTO aperture for each detection giving <span class="hlt">surface</span> <span class="hlt">brightnesses</span> of μr ≳ 25 mag arcsec-2 and magnitudes of r > 19.8 mag. None of these galaxies are <span class="hlt">bright</span> enough to be within the GAMA main survey limit but could be part of future deeper surveys to measure the low-mass end of the galaxy stellar mass function.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006ApJ...642L.115H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006ApJ...642L.115H"><span>Intrinsic <span class="hlt">Brightness</span> Temperatures of AGN Jets</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Homan, D. C.; Kovalev, Y. Y.; Lister, M. L.; Ros, E.; Kellermann, K. I.; Cohen, M. H.; Vermeulen, R. C.; Zensus, J. A.; Kadler, M.</p> <p>2006-05-01</p> <p>We present a new method for studying the intrinsic <span class="hlt">brightness</span> temperatures of the parsec-scale jet cores of active galactic nuclei (AGNs). Our method uses <span class="hlt">observed</span> superluminal motions and <span class="hlt">observed</span> <span class="hlt">brightness</span> temperatures for a large sample of AGNs to constrain the characteristic intrinsic <span class="hlt">brightness</span> temperature of the sample as a whole. To study changes in intrinsic <span class="hlt">brightness</span> temperature, we assume that the Doppler factors of individual jets are constant in time, as justified by their relatively small changes in <span class="hlt">observed</span> flux density. We find that in their median-low <span class="hlt">brightness</span> temperature state, the sources in our sample have a narrow range of intrinsic <span class="hlt">brightness</span> temperatures centered on a characteristic temperature, Tint~=3×1010 K, which is close to the value expected for equipartition, when the energy in the radiating particles equals the energy stored in the magnetic fields. However, in their maximum <span class="hlt">brightness</span> state, we find that sources in our sample have a characteristic intrinsic <span class="hlt">brightness</span> temperature greater than 2×1011 K, which is well in excess of the equipartition temperature. In this state, we estimate that the energy in radiating particles exceeds the energy in the magnetic field by a factor of ~105. We suggest that the excess of particle energy when sources are in their maximum <span class="hlt">brightness</span> state is due to injection or acceleration of particles at the base of the jet. Our results suggest that the common method of estimating jet Doppler factors by using a single measurement of <span class="hlt">observed</span> <span class="hlt">brightness</span> temperature, the assumption of equipartition, or both may lead to large scatter or systematic errors in the derived values.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20130006618&hterms=microwaves+water+structure&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dmicrowaves%2Bwater%2Bstructure','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20130006618&hterms=microwaves+water+structure&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dmicrowaves%2Bwater%2Bstructure"><span><span class="hlt">Surface</span> and Atmospheric Contributions to Passive Microwave <span class="hlt">Brightness</span> Temperatures for Falling Snow Events</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Skofronick-Jackson, Gail; Johnson, Benjamin T.</p> <p>2011-01-01</p> <p>Physically based passive microwave precipitation retrieval algorithms require a set of relationships between satellite -<span class="hlt">observed</span> <span class="hlt">brightness</span> temperatures (TBs) and the physical state of the underlying atmosphere and <span class="hlt">surface</span>. These relationships are nonlinear, such that inversions are ill ]posed especially over variable land <span class="hlt">surfaces</span>. In order to elucidate these relationships, this work presents a theoretical analysis using TB weighting functions to quantify the percentage influence of the TB resulting from absorption, emission, and/or reflection from the <span class="hlt">surface</span>, as well as from frozen hydrometeors in clouds, from atmospheric water vapor, and from other contributors. The percentage analysis was also compared to Jacobians. The results are presented for frequencies from 10 to 874 GHz, for individual snow profiles, and for averages over three cloud-resolving model simulations of falling snow. The bulk structure (e.g., ice water path and cloud depth) of the underlying cloud scene was found to affect the resultant TB and percentages, producing different values for blizzard, lake effect, and synoptic snow events. The slant path at a 53 viewing angle increases the hydrometeor contributions relative to nadir viewing channels. Jacobians provide the magnitude and direction of change in the TB values due to a change in the underlying scene; however, the percentage analysis provides detailed information on how that change affected contributions to the TB from the <span class="hlt">surface</span>, hydrometeors, and water vapor. The TB percentage information presented in this paper provides information about the relative contributions to the TB and supplies key pieces of information required to develop and improve precipitation retrievals over land <span class="hlt">surfaces</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AAS...209.9702W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AAS...209.9702W"><span>A Search for Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Galaxies in the Ultraviolet with GALEX</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wyder, Ted K.; GALEX Science Team</p> <p>2006-12-01</p> <p>Low <span class="hlt">surface</span> <span class="hlt">brightness</span> (LSB) galaxies have traditionally been difficult to detect at visible wavelengths due to their low contrast with the night sky and their low numbers per deg2. We describe a new search for LSB galaxies using UV images from the Galaxy Evolution Explorer (GALEX) satellite. The images are from the GALEX Medium Imaging Survey targeting mainly areas of the sky within the Sloan Digital Sky Survey (SDSS) footprint. Due to the UV sky background at high Galactic latitudes reaching levels of only approximately 28 mag arcsec-2 as well as the relatively large sky coverage from GALEX, we can potentially search for LSB galaxies that would be difficult to detect optically.After first convolving the images with a suitable kernel, we select a diameter limited set of objects which we then inspect manually in order to remove image artifacts and other spurious detections. Red galaxies that have high optical <span class="hlt">surface</span> <span class="hlt">brightness</span> can be identified using either the ratio of far-UV to near-UV flux or via comparison to SDSS images. We quantify our selection limits using a set of artificial galaxy tests. Our goal is to find blue, ultra-LSB galaxies that would be virtually undetectable in large optical imaging surveys. GALEX is a NASA Small Explorer, launched in April 2003. We gratefully acknowledge NASA's support for construction, operation, and science analysis for the GALEX mission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AMT....11..161N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AMT....11..161N"><span>Error sources in the retrieval of aerosol information over <span class="hlt">bright</span> <span class="hlt">surfaces</span> from satellite measurements in the oxygen A band</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nanda, Swadhin; de Graaf, Martin; Sneep, Maarten; de Haan, Johan F.; Stammes, Piet; Sanders, Abram F. J.; Tuinder, Olaf; Pepijn Veefkind, J.; Levelt, Pieternel F.</p> <p>2018-01-01</p> <p><p class="p">Retrieving aerosol optical thickness and aerosol layer height over a <span class="hlt">bright</span> <span class="hlt">surface</span> from measured top-of-atmosphere reflectance spectrum in the oxygen A band is known to be challenging, often resulting in large errors. In certain atmospheric conditions and viewing geometries, a loss of sensitivity to aerosol optical thickness has been reported in the literature. This loss of sensitivity has been attributed to a phenomenon known as critical <span class="hlt">surface</span> albedo regime, which is a range of <span class="hlt">surface</span> albedos for which the top-of-atmosphere reflectance has minimal sensitivity to aerosol optical thickness. This paper extends the concept of critical <span class="hlt">surface</span> albedo for aerosol layer height retrievals in the oxygen A band, and discusses its implications. The underlying physics are introduced by analysing the top-of-atmosphere reflectance spectrum as a sum of atmospheric path contribution and <span class="hlt">surface</span> contribution, obtained using a radiative transfer model. Furthermore, error analysis of an aerosol layer height retrieval algorithm is conducted over dark and <span class="hlt">bright</span> <span class="hlt">surfaces</span> to show the dependence on <span class="hlt">surface</span> reflectance. The analysis shows that the derivative with respect to aerosol layer height of the atmospheric path contribution to the top-of-atmosphere reflectance is opposite in sign to that of the <span class="hlt">surface</span> contribution - an increase in <span class="hlt">surface</span> <span class="hlt">brightness</span> results in a decrease in information content. In the case of aerosol optical thickness, these derivatives are anti-correlated, leading to large retrieval errors in high <span class="hlt">surface</span> albedo regimes. The consequence of this anti-correlation is demonstrated with measured spectra in the oxygen A band from the GOME-2 instrument on board the Metop-A satellite over the 2010 Russian wildfires incident.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22661280-long-slit-spectroscopy-edge-low-surface-brightness-galaxies','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22661280-long-slit-spectroscopy-edge-low-surface-brightness-galaxies"><span>Long-slit Spectroscopy of Edge-on Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Galaxies</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Du, Wei; Wu, Hong; Zhu, Yinan</p> <p>2017-03-10</p> <p>We present long-slit optical spectra of 12 edge-on low <span class="hlt">surface</span> <span class="hlt">brightness</span> galaxies (LSBGs) positioned along their major axes. After performing reddening corrections for the emission-line fluxes measured from the extracted integrated spectra, we measured the gas-phase metallicities of our LSBG sample using both the [N ii]/H α and the R {sub 23} diagnostics. Both sets of oxygen abundances show good agreement with each other, giving a median value of 12 + log(O/H) = 8.26 dex. In the luminosity–metallicity plot, our LSBG sample is consistent with the behavior of normal galaxies. In the mass–metallicity diagram, our LSBG sample has lower metallicitiesmore » for lower stellar mass, similar to normal galaxies. The stellar masses estimated from z -band luminosities are comparable to those of prominent spirals. In a plot of the gas mass fraction versus metallicity, our LSBG sample generally agrees with other samples in the high gas mass fraction space. Additionally, we have studied stellar populations of three LSBGs, which have relatively reliable spectral continua and high signal-to-noise ratios, and qualitatively conclude that they have a potential dearth of stars with ages <1 Gyr instead of being dominated by stellar populations with ages >1 Gyr. Regarding the chemical evolution of our sample, the LSBG data appear to allow for up to 30% metal loss, but we cannot completely rule out the closed-box model. Additionally, we find evidence that our galaxies retain up to about three times as much of their metals compared with dwarfs, consistent with metal retention being related to galaxy mass. In conclusion, our data support the view that LSBGs are probably just normal disk galaxies continuously extending to the low end of <span class="hlt">surface</span> <span class="hlt">brightness</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998ApJ...508..132D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998ApJ...508..132D"><span>Testing Modified Newtonian Dynamics with Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Galaxies: Rotation Curve FITS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>de Blok, W. J. G.; McGaugh, S. S.</p> <p>1998-11-01</p> <p>We present modified Newtonian dynamics (MOND) fits to 15 rotation curves of low <span class="hlt">surface</span> <span class="hlt">brightness</span> (LSB) galaxies. Good fits are readily found, although for a few galaxies minor adjustments to the inclination are needed. Reasonable values for the stellar mass-to-light ratios are found, as well as an approximately constant value for the total (gas and stars) mass-to-light ratio. We show that the LSB galaxies investigated here lie on the one, unique Tully-Fisher relation, as predicted by MOND. The scatter on the Tully-Fisher relation can be completely explained by the <span class="hlt">observed</span> scatter in the total mass-to-light ratio. We address the question of whether MOND can fit any arbitrary rotation curve by constructing a plausible fake model galaxy. While MOND is unable to fit this hypothetical galaxy, a normal dark-halo fit is readily found, showing that dark matter fits are much less selective in producing fits. The good fits to rotation curves of LSB galaxies support MOND, especially because these are galaxies with large mass discrepancies deep in the MOND regime.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.470.1512W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.470.1512W"><span>A population of faint low <span class="hlt">surface</span> <span class="hlt">brightness</span> galaxies in the Perseus cluster core</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wittmann, Carolin; Lisker, Thorsten; Ambachew Tilahun, Liyualem; Grebel, Eva K.; Conselice, Christopher J.; Penny, Samantha; Janz, Joachim; Gallagher, John S.; Kotulla, Ralf; McCormac, James</p> <p>2017-09-01</p> <p>We present the detection of 89 low <span class="hlt">surface</span> <span class="hlt">brightness</span> (LSB), and thus low stellar density galaxy candidates in the Perseus cluster core, of the kind named 'ultra-diffuse galaxies', with mean effective V-band <span class="hlt">surface</span> <span class="hlt">brightnesses</span> 24.8-27.1 mag arcsec-2, total V-band magnitudes -11.8 to -15.5 mag, and half-light radii 0.7-4.1 kpc. The candidates have been identified in a deep mosaic covering 0.3 deg2, based on wide-field imaging data obtained with the William Herschel Telescope. We find that the LSB galaxy population is depleted in the cluster centre and only very few LSB candidates have half-light radii larger than 3 kpc. This appears consistent with an estimate of their tidal radius, which does not reach beyond the stellar extent even if we assume a high dark matter content (M/L = 100). In fact, three of our candidates seem to be associated with tidal streams, which points to their current disruption. Given that published data on faint LSB candidates in the Coma cluster - with its comparable central density to Perseus - show the same dearth of large objects in the core region, we conclude that these cannot survive the strong tides in the centres of massive clusters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25215794','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25215794"><span>Generalized dark-<span class="hlt">bright</span> vector soliton solution to the mixed coupled nonlinear Schrödinger equations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Manikandan, N; Radhakrishnan, R; Aravinthan, K</p> <p>2014-08-01</p> <p>We have constructed a dark-<span class="hlt">bright</span> N-soliton solution with 4N+3 real parameters for the physically interesting system of mixed coupled nonlinear Schrödinger equations. Using this as well as an asymptotic analysis we have investigated the interaction between dark-<span class="hlt">bright</span> vector solitons. Each colliding dark-<span class="hlt">bright</span> one-soliton at the asymptotic limits includes more coupling parameters not only in the polarization vector but also in the amplitude part. Our present solution generalizes the dark-<span class="hlt">bright</span> soliton in the literature with parametric constraints. By exploiting the role of such coupling parameters we are able to control certain interaction effects, namely beating, breathing, bouncing, attraction, jumping, etc., without affecting other soliton parameters. Particularly, the results of the interactions between the bound state dark-<span class="hlt">bright</span> vector solitons reveal oscillations in their amplitudes under certain parametric choices. A similar kind of effect was also <span class="hlt">observed</span> experimentally in the BECs. We have also characterized the solutions with complicated structure and nonobvious wrinkle to define polarization vector, envelope speed, envelope width, envelope amplitude, grayness, and complex modulation. It is interesting to identify that the polarization vector of the dark-<span class="hlt">bright</span> one-soliton evolves on a spherical <span class="hlt">surface</span> instead of a hyperboloid <span class="hlt">surface</span> as in the <span class="hlt">bright-bright</span> case of the mixed coupled nonlinear Schrödinger equations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011A%26A...526A.134A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011A%26A...526A.134A"><span>Hinode <span class="hlt">observations</span> and 3D magnetic structure of an X-ray <span class="hlt">bright</span> point</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alexander, C. E.; Del Zanna, G.; Maclean, R. C.</p> <p>2011-02-01</p> <p>Aims: We present complete Hinode Solar Optical Telescope (SOT), X-Ray Telescope (XRT)and EUV Imaging Spectrometer (EIS) <span class="hlt">observations</span> of an X-ray <span class="hlt">bright</span> point (XBP) <span class="hlt">observed</span> on the 10, 11 of October 2007 over its entire lifetime (~12 h). We aim to show how the measured plasma parameters of the XBP change over time and also what kind of similarities the X-ray emission has to a potential magnetic field model. Methods: Information from all three instruments on-board Hinode was used to study its entire evolution. XRT data was used to investigate the structure of the <span class="hlt">bright</span> point and to measure the X-ray emission. The EIS instrument was used to measure various plasma parameters over the entire lifetime of the XBP. Lastly, the SOT was used to measure the magnetic field strength and provide a basis for potential field extrapolations of the photospheric fields to be made. These were performed and then compared to the <span class="hlt">observed</span> coronal features. Results: The XBP measured ~15´´ in size and was found to be formed directly above an area of merging and cancelling magnetic flux on the photosphere. A good correlation between the rate of X-ray emission and decrease in total magnetic flux was found. The magnetic fragments of the XBP were found to vary on very short timescales (minutes), however the global quasi-bipolar structure remained throughout the lifetime of the XBP. The potential field extrapolations were a good visual fit to the <span class="hlt">observed</span> coronal loops in most cases, meaning that the magnetic field was not too far from a potential state. Electron density measurements were obtained using a line ratio of Fe XII and the average density was found to be 4.95 × 109 cm-3 with the volumetric plasma filling factor calculated to have an average value of 0.04. Emission measure loci plots were then used to infer a steady temperature of log Te [ K] ~ 6.1. The calculated Fe XII Doppler shifts show velocity changes in and around the <span class="hlt">bright</span> point of ±15 km s-1 which are <span class="hlt">observed</span> to change</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130008783','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130008783"><span>Estimating Sea <span class="hlt">Surface</span> Salinity and Wind Using Combined Passive and Active L-Band Microwave <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yueh, Simon H.; Chaubell, Mario J.</p> <p>2012-01-01</p> <p>Several L-band microwave radiometer and radar missions have been, or will be, operating in space for land and ocean <span class="hlt">observations</span>. These include the NASA Aquarius mission and the Soil Moisture Active Passive (SMAP) mission, both of which use combined passive/ active L-band instruments. Aquarius s passive/active L-band microwave sensor has been designed to map the salinity field at the <span class="hlt">surface</span> of the ocean from space. SMAP s primary objectives are for soil moisture and freeze/thaw detection, but it will operate continuously over the ocean, and hence will have significant potential for ocean <span class="hlt">surface</span> research. In this innovation, an algorithm has been developed to retrieve simultaneously ocean <span class="hlt">surface</span> salinity and wind from combined passive/active L-band microwave <span class="hlt">observations</span> of sea <span class="hlt">surfaces</span>. The algorithm takes advantage of the differing response of <span class="hlt">brightness</span> temperatures and radar backscatter to salinity, wind speed, and direction, thus minimizing the least squares error (LSE) measure, which signifies the difference between measurements and model functions of <span class="hlt">brightness</span> temperatures and radar backscatter. The algorithm uses the conjugate gradient method to search for the local minima of the LSE. Three LSE measures with different measurement combinations have been tested. The first LSE measure uses passive microwave data only with retrieval errors reaching 1 to 2 psu (practical salinity units) for salinity, and 1 to 2 m/s for wind speed. The second LSE measure uses both passive and active microwave data for vertical and horizontal polarizations. The addition of active microwave data significantly improves the retrieval accuracy by about a factor of five. To mitigate the impact of Faraday rotation on satellite <span class="hlt">observations</span>, the third LSE measure uses measurement combinations invariant under the Faraday rotation. For Aquarius, the expected RMS SSS (sea <span class="hlt">surface</span> salinity) error will be less than about 0.2 psu for low winds, and increases to 0.3 psu at 25 m/s wind speed</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002ApJ...581.1013B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002ApJ...581.1013B"><span>Exponential Stellar Disks in Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Galaxies: A Critical Test of Viscous Evolution</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bell, Eric F.</p> <p>2002-12-01</p> <p>Viscous redistribution of mass in Milky Way-type galactic disks is an appealing way of generating an exponential stellar profile over many scale lengths, almost independent of initial conditions, requiring only that the viscous timescale and star formation timescale are approximately equal. However, galaxies with solid-body rotation curves cannot undergo viscous evolution. Low <span class="hlt">surface</span> <span class="hlt">brightness</span> (LSB) galaxies have exponential <span class="hlt">surface</span> <span class="hlt">brightness</span> profiles, yet have slowly rising, nearly solid-body rotation curves. Because of this, viscous evolution may be inefficient in LSB galaxies: the exponential profiles, instead, would give important insight into initial conditions for galaxy disk formation. Using star formation laws from the literature and tuning the efficiency of viscous processes to reproduce an exponential stellar profile in Milky Way-type galaxies, I test the role of viscous evolution in LSB galaxies. Under the conservative and not unreasonable condition that LSB galaxies are gravitationally unstable for at least a part of their lives, I find that it is impossible to rule out a significant role for viscous evolution. This type of model still offers an attractive way of producing exponential disks, even in LSB galaxies with slowly rising rotation curves.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApJ...823..123T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApJ...823..123T"><span>Beyond 31 mag arcsec-2: The Frontier of Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Imaging with the Largest Optical Telescopes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Trujillo, Ignacio; Fliri, Jüergen</p> <p>2016-06-01</p> <p>The detection of structures in the sky with optical <span class="hlt">surface</span> <span class="hlt">brightnesses</span> fainter than 30 mag arcsec-2 (3σ in 10 × 10 arcsec boxes; r-band) has remained elusive in current photometric deep surveys. Here we show how present-day telescopes of 10 m class can provide broadband imaging 1.5-2 mag deeper than most previous results within a reasonable amount of time (I.e., <10 hr on-source integration). In particular, we illustrate the ability of the 10.4 m Gran Telescopio de Canarias telescope to produce imaging with a limiting <span class="hlt">surface</span> <span class="hlt">brightness</span> of 31.5 mag arcsec-2 (3σ in 10 × 10 arcsec boxes; r-band) using 8.1 hr on source. We apply this power to explore the stellar halo of the galaxy UGC 00180, a galaxy analogous to M31 located at ˜150 Mpc, by obtaining a radial profile of <span class="hlt">surface</span> <span class="hlt">brightness</span> down to μ r ˜ 33 mag arcsec-2. This depth is similar to that obtained using the star-counts techniques for Local Group galaxies, but is achieved at a distance where this technique is unfeasible. We find that the mass of the stellar halo of this galaxy is ˜4 × 109 M ⊙, I.e., (3 ± 1)% of the total stellar mass of the whole system. This amount of mass in the stellar halo is in agreement with current theoretical expectations for galaxies of this kind.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850041298&hterms=usher&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dusher','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850041298&hterms=usher&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dusher"><span>A medium-<span class="hlt">bright</span> quasar sample - New quasar <span class="hlt">surface</span> densities in the magnitude range from 16.4 to 17.65 for B</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mitchell, K. J.; Warnock, A., III; Usher, P. D.</p> <p>1984-01-01</p> <p>A new medium-<span class="hlt">bright</span> quasar sample (MBQS) is constructed from spectroscopic <span class="hlt">observations</span> of 140 <span class="hlt">bright</span> objects selected for varying degrees of blue and ultraviolet excess (B-UVX) in five Palomar 1.2 m Schmidt fields. The MBQS contains 32 quasars with B less than 17.65 mag. The new integral <span class="hlt">surface</span> densities in the B range from 16.45 to 17.65 mag are approximately 40 percent (or more) higher than expected. The MBQS and its redshift distribution increase the area of the Hubble diagram covered by complete samples of quasars. The general spectroscopic results indicate that the three-color classification process used to catalog the spectroscopic candidates (1) has efficiently separated the intrinsically B-UVX stellar objects from the Population II subdwarfs and (2) has produced samples of B-UVX objects which are more complete than samples selected by (U - B) color alone.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997ApJ...481..735F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997ApJ...481..735F"><span>An Ultraviolet and Near-Infrared View of NGC 4214: A Starbursting Core Embedded in a Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Disk</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fanelli, Michael N.; Waller, William W.; Smith, Denise A.; Freedman, Wendy L.; Madore, Barry; Neff, Susan G.; O'Connell, Robert W.; Roberts, Morton S.; Bohlin, Ralph; Smith, Andrew M.; Stecher, Theodore P.</p> <p>1997-05-01</p> <p>During the Astro-2 Spacelab mission in 1995 March, the Ultraviolet Imaging Telescope (UIT) obtained far-UV (λ = 1500 Å) imagery of the nearby Sm/Im galaxy NGC 4214. The UIT images have a spatial resolution of ~3" and a limiting <span class="hlt">surface</span> <span class="hlt">brightness</span>, μ1500 > 25 mag arcsec-2, permitting detailed investigation of the intensity and spatial distribution of the young, high-mass stellar component. These data provide the first far-UV imagery covering the full spatial extent of NGC 4214. Comparison with a corresponding I-band image reveals the presence of a starbursting core embedded in an extensive low <span class="hlt">surface</span> <span class="hlt">brightness</span> disk. In the far-UV (FUV), NGC 4214 is resolved into several components: a luminous, central knot; an inner region (r <~ 2.5 kpc) with ~15 resolved sources embedded in <span class="hlt">bright</span>, diffuse emission; and a population of fainter knots extending to the edge of the optically defined disk (r ~ 5 kpc). The FUV light, which traces recent massive star formation, is <span class="hlt">observed</span> to be more centrally concentrated than the I-band light, which traces the global stellar population. The FUV radial light profile is remarkably well represented by an R1/4 law, providing evidence that the centrally concentrated massive star formation in NGC 4214 is the result of an interaction, possibly a tidal encounter, with a dwarf companion(s). The brightest FUV source produces ~8% of the global FUV luminosity. This unresolved source, corresponding to the Wolf-Rayet knot described by Sargent & Filippenko, is located at the center of the FUV light distribution, giving NGC 4214 an active galactic nucleus-like morphology. Another strong source is present in the I band, located 19" west, 10" north of the central starburst knot, with no FUV counterpart. The I-band source may be the previously unrecognized nucleus of NGC 4214 or an evolved star cluster with an age greater than ~200 Myr. The global star formation rate derived from the total FUV flux is consistent with rates derived using data at other</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4705039','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4705039"><span>Precise stellar <span class="hlt">surface</span> gravities from the time scales of convectively driven <span class="hlt">brightness</span> variations</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Kallinger, Thomas; Hekker, Saskia; García, Rafael A.; Huber, Daniel; Matthews, Jaymie M.</p> <p>2016-01-01</p> <p>A significant part of the intrinsic <span class="hlt">brightness</span> variations in cool stars of low and intermediate mass arises from <span class="hlt">surface</span> convection (seen as granulation) and acoustic oscillations (p-mode pulsations). The characteristics of these phenomena are largely determined by the stars’ <span class="hlt">surface</span> gravity (g). Detailed photometric measurements of either signal can yield an accurate value of g. However, even with ultraprecise photometry from NASA’s Kepler mission, many stars are too faint for current methods or only moderate accuracy can be achieved in a limited range of stellar evolutionary stages. This means that many of the stars in the Kepler sample, including exoplanet hosts, are not sufficiently characterized to fully describe the sample and exoplanet properties. We present a novel way to measure <span class="hlt">surface</span> gravities with accuracies of about 4%. Our technique exploits the tight relation between g and the characteristic time scale of the combined granulation and p-mode oscillation signal. It is applicable to all stars with a convective envelope, including active stars. It can measure g in stars for which no other analysis is now possible. Because it depends on the time scale (and no other properties) of the signal, our technique is largely independent of the type of measurement (for example, photometry or radial velocity measurements) and the calibration of the instrumentation used. However, the oscillation signal must be temporally resolved; thus, it cannot be applied to dwarf stars <span class="hlt">observed</span> by Kepler in its long-cadence mode. PMID:26767193</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26767193','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26767193"><span>Precise stellar <span class="hlt">surface</span> gravities from the time scales of convectively driven <span class="hlt">brightness</span> variations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kallinger, Thomas; Hekker, Saskia; García, Rafael A; Huber, Daniel; Matthews, Jaymie M</p> <p>2016-01-01</p> <p>A significant part of the intrinsic <span class="hlt">brightness</span> variations in cool stars of low and intermediate mass arises from <span class="hlt">surface</span> convection (seen as granulation) and acoustic oscillations (p-mode pulsations). The characteristics of these phenomena are largely determined by the stars' <span class="hlt">surface</span> gravity (g). Detailed photometric measurements of either signal can yield an accurate value of g. However, even with ultraprecise photometry from NASA's Kepler mission, many stars are too faint for current methods or only moderate accuracy can be achieved in a limited range of stellar evolutionary stages. This means that many of the stars in the Kepler sample, including exoplanet hosts, are not sufficiently characterized to fully describe the sample and exoplanet properties. We present a novel way to measure <span class="hlt">surface</span> gravities with accuracies of about 4%. Our technique exploits the tight relation between g and the characteristic time scale of the combined granulation and p-mode oscillation signal. It is applicable to all stars with a convective envelope, including active stars. It can measure g in stars for which no other analysis is now possible. Because it depends on the time scale (and no other properties) of the signal, our technique is largely independent of the type of measurement (for example, photometry or radial velocity measurements) and the calibration of the instrumentation used. However, the oscillation signal must be temporally resolved; thus, it cannot be applied to dwarf stars <span class="hlt">observed</span> by Kepler in its long-cadence mode.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=426537','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=426537"><span><span class="hlt">Surface</span> Photometry of Celestial Sources from a Space Vehicle: Introduction and <span class="hlt">Observational</span> Procedures*</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Roach, Franklin E.; Carroll, Benjamin; Aller, Lawrence H.; Smith, Leroi</p> <p>1972-01-01</p> <p>Diffuse celestial sources of relatively low <span class="hlt">surface</span> <span class="hlt">brightness</span> such as the Milky Way, zodiacal light, and gegenschein (or contre lumière) can be studied most reliably from above the earth's atmosphere with equipment flown in artificial satellites. We review the techniques used and some of the difficulties encountered in day-time <span class="hlt">observations</span> from satellites by the use of a special photometer and polarimeter flown in the orbiting skylab observatory, OSO-6. PMID:16591970</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017OptLE..88..120F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017OptLE..88..120F"><span>Exploring combined dark and <span class="hlt">bright</span> field illumination to improve the detection of defects on specular <span class="hlt">surfaces</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Forte, Paulo M. F.; Felgueiras, P. E. R.; Ferreira, Flávio P.; Sousa, M. A.; Nunes-Pereira, Eduardo J.; Bret, Boris P. J.; Belsley, Michael S.</p> <p>2017-01-01</p> <p>An automatic optical inspection system for detecting local defects on specular <span class="hlt">surfaces</span> is presented. The system uses an image display to produce a sequence of structured diffuse illumination patterns and a digital camera to acquire the corresponding sequence of images. An image enhancement algorithm, which measures the local intensity variations between <span class="hlt">bright</span>- and dark-field illumination conditions, yields a final image in which the defects are revealed with a high contrast. Subsequently, an image segmentation algorithm, which compares statistically the enhanced image of the inspected <span class="hlt">surface</span> with the corresponding image for a defect-free template, allows separating defects from non-defects with an adjusting decision threshold. The method can be applied to shiny <span class="hlt">surfaces</span> of any material including metal, plastic and glass. The described method was tested on the plastic <span class="hlt">surface</span> of a car dashboard system. We were able to detect not only scratches but also dust and fingerprints. In our experiment we <span class="hlt">observed</span> a detection contrast increase from about 40%, when using an extended light source, to more than 90% when using a structured light source. The presented method is simple, robust and can be carried out with short cycle times, making it appropriate for applications in industrial environments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JASTP.167...66B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JASTP.167...66B"><span>Making limb and nadir measurements comparable: A common volume study of PMC <span class="hlt">brightness</span> <span class="hlt">observed</span> by Odin OSIRIS and AIM CIPS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Benze, Susanne; Gumbel, Jörg; Randall, Cora E.; Karlsson, Bodil; Hultgren, Kristoffer; Lumpe, Jerry D.; Baumgarten, Gerd</p> <p>2018-01-01</p> <p>Combining limb and nadir satellite <span class="hlt">observations</span> of Polar Mesospheric Clouds (PMCs) has long been recognized as problematic due to differences in <span class="hlt">observation</span> geometry, scattering conditions, and retrieval approaches. This study offers a method of comparing PMC <span class="hlt">brightness</span> <span class="hlt">observations</span> from the nadir-viewing Aeronomy of Ice in the Mesosphere (AIM) Cloud Imaging and Particle Size (CIPS) instrument and the limb-viewing Odin Optical Spectrograph and InfraRed Imaging System (OSIRIS). OSIRIS and CIPS measurements are made comparable by defining a common volume for overlapping OSIRIS and CIPS <span class="hlt">observations</span> for two northern hemisphere (NH) PMC seasons: NH08 and NH09. We define a scattering intensity quantity that is suitable for either nadir or limb <span class="hlt">observations</span> and for different scattering conditions. A known CIPS bias is applied, differences in instrument sensitivity are analyzed and taken into account, and effects of cloud inhomogeneity and common volume definition on the comparison are discussed. Not accounting for instrument sensitivity differences or inhomogeneities in the PMC field, the mean relative difference in cloud <span class="hlt">brightness</span> (CIPS - OSIRIS) is -102 ± 55%. The differences are largest for coincidences with very inhomogeneous clouds that are dominated by pixels that CIPS reports as non-cloud points. Removing these coincidences, the mean relative difference in cloud <span class="hlt">brightness</span> reduces to -6 ± 14%. The correlation coefficient between the CIPS and OSIRIS measurements of PMC <span class="hlt">brightness</span> variations in space and time is remarkably high, at 0.94. Overall, the comparison shows excellent agreement despite different retrieval approaches and <span class="hlt">observation</span> geometries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016A%26A...596A..23S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016A%26A...596A..23S"><span>Stellar systems in the direction of the Hickson Compact Group 44. I. Low <span class="hlt">surface</span> <span class="hlt">brightness</span> galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Smith Castelli, A. V.; Faifer, F. R.; Escudero, C. G.</p> <p>2016-11-01</p> <p>Context. In spite of the numerous studies of low-luminosity galaxies in different environments, there is still no consensus about their formation scenario. In particular, a large number of galaxies displaying extremely low-<span class="hlt">surface</span> <span class="hlt">brightnesses</span> have been detected in the last year, and the nature of these objects is under discussion. Aims: In this paper we report the detection of two extended low-<span class="hlt">surface</span> <span class="hlt">brightness</span> (LSB) objects (μeffg' ≃ 27 mag) found, in projection, next to NGC 3193 and in the zone of the Hickson Compact Group (HCG) 44, respectively. Methods: We analyzed deep, high-quality, GEMINI-GMOS images with ELLIPSE within IRAF in order to obtain their <span class="hlt">brightness</span> profiles and structural parameters. We also searched for the presence of globular clusters (GC) in these fields. Results: We have found that, if these LSB galaxies were at the distances of NGC 3193 and HCG 44, they would show sizes and luminosities similar to those of the ultra-diffuse galaxies (UDGs) found in the Coma cluster and other associations. In that case, their sizes would be rather larger than those displayed by the Local Group dwarf spheroidal (dSph) galaxies. We have detected a few unresolved sources in the sky zone occupied by these galaxies showing colors and <span class="hlt">brightnesses</span> typical of blue globular clusters. Conclusions: From the comparison of the properties of the galaxies presented in this work with those of similar objects reported in the literature, we have found that LSB galaxies display sizes covering a quite extended continous range (reff 0.3-4.5 kpc), in contrast to "normal" early-type galaxies, which show reff 1.0 kpc with a low dispersion. This fact might point to different formation processes for both types of galaxies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120013214','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120013214"><span>Synthesizing SMOS Zero-Baselines with Aquarius <span class="hlt">Brightness</span> Temperature Simulator</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Colliander, A.; Dinnat, E.; Le Vine, D.; Kainulainen, J.</p> <p>2012-01-01</p> <p>SMOS [1] and Aquarius [2] are ESA and NASA missions, respectively, to make L-band measurements from the Low Earth Orbit. SMOS makes passive measurements whereas Aquarius measures both passive and active. SMOS was launched in November 2009 and Aquarius in June 2011.The scientific objectives of the missions are overlapping: both missions aim at mapping the global Sea <span class="hlt">Surface</span> Salinity (SSS). Additionally, SMOS mission produces soil moisture product (however, Aquarius data will eventually be used for retrieving soil moisture too). The consistency of the <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> made by the two instruments is essential for long-term studies of SSS and soil moisture. For resolving the consistency, the calibration of the instruments is the key. The basis of the SMOS <span class="hlt">brightness</span> temperature level is the measurements performed with the so-called zero-baselines [3]; SMOS employs an interferometric measurement technique which forms a <span class="hlt">brightness</span> temperature image from several baselines constructed by combination of multiple receivers in an array; zero-length baseline defines the overall <span class="hlt">brightness</span> temperature level. The basis of the Aquarius <span class="hlt">brightness</span> temperature level is resolved from the <span class="hlt">brightness</span> temperature simulator combined with ancillary data such as antenna patterns and environmental models [4]. Consistency between the SMOS zero-baseline measurements and the simulator output would provide a robust basis for establishing the overall comparability of the missions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950031179&hterms=ART+ROCK&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DART%2BROCK','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950031179&hterms=ART+ROCK&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DART%2BROCK"><span><span class="hlt">Surface</span>-induced <span class="hlt">brightness</span> temperature variations and their effects on detecting thin cirrus clouds using IR emission channels in the 8-12 microns region</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gao, Bo-Cai; Wiscombe, W. J.</p> <p>1994-01-01</p> <p>A method for detecting cirrus clouds in terms of <span class="hlt">brightness</span> temperature differences between narrowbands at 8, 11, and 12 microns has been proposed by Ackerman et al. In this method, the variation of emissivity with wavelength for different <span class="hlt">surface</span> targets was not taken into consideration. Based on state-of-the-art laboratory measurements of reflectance spectra of terrestrial materials by Salisbury and D'Aria, it is found that the <span class="hlt">brightness</span> temperature differences between the 8- and 11-microns bands for soils, rocks, and minerals, and dry vegetation can vary between approximately -8 and +8 K due solely to <span class="hlt">surface</span> emissivity variations. The large <span class="hlt">brightness</span> temperature differences are sufficient to cause false detection of cirrus clouds from remote sensing data acquired over certain <span class="hlt">surface</span> targets using the 8-11-12-microns method directly. It is suggested that the 8-11-12-microns method should be improved to include the <span class="hlt">surface</span> emissivity effects. In addition, it is recommended that in the future the variation of <span class="hlt">surface</span> emissivity with wavelength should be taken into account in algorithms for retrieving <span class="hlt">surface</span> temperatures and low-level atmospheric temperature and water vapor profiles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140013215','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140013215"><span>VERITAS <span class="hlt">Observations</span> of Six <span class="hlt">Bright</span>, Hard-Spectrum Fermi-LAT Blazars</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>E. Aliu; Archambault, S.; Arlen, T.; Aune, T.; Beilicke, M.; Benbow, W.; Boettcher, M.; Bouvier, A.; Buckley, J. H.; Bugaev, V.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20140013215'); toggleEditAbsImage('author_20140013215_show'); toggleEditAbsImage('author_20140013215_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20140013215_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20140013215_hide"></p> <p>2012-01-01</p> <p>We report on VERITAS very-high-energy (VHE; E >= 100 GeV) <span class="hlt">observations</span> of six blazars selected from the Fermi Large Area Telescope First Source Catalog (1FGL). The gamma-ray emission from 1FGL sources was extrapolated up to the VHE band, taking gamma-ray absorption by the extragalactic background light into account. This allowed the selection of six <span class="hlt">bright</span>, hard-spectrum blazars that were good candidate TeV emitters. Spectroscopic redshift measurements were attempted with the Keck Telescope for the targets without Sloan Digital Sky Survey (SDSS) spectroscopic data. No VHE emission is detected during the <span class="hlt">observations</span> of the six sources described here. Corresponding TeV upper limits are presented, along with contemporaneous Fermi <span class="hlt">observations</span> and non-concurrent Swift UVOT and XRT data. The blazar broadband spectral energy distributions (SEDs) are assembled and modeled with a single-zone synchrotron self-Compton model. The SED built for each of the six blazars show a synchrotron peak bordering between the intermediate- and high-spectrum-peak classifications, with four of the six resulting in particle-dominated emission region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...851...22M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...851...22M"><span>The Star-forming Main Sequence of Dwarf Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McGaugh, Stacy S.; Schombert, James M.; Lelli, Federico</p> <p>2017-12-01</p> <p>We explore the star-forming properties of late-type, low <span class="hlt">surface</span> <span class="hlt">brightness</span> (LSB) galaxies. The star-forming main sequence ({SFR}-{M}* ) of LSB dwarfs has a steep slope, indistinguishable from unity (1.04 ± 0.06). They form a distinct sequence from more massive spirals, which exhibit a shallower slope. The break occurs around {M}* ≈ {10}10 {M}⊙ , and can also be seen in the gas mass—stellar mass plane. The global Kennicutt-Schmidt law ({SFR}-{M}g) has a slope of 1.47 ± 0.11 without the break seen in the main sequence. There is an ample supply of gas in LSB galaxies, which have gas depletion times well in excess of a Hubble time, and often tens of Hubble times. Only ˜ 3 % of this cold gas needs be in the form of molecular gas to sustain the <span class="hlt">observed</span> star formation. In analogy with the faint, long-lived stars of the lower stellar main sequence, it may be appropriate to consider the main sequence of star-forming galaxies to be defined by thriving dwarfs (with {M}* < {10}10 {M}⊙ ), while massive spirals (with {M}* > {10}10 {M}⊙ ) are weary giants that constitute more of a turn-off population.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25818045','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25818045"><span><span class="hlt">Brightness</span> masking is modulated by disparity structure.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pelekanos, Vassilis; Ban, Hiroshi; Welchman, Andrew E</p> <p>2015-05-01</p> <p>The luminance contrast at the borders of a <span class="hlt">surface</span> strongly influences <span class="hlt">surface</span>'s apparent <span class="hlt">brightness</span>, as demonstrated by a number of classic visual illusions. Such phenomena are compatible with a propagation mechanism believed to spread contrast information from borders to the interior. This process is disrupted by masking, where the perceived <span class="hlt">brightness</span> of a target is reduced by the brief presentation of a mask (Paradiso & Nakayama, 1991), but the exact visual stage that this happens remains unclear. In the present study, we examined whether <span class="hlt">brightness</span> masking occurs at a monocular-, or a binocular-level of the visual hierarchy. We used backward masking, whereby a briefly presented target stimulus is disrupted by a mask coming soon afterwards, to show that <span class="hlt">brightness</span> masking is affected by binocular stages of the visual processing. We manipulated the 3-D configurations (slant direction) of the target and mask and measured the differential disruption that masking causes on <span class="hlt">brightness</span> estimation. We found that the masking effect was weaker when stimuli had a different slant. We suggest that <span class="hlt">brightness</span> masking is partly mediated by mid-level neuronal mechanisms, at a stage where binocular disparity edge structure has been extracted. Copyright © 2015 The Authors. Published by Elsevier Ltd.. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015DPS....4740002H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015DPS....4740002H"><span><span class="hlt">Bright</span> features in Neptune on 2013-2015 from ground-based <span class="hlt">observations</span> with small (40 cm) and large telescopes (10 m)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hueso, Ricardo; Delcroix, Marc; Baranec, Christoph; Sánchez-Lavega, Agustín; María Gómez-Forrellad, Josep; Félix Rojas, Jose; Luszcz-Cook, Statia; de Pater, Imke; de Kleer, Katherine; Colas, François; Guarro, Joan; Goczynski, Peter; Jones, Paul; Kivits, Willem; Maxson, Paul; Phillips, Michael; Sussenbach, John; Wesley, Anthony; Hammel, Heidi B.; Pérez-Hoyos, Santiago; Mendikoa, Iñigo; Riddle, Reed; Law, Nicholas M.; Sayanagi, Kunio</p> <p>2015-11-01</p> <p><span class="hlt">Observations</span> of Neptune over the last few years obtained with small telescopes (30-50 cm) have resulted in several detections of <span class="hlt">bright</span> features on the planet. In 2013, 2014 and 2015, different <span class="hlt">observers</span> have repeatedly <span class="hlt">observed</span> features of high contrast at Neptune’s mid-latitudes using long-pass red filters. This success at <span class="hlt">observing</span> Neptune clouds with such small telescopes is due to the presence of strong methane absorption bands in Neptune’s spectra at red and near infrared wavelengths; these bands provide good contrast for elevated cloud structures. In each case, the atmospheric features identified in the images survived at least a few weeks, but were essentially much more variable and apparently shorter-lived, than the large convective system recently reported on Uranus [de Pater et al. 2015]. The latest and brightest spot on Neptune was first detected on July 13th 2015 with the 2.2m telescope at Calar Alto observatory with the PlanetCam UPV/EHU instrument. The range of wavelengths covered by PlanetCam (from 350 nm to the H band including narrow-band and wide-band filters in and out of methane bands) allows the study of the vertical cloud structure of this <span class="hlt">bright</span> spot. In particular, the spot is particularly well contrasted at the H band where it accounted to a 40% of the total planet <span class="hlt">brightness</span>. <span class="hlt">Observations</span> obtained with small telescopes a few days later provide a good comparison that can be used to scale similar structures in 2013 and 2014 that were <span class="hlt">observed</span> with 30-50 cm telescopes and the Robo-AO instrument at Palomar observatory. Further high-resolution <span class="hlt">observations</span> of the 2015 event were obtained in July 25th with the NIRC2 camera in the Keck 2 10-m telescope. These images show the <span class="hlt">bright</span> spot as a compact <span class="hlt">bright</span> feature in H band with a longitudinal size of 8,300 km and a latitudinal extension of 5,300 km, well separated from a nearby <span class="hlt">bright</span> band. The ensemble of <span class="hlt">observations</span> locate the structure at -41º latitude drifting at about +24.27º/day or</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.470.2133M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.470.2133M"><span>Extragalactic background light: a measurement at 400 nm using dark cloud shadow*†- I. Low <span class="hlt">surface</span> <span class="hlt">brightness</span> spectrophotometry in the area of Lynds 1642</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mattila, K.; Lehtinen, K.; Väisänen, P.; von Appen-Schnur, G.; Leinert, Ch.</p> <p>2017-09-01</p> <p>We present the method and <span class="hlt">observations</span> for the measurement of the Extragalactic Background Light (EBL) utilizing the shadowing effect of a dark cloud. We measure the <span class="hlt">surface</span> <span class="hlt">brightness</span> difference between the opaque cloud core and its unobscured surroundings. In the difference the large atmospheric and Zodiacal light components are eliminated and the only remaining foreground component is the scattered starlight from the cloud itself. Although much smaller, its separation is the key problem in the method. For its separation we use spectroscopy. While the scattered starlight has the characteristic Fraunhofer lines and 400 nm discontinuity, the EBL spectrum is smooth and without these features. Medium resolution spectrophotometry at λ = 380-580 nm was performed with VLT/FORS at ESO of the <span class="hlt">surface</span> <span class="hlt">brightness</span> in and around the high-galactic-latitude dark cloud Lynds 1642. Besides the spectrum for the core with AV ≳ 15 mag, further spectra were obtained for intermediate-opacity cloud positions. They are used as proxy for the spectrum of the impinging starlight spectrum and to facilitate the separation of the scattered starlight (cf. Paper II; Mattila et al.). Our spectra reach a precision of ≲ 0.5 × 10-9 erg cm-2 s-1 sr-1 Å-1 as required to measure an EBL intensity in range of ˜1 to a few times 10-9 erg cm-2 s-1 sr-1 Å-1. Because all <span class="hlt">surface</span> <span class="hlt">brightness</span> components are measured using the same equipment, the method does not require unusually high absolute calibration accuracy, a condition that has been a problem for some previous EBL projects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3231112','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3231112"><span>L Band <span class="hlt">Brightness</span> Temperature <span class="hlt">Observations</span> over a Corn Canopy during the Entire Growth Cycle</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Joseph, Alicia T.; van der Velde, Rogier; O’Neill, Peggy E.; Choudhury, Bhaskar J.; Lang, Roger H.; Kim, Edward J.; Gish, Timothy</p> <p>2010-01-01</p> <p>During a field campaign covering the 2002 corn growing season, a dual polarized tower mounted L-band (1.4 GHz) radiometer (LRAD) provided <span class="hlt">brightness</span> temperature (TB) measurements at preset intervals, incidence and azimuth angles. These radiometer measurements were supported by an extensive characterization of land <span class="hlt">surface</span> variables including soil moisture, soil temperature, vegetation biomass, and <span class="hlt">surface</span> roughness. In the period May 22 to August 30, ten days of radiometer and ground measurements are available for a corn canopy with a vegetation water content (W) range of 0.0 to 4.3 kg m−2. Using this data set, the effects of corn vegetation on <span class="hlt">surface</span> emissions are investigated by means of a semi-empirical radiative transfer model. Additionally, the impact of roughness on the <span class="hlt">surface</span> emission is quantified using TB measurements over bare soil conditions. Subsequently, the estimated roughness parameters, ground measurements and horizontally (H)-polarized TB are employed to invert the H-polarized transmissivity (γh) for the monitored corn growing season. PMID:22163585</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AAS...23221801M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AAS...23221801M"><span>Exploring the <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Breaks and Star Formation in Disk Galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Malko, Bradley Ann; Hunter, Deidre Ann</p> <p>2018-06-01</p> <p>Stellar <span class="hlt">surface</span> <span class="hlt">brightness</span> profiles of both spirals and dwarf irregular galaxies often show breaks in which the exponential fall-off abruptly changes slope. Most often the profile is down-bending (Type II) in the outer disk, but sometimes it is up-bending (Type III). Stellar disks extend a long ways beyond the profile breaks, but we do not understand what happens physically at the breaks. To explore this we are examining the star formation activity, as traced with FUV emission, interior to the break compared to that exterior to the break in both dwarf irregulars and spiral galaxies. We present the results for the spiral galaxy NGC 2500 and compare it to the LITTLE THINGS dwarf irregular galaxies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22661025-interferometric-monitoring-gamma-ray-bright-agns-results-single-epoch-multifrequency-observations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22661025-interferometric-monitoring-gamma-ray-bright-agns-results-single-epoch-multifrequency-observations"><span>INTERFEROMETRIC MONITORING OF GAMMA-RAY <span class="hlt">BRIGHT</span> AGNs. I. THE RESULTS OF SINGLE-EPOCH MULTIFREQUENCY <span class="hlt">OBSERVATIONS</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Lee, Sang-Sung; Wajima, Kiyoaki; Algaba, Juan-Carlos</p> <p>2016-11-01</p> <p>We present results of single-epoch very long baseline interferometry (VLBI) <span class="hlt">observations</span> of gamma-ray <span class="hlt">bright</span> active galactic nuclei (AGNs) using the Korean VLBI Network (KVN) at the 22, 43, 86, and 129 GHz bands, which are part of a KVN key science program, Interferometric Monitoring of Gamma-Ray <span class="hlt">Bright</span> AGNs. We selected a total of 34 radio-loud AGNs of which 30 sources are gamma-ray <span class="hlt">bright</span> AGNs with flux densities of >6 × 10{sup −10} ph cm{sup −2} s{sup −1}. Single-epoch multifrequency VLBI <span class="hlt">observations</span> of the target sources were conducted during a 24 hr session on 2013 November 19 and 20. All <span class="hlt">observed</span> sources weremore » detected and imaged at all frequency bands, with or without a frequency phase transfer technique, which enabled the imaging of 12 faint sources at 129 GHz, except for one source. Many of the target sources are resolved on milliarcsecond scales, yielding a core-jet structure, with the VLBI core dominating the synchrotron emission on a milliarcsecond scale. CLEAN flux densities of the target sources are 0.43–28 Jy, 0.32–21 Jy, 0.18–11 Jy, and 0.35–8.0 Jy in the 22, 43, 86, and 129 GHz bands, respectively. Spectra of the target sources become steeper at higher frequency, with spectral index means of −0.40, −0.62, and −1.00 in the 22–43 GHz, 43–86 GHz and 86–129 GHz bands, respectively, implying that the target sources become optically thin at higher frequencies (e.g., 86–129 GHz).« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22348403-discovery-megaparsec-scale-low-surface-brightness-nonthermal-emission-merging-galaxy-clusters-using-green-bank-telescope','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22348403-discovery-megaparsec-scale-low-surface-brightness-nonthermal-emission-merging-galaxy-clusters-using-green-bank-telescope"><span>Discovery of megaparsec-scale, low <span class="hlt">surface</span> <span class="hlt">brightness</span> nonthermal emission in merging galaxy clusters using the green bank telescope</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Farnsworth, Damon; Rudnick, Lawrence; Brown, Shea</p> <p>2013-12-20</p> <p>We present results from a study of 12 X-ray <span class="hlt">bright</span> clusters at 1.4 GHz with the 100 m Green Bank Telescope. After subtraction of point sources using existing interferometer data, we reach a median (best) 1σ rms sensitivity level of 0.01 (0.006) μJy arcsec{sup –2}, and find a significant excess of diffuse, low <span class="hlt">surface</span> <span class="hlt">brightness</span> emission in 11 of 12 Abell clusters <span class="hlt">observed</span>. We also present initial results at 1.4 GHz of A2319 from the Very Large Array. In particular, we find: (1) four new detections of diffuse structures tentatively classified as two halos (A2065, A2069) and two relics (A2067,more » A2073); (2) the first detection of the radio halo in A2061 at 1.4 GHz, which qualifies this as a possible ultra-steep spectrum halo source with a synchrotron spectral index of α ∼ 1.8 between 327 MHz and 1.4 GHz; (3) a ∼2 Mpc radio halo in the sloshing, minor-merger cluster A2142; (4) a >2× increase of the giant radio halo extent and luminosity in the merging cluster A2319; (5) a ∼7× increase to the integrated radio flux and >4× increase to the <span class="hlt">observed</span> extent of the peripheral radio relic in A1367 to ∼600 kpc, which we also <span class="hlt">observe</span> to be polarized on a similar scale; (6) significant excess emission of ambiguous nature in three clusters with embedded tailed radio galaxies (A119, A400, A3744). Our radio halo detections agree with the well-known X-ray/radio luminosity correlation, but they are larger and fainter than current radio power correlation studies would predict. The corresponding volume-averaged synchrotron emissivities are 1-2 orders of magnitude below the characteristic value found in previous studies. Some of the halo-like detections may be some type of previously unseen, low <span class="hlt">surface</span> <span class="hlt">brightness</span> radio halo or blend of unresolved shock structures and sub-Mpc-scale turbulent regions associated with their respective cluster merging activity. Four of the five tentative halos contain one or more X-ray cold fronts, suggesting a possible connection</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3152653','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3152653"><span>Spatiotemporal analysis of <span class="hlt">brightness</span> induction</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>McCourt, Mark E.</p> <p>2011-01-01</p> <p><span class="hlt">Brightness</span> induction refers to a class of visual illusions in which the perceived intensity of a region of space is influenced by the luminance of surrounding regions. These illusions are significant because they provide insight into the neural organization of the visual system. A novel quadrature-phase motion cancelation technique was developed to measure the magnitude of the grating induction <span class="hlt">brightness</span> illusion across a wide range of spatial frequencies, temporal frequencies and test field heights. Canceling contrast is greatest at low frequencies and declines with increasing frequency in both dimensions, and with increasing test field height. Canceling contrast scales as the product of inducing grating spatial frequency and test field height (the number of inducing grating cycles per test field height). When plotted using a spatial axis which indexes this product, the spatiotemporal induction <span class="hlt">surfaces</span> for four test field heights can be described as four partially overlapping sections of a single larger <span class="hlt">surface</span>. These properties of <span class="hlt">brightness</span> induction are explained in the context of multiscale spatial filtering. The present study is the first to measure the magnitude of grating induction as a function of temporal frequency. Taken in conjunction with several other studies (Blakeslee & McCourt, 2008; Robinson & de Sa, 2008; Magnussen & Glad, 1975) the results of this study illustrate that at least one form of <span class="hlt">brightness</span> induction is very much faster than that reported by DeValois et al. (1986) and Rossi and Paradiso (1996), and are inconsistent with the proposition that <span class="hlt">brightness</span> induction results from a slow “filling in” process. PMID:21763339</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016A%26A...596A..43C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016A%26A...596A..43C"><span>Non-magnetic photospheric <span class="hlt">bright</span> points in 3D simulations of the solar atmosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Calvo, F.; Steiner, O.; Freytag, B.</p> <p>2016-11-01</p> <p>Context. Small-scale <span class="hlt">bright</span> features in the photosphere of the Sun, such as faculae or G-band <span class="hlt">bright</span> points, appear in connection with small-scale magnetic flux concentrations. Aims: Here we report on a new class of photospheric <span class="hlt">bright</span> points that are free of magnetic fields. So far, these are visible in numerical simulations only. We explore conditions required for their <span class="hlt">observational</span> detection. Methods: Numerical radiation (magneto-)hydrodynamic simulations of the near-<span class="hlt">surface</span> layers of the Sun were carried out. The magnetic field-free simulations show tiny <span class="hlt">bright</span> points, reminiscent of magnetic <span class="hlt">bright</span> points, only smaller. A simple toy model for these non-magnetic <span class="hlt">bright</span> points (nMBPs) was established that serves as a base for the development of an algorithm for their automatic detection. Basic physical properties of 357 detected nMBPs were extracted and statistically evaluated. We produced synthetic intensity maps that mimic <span class="hlt">observations</span> with various solar telescopes to obtain hints on their detectability. Results: The nMBPs of the simulations show a mean bolometric intensity contrast with respect to their intergranular surroundings of approximately 20%, a size of 60-80 km, and the isosurface of optical depth unity is at their location depressed by 80-100 km. They are caused by swirling downdrafts that provide, by means of the centripetal force, the necessary pressure gradient for the formation of a funnel of reduced mass density that reaches from the subsurface layers into the photosphere. Similar, frequently occurring funnels that do not reach into the photosphere, do not produce <span class="hlt">bright</span> points. Conclusions: Non-magnetic <span class="hlt">bright</span> points are the <span class="hlt">observable</span> manifestation of vertically extending vortices (vortex tubes) in the photosphere. The resolving power of 4-m-class telescopes, such as the DKIST, is needed for an unambiguous detection of them. The movie associated to Fig. 1 is available at http://www.aanda.org</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA15454.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA15454.html"><span>Apparent <span class="hlt">Brightness</span> and Topography Images of Vibidia Crater</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2012-03-09</p> <p>The left-hand image from NASA Dawn spacecraft shows the apparent <span class="hlt">brightness</span> of asteroid Vesta <span class="hlt">surface</span>. The right-hand image is based on this apparent <span class="hlt">brightness</span> image, with a color-coded height representation of the topography overlain onto it.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016Natur.536...54D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016Natur.536...54D"><span><span class="hlt">Bright</span> carbonate deposits as evidence of aqueous alteration on (1) Ceres</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>de Sanctis, M. C.; Raponi, A.; Ammannito, E.; Ciarniello, M.; Toplis, M. J.; McSween, H. Y.; Castillo-Rogez, J. C.; Ehlmann, B. L.; Carrozzo, F. G.; Marchi, S.; Tosi, F.; Zambon, F.; Capaccioni, F.; Capria, M. T.; Fonte, S.; Formisano, M.; Frigeri, A.; Giardino, M.; Longobardo, A.; Magni, G.; Palomba, E.; McFadden, L. A.; Pieters, C. M.; Jaumann, R.; Schenk, P.; Mugnuolo, R.; Raymond, C. A.; Russell, C. T.</p> <p>2016-08-01</p> <p>The typically dark <span class="hlt">surface</span> of the dwarf planet Ceres is punctuated by areas of much higher albedo, most prominently in the Occator crater. These small <span class="hlt">bright</span> areas have been tentatively interpreted as containing a large amount of hydrated magnesium sulfate, in contrast to the average <span class="hlt">surface</span>, which is a mixture of low-albedo materials and magnesium phyllosilicates, ammoniated phyllosilicates and carbonates. Here we report high spatial and spectral resolution near-infrared <span class="hlt">observations</span> of the <span class="hlt">bright</span> areas in the Occator crater on Ceres. Spectra of these <span class="hlt">bright</span> areas are consistent with a large amount of sodium carbonate, constituting the most concentrated known extraterrestrial occurrence of carbonate on kilometre-wide scales in the Solar System. The carbonates are mixed with a dark component and small amounts of phyllosilicates, as well as ammonium carbonate or ammonium chloride. Some of these compounds have also been detected in the plume of Saturn’s sixth-largest moon Enceladus. The compounds are endogenous and we propose that they are the solid residue of crystallization of brines and entrained altered solids that reached the <span class="hlt">surface</span> from below. The heat source may have been transient (triggered by impact heating). Alternatively, internal temperatures may be above the eutectic temperature of subsurface brines, in which case fluids may exist at depth on Ceres today.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MAP...tmp...10J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MAP...tmp...10J"><span>A case study on large-scale dynamical influence on <span class="hlt">bright</span> band using cloud radar during the Indian summer monsoon</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jha, Ambuj K.; Kalapureddy, M. C. R.; Devisetty, Hari Krishna; Deshpande, Sachin M.; Pandithurai, G.</p> <p>2018-02-01</p> <p>The present study is a first of its kind attempt in exploring the physical features (e.g., height, width, intensity, duration) of tropical Indian <span class="hlt">bright</span> band using a Ka-band cloud radar under the influence of large-scale cyclonic circulation and attempts to explain the abrupt changes in <span class="hlt">bright</span> band features, viz., rise in the <span class="hlt">bright</span> band height by 430 m and deepening of the <span class="hlt">bright</span> band by about 300 m <span class="hlt">observed</span> at around 14:00 UTC on Sep 14, 2016, synoptically as well as locally. The study extends the utility of cloud radar to understand how the <span class="hlt">bright</span> band features are associated with light precipitation, ranging from 0 to 1.5 mm/h. Our analysis of the precipitation event of Sep 14-15, 2016 shows that the <span class="hlt">bright</span> band above (below) 3.7 km, thickness less (more) than 300 m can potentially lead to light drizzle of 0-0.25 mm/h (drizzle/light rain) at the <span class="hlt">surface</span>. It is also seen that the cloud radar may be suitable for <span class="hlt">bright</span> band study within light drizzle limits than under higher rain conditions. Further, the study illustrates that the <span class="hlt">bright</span> band features can be determined using the polarimetric capability of the cloud radar. It is shown that an LDR value of - 22 dB can be associated with the top height of <span class="hlt">bright</span> band in the Ka-band <span class="hlt">observations</span> which is useful in the extraction of the <span class="hlt">bright</span> band top height and its width. This study is useful for understanding the <span class="hlt">bright</span> band phenomenon and could be potentially useful in establishing the <span class="hlt">bright</span> band-<span class="hlt">surface</span> rain relationship through the perspective of a cloud radar, which would be helpful to enhance the cloud radar-based quantitative estimates of precipitation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930030997&hterms=carbon+emissions&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dcarbon%2Bemissions','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930030997&hterms=carbon+emissions&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dcarbon%2Bemissions"><span>A <span class="hlt">surface</span> <span class="hlt">brightness</span> correlation between carbon monoxide and nonthermal radio continuum emission in the galaxy</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Allen, R. J.</p> <p>1992-01-01</p> <p>The relation between the projected face-on velocity-integrated CO (1-0) <span class="hlt">brightness</span> ICO and the 20 cm nonthermal radio continuum <span class="hlt">brightness</span> T20 is examined as a function of radius in the Galactic disk. Averaged in 1 kpc annuli, the ratio ICO/T20 is nearly constant with a mean value of 1.51 +/- 0.34 km/s from 2 to 10 kpc. The manner in which ICO and T20 are derived for the Galaxy is different in several significant respects from the more direct <span class="hlt">observational</span> determinations possible in nearby galaxies. The fact that the Galaxy also follows this correlation further strengthens the generality of the result.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120002989','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120002989"><span>Visible Color and Photometry of <span class="hlt">Bright</span> Materials on Vesta</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schroder, S. E.; Li, J. Y.; Mittlefehldt, D. W.; Pieters, C. M.; De Sanctis, M. C.; Hiesinger, H.; Blewett, D. T.; Russell, C. T.; Raymond, C. A.; Keller, H. U.</p> <p>2012-01-01</p> <p>The Dawn Framing Camera (FC) collected images of the <span class="hlt">surface</span> of Vesta at a pixel scale of 70 m in the High Altitude Mapping Orbit (HAMO) phase through its clear and seven color filters spanning from 430 nm to 980 nm. The <span class="hlt">surface</span> of Vesta displays a large diversity in its <span class="hlt">brightness</span> and colors, evidently related to the diverse geology [1] and mineralogy [2]. Here we report a detailed investigation of the visible colors and photometric properties of the apparently <span class="hlt">bright</span> materials on Vesta in order to study their origin. The global distribution and the spectroscopy of <span class="hlt">bright</span> materials are discussed in companion papers [3, 4], and the synthesis results about the origin of Vestan <span class="hlt">bright</span> materials are reported in [5].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1356665-veritas-observations-six-bright-hard-spectrum-fermi-lat-blazars','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1356665-veritas-observations-six-bright-hard-spectrum-fermi-lat-blazars"><span>VERITAS <span class="hlt">Observations</span> of Six <span class="hlt">Bright</span>, Hard-Spectrum Fermi-LAT Blazars</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Aliu, E.; Archambault, S.; Arlen, T.; ...</p> <p>2012-10-25</p> <p>In this paper, we report on VERITAS very high energy (VHE; E ≥ 100 GeV) <span class="hlt">observations</span> of six blazars selected from the Fermi Large Area Telescope First Source Catalog (1FGL). The gamma-ray emission from 1FGL sources was extrapolated up to the VHE band, taking gamma-ray absorption by the extragalactic background light into account. This allowed the selection of six <span class="hlt">bright</span>, hard-spectrum blazars that were good candidate TeV emitters. Spectroscopic redshift measurements were attempted with the Keck Telescope for the targets without Sloan Digital Sky Survey spectroscopic data. No VHE emission is detected during the <span class="hlt">observations</span> of the six sources describedmore » here. Corresponding TeV upper limits are presented, along with contemporaneous Fermi <span class="hlt">observations</span> and non-concurrent Swift UVOT and X-Ray Telescope data. The blazar broadband spectral energy distributions (SEDs) are assembled and modeled with a single-zone synchrotron self-Compton model. Finally, the SED built for each of the six blazars shows a synchrotron peak bordering between the intermediate- and high-spectrum-peak classifications, with four of the six resulting in particle-dominated emission regions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01465&hterms=Sunlight+cities&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DSunlight%2Bcities','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01465&hterms=Sunlight+cities&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DSunlight%2Bcities"><span>Hubble <span class="hlt">Observes</span> <span class="hlt">Surface</span> of Titan</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>Scientists for the first time have made images of the <span class="hlt">surface</span> of Saturn's giant, haze-shrouded moon, Titan. They mapped light and dark features over the <span class="hlt">surface</span> of the satellite during nearly a complete 16-day rotation. One prominent <span class="hlt">bright</span> area they discovered is a <span class="hlt">surface</span> feature 2,500 miles across, about the size of the continent of Australia.<p/>Titan, larger than Mercury and slightly smaller than Mars, is the only body in the solar system, other than Earth, that may have oceans and rainfall on its <span class="hlt">surface</span>, albeit oceans and rain of ethane-methane rather than water. Scientists suspect that Titan's present environment -- although colder than minus 289 degrees Fahrenheit, so cold that water ice would be as hard as granite -- might be similar to that on Earth billions of years ago, before life began pumping oxygen into the atmosphere.<p/>Peter H. Smith of the University of Arizona Lunar and Planetary Laboratory and his team took the images with the Hubble Space Telescope during 14 <span class="hlt">observing</span> runs between Oct. 4 - 18. Smith announced the team's first results last week at the 26th annual meeting of the American Astronomical Society Division for Planetary Sciences in Bethesda, Md. Co-investigators on the team are Mark Lemmon, a doctoral candidate with the UA Lunar and Planetary Laboratory; John Caldwell of York University, Canada; Larry Sromovsky of the University of Wisconsin; and Michael Allison of the Goddard Institute for Space Studies, New York City.<p/>Titan's atmosphere, about four times as dense as Earth's atmosphere, is primarily nitrogen laced with such poisonous substances as methane and ethane. This thick, orange, hydrocarbon haze was impenetrable to cameras aboard the Pioneer and Voyager spacecraft that flew by the Saturn system in the late 1970s and early 1980s. The haze is formed as methane in the atmosphere is destroyed by sunlight. The hydrocarbons produced by this methane destruction form a smog similar to that found over large cities, but is much</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ASPC..504..273Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ASPC..504..273Y"><span>Modelling Solar and Stellar <span class="hlt">Brightness</span> Variabilities</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yeo, K. L.; Shapiro, A. I.; Krivova, N. A.; Solanki, S. K.</p> <p>2016-04-01</p> <p>Total and spectral solar irradiance, TSI and SSI, have been measured from space since 1978. This is accompanied by the development of models aimed at replicating the <span class="hlt">observed</span> variability by relating it to solar <span class="hlt">surface</span> magnetism. Despite significant progress, there remains persisting controversy over the secular change and the wavelength-dependence of the variation with impact on our understanding of the Sun's influence on the Earth's climate. We highlight the recent progress in TSI and SSI modelling with SATIRE. <span class="hlt">Brightness</span> variations have also been <span class="hlt">observed</span> for Sun-like stars. Their analysis can profit from knowledge of the solar case and provide additional constraints for solar modelling. We discuss the recent effort to extend SATIRE to Sun-like stars.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19740051089&hterms=sparrow&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsparrow','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19740051089&hterms=sparrow&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsparrow"><span>Is the zodiacal light intensity steady. [cloud <span class="hlt">surface</span> <span class="hlt">brightness</span> and polarization from OSO-5 data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Burnett, G. B.; Sparrow, J. G.; Ney, E. P.</p> <p>1974-01-01</p> <p>It is pointed out that conclusions reported by Sparrow and Ney (1972, 1973) could be confirmed in an investigation involving the refinement of OSO-5 data on zodiacal light. It had been found by Sparrow and Ney that the absolute value of both the <span class="hlt">surface</span> <span class="hlt">brightness</span> and polarization of the zodiacal cloud varied by less than 10% over the 4-yr period from January 1969 to January 1973.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22521637-active-region-bright-grains-observed-transition-region-imaging-channels-iris','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22521637-active-region-bright-grains-observed-transition-region-imaging-channels-iris"><span>ON THE ACTIVE REGION <span class="hlt">BRIGHT</span> GRAINS <span class="hlt">OBSERVED</span> IN THE TRANSITION REGION IMAGING CHANNELS OF IRIS</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Skogsrud, H.; Voort, L. Rouppe van der; Pontieu, B. De</p> <p></p> <p>The Interface Region Imaging Spectrograph (IRIS) provides spectroscopy and narrow band slit-jaw (SJI) imaging of the solar chromosphere and transition region at unprecedented spatial and temporal resolutions. Combined with high-resolution context spectral imaging of the photosphere and chromosphere as provided by the Swedish 1 m Solar Telescope (SST), we can now effectively trace dynamic phenomena through large parts of the solar atmosphere in both space and time. IRIS SJI 1400 images from active regions, which primarily sample the transition region with the Si iv 1394 and 1403 Å lines, reveal ubiquitous <span class="hlt">bright</span> “grains” which are short-lived (two to five minute)more » <span class="hlt">bright</span> roundish small patches of sizes 0.″5–1.″7 that generally move limbward with velocities up to about 30 km s{sup −1}. In this paper, we show that many <span class="hlt">bright</span> grains are the result of chromospheric shocks impacting the transition region. These shocks are associated with dynamic fibrils (DFs), most commonly <span class="hlt">observed</span> in Hα. We find that the grains show the strongest emission in the ascending phase of the DF, that the emission is strongest toward the top of the DF, and that the grains correspond to a blueshift and broadening of the Si iv lines. We note that the SJI 1400 grains can also be <span class="hlt">observed</span> in the SJI 1330 channel which is dominated by C ii lines. Our <span class="hlt">observations</span> show that a significant part of the active region transition region dynamics is driven from the chromosphere below rather than from coronal activity above. We conclude that the shocks that drive DFs also play an important role in the heating of the upper chromosphere and lower transition region.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110015331','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110015331"><span>L Band <span class="hlt">Brightness</span> Temperature <span class="hlt">Observations</span> Over a Corn Canopy During the Entire Growth Cycle</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Joseph, Alicia T.; O'Neill, Peggy E.; Choudhury, Bhaskar J.; vanderVelde, Rogier; Lang, Roger H.; Gish, Timothy</p> <p>2011-01-01</p> <p>During a field campaign covering the 2002 corn growing season, a dual polarized tower mounted L-band (1.4 GHz) radiometer (LRAD) provided <span class="hlt">brightness</span> temperature (T(sub B)) measurements at preset intervals, incidence and azimuth angles. These radiometer measurements were supported by an extensive characterization of land <span class="hlt">surface</span> variables including soil moisture, soil temperature, vegetation biomass, and <span class="hlt">surface</span> roughness. During the period from May 22, 2002 to August 30, 2002 a range of vegetation water content (W) of 0.0 to 4.3 kg/square m, ten days of radiometer and ground measurements were available. Using this data set, the effects of corn vegetation on <span class="hlt">surface</span> emissions are investigated by means of a semi-empirical radiative transfer model. Additionally, the impact of roughness on the <span class="hlt">surface</span> emission is quantified using T(sub B) measurements over bare soil conditions. Subsequently, the estimated roughness parameters, ground measurements and horizontally (H)-polarized T(sub B) are employed to invert the H-polarized transmissivity (gamma-h) for the monitored corn growing season.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900042461&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900042461&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour"><span>Coronal <span class="hlt">bright</span> points at 6cm wavelength</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fu, Qijun; Kundu, M. R.; Schmahl, E. J.</p> <p>1988-01-01</p> <p>Results are presented from <span class="hlt">observations</span> of <span class="hlt">bright</span> points at a wavelength of 6-cm using the VLA with a spatial resolution of 1.2 arcsec. During two hours of <span class="hlt">observations</span>, 44 sources were detected with <span class="hlt">brightness</span> temperatures between 2000 and 30,000 K. Of these sources, 27 are associated with weak dark He 10830 A features at distances less than 40 arcsecs. Consideration is given to variations in the source parameters and the relationship between ephemeral regions and <span class="hlt">bright</span> points.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950044608&hterms=Andromeda&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DAndromeda','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950044608&hterms=Andromeda&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DAndromeda"><span>HST <span class="hlt">observations</span> of globular clusters in M 31. 1: <span class="hlt">Surface</span> photometry of 13 objects</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Pecci, F. Fusi; Battistini, P.; Bendinelli, O.; Bonoli, F.; Cacciari, C.; Djorgovski, S.; Federici, L.; Ferraro, F. R.; Parmeggiani, G.; Weir, N.</p> <p>1994-01-01</p> <p>We present the initial results of a study of globular clusters in M 31, using the Faint Object Camera (FOC) on the Hubble Space Telescope (HST). The sample of objects consists of 13 clusters spanning a range of properties. Three independent image deconvolution techniques were used in order to compensate for the optical problems of the HST, leading to mutually fully consistent results. We present detailed tests and comparisons to determine the reliability and limits of these deconvolution methods, and conclude that high-quality <span class="hlt">surface</span> photometry of M 31 globulars is possible with the HST data. <span class="hlt">Surface</span> <span class="hlt">brightness</span> profiles have been extracted, and core radii, half-light radii, and central <span class="hlt">surface</span> <span class="hlt">brightness</span> values have been measured for all of the clusters in the sample. Their comparison with the values from ground-based <span class="hlt">observations</span> indicates the later to be systematically and strongly biased by the seeing effects, as it may be expected. A comparison of the structural parameters with those of the Galactic globulars shows that the structural properties of the M 31 globulars are very similar to those of their Galactic counterparts. A candidate for a post-core-collapse cluster, Bo 343 = G 105, has been already identified from these data; this is the first such detection in the M 31 globular cluster system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870033158&hterms=Dwarf+stars+blue&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DWhy%2BDwarf%2Bstars%2Bblue','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870033158&hterms=Dwarf+stars+blue&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DWhy%2BDwarf%2Bstars%2Bblue"><span><span class="hlt">Surface</span> <span class="hlt">brightness</span> and color distributions in blue compact dwarf galaxies. I - Haro 2, an extreme example of a star-forming young elliptical galaxy</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Loose, Hans-Hermann; Thuan, Trinh X.</p> <p>1986-01-01</p> <p>The first results of a large-scale program to study the morphology and structure of blue compact dwarf galaxies from CCD <span class="hlt">observations</span> are presented. The <span class="hlt">observations</span> and reduction procedures are described, and <span class="hlt">surface</span> <span class="hlt">brightness</span> and color profiles are shown. The results are used to discuss the morphological type of Haro 2 and its stellar populations. It is found that Haro 2 appears to be an extreme example of an elliptical galaxy undergoing intense star formation in its central regions, and that the oldest stars it contains were made only about four million yr ago. The 'missing' mass problem of Haro 2 is also discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19780031984&hterms=Saunders&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D40%26Ntt%3DSaunders%252C%2BM','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19780031984&hterms=Saunders&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D40%26Ntt%3DSaunders%252C%2BM"><span>Geologic interpretation of new <span class="hlt">observations</span> of the <span class="hlt">surface</span> of Venus</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Saunders, R. S.; Malin, M. C.</p> <p>1977-01-01</p> <p>New radar <span class="hlt">observations</span> of the <span class="hlt">surface</span> of Venus provide further evidence of a diverse and complex geologic evolution. The radar <span class="hlt">bright</span> feature 'Beta' (24 deg N, 85 deg W) is seen to be a 700 km diameter region elevated a maximum of approximately 10 km relative to its surroundings with a 60 x 90 km wide depression at its summit. 'Beta' is interpreted to be a large volcanic construct, analogous to terrestrial and Martian shield volcanoes. Two large, quasi-circular areas of low reflectivity, examples of a class of features interpreted to be impact basins by previous investigators who were without the benefit of actual topographic information, are shown in altimetry maps to be depressions. Thus the term 'basin' can be applied, although we urge a non-genetic usage until more complete understanding of their origin is achieved through analysis of future <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150002552','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150002552"><span>Hi-C <span class="hlt">Observations</span> of Penumbral <span class="hlt">Bright</span> Dots: Comparison with the IRIS Results</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Alpert, S. E.; Tiwari, S. K.; Moore, R. L.; Savage, S. L.; Winebarger, A. R.</p> <p>2014-01-01</p> <p>We <span class="hlt">observed</span> <span class="hlt">bright</span> dots (BDs) in a sunspot penumbra by using data acquired by the High Resolution Coronal Imager (Hi-C). The sizes of these BDs are on the order of 1 arcsecond (1') and are therefore hard to identify using the Atmospheric Imaging Assembly's (AIA) 0.6' pixel -1 resolution. These BDs become readily apparent with Hi-C's 0.1' pixel -1 resolution. Tian et al. (2014) found penumbral BDs in the transition region (TR) by using the Interface Region Imaging Spectrograph (IRIS). However, only a few of their dots could be associated with any enhanced <span class="hlt">brightness</span> in AIA channels. In this work, we examine the characteristics of the penumbral BDs <span class="hlt">observed</span> by Hi-C in a sunspot penumbra, including their sizes, lifetimes, speeds, and intensity. We also attempt to relate these BDs to the IRIS BDs. There are fewer Hi-C BDs in the penumbra than seen by IRIS, though different sunspots were studied and Hi-C had a short <span class="hlt">observation</span> time. We use 193 A Hi-C data from July 11, 2012 which <span class="hlt">observed</span> from 18:52:00 UT{18:56:00 UT and supplement it with data from AIA's 193 A passband to see the complete lifetime of the dots that were born before and/or lasted longer than Hi-C's 5-minute <span class="hlt">observation</span> period. We use additional AIA passbands and compare the light curves of the BDs at different temperatures to test whether the Hi-C BDs are TR BDs. We find that most Hi-C BDs show clear movement, and of those that do, they move in a radial direction, toward or away from the sunspot umbra, sometimes doing both. BDs interact with other BDs, combining to fade away or brighten. The BDs that do not interact with other BDs tend to move less and last longer. We examine the properties of the Hi-C BDs and compare them with the IRIS BDs. Our BDs are similar to the exceptional values of the IRIS BDs: they move slower on average and their sizes and lifetimes are on the higher end of the distributions of IRIS BDs. We infer that our penumbral BDs are some of the larger BDs <span class="hlt">observed</span> by IRIS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21301402-observation-tev-gamma-rays-from-fermi-bright-galactic-sources-tibet-air-shower-array','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21301402-observation-tev-gamma-rays-from-fermi-bright-galactic-sources-tibet-air-shower-array"><span><span class="hlt">OBSERVATION</span> OF TeV GAMMA RAYS FROM THE FERMI <span class="hlt">BRIGHT</span> GALACTIC SOURCES WITH THE TIBET AIR SHOWER ARRAY</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Amenomori, M.; Bi, X. J.; Ding, L. K.</p> <p>2010-01-20</p> <p>Using the Tibet-III air shower array, we search for TeV {gamma}-rays from 27 potential Galactic sources in the early list of <span class="hlt">bright</span> sources obtained by the Fermi Large Area Telescope at energies above 100 MeV. Among them, we <span class="hlt">observe</span> seven sources instead of the expected 0.61 sources at a significance of 2{sigma} or more excess. The chance probability from Poisson statistics would be estimated to be 3.8 x 10{sup -6}. If the excess distribution <span class="hlt">observed</span> by the Tibet-III array has a density gradient toward the Galactic plane, the expected number of sources may be enhanced in chance association. Then, themore » chance probability rises slightly, to 1.2 x 10{sup -5}, based on a simple Monte Carlo simulation. These low chance probabilities clearly show that the Fermi <span class="hlt">bright</span> Galactic sources have statistically significant correlations with TeV {gamma}-ray excesses. We also find that all seven sources are associated with pulsars, and six of them are coincident with sources detected by the Milagro experiment at a significance of 3{sigma} or more at the representative energy of 35 TeV. The significance maps <span class="hlt">observed</span> by the Tibet-III air shower array around the Fermi sources, which are coincident with the Milagro {>=}3{sigma} sources, are consistent with the Milagro <span class="hlt">observations</span>. This is the first result of the northern sky survey of the Fermi <span class="hlt">bright</span> Galactic sources in the TeV region.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870011490','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870011490"><span>Simultaneous <span class="hlt">observations</span> of changes in coronal <span class="hlt">bright</span> point emission at the 20 cm radio and He Lambda 10830 wavelengths</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Habbal, Shadia R.; Harvey, Karen L.</p> <p>1986-01-01</p> <p>Preliminary results of <span class="hlt">observations</span> of solar coronal <span class="hlt">bright</span> points acquired simultaneously from ground based observatories at the radio wavelength of 20 cm and in the He I wavelength 10830 line on September 8, 1985, are reported. The impetus for obtaining simultaneous radio and optical data is to identify correlations, if any, in changes of the low transition-coronal signatures of <span class="hlt">bright</span> points with the evolution of the magnetic field, and to distinguish between intermittent heating and changes in the magnetic field topology. Although simultaneous <span class="hlt">observations</span> of H alpha emission and the photospheric magnetic field at Big Bear were also made, as well as radio <span class="hlt">observations</span> from Owen Valley Radio Interferometer and Solar Maximum Mission (SSM) (O VIII line), only the comparison between He 10830 and the Very Large Array (VLA) radio data are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...856..126C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...856..126C"><span>The Next Generation Virgo Cluster Survey (NGVS). XVIII. Measurement and Calibration of <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Fluctuation Distances for <span class="hlt">Bright</span> Galaxies in Virgo (and Beyond)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cantiello, Michele; Blakeslee, John P.; Ferrarese, Laura; Côté, Patrick; Roediger, Joel C.; Raimondo, Gabriella; Peng, Eric W.; Gwyn, Stephen; Durrell, Patrick R.; Cuillandre, Jean-Charles</p> <p>2018-04-01</p> <p>We describe a program to measure <span class="hlt">surface</span> <span class="hlt">brightness</span> fluctuation (SBF) distances to galaxies <span class="hlt">observed</span> in the Next Generation Virgo Cluster Survey (NGVS), a photometric imaging survey covering 104 deg2 of the Virgo cluster in the u*, g, i, and z bandpasses with the Canada–France–Hawaii Telescope. We describe the selection of the sample galaxies, the procedures for measuring the apparent i-band SBF magnitude {\\overline{m}}i, and the calibration of the absolute Mibar as a function of <span class="hlt">observed</span> stellar population properties. The multiband NGVS data set provides multiple options for calibrating the SBF distances, and we explore various calibrations involving individual color indices as well as combinations of two different colors. Within the color range of the present sample, the two-color calibrations do not significantly improve the scatter with respect to wide-baseline, single-color calibrations involving u*. We adopt the ({u}* -z) calibration as a reference for the present galaxy sample, with an <span class="hlt">observed</span> scatter of 0.11 mag. For a few cases that lack good u* photometry, we use an alternative relation based on a combination of (g-i) and (g-z) colors, with only a slightly larger <span class="hlt">observed</span> scatter of 0.12 mag. The agreement of our measurements with the best existing distance estimates provides confidence that our measurements are accurate. We present a preliminary catalog of distances for 89 galaxies brighter than B T ≈ 13.0 mag within the survey footprint, including members of the background M and W Clouds at roughly twice the distance of the main body of the Virgo cluster. The extension of the present work to fainter and bluer galaxies is in progress.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22086254-brightness-fluctuation-mid-infrared-sky-from-akari-observations-toward-north-ecliptic-pole','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22086254-brightness-fluctuation-mid-infrared-sky-from-akari-observations-toward-north-ecliptic-pole"><span><span class="hlt">BRIGHTNESS</span> AND FLUCTUATION OF THE MID-INFRARED SKY FROM AKARI <span class="hlt">OBSERVATIONS</span> TOWARD THE NORTH ECLIPTIC POLE</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Pyo, Jeonghyun; Jeong, Woong-Seob; Matsumoto, Toshio</p> <p>2012-12-01</p> <p>We present the smoothness of the mid-infrared sky from <span class="hlt">observations</span> by the Japanese infrared astronomical satellite AKARI. AKARI monitored the north ecliptic pole (NEP) during its cold phase with nine wave bands covering from 2.4 to 24 {mu}m, out of which six mid-infrared bands were used in this study. We applied power-spectrum analysis to the images in order to search for the fluctuation of the sky <span class="hlt">brightness</span>. <span class="hlt">Observed</span> fluctuation is explained by fluctuation of photon noise, shot noise of faint sources, and Galactic cirrus. The fluctuations at a few arcminutes scales at short mid-infrared wavelengths (7, 9, and 11 {mu}m)more » are largely caused by the diffuse Galactic light of the interstellar dust cirrus. At long mid-infrared wavelengths (15, 18, and 24 {mu}m), photon noise is the dominant source of fluctuation over the scale from arcseconds to a few arcminutes. The residual fluctuation amplitude at 200'' after removing these contributions is at most 1.04 {+-} 0.23 nW m{sup -2} sr{sup -1} or 0.05% of the <span class="hlt">brightness</span> at 24 {mu}m and at least 0.47 {+-} 0.14 nW m{sup -2} sr{sup -1} or 0.02% at 18 {mu}m. We conclude that the upper limit of the fluctuation in the zodiacal light toward the NEP is 0.03% of the sky <span class="hlt">brightness</span>, taking 2{sigma} error into account.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21301418-brightness-waiting-time-distributions-type-iii-radio-storm-observed-stereo-waves','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21301418-brightness-waiting-time-distributions-type-iii-radio-storm-observed-stereo-waves"><span>ON THE <span class="hlt">BRIGHTNESS</span> AND WAITING-TIME DISTRIBUTIONS OF A TYPE III RADIO STORM <span class="hlt">OBSERVED</span> BY STEREO/WAVES</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Eastwood, J. P.; Hudson, H. S.; Krucker, S.</p> <p>2010-01-10</p> <p>Type III solar radio storms, <span class="hlt">observed</span> at frequencies below {approx}16 MHz by space-borne radio experiments, correspond to the quasi-continuous, bursty emission of electron beams onto open field lines above active regions. The mechanisms by which a storm can persist in some cases for more than a solar rotation whilst exhibiting considerable radio activity are poorly understood. To address this issue, the statistical properties of a type III storm <span class="hlt">observed</span> by the STEREO/WAVES radio experiment are presented, examining both the <span class="hlt">brightness</span> distribution and (for the first time) the waiting-time distribution (WTD). Single power-law behavior is <span class="hlt">observed</span> in the number distribution asmore » a function of <span class="hlt">brightness</span>; the power-law index is {approx}2.1 and is largely independent of frequency. The WTD is found to be consistent with a piecewise-constant Poisson process. This indicates that during the storm individual type III bursts occur independently and suggests that the storm dynamics are consistent with avalanche-type behavior in the underlying active region.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20150007372&hterms=bias+correction&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dbias%2Bcorrection','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20150007372&hterms=bias+correction&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dbias%2Bcorrection"><span>Characterization and Correction of Aquarius Long Term Calibration Drift Using On-Earth <span class="hlt">Brightness</span> Temperature Refernces</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Brown, Shannon; Misra, Sidharth</p> <p>2013-01-01</p> <p>The Aquarius/SAC-D mission was launched on June 10, 2011 from Vandenberg Air Force Base. Aquarius consists of an L-band radiometer and scatterometer intended to provide global maps of sea <span class="hlt">surface</span> salinity. One of the main mission objectives is to provide monthly global salinity maps for climate studies of ocean circulation, <span class="hlt">surface</span> evaporation and precipitation, air/sea interactions and other processes. Therefore, it is critical that any spatial or temporal systematic biases be characterized and corrected. One of the main mission requirements is to measure salinity with an accuracy of 0.2 psu on montly time scales which requires a <span class="hlt">brightness</span> temperature stability of about 0.1K, which is a challenging requirement for the radiometer. A secondary use of the Aquarius data is for soil moisture applications, which requires <span class="hlt">brightness</span> temperature stability at the warmer end of the <span class="hlt">brightness</span> temperature dynamic range. Soon after launch, time variable drifts were <span class="hlt">observed</span> in the Aquarius data compared to in-situ data from ARGO and models for the ocean <span class="hlt">surface</span> salinity. These drifts could arise from a number of sources, including the various components of the retrieval algorithm, such as the correction for direct and reflected galactic emission, or from the instrument <span class="hlt">brightness</span> temperature calibration. If arising from the <span class="hlt">brightness</span> temperature calibration, they could have gain and offset components. It is critical that the nature of the drifts be understood before a suitable correction can be implemented. This paper describes the approach that was used to detect and characterize the components of the drift that were in the <span class="hlt">brightness</span> temperature calibration using on-Earth reference targets that were independent of the ocean model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22666232-beyond-mag-arcsec-sup-frontier-low-surface-brightness-imaging-largest-optical-telescopes','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22666232-beyond-mag-arcsec-sup-frontier-low-surface-brightness-imaging-largest-optical-telescopes"><span>BEYOND 31 mag arcsec{sup −2}: THE FRONTIER OF LOW <span class="hlt">SURFACE</span> <span class="hlt">BRIGHTNESS</span> IMAGING WITH THE LARGEST OPTICAL TELESCOPES</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Trujillo, Ignacio; Fliri, Jüergen, E-mail: trujillo@iac.es; Departamento de Astrofísica, Universidad de La Laguna, E-38206, La Laguna, Tenerife</p> <p>2016-06-01</p> <p>The detection of structures in the sky with optical <span class="hlt">surface</span> <span class="hlt">brightnesses</span> fainter than 30 mag arcsec{sup −2} (3 σ in 10 × 10 arcsec boxes; r -band) has remained elusive in current photometric deep surveys. Here we show how present-day telescopes of 10 m class can provide broadband imaging 1.5–2 mag deeper than most previous results within a reasonable amount of time (i.e., <10 hr on-source integration). In particular, we illustrate the ability of the 10.4 m Gran Telescopio de Canarias telescope to produce imaging with a limiting <span class="hlt">surface</span> <span class="hlt">brightness</span> of 31.5 mag arcsec{sup −2} (3 σ in 10 ×more » 10 arcsec boxes; r -band) using 8.1 hr on source. We apply this power to explore the stellar halo of the galaxy UGC 00180, a galaxy analogous to M31 located at ∼150 Mpc, by obtaining a radial profile of <span class="hlt">surface</span> <span class="hlt">brightness</span> down to μ{sub r} ∼ 33 mag arcsec{sup −2}. This depth is similar to that obtained using the star-counts techniques for Local Group galaxies, but is achieved at a distance where this technique is unfeasible. We find that the mass of the stellar halo of this galaxy is ∼4 × 10{sup 9} M {sub ⊙}, i.e., (3 ± 1)% of the total stellar mass of the whole system. This amount of mass in the stellar halo is in agreement with current theoretical expectations for galaxies of this kind.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150002873','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150002873"><span>Extremely Low Passive Microwave <span class="hlt">Brightness</span> Temperatures Due to Thunderstorms</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cecil, Daniel J.</p> <p>2015-01-01</p> <p>Extreme events by their nature fall outside the bounds of routine experience. With imperfect or ambiguous measuring systems, it is appropriate to question whether an unusual measurement represents an extreme event or is the result of instrument errors or other sources of noise. About three weeks after the Tropical Rainfall Measuring Mission (TRMM) satellite began collecting data in Dec 1997, a thunderstorm was <span class="hlt">observed</span> over northern Argentina with 85 GHz <span class="hlt">brightness</span> temperatures below 50 K and 37 GHz <span class="hlt">brightness</span> temperatures below 70 K (Zipser et al. 2006). These values are well below what had previously been <span class="hlt">observed</span> from satellite sensors with lower resolution. The 37 GHz <span class="hlt">brightness</span> temperatures are also well below those measured by TRMM for any other storm in the subsequent 16 years. Without corroborating evidence, it would be natural to suspect a problem with the instrument, or perhaps an irregularity with the platform during the first weeks of the satellite mission. Automated quality control flags or other procedures in retrieval algorithms could treat these measurements as errors, because they fall outside the expected bounds. But the TRMM satellite also carries a radar and a lightning sensor, both confirming the presence of an intense thunderstorm. The radar recorded 40+ dBZ reflectivity up to about 19 km altitude. More than 200 lightning flashes per minute were recorded. That same storm's 19 GHz <span class="hlt">brightness</span> temperatures below 150 K would normally be interpreted as the result of a low-emissivity water <span class="hlt">surface</span> (e.g., a lake, or flood waters) if not for the simultaneous measurements of such intense convection. This paper will examine records from TRMM and related satellite sensors including SSMI, AMSR-E, and the new GMI to find the strongest signatures resulting from thunderstorms, and distinguishing those from sources of noise. The lowest <span class="hlt">brightness</span> temperatures resulting from thunderstorms as seen by TRMM have been in Argentina in November and December. For</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_9 --> <div id="page_10" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="181"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012cosp...39.1297M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012cosp...39.1297M"><span>Assimilation of SMOS <span class="hlt">brightness</span> temperatures in the ECMWF EKF for the analysis of soil moisture</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Munoz-Sabater, Joaquin</p> <p>2012-07-01</p> <p>Since November 2nd 2009, the European Centre for Medium-Range Weather Forecasts (ECMWF) has being monitoring, in Near Real Time (NRT), L-band <span class="hlt">brightness</span> temperatures measured by the Soil Moisture and Ocean Salinity (SMOS) satellite mission of the European Space Agency (ESA). The main objective of the monitoring suite for SMOS data is to systematically monitor the difference between SMOS <span class="hlt">observed</span> <span class="hlt">brightness</span> temperatures and the corresponding model equivalent simulated by the Community Microwave Emission Model (CMEM), the so-called first guess departures. This is a crucial step, as first guess departures is the quantity used in the analysis. The ultimate goal is to investigate how the assimilation of SMOS <span class="hlt">brightness</span> temperatures over land improves the weather forecast skill, through a more accurate initialization of the global soil moisture state. In this presentation, some significant results from the activities preparing for the assimilation of SMOS data are shown. Among these activities, an effective data thinning strategy, a practical approach to reduce noise from the <span class="hlt">observed</span> <span class="hlt">brightness</span> temperatures and a bias correction scheme are of special interest. Firstly, SMOS data needs to be significantly thinned as the data volume delivered for a single orbit is too large for the current operational capabilities in any Numerical Weather Prediction system. Different thinning strategies have been analysed and tested. The most suitable one is the assimilation of SMOS <span class="hlt">brightness</span> temperatures which match the ECMWF T511 (~40 km) reduced Gaussian Grid. Secondly, SMOS <span class="hlt">observational</span> noise is reduced significantly by averaging the data in angular bins. In addition, this methodology contributes to further thinning of the SMOS data before the analysis. Finally, a bias correction scheme based on a CDF matching is applied to the <span class="hlt">observations</span> to ensure an unbiased dataset ready for assimilation in the ECMWF <span class="hlt">surface</span> analysis system. The current ECMWF operational soil moisture analysis</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22357255-sparkling-extreme-ultraviolet-bright-dots-observed-hi','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22357255-sparkling-extreme-ultraviolet-bright-dots-observed-hi"><span>Sparkling extreme-ultraviolet <span class="hlt">bright</span> dots <span class="hlt">observed</span> with Hi-C</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Régnier, S.; Alexander, C. E.; Walsh, R. W.</p> <p></p> <p><span class="hlt">Observing</span> the Sun at high time and spatial scales is a step toward understanding the finest and fundamental scales of heating events in the solar corona. The high-resolution coronal (Hi-C) instrument has provided the highest spatial and temporal resolution images of the solar corona in the EUV wavelength range to date. Hi-C <span class="hlt">observed</span> an active region on 2012 July 11 that exhibits several interesting features in the EUV line at 193 Å. One of them is the existence of short, small brightenings 'sparkling' at the edge of the active region; we call these EUV <span class="hlt">bright</span> dots (EBDs). Individual EBDs havemore » a characteristic duration of 25 s with a characteristic length of 680 km. These brightenings are not fully resolved by the SDO/AIA instrument at the same wavelength; however, they can be identified with respect to the Hi-C location of the EBDs. In addition, EBDs are seen in other chromospheric/coronal channels of SDO/AIA, which suggests a temperature between 0.5 and 1.5 MK. Based on their frequency in the Hi-C time series, we define four different categories of EBDs: single peak, double peak, long duration, and bursty. Based on a potential field extrapolation from an SDO/HMI magnetogram, the EBDs appear at the footpoints of large-scale, trans-equatorial coronal loops. The Hi-C <span class="hlt">observations</span> provide the first evidence of small-scale EUV heating events at the base of these coronal loops, which have a free magnetic energy of the order of 10{sup 26} erg.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016A%26A...593A..32G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016A%26A...593A..32G"><span>Height formation of <span class="hlt">bright</span> points <span class="hlt">observed</span> by IRIS in Mg II line wings during flux emergence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Grubecka, M.; Schmieder, B.; Berlicki, A.; Heinzel, P.; Dalmasse, K.; Mein, P.</p> <p>2016-09-01</p> <p>Context. A flux emergence in the active region AR 111850 was <span class="hlt">observed</span> on September 24, 2013 with the Interface Region Imaging Spectrograph (IRIS). Many <span class="hlt">bright</span> points are associated with the new emerging flux and show enhancement brightening in the UV spectra. Aims: The aim of this work is to compute the altitude formation of the compact <span class="hlt">bright</span> points (CBs) <span class="hlt">observed</span> in Mg II lines in the context of searching Ellerman bombs (EBs). Methods: IRIS provided two large dense rasters of spectra in Mg II h and k lines, Mg II triplet, C II and Si IV lines covering all the active region and slit jaws in the two bandpasses (1400 Å and 2796 Å) starting at 11:44 UT and 15:39 UT, and lasting 20 min each. Synthetic profiles of Mg II and Hα lines are computed with non-local thermodynamic equlibrium (NLTE) radiative transfer treatment in 1D solar atmosphere model including a hotspot region defined by three parameters: temperature, altitude, and width. Results: Within the two IRIS rasters, 74 CBs are detected in the far wings of the Mg II lines (at +/-1 Å and 3.5 Å). Around 10% of CBs have a signature in Si IV and CII. NLTE models with a hotspot located in the low atmosphere were found to fit a sample of Mg II profiles in CBs. The Hα profiles computed with these Mg II CB models are consistent with typical EB profiles <span class="hlt">observed</span> from ground based telescopes e.g. THEMIS. A 2D NLTE modelling of fibrils (canopy) demonstrates that the Mg II line centres can be significantly affected but not the peaks and the wings of Mg II lines. Conclusions: We conclude that the <span class="hlt">bright</span> points <span class="hlt">observed</span> in Mg II lines can be formed in an extended domain of altitudes in the photosphere and/or the chromosphere (400 to 750 km). Our results are consistent with the theory of heating by Joule dissipation in the atmosphere produced by magnetic field reconnection during flux emergence.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120002744','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120002744"><span>Colors and Photometry of <span class="hlt">Bright</span> Materials on Vesta as Seen by the Dawn Framing Camera</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schroeder, S. E.; Li, J.-Y.; Mittlefehldt, D. W.; Pieters, C. M.; De Sanctis, M. C.; Hiesinger, H.; Blewett, D. T.; Russell, C. T.; Raymond, C. A.; Keller, H. U.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20120002744'); toggleEditAbsImage('author_20120002744_show'); toggleEditAbsImage('author_20120002744_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20120002744_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20120002744_hide"></p> <p>2012-01-01</p> <p>The Dawn spacecraft has been in orbit around the asteroid Vesta since July, 2011. The on-board Framing Camera has acquired thousands of high-resolution images of the regolith-covered <span class="hlt">surface</span> through one clear and seven narrow-band filters in the visible and near-IR wavelength range. It has <span class="hlt">observed</span> <span class="hlt">bright</span> and dark materials that have a range of reflectance that is unusually wide for an asteroid. Material brighter than average is predominantly found on crater walls, and in ejecta surrounding caters in the southern hemisphere. Most likely, the brightest material identified on the Vesta <span class="hlt">surface</span> so far is located on the inside of a crater at 64.27deg S, 1.54deg . The apparent <span class="hlt">brightness</span> of a regolith is influenced by factors such as particle size, mineralogical composition, and viewing geometry. As such, the presence of <span class="hlt">bright</span> material can indicate differences in lithology and/or degree of space weathering. We retrieve the spectral and photometric properties of various <span class="hlt">bright</span> terrains from false-color images acquired in the High Altitude Mapping Orbit (HAMO). We find that most <span class="hlt">bright</span> material has a deeper 1-m pyroxene band than average. However, the aforementioned brightest material appears to have a 1-m band that is actually less deep, a result that awaits confirmation by the on-board VIR spectrometer. This site may harbor a class of material unique for Vesta. We discuss the implications of our spectral findings for the origin of <span class="hlt">bright</span> materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSH43A2806I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSH43A2806I"><span>ALMA Discovery of Solar Umbral <span class="hlt">Brightness</span> Enhancement at λ = 3 mm</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Iwai, K.; Loukitcheva, M.; Shimojo, M.; Solanki, S. K.; White, S. M.</p> <p>2017-12-01</p> <p>We report the discovery of a <span class="hlt">brightness</span> enhancement in the center of a large sunspot umbra at a wavelength of 3 mm using the Atacama Large Millimeter/sub-millimeter Array (ALMA). Sunspots are among the most prominent features on the solar <span class="hlt">surface</span>, but many of their aspects are surprisingly poorly understood. We analyzed a λ = 3 mm (100 GHz) mosaic image obtained by ALMA that includes a large sunspot within the active region AR12470, on 2015 December 16. The 3 mm map has a 300''×300'' field of view and 4.9''×2.2'' spatial resolution, which is the highest spatial resolution map of an entire sunspot in this frequency range. We find a gradient of 3 mm <span class="hlt">brightness</span> from a high value in the outer penumbra to a low value in the inner penumbra/outer umbra. Within the inner umbra, there is a marked increase in 3 mm <span class="hlt">brightness</span> temperature, which we call an umbral <span class="hlt">brightness</span> enhancement. This enhanced emission corresponds to a temperature excess of 800 K relative to the surrounding inner penumbral region and coincides with excess <span class="hlt">brightness</span> in the 1330 and 1400 Å slit-jaw images of the Interface Region Imaging Spectrograph (IRIS), adjacent to a partial lightbridge. This λ = 3 mm <span class="hlt">brightness</span> enhancement may be an intrinsic feature of the sunspot umbra at chromospheric heights, such as a manifestation of umbral flashes, or it could be related to a coronal plume, since the <span class="hlt">brightness</span> enhancement was coincident with the footpoint of a coronal loop <span class="hlt">observed</span> at 171 Å.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.472L..30P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.472L..30P"><span>Spot distribution and fast <span class="hlt">surface</span> evolution on Vega</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Petit, P.; Hébrard, E. M.; Böhm, T.; Folsom, C. P.; Lignières, F.</p> <p>2017-11-01</p> <p>Spectral signatures of <span class="hlt">surface</span> spots were recently discovered from high cadence <span class="hlt">observations</span> of the A star Vega. We aim at constraining the <span class="hlt">surface</span> distribution of these photospheric inhomogeneities and investigating a possible short-term evolution of the spot pattern. Using data collected over five consecutive nights, we employ the Doppler imaging method to reconstruct three different maps of the stellar <span class="hlt">surface</span>, from three consecutive subsets of the whole time series. The <span class="hlt">surface</span> maps display a complex distribution of dark and <span class="hlt">bright</span> spots, covering most of the visible fraction of the stellar <span class="hlt">surface</span>. A number of <span class="hlt">surface</span> features are consistently recovered in all three maps, but other features seem to evolve over the time span of <span class="hlt">observations</span>, suggesting that fast changes can affect the <span class="hlt">surface</span> of Vega within a few days at most. The short-term evolution is <span class="hlt">observed</span> as emergence or disappearance of individual spots, and may also show up as zonal flows, with low- and high-latitude belts rotating faster than intermediate latitudes. It is tempting to relate the <span class="hlt">surface</span> <span class="hlt">brightness</span> activity to the complex magnetic field topology previously reconstructed for Vega, although strictly simultaneous <span class="hlt">brightness</span> and magnetic maps will be necessary to assess this potential link.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950034096&hterms=correlation+coefficient&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcorrelation%2Bcoefficient','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950034096&hterms=correlation+coefficient&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcorrelation%2Bcoefficient"><span><span class="hlt">Observations</span> of copolar correlation coefficient through a <span class="hlt">bright</span> band at vertical incidence</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zrnic, D. S.; Raghavan, R.; Chandrasekar, V.</p> <p>1994-01-01</p> <p>This paper discusses an application of polarimetric measurements at vertical incidence. In particular, the correlation coefficients between linear copolar components are examined, and measurements obtained with the National Severe Storms Laboratory (NSSL)'s and National Center for Atmospheric Research (NCAR)'s polarimetric radars are presented. The data are from two well-defined <span class="hlt">bright</span> bands. A sharp decrease of the correlation coefficient, confined to a height interval of a few hundred meters, marks the bottom of the <span class="hlt">bright</span> band.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22662986-hi-observations-sunspot-penumbral-bright-dots','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22662986-hi-observations-sunspot-penumbral-bright-dots"><span>Hi-C <span class="hlt">OBSERVATIONS</span> OF SUNSPOT PENUMBRAL <span class="hlt">BRIGHT</span> DOTS</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Alpert, Shane E.; Tiwari, Sanjiv K.; Moore, Ronald L.</p> <p></p> <p>We report <span class="hlt">observations</span> of <span class="hlt">bright</span> dots (BDs) in a sunspot penumbra using High Resolution Coronal Imager (Hi-C) data in 193 Å and examine their sizes, lifetimes, speeds, and intensities. The sizes of the BDs are on the order of 1″ and are therefore hard to identify in the Atmospheric Imaging Assembly (AIA) 193 Å images, which have a 1.″2 spatial resolution, but become readily apparent with Hi-C's spatial resolution, which is five times better. We supplement Hi-C data with data from AIA's 193 Å passband to see the complete lifetime of the BDs that appeared before and/or lasted longer thanmore » Hi-C's three-minute <span class="hlt">observation</span> period. Most Hi-C BDs show clear lateral movement along penumbral striations, either toward or away from the sunspot umbra. Single BDs often interact with other BDs, combining to fade away or brighten. The BDs that do not interact with other BDs tend to have smaller displacements. These BDs are about as numerous but move slower on average than Interface Region Imaging Spectrograph (IRIS) BDs, which was recently reported by Tian et al., and the sizes and lifetimes are on the higher end of the distribution of IRIS BDs. Using additional AIA passbands, we compare the light curves of the BDs to test whether the Hi-C BDs have transition region (TR) temperatures like those of the IRIS BDs. The light curves of most Hi-C BDs peak together in different AIA channels, indicating that their temperatures are likely in the range of the cooler TR (1−4 × 10{sup 5} K).« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018eMetN...3...51K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018eMetN...3...51K"><span>Two <span class="hlt">bright</span> fireballs over Great Britain</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Koukal, Jakub; Káčerek, Richard</p> <p>2018-02-01</p> <p>On November 24, 2017 shortly before midnight and on November 25, 2017 shortly before sunrise, two very <span class="hlt">bright</span> fireballs lit up the sky over the United Kingdom. The UKMON (United Kingdom Meteor <span class="hlt">Observation</span> Network) cameras and onboard cameras in the automobiles recorded their flight. The fireballs paths in the Earth's atmosphere were calculated, as well as the orbits of bodies in the Solar System. The flight of both bodies, the absolute magnitude of which approached the <span class="hlt">brightness</span> of the full Moon, was also <span class="hlt">observed</span> by numerous random <span class="hlt">observers</span> from the public in Great Britain, Ireland and France.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DMP.Q1100K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DMP.Q1100K"><span>Dark-<span class="hlt">Bright</span> Soliton Dynamics Beyond the Mean-Field Approximation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Katsimiga, Garyfallia; Koutentakis, Georgios; Mistakidis, Simeon; Kevrekidis, Panagiotis; Schmelcher, Peter; Theory Group of Fundamental Processes in Quantum Physics Team</p> <p>2017-04-01</p> <p>The dynamics of dark <span class="hlt">bright</span> solitons beyond the mean-field approximation is investigated. We first examine the case of a single dark-<span class="hlt">bright</span> soliton and its oscillations within a parabolic trap. Subsequently, we move to the setting of collisions, comparing the mean-field approximation to that involving multiple orbitals in both the dark and the <span class="hlt">bright</span> component. Fragmentation is present and significantly affects the dynamics, especially in the case of slower solitons and in that of lower atom numbers. It is shown that the presence of fragmentation allows for bipartite entanglement between the distinguishable species. Most importantly the interplay between fragmentation and entanglement leads to the decay of each of the initial mean-field dark-<span class="hlt">bright</span> solitons into fast and slow fragmented dark-<span class="hlt">bright</span> structures. A variety of excitations including dark-<span class="hlt">bright</span> solitons in multiple (concurrently populated) orbitals is <span class="hlt">observed</span>. Dark-antidark states and domain-wall-<span class="hlt">bright</span> soliton complexes can also be <span class="hlt">observed</span> to arise spontaneously in the beyond mean-field dynamics. Deutsche Forschungsgemeinschaft (DFG) in the framework of the SFB 925 ``Light induced dynamics and control of correlated quantum systems''.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24323112','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24323112"><span><span class="hlt">Brightness</span> perception of unrelated self-luminous colors.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Withouck, Martijn; Smet, Kevin A G; Ryckaert, Wouter R; Pointer, Michael R; Deconinck, Geert; Koenderink, Jan; Hanselaer, Peter</p> <p>2013-06-01</p> <p>The perception of <span class="hlt">brightness</span> of unrelated self-luminous colored stimuli of the same luminance has been investigated. The Helmholtz-Kohlrausch (H-K) effect, i.e., an increase in <span class="hlt">brightness</span> perception due to an increase in saturation, is clearly <span class="hlt">observed</span>. This <span class="hlt">brightness</span> perception is compared with the calculated <span class="hlt">brightness</span> according to six existing vision models, color appearance models, and models based on the concept of equivalent luminance. Although these models included the H-K effect and half of them were developed to work with unrelated colors, none of the models seemed to be able to fully predict the perceived <span class="hlt">brightness</span>. A tentative solution to increase the prediction accuracy of the color appearance model CAM97u, developed by Hunt, is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA17471.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA17471.html"><span>Dark Lakes on a <span class="hlt">Bright</span> Landscape</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2013-10-23</p> <p>Ultracold hydrocarbon lakes and seas dark shapes near the north pole of Saturn moon Titan can be seen embedded in some kind of <span class="hlt">bright</span> <span class="hlt">surface</span> material in this infrared mosaic from NASA Cassini mission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040121126&hterms=landcover&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dlandcover','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040121126&hterms=landcover&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dlandcover"><span>Landcover Based Optimal Deconvolution of PALS L-band Microwave <span class="hlt">Brightness</span> Temperature</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Limaye, Ashutosh S.; Crosson, William L.; Laymon, Charles A.; Njoku, Eni G.</p> <p>2004-01-01</p> <p>An optimal de-convolution (ODC) technique has been developed to estimate microwave <span class="hlt">brightness</span> temperatures of agricultural fields using microwave radiometer <span class="hlt">observations</span>. The technique is applied to airborne measurements taken by the Passive and Active L and S band (PALS) sensor in Iowa during Soil Moisture Experiments in 2002 (SMEX02). Agricultural fields in the study area were predominantly soybeans and corn. The <span class="hlt">brightness</span> temperatures of corn and soybeans were <span class="hlt">observed</span> to be significantly different because of large differences in vegetation biomass. PALS <span class="hlt">observations</span> have significant over-sampling; <span class="hlt">observations</span> were made about 100 m apart and the sensor footprint extends to about 400 m. Conventionally, <span class="hlt">observations</span> of this type are averaged to produce smooth spatial data fields of <span class="hlt">brightness</span> temperatures. However, the conventional approach is in contrast to reality in which the <span class="hlt">brightness</span> temperatures are in fact strongly dependent on landcover, which is characterized by sharp boundaries. In this study, we mathematically de-convolve the <span class="hlt">observations</span> into <span class="hlt">brightness</span> temperature at the field scale (500-800m) using the sensor antenna response function. The result is more accurate spatial representation of field-scale <span class="hlt">brightness</span> temperatures, which may in turn lead to more accurate soil moisture retrieval.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110008803','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110008803"><span><span class="hlt">Observations</span> During GRIP from HIRAD: Ocean <span class="hlt">Surface</span> Wind Speed and Rain Rate</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Miller, Timothy L.; James, M. W.; Jones, L.; Ruf, C. S.; Uhlhorn, E. W.; Bailey, M. C.; Buckley, C. D.; Simmons, D. E.; Johnstone, S.; Peterson, A.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20110008803'); toggleEditAbsImage('author_20110008803_show'); toggleEditAbsImage('author_20110008803_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20110008803_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20110008803_hide"></p> <p>2011-01-01</p> <p>HIRAD (Hurricane Imaging Radiometer) flew on the WB-57 during NASA's GRIP (Genesis and Rapid Intensification Processes) campaign in August - September of 2010. HIRAD is a new C-band radiometer using a synthetic thinned array radiometer (STAR) technology to obtain cross-track resolution of approximately 3 degrees, out to approximately 60 degrees to each side of nadir. By obtaining measurements of emissions at 4, 5, 6, and 6.6 GHz, <span class="hlt">observations</span> of ocean <span class="hlt">surface</span> wind speed and rain rate can be inferred. This technique has been used for many years by precursor instruments, including the Stepped Frequency Microwave Radiometer (SFMR), which has been flying on the NOAA and USAF hurricane reconnaissance aircraft for several years. The advantage of HIRAD over SFMR is that HIRAD can <span class="hlt">observe</span> a +/- 60-degree swath, rather than a single footprint at nadir angle. Results from the flights during the GRIP campaign will be shown, including images of <span class="hlt">brightness</span> temperatures, wind speed, and rain rate. To the extent possible, comparisons will be made with <span class="hlt">observations</span> from other instruments on the GRIP campaign, for which HIRAD <span class="hlt">observations</span> are either directly comparable or are complementary. Potential impacts on operational ocean <span class="hlt">surface</span> wind analyses and on numerical weather forecasts will also be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22365802-larger-planet-radii-inferred-from-stellar-flicker-brightness-variations-bright-planet-host-stars','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22365802-larger-planet-radii-inferred-from-stellar-flicker-brightness-variations-bright-planet-host-stars"><span>LARGER PLANET RADII INFERRED FROM STELLAR ''FLICKER'' <span class="hlt">BRIGHTNESS</span> VARIATIONS OF <span class="hlt">BRIGHT</span> PLANET-HOST STARS</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Bastien, Fabienne A.; Stassun, Keivan G.; Pepper, Joshua</p> <p>2014-06-10</p> <p>Most extrasolar planets have been detected by their influence on their parent star, typically either gravitationally (the Doppler method) or by the small dip in <span class="hlt">brightness</span> as the planet blocks a portion of the star (the transit method). Therefore, the accuracy with which we know the masses and radii of extrasolar planets depends directly on how well we know those of the stars, the latter usually determined from the measured stellar <span class="hlt">surface</span> gravity, log g. Recent work has demonstrated that the short-timescale <span class="hlt">brightness</span> variations ({sup f}licker{sup )} of stars can be used to measure log g to a high accuracymore » of ∼0.1-0.2 dex. Here, we use flicker measurements of 289 <span class="hlt">bright</span> (Kepmag < 13) candidate planet-hosting stars with T {sub eff} = 4500-6650 K to re-assess the stellar parameters and determine the resulting impact on derived planet properties. This re-assessment reveals that for the brightest planet-host stars, Malmquist bias contaminates the stellar sample with evolved stars: nearly 50% of the <span class="hlt">bright</span> planet-host stars are subgiants. As a result, the stellar radii, and hence the radii of the planets orbiting these stars, are on average 20%-30% larger than previous measurements had suggested.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011ASPC..445..171S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ASPC..445..171S"><span>Interferometric Constraints on <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Asymmetries in Long-Period Variable Stars: A Threat to Accurate Gaia Parallaxes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sacuto, S.; Jorissen, A.; Cruzalèbes, P.; Pasquato, E.; Chiavassa, A.; Spang, A.; Rabbia, Y.; Chesneau, O.</p> <p>2011-09-01</p> <p>A monitoring of <span class="hlt">surface</span> <span class="hlt">brightness</span> asymmetries in evolved giants and supergiants is necessary to estimate the threat that they represent to accurate Gaia parallaxes. Closure-phase measurements obtained with AMBER/VISA in a 3-telescope configuration are fitted by a simple model to constrain the photocenter displacement. The results for the C-type star TX Psc show a large deviation of the photocenter displacement that could bias the Gaia parallax.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22654467-alma-discovery-solar-umbral-brightness-enhancement-mm','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22654467-alma-discovery-solar-umbral-brightness-enhancement-mm"><span>ALMA Discovery of Solar Umbral <span class="hlt">Brightness</span> Enhancement at λ = 3 mm</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Iwai, Kazumasa; Loukitcheva, Maria; Shimojo, Masumi</p> <p></p> <p>We report the discovery of a <span class="hlt">brightness</span> enhancement in the center of a large sunspot umbra at a wavelength of 3 mm using the Atacama Large Millimeter/sub-millimeter Array (ALMA). Sunspots are among the most prominent features on the solar <span class="hlt">surface</span>, but many of their aspects are surprisingly poorly understood. We analyzed a λ = 3 mm (100 GHz) mosaic image obtained by ALMA that includes a large sunspot within the active region AR12470, on 2015 December 16. The 3 mm map has a 300″ × 300″ field of view and 4.″9 × 2.″2 spatial resolution, which is the highest spatialmore » resolution map of an entire sunspot in this frequency range. We find a gradient of 3 mm <span class="hlt">brightness</span> from a high value in the outer penumbra to a low value in the inner penumbra/outer umbra. Within the inner umbra, there is a marked increase in 3 mm <span class="hlt">brightness</span> temperature, which we call an umbral <span class="hlt">brightness</span> enhancement. This enhanced emission corresponds to a temperature excess of 800 K relative to the surrounding inner penumbral region and coincides with excess <span class="hlt">brightness</span> in the 1330 and 1400 Å slit-jaw images of the Interface Region Imaging Spectrograph ( IRIS ), adjacent to a partial lightbridge. This λ = 3 mm <span class="hlt">brightness</span> enhancement may be an intrinsic feature of the sunspot umbra at chromospheric heights, such as a manifestation of umbral flashes, or it could be related to a coronal plume, since the <span class="hlt">brightness</span> enhancement was coincident with the footpoint of a coronal loop <span class="hlt">observed</span> at 171 Å.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2808153','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2808153"><span>CD94 <span class="hlt">surface</span> density identifies a functional intermediary between the CD56<span class="hlt">bright</span> and CD56dim human NK-cell subsets</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Mao, Hsiaoyin C.; Wei, Min; Hughes, Tiffany; Zhang, Jianying; Park, Il-kyoo; Liu, Shujun; McClory, Susan; Marcucci, Guido; Trotta, Rossana</p> <p>2010-01-01</p> <p>Human CD56<span class="hlt">bright</span> natural killer (NK) cells possess little or no killer immunoglobulin-like receptors (KIRs), high interferon-γ (IFN-γ) production, but little cytotoxicity. CD56dim NK cells have high KIR expression, produce little IFN-γ, yet display high cytotoxicity. We hypothesized that, if human NK maturation progresses from a CD56<span class="hlt">bright</span> to a CD56dim phenotype, an intermediary NK cell must exist, which demonstrates more functional overlap than these 2 subsets, and we used CD94 expression to test our hypothesis. CD94highCD56dim NK cells express CD62L, CD2, and KIR at levels between CD56<span class="hlt">bright</span> and CD94lowCD56dim NK cells. CD94highCD56dim NK cells produce less monokine-induced IFN-γ than CD56<span class="hlt">bright</span> NK cells but much more than CD94lowCD56dim NK cells because of differential interleukin-12–mediated STAT4 phosphorylation. CD94highCD56dim NK cells possess a higher level of granzyme B and perforin expression and CD94-mediated redirected killing than CD56<span class="hlt">bright</span> NK cells but lower than CD94lowCD56dim NK cells. Collectively, our data suggest that the density of CD94 <span class="hlt">surface</span> expression on CD56dim NK cells identifies a functional and likely developmental intermediary between CD56<span class="hlt">bright</span> and CD94lowCD56dim NK cells. This supports the notion that, in vivo, human CD56<span class="hlt">bright</span> NK cells progress through a continuum of differentiation that ends with a CD94lowCD56dim phenotype. PMID:19897577</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19897577','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19897577"><span>CD94 <span class="hlt">surface</span> density identifies a functional intermediary between the CD56<span class="hlt">bright</span> and CD56dim human NK-cell subsets.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yu, Jianhua; Mao, Hsiaoyin C; Wei, Min; Hughes, Tiffany; Zhang, Jianying; Park, Il-kyoo; Liu, Shujun; McClory, Susan; Marcucci, Guido; Trotta, Rossana; Caligiuri, Michael A</p> <p>2010-01-14</p> <p>Human CD56(<span class="hlt">bright</span>) natural killer (NK) cells possess little or no killer immunoglobulin-like receptors (KIRs), high interferon-gamma (IFN-gamma) production, but little cytotoxicity. CD56(dim) NK cells have high KIR expression, produce little IFN-gamma, yet display high cytotoxicity. We hypothesized that, if human NK maturation progresses from a CD56(<span class="hlt">bright</span>) to a CD56(dim) phenotype, an intermediary NK cell must exist, which demonstrates more functional overlap than these 2 subsets, and we used CD94 expression to test our hypothesis. CD94(high)CD56(dim) NK cells express CD62L, CD2, and KIR at levels between CD56(<span class="hlt">bright</span>) and CD94(low)CD56(dim) NK cells. CD94(high)CD56(dim) NK cells produce less monokine-induced IFN-gamma than CD56(<span class="hlt">bright</span>) NK cells but much more than CD94(low)CD56(dim) NK cells because of differential interleukin-12-mediated STAT4 phosphorylation. CD94(high)CD56(dim) NK cells possess a higher level of granzyme B and perforin expression and CD94-mediated redirected killing than CD56(<span class="hlt">bright</span>) NK cells but lower than CD94(low)CD56(dim) NK cells. Collectively, our data suggest that the density of CD94 <span class="hlt">surface</span> expression on CD56(dim) NK cells identifies a functional and likely developmental intermediary between CD56(<span class="hlt">bright</span>) and CD94(low)CD56(dim) NK cells. This supports the notion that, in vivo, human CD56(<span class="hlt">bright</span>) NK cells progress through a continuum of differentiation that ends with a CD94(low)CD56(dim) phenotype.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19760060739&hterms=sparrow&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dsparrow','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19760060739&hterms=sparrow&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dsparrow"><span>Pioneer 10 <span class="hlt">observations</span> of zodiacal light <span class="hlt">brightness</span> near the ecliptic - Changes with heliocentric distance</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hanner, M. S.; Weinberg, J. L.; Beeson, D. E.; Sparrow, J. G.</p> <p>1976-01-01</p> <p>Sky maps made by the Pioneer 10 Imaging Photopolarimeter (IPP) at sun-spacecraft distances from 1 to 3 AU have been analyzed to derive the <span class="hlt">brightness</span> of the zodiacal light near the ecliptic at elongations greater than 90 degrees. The change in zodiacal light <span class="hlt">brightness</span> with heliocentric distance is compared with models of the spatial distribution of the dust. Use of background starlight <span class="hlt">brightnesses</span> derived from IPP measurements beyond the asteroid belt, where the zodiacal light is not detected, and, especially, use of a corrected calibration lead to considerably lower values for zodiacal light than those reported by us previously.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_10 --> <div id="page_11" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="201"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014A%26A...568A..13R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014A%26A...568A..13R"><span>Comparison of solar photospheric <span class="hlt">bright</span> points between Sunrise <span class="hlt">observations</span> and MHD simulations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Riethmüller, T. L.; Solanki, S. K.; Berdyugina, S. V.; Schüssler, M.; Martínez Pillet, V.; Feller, A.; Gandorfer, A.; Hirzberger, J.</p> <p>2014-08-01</p> <p><span class="hlt">Bright</span> points (BPs) in the solar photosphere are thought to be the radiative signatures (small-scale <span class="hlt">brightness</span> enhancements) of magnetic elements described by slender flux tubes or sheets located in the darker intergranular lanes in the solar photosphere. They contribute to the ultraviolet (UV) flux variations over the solar cycle and hence may play a role in influencing the Earth's climate. Here we aim to obtain a better insight into their properties by combining high-resolution UV and spectro-polarimetric <span class="hlt">observations</span> of BPs by the Sunrise Observatory with 3D compressible radiation magnetohydrodynamical (MHD) simulations. To this end, full spectral line syntheses are performed with the MHD data and a careful degradation is applied to take into account all relevant instrumental effects of the <span class="hlt">observations</span>. In a first step it is demonstrated that the selected MHD simulations reproduce the measured distributions of intensity at multiple wavelengths, line-of-sight velocity, spectral line width, and polarization degree rather well. The simulated line width also displays the correct mean, but a scatter that is too small. In the second step, the properties of <span class="hlt">observed</span> BPs are compared with synthetic ones. Again, these are found to match relatively well, except that the <span class="hlt">observations</span> display a tail of large BPs with strong polarization signals (most likely network elements) not found in the simulations, possibly due to the small size of the simulation box. The higher spatial resolution of the simulations has a significant effect, leading to smaller and more numerous BPs. The <span class="hlt">observation</span> that most BPs are weakly polarized is explained mainly by the spatial degradation, the stray light contamination, and the temperature sensitivity of the Fe i line at 5250.2 Å. Finally, given that the MHD simulations are highly consistent with the <span class="hlt">observations</span>, we used the simulations to explore the properties of BPs further. The Stokes V asymmetries increase with the distance to the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130001913','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130001913"><span><span class="hlt">Observations</span> of C-Band <span class="hlt">Brightness</span> Temperature and Ocean <span class="hlt">Surface</span> Wind Speed and Rain Rate in Hurricanes Earl And Karl (2010)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Miller, Timothy; James, Mark; Roberts, Brent J.; Biswax, Sayak; Uhlhorn, Eric; Black, Peter; Linwood Jones, W.; Johnson, Jimmy; Farrar, Spencer; Sahawneh, Saleem</p> <p>2012-01-01</p> <p>Ocean <span class="hlt">surface</span> emission is affected by: a) Sea <span class="hlt">surface</span> temperature. b) Wind speed (foam fraction). c) Salinity After production of calibrated Tb fields, geophysical fields wind speed and rain rate (or column) are retrieved. HIRAD utilizes NASA Instrument Incubator Technology: a) Provides unique <span class="hlt">observations</span> of sea <span class="hlt">surface</span> wind, temp and rain b) Advances understanding & prediction of hurricane intensity c) Expands Stepped Frequency Microwave Radiometer capabilities d) Uses synthetic thinned array and RFI mitigation technology of Lightweight Rain Radiometer (NASA Instrument Incubator) Passive Microwave C-Band Radiometer with Freq: 4, 5, 6 & 6.6 GHz: a) Version 1: H-pol for ocean wind speed, b) Version 2: dual ]pol for ocean wind vectors. Performance Characteristics: a) Earth Incidence angle: 0deg - 60deg, b) Spatial Resolution: 2-5 km, c) Swath: approx.70 km for 20 km altitude. <span class="hlt">Observational</span> Goals: WS 10 - >85 m/s RR 5 - > 100 mm/hr.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19547213','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19547213"><span>Dark-<span class="hlt">bright</span> soliton pairs in nonlocal nonlinear media.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lin, Yuan Yao; Lee, Ray-Kuang</p> <p>2007-07-09</p> <p>We study the formation of dark-<span class="hlt">bright</span> vector soliton pairs in nonlocal Kerr-type nonlinear medium. We show, by analytical analysis and direct numerical calculation, that in addition to stabilize of vector soliton pairs nonlocal nonlinearity also helps to reduce the threshold power for forming a guided <span class="hlt">bright</span> soliton. With help of the nonlocality, it is expected that the <span class="hlt">observation</span> of dark-<span class="hlt">bright</span> vector soliton pairs in experiments becomes more workable.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22370436-investigation-moving-structures-coronal-bright-point','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22370436-investigation-moving-structures-coronal-bright-point"><span>Investigation of the moving structures in a coronal <span class="hlt">bright</span> point</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Ning, Zongjun; Guo, Yang, E-mail: ningzongjun@pmo.ac.cn</p> <p>2014-10-10</p> <p>We have explored the moving structures in a coronal <span class="hlt">bright</span> point (CBP) <span class="hlt">observed</span> by the Solar Dynamic Observatory Atmospheric Imaging Assembly (AIA) on 2011 March 5. This CBP event has a lifetime of ∼20 minutes and is <span class="hlt">bright</span> with a curved shape along a magnetic loop connecting a pair of negative and positive fields. AIA imaging <span class="hlt">observations</span> show that a lot of <span class="hlt">bright</span> structures are moving intermittently along the loop legs toward the two footpoints from the CBP <span class="hlt">brightness</span> core. Such moving <span class="hlt">bright</span> structures are clearly seen at AIA 304 Å. In order to analyze their features, the CBP ismore » cut along the motion direction with a curved slit which is wide enough to cover the bulk of the CBP. After integrating the flux along the slit width, we get the spacetime slices at nine AIA wavelengths. The oblique streaks starting from the edge of the CBP <span class="hlt">brightness</span> core are identified as moving <span class="hlt">bright</span> structures, especially on the derivative images of the <span class="hlt">brightness</span> spacetime slices. They seem to originate from the same position near the loop top. We find that these oblique streaks are bi-directional, simultaneous, symmetrical, and periodic. The average speed is about 380 km s{sup –1}, and the period is typically between 80 and 100 s. Nonlinear force-free field extrapolation shows the possibility that magnetic reconnection takes place during the CBP, and our findings indicate that these moving <span class="hlt">bright</span> structures could be the <span class="hlt">observational</span> outflows after magnetic reconnection in the CBP.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018OptCo.414...29Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018OptCo.414...29Y"><span>The <span class="hlt">bright-bright</span> and <span class="hlt">bright</span>-dark mode coupling-based planar metamaterial for plasmonic EIT-like effect</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, Wei; Meng, Hongyun; Chen, Zhangjie; Li, Xianping; Zhang, Xing; Wang, Faqiang; Wei, Zhongchao; Tan, Chunhua; Huang, Xuguang; Li, Shuti</p> <p>2018-05-01</p> <p>In this paper, we propose a novel planar metamaterial structure for the electromagnetically induced transparency (EIT)-like effect, which consists of a split-ring resonator (SRR) and a pair of metal strips. The simulated results indicate that a single transparency window can be realized in the symmetry situation, which originates from the <span class="hlt">bright-bright</span> mode coupling. Further, a dual-band EIT-like effect can be achieved in the asymmetry situation, which is due to the <span class="hlt">bright-bright</span> mode coupling and <span class="hlt">bright</span>-dark mode coupling, respectively. Different EIT-like effect can be simultaneously achieved in the proposed structure with the different situations. It is of certain significance for the study of EIT-like effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.P53A2092M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.P53A2092M"><span><span class="hlt">Observations</span> of Mercury's <span class="hlt">Surface</span>-Bounded Exosphere from Orbit: Results from the Mercury Atmospheric and <span class="hlt">Surface</span> Composition Spectrometer aboard the MESSENGER Spacecraft</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McClintock, W. E.; Burger, M. H.; Cassidy, T. A.; Killen, R. M.; Merkel, A. W.; Sarantos, M.; Solomon, S. C.; Vervack, R. J., Jr.</p> <p>2015-12-01</p> <p>The Mercury Atmospheric and <span class="hlt">Surface</span> Composition Spectrometer (MASCS), on the MErcury <span class="hlt">Surface</span>, Space ENvironment, GEochemistry, and Ranging (MESSENGER) spacecraft, conducted orbital <span class="hlt">observations</span> of Mercury's dayside and nightside exosphere from 29 March 2011 to the end of the mission on 30 April 2015. Over slightly more than four Earth-years, MASCS measured emission profiles versus altitude for calcium (Ca), sodium (Na), and magnesium (Mg) at a daily cadence. These species exhibit different spatial distributions, suggesting distinct source processes. MASCS <span class="hlt">observed</span> seasonal variations in all three species that are remarkably repeatable from one Mercury year to the next, and did so consistently during the entire 17-Mercury-year duration of the orbital phase of the mission. Whereas MASCS has characterized the seasonal variation, it has provided, at best, only weak evidence for the episodic behavior <span class="hlt">observed</span> in ground-based studies of Na. Joint analyses of MASCS <span class="hlt">observations</span> and <span class="hlt">surface</span> precipitation patterns for energetic particles inferred from <span class="hlt">observations</span> by the Energetic Particle Spectrometer (EPS) and the Fast Imaging Plasma Spectrometer (FIPS) on MESSENGER have not yielded clear correlations. This lack of correlation may be due in part to the MASCS <span class="hlt">observational</span> geometries. MASCS has conducted a number of searches for other, weakly emitting species. Hydrogen data from the orbital phase are consistent with profiles <span class="hlt">observed</span> during MESSENGER's flybys of Mercury. Oxygen detections have proven elusive, and the previously reported <span class="hlt">observation</span> with a <span class="hlt">brightness</span> of 4 R may only be an upper limit. Ongoing analysis of weak species data suggests that additional species are present.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890031691&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890031691&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour"><span>A study of coronal <span class="hlt">bright</span> points at 20 cm wavelength</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nitta, N.; Kundu, M. R.</p> <p>1988-01-01</p> <p>The paper presents the results of a study of coronal <span class="hlt">bright</span> points <span class="hlt">observed</span> at 20 cm with the VLA on a day when the sun was exceptionally quiet. Microwave maps of <span class="hlt">bright</span> points were obtained using data for the entire <span class="hlt">observing</span> period of 5 hours, as well as for shorter periods of a few minutes. Most <span class="hlt">bright</span> points, especially those appearing in the full-period maps, appear to be associated with small bipolar structures on the photospheric magnetogram. Overlays of <span class="hlt">bright</span> point (BP) maps on the Ca(+) K picture, show that the brightest part of BP tends to lie on the boundary of a supergranulation network.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JARS....7.3469H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JARS....7.3469H"><span>Microwave <span class="hlt">brightness</span> temperature features of lunar craters: <span class="hlt">observation</span> from Chang'E-1 mission</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hu, Guo-Ping; Chen, Ke; Guo, Wei; Li, Qing-Xia; Su, Hong-Yan</p> <p>2013-01-01</p> <p>Topographic features of lunar craters have been found from the <span class="hlt">brightness</span> temperature (TB) <span class="hlt">observed</span> by the multichannel (3.0, 7.8, 19.35, and 37 GHz) microwave radiometer (MRM) aboard Chang'E-1 (CE-1) in a single track view. As the topographic effect is more obvious at 37 GHz, 37 GHz TB has been focused on in this work. The variation of 37 GHz daytime (nighttime) TB along the profile of a crater is found to show an oscillatory behavior. The amplitude of daytime TB is significantly affected by the <span class="hlt">observation</span> time and the shape of the crater, whose diameter is bigger than the spatial resolution of MRM onboard CE-1. The large and typical diurnal TB difference (nighttime TB minus daytime TB) at 37 GHz over selected young craters due to the large rock abundance in craters, have been discussed and compared with the altitude profile.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...856...17L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...856...17L"><span>Studies of Isolated and Non-isolated Photospheric <span class="hlt">Bright</span> Points in an Active Region <span class="hlt">Observed</span> by the New Vacuum Solar Telescope</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Yanxiao; Xiang, Yongyuan; Erdélyi, Robertus; Liu, Zhong; Li, Dong; Ning, Zongjun; Bi, Yi; Wu, Ning; Lin, Jun</p> <p>2018-03-01</p> <p>Properties of photospheric <span class="hlt">bright</span> points (BPs) near an active region have been studied in TiO λ 7058 Å images <span class="hlt">observed</span> by the New Vacuum Solar Telescope of the Yunnan Observatories. We developed a novel recognition method that was used to identify and track 2010 BPs. The <span class="hlt">observed</span> evolving BPs are classified into isolated (individual) and non-isolated (where multiple BPs are <span class="hlt">observed</span> to display splitting and merging behaviors) sets. About 35.1% of BPs are non-isolated. For both isolated and non-isolated BPs, the <span class="hlt">brightness</span> varies from 0.8 to 1.3 times the average background intensity and follows a Gaussian distribution. The lifetimes of BPs follow a log-normal distribution, with characteristic lifetimes of (267 ± 140) s and (421 ± 255) s, respectively. Their size also follows log-normal distribution, with an average size of about (2.15 ± 0.74) × 104 km2 and (3.00 ± 1.31) × 104 km2 for area, and (163 ± 27) km and (191 ± 40) km for diameter, respectively. Our results indicate that regions with strong background magnetic field have higher BP number density and higher BP area coverage than regions with weak background field. Apparently, the <span class="hlt">brightness</span>/size of BPs does not depend on the background field. Lifetimes in regions with strong background magnetic field are shorter than those in regions with weak background field, on average.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01515&hterms=gardening&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dgardening','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01515&hterms=gardening&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dgardening"><span><span class="hlt">Bright</span> Ray Craters in Ganymede's Northern Hemisphere</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1979-01-01</p> <p>GANYMEDE COLOR PHOTOS: This color picture as acquired by Voyager 1 during its approach to Ganymede on Monday afternoon (the 5th of March). At ranges between about 230 to 250 thousand km. The images show detail on the <span class="hlt">surface</span> with a resolution of four and a half km. This picture is of a region in the northern hemisphere near the terminator. It shows a variety of impact structures, including both razed and unrazed craters, and the odd, groove-like structures discovered by Voyager in the lighter regions. The most striking features are the <span class="hlt">bright</span> ray craters which have a distinctly 'bluer' color appearing white against the redder background. Ganymede's <span class="hlt">surface</span> is known to contain large amounts of <span class="hlt">surface</span> ice and it appears that these relatively young craters have spread <span class="hlt">bright</span> fresh ice materials over the <span class="hlt">surface</span>. Likewise, the lighter color and reflectivity of the grooved areas suggests that here, too, there is cleaner ice. We see ray craters with all sizes of ray patterns, ranging from extensive systems of the crater in the southern part of this picture, which has rays at least 300-500 kilometers long, down to craters which have only faint remnants of <span class="hlt">bright</span> ejects patterns (such as several of the craters in the southern half of PIA01516; P21262). This variation suggests that, as on the Moon, there are processes which act to darken ray material, probably 'gardening' by micrometeoroid impact. JPL manages and controls the Voyager project for NASA's Office of Space Science.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22518715-detection-ultra-faint-low-surface-brightness-dwarf-galaxies-virgo-cluster-probe-dark-matter-baryonic-physics','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22518715-detection-ultra-faint-low-surface-brightness-dwarf-galaxies-virgo-cluster-probe-dark-matter-baryonic-physics"><span>THE DETECTION OF ULTRA-FAINT LOW <span class="hlt">SURFACE</span> <span class="hlt">BRIGHTNESS</span> DWARF GALAXIES IN THE VIRGO CLUSTER: A PROBE OF DARK MATTER AND BARYONIC PHYSICS</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Giallongo, E.; Menci, N.; Grazian, A.</p> <p>2015-11-01</p> <p>We have discovered 11 ultra-faint (r ≲ 22.1) low <span class="hlt">surface</span> <span class="hlt">brightness</span> (LSB, central <span class="hlt">surface</span> <span class="hlt">brightness</span> 23 ≲ μ{sub r} ≲ 26) dwarf galaxy candidates in one deep Virgo field of just 576 arcmin{sup 2} obtained by the Large Binocular Camera at the Large Binocular Telescope. Their association with the Virgo cluster is supported by their distinct position in the central <span class="hlt">surface</span> brightness—total magnitude plane with respect to the background galaxies of similar total magnitude. They have typical absolute magnitudes and scale sizes, if at the distance of Virgo, in the range −13 ≲ M{sub r} ≲ −9 and 250 ≲more » r{sub s} ≲ 850 pc, respectively. Their colors are consistent with a gradually declining star formation history with a specific star formation rate of the order of 10{sup −11} yr{sup −1}, i.e., 10 times lower than that of main sequence star-forming galaxies. They are older than the cluster formation age and appear to be regular in morphology. They represent the faintest extremes of the population of low luminosity LSB dwarfs that has recently been detected in wider surveys of the Virgo cluster. Thanks to the depth of our <span class="hlt">observations</span>, we are able to extend the Virgo luminosity function down to M{sub r} ∼ −9.3 (corresponding to total masses M ∼ 10{sup 7} M{sub ⊙}), finding an average faint-end slope α ≃ −1.4. This relatively steep slope puts interesting constraints on the nature of the dark matter and, in particular, on warm dark matter (WDM) often invoked to solve the overprediction of the dwarf number density by the standard cold dark matter scenario. We derive a lower limit on the WDM particle mass >1.5 keV.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.3690Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.3690Z"><span><span class="hlt">Surface</span> Soil Moisture Estimates Across China Based on Multi-satellite <span class="hlt">Observations</span> and A Soil Moisture Model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Ke; Yang, Tao; Ye, Jinyin; Li, Zhijia; Yu, Zhongbo</p> <p>2017-04-01</p> <p>Soil moisture is a key variable that regulates exchanges of water and energy between land <span class="hlt">surface</span> and atmosphere. Soil moisture retrievals based on microwave satellite remote sensing have made it possible to estimate global <span class="hlt">surface</span> (up to about 10 cm in depth) soil moisture routinely. Although there are many satellites operating, including NASA's Soil Moisture Acitive Passive mission (SMAP), ESA's Soil Moisture and Ocean Salinity mission (SMOS), JAXA's Advanced Microwave Scanning Radiometer 2 mission (AMSR2), and China's Fengyun (FY) missions, key differences exist between different satellite-based soil moisture products. In this study, we applied a single-channel soil moisture retrieval model forced by multiple sources of satellite <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> to estimate consistent daily <span class="hlt">surface</span> soil moisture across China at a spatial resolution of 25 km. By utilizing <span class="hlt">observations</span> from multiple satellites, we are able to estimate daily soil moisture across the whole domain of China. We further developed a daily soil moisture accounting model and applied it to downscale the 25-km satellite-based soil moisture to 5 km. By comparing our estimated soil moisture with <span class="hlt">observations</span> from a dense <span class="hlt">observation</span> network implemented in Anhui Province, China, our estimated soil moisture results show a reasonably good agreement with the <span class="hlt">observations</span> (RMSE < 0.1 and r > 0.8).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJS..235...18L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJS..235...18L"><span>An Hα Imaging Survey of the Low-<span class="hlt">surface-brightness</span> Galaxies Selected from the Fall Sky Region of the 40% ALFALFA H I Survey</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lei, Feng-Jie; Wu, Hong; Du, Wei; Zhu, Yi-Nan; Lam, Man-I.; Zhou, Zhi-Min; He, Min; Jin, Jun-Jie; Cao, Tian-Wen; Zhao, Pin-Song; Yang, Fan; Wu, Chao-Jian; Li, Hong-Bin; Ren, Juan-Juan</p> <p>2018-03-01</p> <p>We present the <span class="hlt">observed</span> Hα flux and derived star formation rates (SFRs) for a fall sample of low-<span class="hlt">surface-brightness</span> galaxies (LSBGs). The sample is selected from the fall sky region of the 40% ALFALFA H I Survey–SDSS DR7 photometric data, and all the Hα images were obtained using the 2.16 m telescope, operated by the National Astronomy Observatories, Chinese Academy of Sciences. A total of 111 LSBGs were <span class="hlt">observed</span> and Hα flux was measured in 92 of them. Though almost all the LSBGs in our sample are H I-rich, their SFRs, derived from the extinction and filter-transmission-corrected Hα flux, are less than 1 M ⊙ yr‑1. LSBGs and star-forming galaxies have similar H I <span class="hlt">surface</span> densities, but LSBGs have much lower SFRs and SFR <span class="hlt">surface</span> densities than star-forming galaxies. Our results show that LSBGs deviate from the Kennicutt–Schmidt law significantly, which indicates that they have low star formation efficiency. The SFRs of LSBGs are close to average SFRs in Hubble time and support previous arguments that most of the LSBGs are stable systems and they tend to seldom contain strong interactions or major mergers in their star formation histories.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140006632','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140006632"><span>Three-Dimensional Structure and Evolution of Extreme-Ultraviolet <span class="hlt">Bright</span> Points <span class="hlt">Observed</span> by STEREO/SECCHI/EUVI</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kwon, Ryun Young; Chae, Jongchul; Davila, Joseph M.; Zhang, Jie; Moon, Yong-Jae; Poomvises, Watanachak; Jones, Shaela I.</p> <p>2012-01-01</p> <p>We unveil the three-dimensional structure of quiet-Sun EUV <span class="hlt">bright</span> points and their temporal evolution by applying a triangulation method to time series of images taken by SECCHI/EUVI on board the STEREO twin spacecraft. For this study we examine the heights and lengths as the components of the three-dimensional structure of EUV <span class="hlt">bright</span> points and their temporal evolutions. Among them we present three <span class="hlt">bright</span> points which show three distinct changes in the height and length: decreasing, increasing, and steady. We show that the three distinct changes are consistent with the motions (converging, diverging, and shearing, respectively) of their photospheric magnetic flux concentrations. Both growth and shrinkage of the magnetic fluxes occur during their lifetimes and they are dominant in the initial and later phases, respectively. They are all multi-temperature loop systems which have hot loops (approx. 10(exp 6.2) K) overlying cooler ones (approx 10(exp 6.0) K) with cool legs (approx 10(exp 4.9) K) during their whole evolutionary histories. Our results imply that the multi-thermal loop system is a general character of EUV <span class="hlt">bright</span> points. We conclude that EUV <span class="hlt">bright</span> points are flaring loops formed by magnetic reconnection and their geometry may represent the reconnected magnetic field lines rather than the separator field lines.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120013540','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120013540"><span>Assimilation of Smos <span class="hlt">Observations</span> to Generate a Prototype SMAP Level 4 <span class="hlt">Surface</span> and Root-Zone Soil Moisture Product</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Reichle, Rolf H.; De Lannoy, Gabrielle J. M.; Crow, Wade T.; Koster, Randal D.; Kimball, John</p> <p>2012-01-01</p> <p>The Soil Moisture Active and Passive (SMAP; [1]) mission is being implemented by NASA for launch in October 2014. The primary science objectives of SMAP are to enhance understanding of land <span class="hlt">surface</span> controls on the water, energy and carbon cycles, and to determine their linkages. Moreover, the high-resolution soil moisture mapping provided by SMAP has practical applications in weather and seasonal climate prediction, agriculture, human health, drought and flood decision support. The Soil Moisture and Ocean Salinity (SMOS; [2]) mission was launched by ESA in November 2009 and has since been <span class="hlt">observing</span> L-band (1.4 GHz) upwelling passive microwaves. In this paper we describe our use of SMOS <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> to generate a prototype of the planned SMAP Level 4 <span class="hlt">Surface</span> and Root-zone Soil Moisture (L4_SM) product [5].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA19047.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA19047.html"><span><span class="hlt">Bright</span> Feature Appears in Titan Kraken Mare</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2014-11-10</p> <p>Two Synthetic Aperture Radar (SAR) images from the radar experiment on NASA's Cassini spacecraft show that, between May 2013 and August 2014, a <span class="hlt">bright</span> feature appeared in Kraken Mare, the largest hydrocarbon sea on Saturn's moon Titan. Researchers think the <span class="hlt">bright</span> feature is likely representative of something on the hydrocarbon sea's <span class="hlt">surface</span>, such as waves or floating debris. A similar feature appeared in Ligea Mare, another Titan sea, and was seen to evolve in appearance between 2013 and 2014 (see PIA18430). The image at left was taken on May 23, 2013 at an incidence angle of 56 degrees; the image at right was taken on August 21, 2014 at an incidence angle of 5 degrees. Incidence angle refers to the angle at which the radar beam strikes the <span class="hlt">surface</span>. http://photojournal.jpl.nasa.gov/catalog/PIA19047</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940017868','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940017868"><span><span class="hlt">Surface</span>-induced <span class="hlt">brightness</span> temperature variations and their effects on detecting thin cirrus clouds using IR emission channels in the 8-12 micrometer region</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gao, Bo-Cai; Wiscombe, W. J.</p> <p>1993-01-01</p> <p>A method for detecting cirrus clouds in terms of <span class="hlt">brightness</span> temperature differences between narrow bands at 8, 11, and 12 mu m has been proposed by Ackerman et al. (1990). In this method, the variation of emissivity with wavelength for different <span class="hlt">surface</span> targets was not taken into consideration. Based on state-of-the-art laboratory measurements of reflectance spectra of terrestrial materials by Salisbury and D'Aria (1992), we have found that the <span class="hlt">brightness</span> temperature differences between the 8 and 11 mu m bands for soils, rocks and minerals, and dry vegetation can vary between approximately -8 K and +8 K due solely to <span class="hlt">surface</span> emissivity variations. We conclude that although the method of Ackerman et al. is useful for detecting cirrus clouds over areas covered by green vegetation, water, and ice, it is less effective for detecting cirrus clouds over areas covered by bare soils, rocks and minerals, and dry vegetation. In addition, we recommend that in future the variation of <span class="hlt">surface</span> emissivity with wavelength should be taken into account in algorithms for retrieving <span class="hlt">surface</span> temperatures and low-level atmospheric temperature and water vapor profiles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22047386-wave-properties-coronal-bright-fronts-observed-using-sdo-aia','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22047386-wave-properties-coronal-bright-fronts-observed-using-sdo-aia"><span>THE WAVE PROPERTIES OF CORONAL <span class="hlt">BRIGHT</span> FRONTS <span class="hlt">OBSERVED</span> USING SDO/AIA</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Long, David M.; DeLuca, Edward E.; Gallagher, Peter T., E-mail: longda@tcd.ie</p> <p>2011-11-15</p> <p>Coronal <span class="hlt">bright</span> fronts (CBFs) are large-scale wavefronts that propagate through the solar corona at hundreds of kilometers per second. While their kinematics have been studied in detail, many questions remain regarding the temporal evolution of their amplitude and pulse width. Here, contemporaneous high cadence, multi-thermal <span class="hlt">observations</span> of the solar corona from the Solar Dynamic Observatory (SDO) and Solar TErrestrial RElations Observatory (STEREO) spacecraft are used to determine the kinematics and expansion rate of a CBF wavefront <span class="hlt">observed</span> on 2010 August 14. The CBF was found to have a lower initial velocity with weaker deceleration in STEREO <span class="hlt">observations</span> compared to SDOmore » <span class="hlt">observations</span> ({approx}340 km s{sup -1} and -72 m s{sup -2} as opposed to {approx}410 km s{sup -1} and -279 m s{sup -2}). The CBF kinematics from SDO were found to be highly passband-dependent, with an initial velocity ranging from 379 {+-} 12 km s{sup -1} to 460 {+-} 28 km s{sup -1} and acceleration ranging from -128 {+-} 28 m s{sup -2} to -431 {+-} 86 m s{sup -2} in the 335 A and 304 A passbands, respectively. These kinematics were used to estimate a quiet coronal magnetic field strength range of {approx}1-2 G. Significant pulse broadening was also <span class="hlt">observed</span>, with expansion rates of {approx}130 km s{sup -1} (STEREO) and {approx}220 km s{sup -1} (SDO). By treating the CBF as a linear superposition of sinusoidal waves within a Gaussian envelope, the resulting dispersion rate of the pulse was found to be {approx}8-13 Mm{sup 2} s{sup -1}. These results are indicative of a fast-mode magnetoacoustic wave pulse propagating through an inhomogeneous medium.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.478..667P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.478..667P"><span>Automated detectionof very low <span class="hlt">surface</span> <span class="hlt">brightness</span> galaxiesin the Virgo cluster</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Prole, D. J.; Davies, J. I.; Keenan, O. C.; Davies, L. J. M.</p> <p>2018-07-01</p> <p>We report the automatic detection of a new sample of very low <span class="hlt">surface</span> <span class="hlt">brightness</span> (LSB) galaxies, likely members of the Virgo cluster. We introduce our new software, DeepScan, that has been designed specifically to detect extended LSB features automatically using the DBSCAN algorithm. We demonstrate the technique by applying it over a 5 deg2 portion of the Next Generation Virgo Survey (NGVS) data to reveal 53 LSB galaxies that are candidate cluster members based on their sizes and colours. 30 of these sources are new detections despite the region being searched specifically for LSB galaxies previously. Our final sample contains galaxies with 26.0 ≤ ⟨μe⟩ ≤ 28.5 and 19 ≤ mg ≤ 21, making them some of the faintest known in Virgo. The majority of them have colours consistent with the red sequence, and have a mean stellar mass of 106.3 ± 0.5 M⊙ assuming cluster membership. After using ProFit to fit Sérsic profiles to our detections, none of the new sources have effective radii larger than 1.5 Kpc and do not meet the criteria for ultra-diffuse galaxy (UDG) classification, so we classify them as ultra-faint dwarfs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830053990&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830053990&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour"><span><span class="hlt">Bright</span> point study. [of solar corona</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tang, F.; Harvey, K.; Bruner, M.; Kent, B.; Antonucci, E.</p> <p>1982-01-01</p> <p>Transition region and coronal <span class="hlt">observations</span> of <span class="hlt">bright</span> points by instruments aboard the Solar Maximum Mission and high resolution photospheric magnetograph <span class="hlt">observations</span> on September 11, 1980 are presented. A total of 31 bipolar ephemeral regions were found in the photosphere from birth in 9.3 hours of combined magnetograph <span class="hlt">observations</span> from three observatories. Two of the three ephemeral regions present in the field of view of the Ultraviolet Spectrometer-Polarimeter were <span class="hlt">observed</span> in the C IV 1548 line. The unobserved ephemeral region was determined to be the shortest-lived (2.5 hr) and lowest in magnetic flux density (13G) of the three regions. The Flat Crystal Spectrometer <span class="hlt">observed</span> only low level signals in the O VIII 18.969 A line, which were not statistically significant to be positively identified with any of the 16 ephemeral regions detected in the photosphere. In addition, the data indicate that at any given time there lacked a one-to-one correspondence between <span class="hlt">observable</span> <span class="hlt">bright</span> points and photospheric ephemeral regions, while more ephemeral regions were <span class="hlt">observed</span> than their counterparts in the transition region and the corona.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_11 --> <div id="page_12" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="221"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...857...39M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...857...39M"><span>Possible <span class="hlt">Bright</span> Starspots on TRAPPIST-1</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Morris, Brett M.; Agol, Eric; Davenport, James R. A.; Hawley, Suzanne L.</p> <p>2018-04-01</p> <p>The M8V star TRAPPIST-1 hosts seven roughly Earth-sized planets and is a promising target for exoplanet characterization. Kepler/K2 Campaign 12 <span class="hlt">observations</span> of TRAPPIST-1 in the optical show an apparent rotational modulation with a 3.3-day period, though that rotational signal is not readily detected in the Spitzer light curve at 4.5 μm. If the rotational modulation is due to starspots, persistent dark spots can be excluded from the lack of photometric variability in the Spitzer light curve. We construct a photometric model for rotational modulation due to photospheric <span class="hlt">bright</span> spots on TRAPPIST-1 that is consistent with both the Kepler and Spitzer light curves. The maximum-likelihood model with three spots has typical spot sizes of R spot/R ⋆ ≈ 0.004 at temperature T spot ≳ 5300 ± 200 K. We also find that large flares are <span class="hlt">observed</span> more often when the brightest spot is facing the <span class="hlt">observer</span>, suggesting a correlation between the position of the <span class="hlt">bright</span> spots and flare events. In addition, these flares may occur preferentially when the spots are increasing in <span class="hlt">brightness</span>, which suggests that the 3.3-day periodicity may not be a rotational signal, but rather a characteristic timescale of active regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017A%26A...608A.126R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...608A.126R"><span>The role of stellar radial motions in shaping galaxy <span class="hlt">surface</span> <span class="hlt">brightness</span> profiles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ruiz-Lara, T.; Few, C. G.; Florido, E.; Gibson, B. K.; Pérez, I.; Sánchez-Blázquez, P.</p> <p>2017-12-01</p> <p>Aims: The physics driving features such as breaks <span class="hlt">observed</span> in galaxy <span class="hlt">surface</span> <span class="hlt">brightness</span> (SB) profiles remains contentious. Here, we assess the importance of stellar radial motions in shaping their characteristics. Methods: We use the simulated Milky Way-mass cosmological discs from the Ramses Disc Environment Study (RaDES) to characterise the radial redistribution of stars in galaxies displaying type-I (pure exponentials), II (downbending), and III (upbending) SB profiles. We compare radial profiles of the mass fractions and the velocity dispersions of different sub-populations of stars according to their birth and current location. Results: Radial redistribution of stars is important in all galaxies regardless of their light profiles. Type-II breaks seem to be a consequence of the combined effects of outward-moving and accreted stars. The former produce shallower inner profiles (lack of stars in the inner disc) and accumulate material around the break radius and beyond, strengthening the break; the latter can weaken or even convert the break into a pure exponential. Further accretion from satellites can concentrate material in the outermost parts, leading to type-III breaks that can coexist with type-II breaks, but situated further out. Type-III galaxies would be the result of an important radial redistribution of material throughout the entire disc, as well as a concentration of accreted material in the outskirts. In addition, type-III galaxies display the most efficient radial redistribution and the largest number of accreted stars, followed by type-I and II systems, suggesting that type-I galaxies may be an intermediate case between types II and III. In general, the velocity dispersion profiles of all galaxies tend to flatten or even increase around the locations where the breaks are found. The age and metallicity profiles are also affected, exhibiting different inner gradients depending on their SB profile, being steeper in the case of type-II systems (as</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2659800','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2659800"><span>The <span class="hlt">Brightness</span> of Colour</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Corney, David; Haynes, John-Dylan; Rees, Geraint; Lotto, R. Beau</p> <p>2009-01-01</p> <p>Background The perception of <span class="hlt">brightness</span> depends on spatial context: the same stimulus can appear light or dark depending on what surrounds it. A less well-known but equally important contextual phenomenon is that the colour of a stimulus can also alter its <span class="hlt">brightness</span>. Specifically, stimuli that are more saturated (i.e. purer in colour) appear brighter than stimuli that are less saturated at the same luminance. Similarly, stimuli that are red or blue appear brighter than equiluminant yellow and green stimuli. This non-linear relationship between stimulus intensity and <span class="hlt">brightness</span>, called the Helmholtz-Kohlrausch (HK) effect, was first described in the nineteenth century but has never been explained. Here, we take advantage of the relative simplicity of this ‘illusion’ to explain it and contextual effects more generally, by using a simple Bayesian ideal <span class="hlt">observer</span> model of the human visual ecology. We also use fMRI brain scans to identify the neural correlates of <span class="hlt">brightness</span> without changing the spatial context of the stimulus, which has complicated the interpretation of related fMRI studies. Results Rather than modelling human vision directly, we use a Bayesian ideal <span class="hlt">observer</span> to model human visual ecology. We show that the HK effect is a result of encoding the non-linear statistical relationship between retinal images and natural scenes that would have been experienced by the human visual system in the past. We further show that the complexity of this relationship is due to the response functions of the cone photoreceptors, which themselves are thought to represent an efficient solution to encoding the statistics of images. Finally, we show that the locus of the response to the relationship between images and scenes lies in the primary visual cortex (V1), if not earlier in the visual system, since the <span class="hlt">brightness</span> of colours (as opposed to their luminance) accords with activity in V1 as measured with fMRI. Conclusions The data suggest that perceptions of <span class="hlt">brightness</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1714659H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1714659H"><span>Lunar <span class="hlt">Surface</span> Properties from Diviner Eclipse <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hayne, Paul; Paige, David; Greenhagen, Benjamin; Bandfield, Joshua; Siegler, Matthew; Lucey, Paul</p> <p>2015-04-01</p> <p> presence of blocky ejecta material. Comparisons of the rock abundance derived from the eclipse measurements can be made to those derived from the standard Diviner diurnal data [2] in order to constrain the rock size distribution. At a small nighttime cold spot, we <span class="hlt">observed</span> <span class="hlt">brightness</span> temperatures during the eclipse that were more than 10K higher than those <span class="hlt">observed</span> in surrounding non-cold-spot regions. This seemingly paradoxical result implies that the vertical stratigraphy of the Moon's near-<span class="hlt">surface</span> regolith may be more complex than has been previously appreciated. We are in the process of evaluating several possible explanations for this phenomenon quantitatively. References: [1] Paige D. A., et al. (2010) Space Sci. Rev. 150, 125-160. [2] Bandfield J. L., et al. (2011) J. Geophys. Res., 116, E12. [3] Ghent R. R., et al. (2014) Geology, 42 (12), 1059-1062. [4] Paige D. A., et al. (2010) Science, 330, 479-482. [5] Hayne P. O., et al. (2015) Icarus, submitted. [6] Greenhagen B. T., et al. (2010) Science, 329, 1507-1509. [7] Glotch T. D., et al. (2010) Science, 329, 1510-1513. [8] Allen C. C., et al. (2012) J. Geophys. Res., 117, E12. [9] Bandfield J. L., et al. (2014) Icarus, 231, 221-231. [10] Hayne P. O., et al. (2011) AGU Fall Meeting Abstracts, p. 1712. [11] Hayne P. O., et al. (2010) Science, 330, 477-479. Acknowledgement: Part of this work was performed at the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120002980','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120002980"><span>Investigating the Origin of <span class="hlt">Bright</span> Materials on Vesta: Synthesis, Conclusions, and Implications</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Li, Jian-Yang; Mittlefehldt, D. W.; Pieters, C. M.; De Sanctis, M. C.; Schroder, S. E.; Hiesinger, H.; Blewett, D. T.; Russell, C. T.; Raymond, C. A.; Keller, H. U.</p> <p>2012-01-01</p> <p>The Dawn spacecraft started orbiting the second largest asteroid (4) Vesta in August 2011, revealing the details of its <span class="hlt">surface</span> at an unprecedented pixel scale as small as approx.70 m in Framing Camera (FC) clear and color filter images and approx.180 m in the Visible and Infrared Spectrometer (VIR) data in its first two science orbits, the Survey Orbit and the High Altitude Mapping Orbit (HAMO) [1]. The <span class="hlt">surface</span> of Vesta displays the greatest diversity in terms of geology and mineralogy of all asteroids studied in detail [2, 3]. While the albedo of Vesta of approx.0.38 in the visible wavelengths [4, 5] is one of the highest among all asteroids, the <span class="hlt">surface</span> of Vesta shows the largest variation of albedos found on a single asteroid, with geometric albedos ranging at least from approx.0.10 to approx.0.67 in HAMO images [5]. There are many distinctively <span class="hlt">bright</span> and dark areas <span class="hlt">observed</span> on Vesta, associated with various geological features and showing remarkably different forms. Here we report our initial attempt to understand the origin of the areas that are distinctively brighter than their surroundings. The dark materials on Vesta clearly are different in origin from <span class="hlt">bright</span> materials and are reported in a companion paper [6].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss024e009404.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss024e009404.html"><span>Earth <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2010-07-23</p> <p>ISS024-E-009404 (23 July 2010) --- This photo taken from the International Space Station on July 23, 2010, shows the Gulf of Mexico oil spill as part of ongoing <span class="hlt">observations</span> of the region. When this image was taken, three months after the explosion on the Deepwater Horizon oil rig, the leak had been plugged for eight days. Water <span class="hlt">surfaces</span> appear <span class="hlt">bright</span> and land <span class="hlt">surfaces</span> appear dark in the image. The stark contrast is due to sun glint, in which the sun is reflected brilliantly off all water <span class="hlt">surfaces</span> back towards the astronaut <span class="hlt">observer</span> on board the station. The sun glint reveals various features in the Gulf of Mexico, especially sheens of oil as packets of long <span class="hlt">bright</span> streaks seen on the left side of the image. Sediments carried by the Mississippi River have a light-yellow coloration in this image, with distinct margins between plumes that likely mark tidal pulses of river water into the Gulf of Mexico. A boat wake cuts across one of the oil packets at image lower left. Daily National Oceanic and Atmospheric Administration (NOAA) maps of oil distribution show predicted heavier and lighter oil movement near the Gulf coastline. The maps show that on the day this image was taken, the north edge of the ?oiled? zone was expected to bank up against the delta. The <span class="hlt">observed</span> spread of the <span class="hlt">surface</span> oil in the approximately 100 days since the explosion highlights the connectivity between the deepwater areas and coastlines of the Gulf of Mexico.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.473.1751B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.473.1751B"><span>The Herschel <span class="hlt">Bright</span> Sources (HerBS): sample definition and SCUBA-2 <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bakx, Tom J. L. C.; Eales, S. A.; Negrello, M.; Smith, M. W. L.; Valiante, E.; Holland, W. S.; Baes, M.; Bourne, N.; Clements, D. L.; Dannerbauer, H.; De Zotti, G.; Dunne, L.; Dye, S.; Furlanetto, C.; Ivison, R. J.; Maddox, S.; Marchetti, L.; Michałowski, M. J.; Omont, A.; Oteo, I.; Wardlow, J. L.; van der Werf, P.; Yang, C.</p> <p>2018-01-01</p> <p>We present the Herschel <span class="hlt">Bright</span> Sources (HerBS) sample, a sample of <span class="hlt">bright</span>, high-redshift Herschel sources detected in the 616.4 deg2 Herschel Astrophysical Terahertz Large Area Survey. The HerBS sample contains 209 galaxies, selected with a 500 μm flux density greater than 80 mJy and an estimated redshift greater than 2. The sample consists of a combination of hyperluminous infrared galaxies and lensed ultraluminous infrared galaxies during the epoch of peak cosmic star formation. In this paper, we present Submillimetre Common-User Bolometer Array 2 (SCUBA-2) <span class="hlt">observations</span> at 850 μm of 189 galaxies of the HerBS sample, 152 of these sources were detected. We fit a spectral template to the Herschel-Spectral and Photometric Imaging Receiver (SPIRE) and 850 μm SCUBA-2 flux densities of 22 sources with spectroscopically determined redshifts, using a two-component modified blackbody spectrum as a template. We find a cold- and hot-dust temperature of 21.29_{-1.66}^{+1.35} and 45.80_{-3.48}^{+2.88} K, a cold-to-hot dust mass ratio of 26.62_{-6.74}^{+5.61} and a β of 1.83_{-0.28}^{+0.14}. The poor quality of the fit suggests that the sample of galaxies is too diverse to be explained by our simple model. Comparison of our sample to a galaxy evolution model indicates that the fraction of lenses are high. Out of the 152 SCUBA-2 detected galaxies, the model predicts 128.4 ± 2.1 of those galaxies to be lensed (84.5 per cent). The SPIRE 500 μm flux suggests that out of all 209 HerBS sources, we expect 158.1 ± 1.7 lensed sources, giving a total lensing fraction of 76 per cent.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001A%26A...369..736Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001A%26A...369..736Y"><span><span class="hlt">Surface-brightness</span> profiles of dwarf galaxies in the NGC 5044 Group: Implications for the luminosity-shape and scalelength-shape relationships as distance indicators</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Young, C. K.; Currie, M. J.</p> <p>2001-04-01</p> <p>In a recent paper, which presents CCD photometry for fifteen dwarf and intermediate early-type galaxies in the NGC 5044 Group, it has been claimed that ``a few relatively <span class="hlt">bright</span> galaxies with ``convex'' profiles destroy the known relation between total magnitude and the ``shape'' parameter... thus ruling out the use of this relation as a distance indicator for individual galaxies''. In the same paper, further reasons were cited supposedly ``limiting also its use as a distance indicator for groups of galaxies''. We demonstrate that none of the three relatively <span class="hlt">bright</span> galaxies cited as possessing ``convex'' profiles actually has a convex profile, and that one of these objects should be excluded because it is a late-type galaxy. Of the two remaining objects, one has an anomalous profile shape whilst the other is brighter than one might expect from its colour alone. However, we show that all of the other issues raised have already been accounted for by Young & Currie (\\cite{you94}, \\cite{you95} & \\cite{you98}). The main implications of the new <span class="hlt">observations</span> are: (1) that the case of one galaxy with an anomalous profile shape, N42, highlights the need for some a priori criteria to be defined in order to establish objectively which objects are not suitable for distance determinations; and (2) on the basis of another unusual galaxy, N50, colour has now been shown to be a poorer discriminant between objects of the same profile shape and scalelength (but of different central <span class="hlt">surface</span> <span class="hlt">brightness</span>) than previously thought. How significant this latter problem is depends on how common N50-like objects are. This consideration reinforces the case for always using the more general scalelength-shape relationship of Young & Currie (\\cite{you95}) in preference to the luminosity-shape one of Young & Currie (\\cite{you94}). Reassuringly, through a re-analysis of the same CCD photometry, we find that NGC 5044 Group galaxies <span class="hlt">observe</span> a tight scalelength-shape relationship. This finding</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2000fuse.prop.P187M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000fuse.prop.P187M"><span>Pulsar and CV <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Malina, R. F.</p> <p></p> <p>PSR_0656+14: Measurement of <span class="hlt">surface</span> thermal emission from neutron stars (NS) is essential to theories regarding the condensed matter state equation, the thermal evolution of NS, and of NS atmospheres. We propose to conduct 50 Ang band FUV photometric <span class="hlt">observations</span> of PSR B0656+14, an X-ray, SXR and EUV <span class="hlt">bright</span> isolated NS with an optical counterpart. FUV photometry will provide critical characterization of the NS's <span class="hlt">surface</span> thermal radiation. Higher energy <span class="hlt">observations</span> may be effected by poorly established effects including magnetized atmospheres, chemical compositions, temperature gradients and gravitational effects. Optical <span class="hlt">observations</span> may be subject to non-thermal effects. V3885 Sgr: V3885 Sgr is one of the brightest nonmagnetic cataclysmic variables. We propose to <span class="hlt">observe</span> V3885 Sgr for 5 to 6 contiguous FUSE orbits, achieving a S/N of about 12 at full resolution even at the troughs of the source's O VI absorption lines in each spectrum (assuming 2000 sec visibility per orbit). The primary purpose of the <span class="hlt">observations</span> is to use the source as a <span class="hlt">bright</span> continuum against which to study local interstellar absorption lines. Although <span class="hlt">observed</span> on Malina's Co-I Program, the data will be analyzed in collaboration with members of the O VI Project.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008sptz.prop50253M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008sptz.prop50253M"><span>IR <span class="hlt">Observations</span> of a Complete Unbiased Sample of <span class="hlt">Bright</span> Seyfert Galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Malkan, Matthew; Bendo, George; Charmandaris, Vassilis; Smith, Howard; Spinoglio, Luigi; Tommasin, Silvia</p> <p>2008-03-01</p> <p>IR spectra will measure the 2 main energy-generating processes by which galactic nuclei shine: black hole accretion and star formation. Both of these play roles in galaxy evolution, and they appear connected. To obtain a complete sample of AGN, covering the range of luminosities and column-densities, we will combine 2 complete all-sky samples with complementary selections, minimally biased by dust obscuration: the 116 IRAS 12um AGN and the 41 Swift/BAT hard Xray AGN. These galaxies have been extensively studied across the entire EM spectrum. Herschel <span class="hlt">observations</span> have been requested and will be synergistic with the Spitzer database. IRAC and MIPS imaging will allow us to separate the nuclear and galactic continua. We are completing full IR <span class="hlt">observations</span> of the local AGN population, most of which have already been done. The only remaining <span class="hlt">observations</span> we request are 10 IRS/HIRES, 57 MIPS-24 and 30 IRAC pointings. These high-quality <span class="hlt">observations</span> of <span class="hlt">bright</span> AGN in the bolometric-flux-limited samples should be completed, for the high legacy value of complete uniform datasets. We will measure quantitatively the emission at each wavelength arising from stars and from accretion in each galactic center. Since our complete samples come from flux-limited all-sky surveys in the IR and HX, we will calculate the bi-variate AGN and star formation Luminosity Functions for the local population of active galaxies, for comparison with higher redshifts.Our second aim is to understand the physical differences between AGN classes. This requires statistical comparisons of full multiwavelength <span class="hlt">observations</span> of complete representative samples. If the difference between Sy1s and Sy2s is caused by orientation, their isotropic properties, including those of the surrounding galactic centers, should be similar. In contrast, if they are different evolutionary stages following a galaxy encounter, then we may find <span class="hlt">observational</span> evidence that the circumnuclear ISM of Sy2s is relatively younger.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29215851','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29215851"><span>[Monitoring of <span class="hlt">brightness</span> temperature fluctuation of water in SHF range].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ivanov, Yu D; Kozlov, A F; Galiullin, R A; Tatu, V Yu; Vesnin, S G; Ziborov, V S; Ivanova, N D; Pleshakova, T O</p> <p>2017-01-01</p> <p>The purpose of the research consisted in detection of fluctuation of <span class="hlt">brightness</span> temperature (TSHF) of water in the area of the temperature Т = 42°С (that is critical for human) during its evaporation by SHF radiometry. Methods: Monitoring of the changes in <span class="hlt">brightness</span> temperature of water in superhigh frequency (SHF) range (3.8-4.2 GHz) near the phase transition temperature of water Т = 42°С during its evaporation in the cone dielectric cell. The <span class="hlt">brightness</span> temperature measurements were carried out using radiometer. Results: Fluctuation with maximum of <span class="hlt">brightness</span> temperature was detected in 3.8-4.2 GHz frequency range near at the temperature of water Т = 42°С. It was characteristic for these TSHF fluctuations that <span class="hlt">brightness</span> temperature rise time in this range of frequencies in ~4°С temperature range with 0.05-15°С/min gradient and a sharp decrease during 10 s connected with measuring vapor conditions. Then nonintensive fluctuation series was <span class="hlt">observed</span>. At that, the environment temperature remained constant. Conclusion: The significant increasing in <span class="hlt">brightness</span> temperature of water during its evaporation in SHF range near the temperature of Т ~42°С were detected. It was shown that for water, ТSHF pull with the amplitude DТSHF ~4°C are <span class="hlt">observed</span>. At the same time, thermodynamic temperature virtually does not change. The <span class="hlt">observed</span> effects can be used in the development of the systems for diadnostics of pathologies in human and analytical system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16149755','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16149755"><span>[<span class="hlt">Observations</span> of tolerance of <span class="hlt">bright</span> light treatment in psychiatry].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Krzystanek, Marek; Krupka-Matuszczyk, Irena; Bargiel-Matusiewicz, Kamilla</p> <p>2005-01-01</p> <p><span class="hlt">Bright</span> light (BL) treatment is a new biological treatment used in psychiatry. The probable mechanisms of action of BL treatment are synchronisation of biological rhythms and increase of serotonin transmission in the human brain. The main indication for BL treatment is seasonal affective disorder (SAD). Indications, tolerance and mechanism of action of BL treatment are still under exploration. To present 3 years of experience from the treatment of different psychiatric disorders with BL. The examined group consisted of 104 out-patients with different diagnoses. The mean age was 41.1 and the mean number of sessions of BL treatment was 17.2. Besides-of BL treatment (1 hour, 5000 lux) the patients were treated with psychotropic drugs. Side effects and BL tolerance were <span class="hlt">observed</span>. Side effects were present in 34 (32.6%) patients. They were: tearsing (11.5%), headaches (6.7%), restlessness and agitation (5.7%), eyeball pain (3.8%) and eye burning (4.8%). Tearsing and eyeball pain subsided in the first 15 minutes, the other symptoms subsided by 1 hour after a session. Six patients discontinued the BL treatment due to intolerance of a side effect. BL treatment is a safe and well-tolerated form of biological treatment in psychiatry. The absence of a control group limits the specificity of these side effects. New indications for BL treatment may include psychiatric disorders with brain serotoninergic system or biological rhythms disturbances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28612080','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28612080"><span>Color and emotion: effects of hue, saturation, and <span class="hlt">brightness</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wilms, Lisa; Oberfeld, Daniel</p> <p>2017-06-13</p> <p>Previous studies on emotional effects of color often failed to control all the three perceptual dimensions of color: hue, saturation, and <span class="hlt">brightness</span>. Here, we presented a three-dimensional space of chromatic colors by independently varying hue (blue, green, red), saturation (low, medium, high), and <span class="hlt">brightness</span> (dark, medium, <span class="hlt">bright</span>) in a factorial design. The 27 chromatic colors, plus 3 <span class="hlt">brightness</span>-matched achromatic colors, were presented via an LED display. Participants (N = 62) viewed each color for 30 s and then rated their current emotional state (valence and arousal). Skin conductance and heart rate were measured continuously. The emotion ratings showed that saturated and <span class="hlt">bright</span> colors were associated with higher arousal. The hue also had a significant effect on arousal, which increased from blue and green to red. The ratings of valence were the highest for saturated and <span class="hlt">bright</span> colors, and also depended on the hue. Several interaction effects of the three color dimensions were <span class="hlt">observed</span> for both arousal and valence. For instance, the valence ratings were higher for blue than for the remaining hues, but only for highly saturated colors. Saturated and <span class="hlt">bright</span> colors caused significantly stronger skin conductance responses. Achromatic colors resulted in a short-term deceleration in the heart rate, while chromatic colors caused an acceleration. The results confirm that color stimuli have effects on the emotional state of the <span class="hlt">observer</span>. These effects are not only determined by the hue of a color, as is often assumed, but by all the three color dimensions as well as their interactions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018eMetN...3...28G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018eMetN...3...28G"><span>Leonids 2017 from Norway – A <span class="hlt">bright</span> surprise!</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gaarder, K.</p> <p>2018-01-01</p> <p>I am very pleased to have been able to <span class="hlt">observe</span> near maximum activity of the Leonids, and clearly witnessed the unequal mass distribution during these hours. A lot of <span class="hlt">bright</span> Leonids were seen, followed by a short period of high activity of fainter meteors, before a sharp drop in activity. The Leonids is undoubtedly a shower to watch closely, with its many variations in activity level and magnitude distribution. I already look forward to <span class="hlt">observing</span> the next years’ display, hopefully under a dark and clear sky, filled with <span class="hlt">bright</span> meteors!</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930036499&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930036499&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour"><span><span class="hlt">Observations</span> of the variability of coronal <span class="hlt">bright</span> points by the Soft X-ray Telescope on Yohkoh</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Strong, Keith T.; Harvey, Karen; Hirayama, Tadashi; Nitta, Nariaki; Shimizu, Toshifumi; Tsuneta, Saku</p> <p>1992-01-01</p> <p>We present the initial results of a study of X-ray <span class="hlt">bright</span> points (XBPs) made with data from the Yohkoh Soft X-ray Telescope. High temporal and spatial resolution <span class="hlt">observations</span> of several XBPs illustrate their intensity variability over a wide variety of time scales from a few minutes to hours, as well as rapid changes in their morphology. Several XBPs produced flares during their lifetime. These XBP flares often involve magnetic loops, which are considerably larger than the XBP itself, and which brighten along their lengths at speeds of up to 1100 km/s.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1995yCat.7016....0D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1995yCat.7016....0D"><span>VizieR Online Data Catalog: Reference Catalogue of <span class="hlt">Bright</span> Galaxies (RC1; de Vaucouleurs+ 1964)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>de Vaucouleurs, G.; de Vaucouleurs, A.</p> <p>1995-11-01</p> <p>The Reference Catalogue of <span class="hlt">Bright</span> Galaxies lists for each entry the following information: NGC number, IC number, or A number; A, B, or C designation; B1950.0 positions, position at 100 year precession; galactic and supergalactic positions; revised morphological type and source; type and color class in Yerkes list 1 and 2; Hubble-Sandage type; revised Hubble type according to Holmberg; logarithm of mean major diameter (log D) and ratio of major to minor diameter (log R) and their weights; logarithm of major diameter; sources of the diameters; David Dunlap Observatory type and luminosity class; Harvard photographic apparent magnitude; weight of V, B-V(0), U-B(0); integrated magnitude B(0) and its weight in the B system; mean <span class="hlt">surface</span> <span class="hlt">brightness</span> in magnitude per square minute of arc and sources for the B magnitude; mean B <span class="hlt">surface</span> <span class="hlt">brightness</span> derived from corrected Harvard magnitude; the integrated color index in the standard B-V system; "intrinsic" color index; sources of B-V and/or U-B; integrated color in the standard U-B system; <span class="hlt">observed</span> radial velocity in km/sec; radial velocity corrected for solar motion in km/sec; sources of radial velocities; solar motion correction; and direct photographic source. The catalog was created by concatenating four files side by side. (1 data file).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018A%26A...609A..53C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018A%26A...609A..53C"><span>Tracing the stellar component of low <span class="hlt">surface</span> <span class="hlt">brightness</span> Milky Way dwarf galaxies to their outskirts. I. Sextans</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cicuéndez, L.; Battaglia, G.; Irwin, M.; Bermejo-Climent, J. R.; McMonigal, B.; Bate, N. F.; Lewis, G. F.; Conn, A. R.; de Boer, T. J. L.; Gallart, C.; Guglielmo, M.; Ibata, R.; McConnachie, A.; Tolstoy, E.; Fernando, N.</p> <p>2018-01-01</p> <p>Aims: We present results from deep and very spatially extended CTIO/DECam g and r photometry (reaching out to 2 mag below the oldest main-sequence turn-off and covering 20 deg2) around the Sextans dwarf spheroidal galaxy. We aim to use this dataset to study the structural properties of Sextans overall stellar population and its member stars in different evolutionary phases, as well as to search for possible signs of tidal disturbance from the Milky Way, which would indicate departure from dynamical equilibrium. Methods: We performed the most accurate and quantitative structural analysis to-date of Sextans' stellar components by applying Bayesian Monte Carlo Markov chain methods to the individual stars' positions. <span class="hlt">Surface</span> density maps are built by statistically decontaminating the sample through a matched filter analysis of the colour-magnitude diagram, and then analysed for departures from axisymmetry. Results: Sextans is found to be significantly less spatially extended and more centrally concentrated than early studies suggested. No statistically significant distortions or signs of tidal disturbances were found down to a <span class="hlt">surface</span> <span class="hlt">brightness</span> limit of 31.8 mag/arcsec2 in V-band. We identify an overdensity in the central regions that may correspond to previously reported kinematic substructure(s). In agreement with previous findings, old and metal-poor stars such as Blue Horizontal Branch stars cover a much larger area than stars in other evolutionary phases, and <span class="hlt">bright</span> Blue Stragglers (BSs) are less spatially extended than faint ones. However, the different spatial distribution of <span class="hlt">bright</span> and faint BSs appears consistent with the general age and metallicity gradients found in Sextans' stellar component. This is compatible with Sextans BSs having formed by evolution of binaries and not necessarily due to the presence of a central disrupted globular cluster, as suggested in the literature. We provide structural parameters for the various populations analysed and make</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19990023305','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990023305"><span>Microwave <span class="hlt">Brightness</span> Temperatures of Tilted Convective Systems</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hong, Ye; Haferman, Jeffrey L.; Olson, William S.; Kummerow, Christian D.</p> <p>1998-01-01</p> <p>Aircraft and ground-based radar data from the Tropical Ocean and Global Atmosphere Coupled-Ocean Atmosphere Response Experiment (TOGA COARE) show that convective systems are not always vertical. Instead, many are tilted from vertical. Satellite passive microwave radiometers <span class="hlt">observe</span> the atmosphere at a viewing angle. For example, the Special Sensor Microwave/Imager (SSM/I) on Defense Meteorological Satellite Program (DMSP) satellites and the Tropical Rainfall Measurement Mission (TRMM) Microwave Imager (TMI) on the TRMM satellite have an incident angle of about 50deg. Thus, the <span class="hlt">brightness</span> temperature measured from one direction of tilt may be different than that viewed from the opposite direction due to the different optical depth. This paper presents the investigation of passive microwave <span class="hlt">brightness</span> temperatures of tilted convective systems. To account for the effect of tilt, a 3-D backward Monte Carlo radiative transfer model has been applied to a simple tilted cloud model and a dynamically evolving cloud model to derive the <span class="hlt">brightness</span> temperature. The radiative transfer results indicate that <span class="hlt">brightness</span> temperature varies when the viewing angle changes because of the different optical depth. The tilt increases the displacements between high 19 GHz <span class="hlt">brightness</span> temperature (Tb(sub 19)) due to liquid emission from lower level of cloud and the low 85 GHz <span class="hlt">brightness</span> temperature (Tb(sub 85)) due to ice scattering from upper level of cloud. As the resolution degrades, the difference of <span class="hlt">brightness</span> temperature due to the change of viewing angle decreases dramatically. The dislocation between Tb(sub 19) and Tb(sub 85), however, remains prominent.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003ChPhy..12.1124L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003ChPhy..12.1124L"><span>Self-deflection of a <span class="hlt">bright</span> soliton in a separate <span class="hlt">bright</span>-dark spatial soliton pair based on a higher-order space charge field</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Jin-Song; Hao, Zhong-Hua</p> <p>2003-10-01</p> <p>The self-deflection of a <span class="hlt">bright</span> solitary beam can be controlled by a dark solitary beam via a parametric coupling effect between the <span class="hlt">bright</span> and dark solitary beams in a separate <span class="hlt">bright</span>-dark spatial soliton pair supported by an unbiased series photorefractive crystal circuit. The spatial shift of the <span class="hlt">bright</span> solitary beam centre as a function of the input intensity of the dark solitary beam (hat rho) is investigated by taking into account the higher-order space charge field in the dynamics of the <span class="hlt">bright</span> solitary beam via both numerical and perturbation methods under steady-state conditions. The deflection amount (Deltas0), defined as the value of the spatial shift at the output <span class="hlt">surface</span> of the crystal, is a monotonic and nonlinear function of hat rho. When hat rho is weak or strong enough, Deltas0 is, in fact, unchanged with hat rho, whereas Deltas0 increases or decreases monotonically with hat rho in a middle range of hat rho. The corresponding variation range (deltas) depends strongly on the value of the input intensity of the <span class="hlt">bright</span> solitary beam (r). There are some peak and valley values in the curve of deltas versus r under some conditions. When hat rho increases, the <span class="hlt">bright</span> solitary beam can scan toward both the direction same as and opposite to the crystal's c-axis. Whether the direction is the same as or opposite to the c-axis depends on the parameter values and configuration of the crystal circuit, as well as the value of r. Some potential applications are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.4928B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.4928B"><span>Tropical Pacific sea <span class="hlt">surface</span> salinity variability derived from SMOS data: Comparison with in-situ <span class="hlt">observations</span>.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ballabrera, Joaquim; Hoareau, Nina; Umbert, Marta; Martínez, Justino; Turiel, Antonio</p> <p>2013-04-01</p> <p>Prediction of El Niño/Southern Oscillation (ENSO), and its relation with global climate anomalies, continues to be an important research effort in short-term climate forecasting. This task has become even more challenging as researchers are becoming more and more convinced that there is not a single archetypical El Niño (or La Niña) pattern, but several. During some events (called now Standard or East Pacific), the largest temperature anomalies are located at the eastern part of the Pacific. However, during some of the most recent events, the largest anomalies are restricted to the central part of the Pacific Ocean, and are now called Central Pacific or Modoki (a Japanese word for "almost") events. Although the role of salinity in operational ENSO forecasting was initially neglected (in contrast with temperature, sea level, or <span class="hlt">surface</span> winds), recent studies have shown that salinity does play a role in the preconditioning of ENSO. Moreover, some researchers suggest that sea <span class="hlt">surface</span> salinity might play a role (through the modulation of the western Pacific barrier layer) to favor the Standard or the Modoki nature of each event. Sea <span class="hlt">Surface</span> Salinity maps are being operationally generated from microwave (L-band, 1.4 Ghz) <span class="hlt">brightness</span> temperature maps. The L-band frequency was chosen because is the optimal one for ocean salinity measurements. However, after three years of satellite data, it has been found that noise in <span class="hlt">brightness</span> temperatures (due to natural and artificial sources) is larger than expected. Moreover, the retrieval of SSS information requires special care because of the low sensitivity of the <span class="hlt">brightness</span> temperature to SSS: from 0.2-0.8 K per salinity unit. Despite of all these facts, current accuracy of SS maps ranges from 0.2-0.4, depending on the processing level and the region being considered. We present here our study about the salinity variability in the tropical Pacific Ocean from the 9-day, 0.25 bins salinity maps derived from the SMOS</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_12 --> <div id="page_13" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="241"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040171647','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040171647"><span>X-ray <span class="hlt">Observations</span> of the <span class="hlt">Bright</span> Old Nova V603 Aquilae</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mukai, K.; Orio, M.</p> <p>2004-01-01</p> <p>We report on our Chandra and RXTE <span class="hlt">observations</span> of the <span class="hlt">bright</span> old nova, V603 Aql, performed in 2001 April, supplemented by our analysis of archival X-ray data on this object. We find that the RXTE data are contaminated by the Galactic Ridge X-ray emission. After accounting for this effect, we find a high level of aperiodic variability in the RXTE data, at a level consistent with the uncontaminated Chandra data. The Chandra HETG spectrum clearly originates in a multi-temperature plasma. We constrain the possible emission measure distribution of the plasma through a combination of global and local fits. The X-ray luminosity and the spectral shape of V603 Aql resemble those of SS Cyg in transition between quiescence and outburst. The fact that the X-ray flux variability is only weakly energy dependent can be interpreted by supposing that the variability is due to changes in the maximum temperature of the plasma. The plasma density is likely to be high, and the emission region is likely to be compact. Finally, the apparent overabundance of Ne is consistent with V603 Aql being a young system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SSRv..212.1511G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SSRv..212.1511G"><span>Venus <span class="hlt">Surface</span> Composition Constrained by <span class="hlt">Observation</span> and Experiment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gilmore, Martha; Treiman, Allan; Helbert, Jörn; Smrekar, Suzanne</p> <p>2017-11-01</p> <p>New <span class="hlt">observations</span> from the Venus Express spacecraft as well as theoretical and experimental investigation of Venus analogue materials have advanced our understanding of the petrology of Venus melts and the mineralogy of rocks on the <span class="hlt">surface</span>. The VIRTIS instrument aboard Venus Express provided a map of the southern hemisphere of Venus at ˜1 μm allowing, for the first time, the definition of <span class="hlt">surface</span> units in terms of their 1 μm emissivity and derived mineralogy. Tessera terrain has lower emissivity than the presumably basaltic plains, consistent with a more silica-rich or felsic mineralogy. Thermodynamic modeling and experimental production of melts with Venera and Vega starting compositions predict derivative melts that range from mafic to felsic. Large volumes of felsic melts require water and may link the formation of tesserae to the presence of a Venus ocean. Low emissivity rocks may also be produced by atmosphere-<span class="hlt">surface</span> weathering reactions unlike those seen presently. High 1 μm emissivity values correlate to stratigraphically recent flows and have been used with theoretical and experimental predictions of basalt weathering to identify regions of recent volcanism. The timescale of this volcanism is currently constrained by the weathering of magnetite (higher emissivity) in fresh basalts to hematite (lower emissivity) in Venus' oxidizing environment. Recent volcanism is corroborated by transient thermal anomalies identified by the VMC instrument aboard Venus Express. The interpretation of all emissivity data depends critically on understanding the composition of <span class="hlt">surface</span> materials, kinetics of rock weathering and their measurement under Venus conditions. Extended theoretical studies, continued analysis of earlier spacecraft results, new atmospheric data, and measurements of mineral stability under Venus conditions have improved our understanding atmosphere-<span class="hlt">surface</span> interactions. The calcite-wollastonite CO2 buffer has been discounted due, among other things, to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20000080169&hterms=sensitivity+scale&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dsensitivity%2Bscale','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20000080169&hterms=sensitivity+scale&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dsensitivity%2Bscale"><span>Sensitivity of Spacebased Microwave Radiometer <span class="hlt">Observations</span> to Ocean <span class="hlt">Surface</span> Evaporation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Liu, Timothy W.; Li, Li</p> <p>2000-01-01</p> <p>Ocean <span class="hlt">surface</span> evaporation and the latent heat it carries are the major components of the hydrologic and thermal forcing on the global oceans. However, there is practically no direct in situ measurements. Evaporation estimated from bulk parameterization methods depends on the quality and distribution of volunteer-ship reports which are far less than satisfactory. The only way to monitor evaporation with sufficient temporal and spatial resolutions to study global environment changes is by spaceborne sensors. The estimation of seasonal-to-interannual variation of ocean evaporation, using spacebased measurements of wind speed, sea <span class="hlt">surface</span> temperature (SST), and integrated water vapor, through bulk parameterization method,s was achieved with reasonable success over most of the global ocean, in the past decade. Because all the three geophysical parameters can be retrieved from the radiance at the frequencies measured by the Scanning Multichannel Microwave Radiometer (SMMR) on Nimbus-7, the feasibility of retrieving evaporation directly from the measured radiance was suggested and demonstrated using coincident <span class="hlt">brightness</span> temperatures <span class="hlt">observed</span> by SMMR and latent heat flux computed from ship data, in the monthly time scale. However, the operational microwave radiometers that followed SMMR, the Special Sensor Microwave/Imager (SSM/I), lack the low frequency channels which are sensitive to SST. This low frequency channels are again included in the microwave imager (TMI) of the recently launched Tropical Rain Measuring Mission (TRMM). The radiance at the frequencies <span class="hlt">observed</span> by both TMI and SSM/I were simulated through an atmospheric radiative transfer model using ocean <span class="hlt">surface</span> parameters and atmospheric temperature and humidity profiles produced by the reanalysis of the European Center for Medium Range Weather Forecast (ECMWF). From the same ECMWF data set, coincident evaporation is computed using a <span class="hlt">surface</span> layer turbulent transfer model. The sensitivity of the radiance to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018A%26A...611A..90B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018A%26A...611A..90B"><span>Constraints on <span class="hlt">observing</span> <span class="hlt">brightness</span> asymmetries in protoplanetary disks at solar system scale</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brunngräber, Robert; Wolf, Sebastian</p> <p>2018-04-01</p> <p>We have quantified the potential capabilities of detecting local <span class="hlt">brightness</span> asymmetries in circumstellar disks with the Very Large Telescope Interferometer (VLTI) in the mid-infrared wavelength range. The study is motivated by the need to evaluate theoretical models of planet formation by direct <span class="hlt">observations</span> of protoplanets at early evolutionary stages, when they are still embedded in their host disk. Up to now, only a few embedded candidate protoplanets have been detected with semi-major axes of 20-50 au. Due to the small angular separation from their central star, only long-baseline interferometry provides the angular resolving power to detect disk asymmetries associated to protoplanets at solar system scales in nearby star-forming regions. In particular, infrared <span class="hlt">observations</span> are crucial to <span class="hlt">observe</span> scattered stellar radiation and thermal re-emission in the vicinity of embedded companions directly. For this purpose we performed radiative transfer simulations to calculate the thermal re-emission and scattered stellar flux from a protoplanetary disk hosting an embedded companion. Based on that, visibilities and closure phases are calculated to simulate <span class="hlt">observations</span> with the future beam combiner MATISSE, operating at the L, M and N bands at the VLTI. We find that the flux ratio of the embedded source to the central star can be as low as 0.5 to 0.6% for a detection at a feasible significance level due to the heated dust in the vicinity of the embedded source. Furthermore, we find that the likelihood for detection is highest for sources at intermediate distances r ≈ 2-5 au and disk masses not higher than ≈10-4 M⊙.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27103935','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27103935"><span>Circadian Phase-Shifting Effects of <span class="hlt">Bright</span> Light, Exercise, and <span class="hlt">Bright</span> Light + Exercise.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Youngstedt, Shawn D; Kline, Christopher E; Elliott, Jeffrey A; Zielinski, Mark R; Devlin, Tina M; Moore, Teresa A</p> <p>2016-02-26</p> <p>Limited research has compared the circadian phase-shifting effects of <span class="hlt">bright</span> light and exercise and additive effects of these stimuli. The aim of this study was to compare the phase-delaying effects of late night <span class="hlt">bright</span> light, late night exercise, and late evening <span class="hlt">bright</span> light followed by early morning exercise. In a within-subjects, counterbalanced design, 6 young adults completed each of three 2.5-day protocols. Participants followed a 3-h ultra-short sleep-wake cycle, involving wakefulness in dim light for 2h, followed by attempted sleep in darkness for 1 h, repeated throughout each protocol. On night 2 of each protocol, participants received either (1) <span class="hlt">bright</span> light alone (5,000 lux) from 2210-2340 h, (2) treadmill exercise alone from 2210-2340 h, or (3) <span class="hlt">bright</span> light (2210-2340 h) followed by exercise from 0410-0540 h. Urine was collected every 90 min. Shifts in the 6-sulphatoxymelatonin (aMT6s) cosine acrophase from baseline to post-treatment were compared between treatments. Analyses revealed a significant additive phase-delaying effect of <span class="hlt">bright</span> light + exercise (80.8 ± 11.6 [SD] min) compared with exercise alone (47.3 ± 21.6 min), and a similar phase delay following <span class="hlt">bright</span> light alone (56.6 ± 15.2 min) and exercise alone administered for the same duration and at the same time of night. Thus, the data suggest that late night <span class="hlt">bright</span> light followed by early morning exercise can have an additive circadian phase-shifting effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NaPho..12..159L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NaPho..12..159L"><span><span class="hlt">Bright</span> colloidal quantum dot light-emitting diodes enabled by efficient chlorination</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Xiyan; Zhao, Yong-Biao; Fan, Fengjia; Levina, Larissa; Liu, Min; Quintero-Bermudez, Rafael; Gong, Xiwen; Quan, Li Na; Fan, James; Yang, Zhenyu; Hoogland, Sjoerd; Voznyy, Oleksandr; Lu, Zheng-Hong; Sargent, Edward H.</p> <p>2018-03-01</p> <p>The external quantum efficiencies of state-of-the-art colloidal quantum dot light-emitting diodes (QLEDs) are now approaching the limit set by the out-coupling efficiency. However, the <span class="hlt">brightness</span> of these devices is constrained by the use of poorly conducting emitting layers, a consequence of the present-day reliance on long-chain organic capping ligands. Here, we report how conductive and passivating halides can be implemented in Zn chalcogenide-shelled colloidal quantum dots to enable high-<span class="hlt">brightness</span> green QLEDs. We use a <span class="hlt">surface</span> management reagent, thionyl chloride (SOCl2), to chlorinate the carboxylic group of oleic acid and graft the <span class="hlt">surfaces</span> of the colloidal quantum dots with passivating chloride anions. This results in devices with an improved mobility that retain high external quantum efficiencies in the high-injection-current region and also feature a reduced turn-on voltage of 2.5 V. The treated QLEDs operate with a <span class="hlt">brightness</span> of 460,000 cd m-2, significantly exceeding that of all previously reported solution-processed LEDs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015Msngr.159...46C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015Msngr.159...46C"><span>VEGAS-SSS: A VST Programme to Study the Satellite Stellar Systems around <span class="hlt">Bright</span> Early-type Galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cantiello, M.; Capaccioli, M.; Napolitano, N.; Grado, A.; Limatola, L.; Paolillo, M.; Iodice, E.; Romanowsky, A. J.; Forbes, D. A.; Raimondo, G.; Spavone, M.; La Barbera, F.; Puzia, T. H.; Schipani, P.</p> <p>2015-03-01</p> <p>The VEGAS-SSS programme is devoted to studying the properties of small stellar systems (SSSs) in and around <span class="hlt">bright</span> galaxies, built on the VLT Survey Telescope early-type galaxy survey (VEGAS), an ongoing guaranteed time imaging survey distributed over many semesters (Principal Investigator: Capaccioli). On completion, the VEGAS survey will have collected detailed photometric information of ~ 100 <span class="hlt">bright</span> early-type galaxies to study the properties of diffuse light (<span class="hlt">surface</span> <span class="hlt">brightness</span>, colours, <span class="hlt">surface</span> <span class="hlt">brightness</span> fluctuations, etc.) and the distribution of clustered light (compact ''small'' stellar systems) out to previously unreached projected galactocentric radii. VEGAS-SSS will define an accurate and homogeneous dataset that will have an important legacy value for studies of the evolution and transformation processes taking place in galaxies through the fossil information provided by SSSs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Icar..296..289Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Icar..296..289Z"><span>Aqueous origins of <span class="hlt">bright</span> salt deposits on Ceres</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zolotov, Mikhail Yu.</p> <p>2017-11-01</p> <p><span class="hlt">Bright</span> materials have been reported in association with impact craters on Ceres. The abundant Na2CO3 and some ammonium salts, NH4HCO3 and/or NH4Cl, were detected in <span class="hlt">bright</span> deposits within Occator crater with Dawn near infrared spectroscopy. The composition and appearance of the salts suggest their aqueous mobilization and emplacement after formation of the crater. Here we consider origins of the <span class="hlt">bright</span> deposits through calculation of speciation in the H-C-N-O-Na-Cl water-salt type system constrained by the mass balance of <span class="hlt">observed</span> salts. Calculations of chemical equilibria show that initial solutions had the pH of ∼10. The temperature and salinity of solutions could have not exceeded ∼273 K and ∼100 g per kg H2O, respectively. Freezing models reveal an early precipitation of Na2CO3·10H2O followed by minor NaHCO3. Ammonium salts precipitate near eutectic from brines enriched in NH4+, Cl- and Na+. A late-stage precipitation of NaCl·2H2O is modeled for solution compositions with added NaCl. Calculated eutectics are above 247 K. The apparently unabundant ammonium and chloride salts in Occator's deposits imply a rapid emplacement without a compositional evolution of solution. Salty ice grains could have deposited from post-impact ballistic plumes formed through low-pressure boiling of subsurface solutions. Hydrated and ammonium salts are unstable at maximum temperatures of Ceres' <span class="hlt">surface</span> and could decompose through space weathering. Occator's ice-free salt deposits formed through a post-depositional sublimation of ice followed by dehydration of Na2CO3·10H2O and NaHCO3 to Na2CO3. In other regions, excavated and exposed <span class="hlt">bright</span> materials could be salts initially deposited from plumes and accumulated at depth via post-impact boiling. The lack of detection of sulfates and an elevated carbonate/chloride ratio in Ceres' materials suggest an involvement of compounds abundant in the outer solar system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1992AAS...181.4913C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1992AAS...181.4913C"><span>Three <span class="hlt">Bright</span> X-ray Sources in NGC 1313</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Colbert, E.; Petre, R.; Schlegel, E.</p> <p>1992-12-01</p> <p>Three <span class="hlt">bright</span> X-ray sources were detected in a recent (April/May 1991) ROSAT PSPC <span class="hlt">observation</span> of the nearby (D ~ 4.5 Mpc) face--on barred spiral galaxy NGC 1313. Two of the sources were at positions coincident with X-ray sources detected by Fabbiano & Trinchieri (ApJ 315, 1987) in a previous (Jan 1980) Einstein IPC <span class="hlt">observation</span>. The position of the brightest Einstein source is near the center of NGC 1313, and the second Einstein source is ~ 7' south of the ``nuclear'' source, in the outskirts of the spiral arms. A third <span class="hlt">bright</span> X-ray source was detected in the ROSAT <span class="hlt">observation</span> ~ 7' southwest of the ``nuclear'' source. We present X-ray spectra and X-ray images for the three <span class="hlt">bright</span> sources found in the ROSAT <span class="hlt">observation</span> of NGC 1313, and compare with previous Einstein results. Spectral analysis of these sources require them to have very large soft X-ray luminosities ( ~ 10(40) erg s(-1) ) when compared with typical X-ray sources in our Galaxy. Feasible explanations for the X-ray emission are presented. The third X-ray source is positively identified with the recently discovered (Ryder et. al., ApJ 1992) peculiar type-II supernova 1978K.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSH41A2754S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSH41A2754S"><span>The First ALMA <span class="hlt">Observation</span> of a Solar Plasmoid Ejection from an X-Ray <span class="hlt">Bright</span> Point</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shimojo, M.; Hudson, H. S.; White, S. M.; Bastian, T.; Iwai, K.</p> <p>2017-12-01</p> <p>Eruptive phenomena are important features of energy releases events, such solar flares, and have the potential to improve our understanding of the dynamics of the solar atmosphere. The 304 A EUV line of helium, formed at around 10^5 K, is found to be a reliable tracer of such phenomena, but the determination of physical parameters from such <span class="hlt">observations</span> is not straightforward. We have <span class="hlt">observed</span> a plasmoid ejection from an X-ray <span class="hlt">bright</span> point simultaneously with ALMA, SDO/AIA, and Hinode/XRT. This paper reports the physical parameters of the plasmoid obtained by combining the radio, EUV, and X-ray data. As a result, we conclude that the plasmoid can consist either of (approximately) isothermal ˜10^5 K plasma that is optically thin at 100 GHz, or a ˜10^4 K core with a hot envelope. The analysis demonstrates the value of the additional temperature and density constraints that ALMA provides, and future science <span class="hlt">observations</span> with ALMA will be able to match the spatial resolution of space-borne and other high-resolution telescopes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28488591','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28488591"><span>Red-emission phosphor's <span class="hlt">brightness</span> deterioration by x-ray and <span class="hlt">brightness</span> recovery phenomenon by heating.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nakamura, Masaaki; Chida, Koichi; Inaba, Yohei; Kobayashi, Ryota; Zuguchi, Masayuki</p> <p>2017-06-26</p> <p>There are no feasible real-time and direct skin dosimeters for interventional radiology. One would be available if there were x-ray phosphors that had no <span class="hlt">brightness</span> change caused by x-ray irradiation, but the emission of the Y 2 O 3 :Eu, (Y, Gd, Eu)BO 3 , and YVO 4 :Eu phosphors investigated in our previous study was reduced by x-ray irradiation. We found that the <span class="hlt">brightness</span> of those phosphors recovered, and the purpose of this study is to investigate their recovery phenomena. It is expected that more kinds of phosphors could be used in x-ray dosimeters if the <span class="hlt">brightness</span> changes caused by x-rays are elucidated and prevented. Three kinds of phosphors-Y 2 O 3 :Eu, (Y, Gd, Eu)BO 3 , and YVO 4 :Eu-were irradiated by x-rays (2 Gy) to reduce their <span class="hlt">brightness</span>. After the irradiation, <span class="hlt">brightness</span> changes occurring at room temperature and at 80 °C were investigated. The irradiation reduced the <span class="hlt">brightness</span> of all the phosphors by 5%-10%, but the <span class="hlt">brightness</span> of each recovered immediately both at room temperature and at 80 °C. The recovery at 80 °C was faster than that at room temperature, and at both temperatures the recovered <span class="hlt">brightness</span> remained at 95%-98% of the <span class="hlt">brightness</span> before the x-ray irradiation. The <span class="hlt">brightness</span> recovery phenomena of Y 2 O 3 :Eu, (Y, Gd, Eu)BO 3 , and YVO 4 :Eu phosphors occurring after <span class="hlt">brightness</span> deterioration due to x-ray irradiation were found to be more significant at 80 °C than at room temperature. More kinds of phosphors could be used in x-ray scintillation dosimeters if the reasons for the <span class="hlt">brightness</span> changes caused by x-rays were elucidated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ATel11546....1J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ATel11546....1J"><span>Fermi-LAT <span class="hlt">Bright</span> Gamma-ray Detection of Nova ASASSN-18fv</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jean, P.; Cheung, C. C.; Ojha, R.; van Zyl, P.; Angioni, R.</p> <p>2018-04-01</p> <p>The Large Area Telescope (LAT), one of two instruments on the Fermi Gamma-ray Space Telescope, has <span class="hlt">observed</span> <span class="hlt">bright</span> gamma-ray emission from a source positionally consistent with the <span class="hlt">bright</span> optical nova ASASSN-18fv (ATel #11454, #11456, #11460, #11467, #11508).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.P31D2856M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.P31D2856M"><span>Discovery of a <span class="hlt">Bright</span> Equatorial Storm on Neptune</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Molter, E. M.; De Pater, I.; Alvarez, C.; Tollefson, J.; Luszcz-Cook, S.</p> <p>2017-12-01</p> <p>Images of Neptune, taken with the NIRC2 instrument during testing of the new Twilight Zone <span class="hlt">observing</span> program at Keck Observatory, revealed an extremely large <span class="hlt">bright</span> storm system near Neptune's equator. The storm complex is ≈9,000 km across and brightened considerably between June 26 and July 2. Historically, very <span class="hlt">bright</span> clouds have occasionally been seen on Neptune, but always in the midlatitude regions between ≈15° and ≈60° North or South. Voyager and HST <span class="hlt">observations</span> have shown that cloud features large enough to dominate near-IR photometry are often "companion" clouds of dark anti-cyclonic vortices similar to Jupiter's Great Red Spot, interpreted as orographic clouds. In the past such clouds and their coincident dark vortices often persisted for one up to several years. However, the cloud complex we detect is unique: never before has a <span class="hlt">bright</span> cloud been seen at, or so close to, the equator. The discovery points to a drastic departure in the dynamics of Neptune's atmosphere from what has been <span class="hlt">observed</span> for the past several decades. Detections of the complex in multiple NIRC2 filters allows radiative transfer modeling to constrain the cloud's altitude and vertical extent.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4834751','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4834751"><span>Circadian Phase-Shifting Effects of <span class="hlt">Bright</span> Light, Exercise, and <span class="hlt">Bright</span> Light + Exercise</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Kline, Christopher E.; Elliott, Jeffrey A.; Zielinski, Mark R.; Devlin, Tina M.; Moore, Teresa A.</p> <p>2016-01-01</p> <p>Limited research has compared the circadian phase-shifting effects of <span class="hlt">bright</span> light and exercise and additive effects of these stimuli. The aim of this study was to compare the phase-delaying effects of late night <span class="hlt">bright</span> light, late night exercise, and late evening <span class="hlt">bright</span> light followed by early morning exercise. In a within-subjects, counterbalanced design, 6 young adults completed each of three 2.5-day protocols. Participants followed a 3-h ultra-short sleep-wake cycle, involving wakefulness in dim light for 2h, followed by attempted sleep in darkness for 1 h, repeated throughout each protocol. On night 2 of each protocol, participants received either (1) <span class="hlt">bright</span> light alone (5,000 lux) from 2210–2340 h, (2) treadmill exercise alone from 2210–2340 h, or (3) <span class="hlt">bright</span> light (2210–2340 h) followed by exercise from 0410–0540 h. Urine was collected every 90 min. Shifts in the 6-sulphatoxymelatonin (aMT6s) cosine acrophase from baseline to post-treatment were compared between treatments. Analyses revealed a significant additive phase-delaying effect of <span class="hlt">bright</span> light + exercise (80.8 ± 11.6 [SD] min) compared with exercise alone (47.3 ± 21.6 min), and a similar phase delay following <span class="hlt">bright</span> light alone (56.6 ± 15.2 min) and exercise alone administered for the same duration and at the same time of night. Thus, the data suggest that late night <span class="hlt">bright</span> light followed by early morning exercise can have an additive circadian phase-shifting effect. PMID:27103935</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.476.4488H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.476.4488H"><span>A study of the H I and optical properties of Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> galaxies: spirals, dwarfs, and irregulars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Honey, M.; van Driel, W.; Das, M.; Martin, J.-M.</p> <p>2018-06-01</p> <p>We present a study of the H I and optical properties of nearby (z ≤ 0.1) Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> galaxies (LSBGs). We started with a literature sample of ˜900 LSBGs and divided them into three morphological classes: spirals, irregulars, and dwarfs. Of these, we could use ˜490 LSBGs to study their H I and stellar masses, colours, and colour-magnitude diagrams, and local environment, compare them with normal, High <span class="hlt">Surface</span> <span class="hlt">Brightness</span> (HSB) galaxies and determine the differences between the three morphological classes. We found that LSB and HSB galaxies span a similar range in H I and stellar masses, and have a similar M_{H I}/M⋆-M⋆ relationship. Among the LSBGs, as expected, the spirals have the highest average H I and stellar masses, both of about 109.8 M⊙. The LSGBs' (g - r) integrated colour is nearly constant as function of H I mass for all classes. In the colour-magnitude diagram, the spirals are spread over the red and blue regions whereas the irregulars and dwarfs are confined to the blue region. The spirals also exhibit a steeper slope in the M_{H I}/M⋆-M⋆ plane. Within their local environment, we confirmed that LSBGs are more isolated than HSB galaxies, and LSB spirals more isolated than irregulars and dwarfs. Kolmogorov-Smirnov statistical tests on the H I mass, stellar mass, and number of neighbours indicate that the spirals are a statistically different population from the dwarfs and irregulars. This suggests that the spirals may have different formation and H I evolution than the dwarfs and irregulars.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/792749','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/792749"><span>Neptune and Titan <span class="hlt">Observed</span> with Keck Telescope Adaptive Optics</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Max, C.E.; Macintosh, B.A.; Gibbard, S.</p> <p>2000-05-05</p> <p>The authors report on <span class="hlt">observations</span> taken during engineering science validation time using the new adaptive optics system at the 10-m Keck II Telescope. They <span class="hlt">observe</span> Neptune and Titan at near-infrared wavelengths. These objects are ideal for adaptive optics imaging because they are <span class="hlt">bright</span> and small, yet have many diffraction-limited resolution elements across their disks. In addition Neptune and Titan have prominent physical features, some of which change markedly with time. They have <span class="hlt">observed</span> infrared-<span class="hlt">bright</span> storms on Neptune, and very low-albedo <span class="hlt">surface</span> regions on Titan, Saturn's largest moon, Spatial resolution on Neptune and Titan was 0.05-0.06 and 0.04-0.05 arc sec, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA00348&hterms=asphalt&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dasphalt','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA00348&hterms=asphalt&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dasphalt"><span>Iapetus <span class="hlt">Bright</span> and Dark Terrains</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1990-01-01</p> <p>Saturn's outermost large moon, Iapetus, has a <span class="hlt">bright</span>, heavily cratered icy terrain and a dark terrain, as shown in this Voyager 2 image taken on August 22, 1981. Amazingly, the dark material covers precisely the side of Iapetus that leads in the direction of orbital motion around Saturn (except for the poles), whereas the <span class="hlt">bright</span> material occurs on the trailing hemisphere and at the poles. The <span class="hlt">bright</span> terrain is made of dirty ice, and the dark terrain is <span class="hlt">surfaced</span> by carbonaceous molecules, according to measurements made with Earth-based telescopes. Iapetus' dark hemisphere has been likened to tar or asphalt and is so dark that no details within this terrain were visible to Voyager 2. The <span class="hlt">bright</span> icy hemisphere, likened to dirty snow, shows many large impact craters. The closest approach by Voyager 2 to Iapetus was a relatively distant 600,000 miles, so that our best images, such as this, have a resolution of about 12 miles. The dark material is made of organic substances, probably including poisonous cyano compounds such as frozen hydrogen cyanide polymers. Though we know a little about the dark terrain's chemical nature, we do not understand its origin. Two theories have been developed, but neither is fully satisfactory--(1) the dark material may be organic dust knocked off the small neighboring satellite Phoebe and 'painted' onto the leading side of Iapetus as the dust spirals toward Saturn and Iapetus hurtles through the tenuous dust cloud, or (2) the dark material may be made of icy-cold carbonaceous 'cryovolcanic' lavas that were erupted from Iapetus' interior and then blackened by solar radiation, charged particles, and cosmic rays. A determination of the actual cause, as well as discovery of any other geologic features smaller than 12 miles across, awaits the Cassini Saturn orbiter to arrive in 2004.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010037596','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010037596"><span>The Influence of Microphysical Cloud Parameterization on Microwave <span class="hlt">Brightness</span> Temperatures</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Skofronick-Jackson, Gail M.; Gasiewski, Albin J.; Wang, James R.; Zukor, Dorothy J. (Technical Monitor)</p> <p>2000-01-01</p> <p>The microphysical parameterization of clouds and rain-cells plays a central role in atmospheric forward radiative transfer models used in calculating passive microwave <span class="hlt">brightness</span> temperatures. The absorption and scattering properties of a hydrometeor-laden atmosphere are governed by particle phase, size distribution, aggregate density., shape, and dielectric constant. This study identifies the sensitivity of <span class="hlt">brightness</span> temperatures with respect to the microphysical cloud parameterization. Cloud parameterizations for wideband (6-410 GHz <span class="hlt">observations</span> of baseline <span class="hlt">brightness</span> temperatures were studied for four evolutionary stages of an oceanic convective storm using a five-phase hydrometeor model in a planar-stratified scattering-based radiative transfer model. Five other microphysical cloud parameterizations were compared to the baseline calculations to evaluate <span class="hlt">brightness</span> temperature sensitivity to gross changes in the hydrometeor size distributions and the ice-air-water ratios in the frozen or partly frozen phase. The comparison shows that, enlarging the rain drop size or adding water to the partly Frozen hydrometeor mix warms <span class="hlt">brightness</span> temperatures by up to .55 K at 6 GHz. The cooling signature caused by ice scattering intensifies with increasing ice concentrations and at higher frequencies. An additional comparison to measured Convection and Moisture LA Experiment (CAMEX 3) <span class="hlt">brightness</span> temperatures shows that in general all but, two parameterizations produce calculated T(sub B)'s that fall within the <span class="hlt">observed</span> clear-air minima and maxima. The exceptions are for parameterizations that, enhance the scattering characteristics of frozen hydrometeors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19940033949&hterms=currie&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dcurrie','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19940033949&hterms=currie&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dcurrie"><span>Planetary camera <span class="hlt">observations</span> of the double nucleus of M31</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lauer, Tod R.; Faber, S. M.; Groth, Edward J.; Shaya, Edward J.; Campbell, Bel; Code, Arthur; Currie, Douglas G.; Baum, William A.; Ewald, S. P.; Hester, J. J.</p> <p>1993-01-01</p> <p>HST Planetary Camera images obtained in the V and I band for M31 show its inner nucleus to consist of two components that are separated by 0.49 arcsec. The nuclear component with lower <span class="hlt">surface</span> <span class="hlt">brightness</span> closely coincides with the bulge photocenter and is argued to be at the kinematic center of the galaxy. It is surmised that, if dust absorption generates the asymmetric nuclear morphology <span class="hlt">observed</span>, the dust grain size must either be exceptionally large, or the dust optical depth must be extremely high; the higher <span class="hlt">surface-brightness</span> and off-center nuclear component may alternatively be a separate stellar system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990045714&hterms=astronomia+espacio&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dastronomia%2By%2Bespacio','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990045714&hterms=astronomia+espacio&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dastronomia%2By%2Bespacio"><span><span class="hlt">Bright</span> Points and Subflares in Ultraviolet Lines and X-Rays</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rovira, M.; Schmieder, B.; Demoulin, P.; Simnett, G. M.; Hagyard, M. J.; Reichmann, E.; Reichmann, E.; Tandberg-Hanssen, E.</p> <p>1999-01-01</p> <p>We have analyzed an active region which was <span class="hlt">observed</span> in H.alpha (Multichannel Subtractive Double Pass Spectrograph), in UV lines (SMM/UVSP), and in X-rays (SMM/HXIS). In this active region there were only a few subflares and many small <span class="hlt">bright</span> points visible in UV and in X-rays. Using an extrapolation based on the Fourier transform, we have computed magnetic field lines connecting different photospheric magnetic polarities from ground-based magnetograms. Along the magnetic inversion lines we find two different zones: (1) a high-shear region (> 70 deg) where subflares occur, and (2) a low-shear region along the magnetic inversion line where UV <span class="hlt">bright</span> points are <span class="hlt">observed</span>. In these latter regions the magnetic topology is complex with a mixture of polarities. According to the velocity field <span class="hlt">observed</span> in the Si IV lamda.1402 line and the extrapolation of the magnetic field, we notice that each UV <span class="hlt">bright</span> point is consistent with emission from low-rising loops with downflows at both ends. We notice some hard X-ray emissions above the <span class="hlt">bright</span>-point regions with temperatures up to 8 x 10(exp 6) K, which suggests some induced reconnection due to continuous emergence of new flux. This reconnection is also enhanced by neighboring subflares.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_13 --> <div id="page_14" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="261"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016A%26A...589A.114L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016A%26A...589A.114L"><span>Subarcsecond <span class="hlt">bright</span> points and quasi-periodic upflows below a quiescent filament <span class="hlt">observed</span> by IRIS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, T.; Zhang, J.</p> <p>2016-05-01</p> <p>Context. The new Interface Region Imaging Spectrograph (IRIS) mission provides high-resolution <span class="hlt">observations</span> of UV spectra and slit-jaw images (SJIs). These data have become available for investigating the dynamic features in the transition region (TR) below the on-disk filaments. Aims: The driver of "counter-streaming" flows along the filament spine is still unknown yet. The magnetic structures and the upflows at the footpoints of the filaments and their relations with the filament mainbody have not been well understood. We study the dynamic evolution at the footpoints of filaments in order to find some clues for solving these questions. Methods: Using UV spectra and SJIs from the IRIS, along with coronal images and magnetograms from the Solar Dynamics Observatory (SDO), we present the new features in a quiescent filament channel: subarcsecond <span class="hlt">bright</span> points (BPs) and quasi-periodic upflows. Results: The BPs in the TR have a spatial scale of about 350-580 km and lifetimes of more than several tens of minutes. They are located at stronger magnetic structures in the filament channel with a magnetic flux of about 1017-1018 Mx. Quasi-periodic brightenings and upflows are <span class="hlt">observed</span> in the BPs, and the period is about 4-5 min. The BP and the associated jet-like upflow comprise a "tadpole-shaped" structure. The upflows move along <span class="hlt">bright</span> filament threads, and their directions are almost parallel to the spine of the filament. The upflows initiated from the BPs with opposite polarity magnetic fields have opposite directions. The velocity of the upflows in the plane of sky is about 5-50 km s-1. The emission line of Si IV 1402.77 Å at the locations of upflows exhibits obvious blueshifts of about 5-30 km s-1, and the line profile is broadened with the width of more than 20 km s-1. Conclusions: The BPs seem to be the bases of filament threads, and the upflows are able to convey mass for the dynamic balance of the filament. The "counter-streaming" flows in previous <span class="hlt">observations</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011ACPD...1130949D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ACPD...1130949D"><span>Lidar and radar measurements of the melting layer in the frame of the Convective and Orographically-induced Precipitation Study: <span class="hlt">observations</span> of dark and <span class="hlt">bright</span> band phenomena</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>di Girolamo, P.; Summa, D.; Bhawar, R.; di Iorio, T.; Norton, E. G.; Peters, G.; Dufournet, Y.</p> <p>2011-11-01</p> <p>During the Convective and Orographically-induced Precipitation Study (COPS), lidar dark and <span class="hlt">bright</span> bands were <span class="hlt">observed</span> by the University of BASILicata Raman lidar system (BASIL) during several intensive (IOPs) and special (SOPs) <span class="hlt">observation</span> periods (among others, 23 July, 15 August, and 17 August 2007). Lidar data were supported by measurements from the University of Hamburg cloud radar MIRA 36 (36 GHz), the University of Hamburg dual-polarization micro rain radars (24.1 GHz) and the University of Manchester UHF wind profiler (1.29 GHz). Results from BASIL and the radars for 23 July 2007 are illustrated and discussed to support the comprehension of the microphysical and scattering processes responsible for the appearance of the lidar and radar dark and <span class="hlt">bright</span> bands. Simulations of the lidar dark and <span class="hlt">bright</span> band based on the application of concentric/eccentric sphere Lorentz-Mie codes and a melting layer model are also provided. Lidar and radar measurements and model results are also compared with measurements from a disdrometer on ground and a two-dimensional cloud (2DC) probe on-board the ATR42 SAFIRE.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AAS...20915406C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AAS...20915406C"><span>A New Sky <span class="hlt">Brightness</span> Monitor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Crawford, David L.; McKenna, D.</p> <p>2006-12-01</p> <p>A good estimate of sky <span class="hlt">brightness</span> and its variations throughout the night, the months, and even the years is an essential bit of knowledge both for good <span class="hlt">observing</span> and especially as a tool in efforts to minimize sky <span class="hlt">brightness</span> through local action. Hence a stable and accurate monitor can be a valuable and necessary tool. We have developed such a monitor, with the financial help of Vatican Observatory and Walker Management. The device is now undergoing its Beta test in preparation for production. It is simple, accurate, well calibrated, and automatic, sending its data directly to IDA over the internet via E-mail . Approximately 50 such monitors will be ready soon for deployment worldwide including most major observatories. Those interested in having one should enquire of IDA about details.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19243743','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19243743"><span>A selective deficit in the appreciation and recognition of <span class="hlt">brightness</span>: <span class="hlt">brightness</span> agnosia?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nijboer, Tanja C W; Nys, Gudrun M S; van der Smagt, Maarten J; de Haan, Edward H F</p> <p>2009-01-01</p> <p>We report a patient with extensive brain damage in the right hemisphere who demonstrated a severe impairment in the appreciation of <span class="hlt">brightness</span>. Acuity, contrast sensitivity as well as luminance discrimination were normal, suggesting her <span class="hlt">brightness</span> impairment is not a mere consequence of low-level sensory impairments. The patient was not able to indicate the darker or the lighter of two grey squares, even though she was able to see that they differed. In addition, she could not indicate whether the lights in a room were switched on or off, nor was she able to differentiate between normal greyscale images and inverted greyscale images. As the patient recognised objects, colours, and shapes correctly, the impairment is specific for <span class="hlt">brightness</span>. As low-level, sensory processing is normal, this specific deficit in the recognition and appreciation of <span class="hlt">brightness</span> appears to be of a higher, cognitive level, the level of semantic knowledge. This appears to be the first report of '<span class="hlt">brightness</span> agnosia'.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.884a2052U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.884a2052U"><span>Tolerance limit value of <span class="hlt">brightness</span> and contrast adjustment on digitized radiographs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Utami, S. N.; Kiswanjaya, B.; Syahraini, S. I.; Ustriyana, P.</p> <p>2017-08-01</p> <p>The aim of this study was to measure the tolerance limit value of <span class="hlt">brightness</span> and contrast adjustment on digitized radiograph with apical periodontitis and early apical abscess. <span class="hlt">Brightness</span> and contrast adjustment on 60 periapical radiograph with apical periodontitis and early apical abscess made by 2 <span class="hlt">observers</span>. Reliabilities tested by Cohen’s Kappa Coefficient and significance tested by wilcoxon test. Tolerance limit value of <span class="hlt">brightness</span> and contrast adjustment for apical periodontitis is -5 and +5, early apical abscess is -10 and +10, and both is -5 and +5. <span class="hlt">Brightness</span> and contrast adjustment which not appropriate can alter the evaluation and differential diagnosis of periapical lesion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=10486&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=10486&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour"><span><span class="hlt">Bright</span> Loops at 171</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2007-01-01</p> <p>STEREO was able to capture <span class="hlt">bright</span> loops in exquisite detail as they were arcing above an active region (May 26, 2007) over an 18 hour period. What we are actually seeing are charged particles spinning along magnetic field lines that extend above the Sun's <span class="hlt">surface</span>. Active regions are areas of intense magnetic activity and often the source of solar storms. In fact, the clip ends with a flourish in which a small coronal mass ejection (CME) blows out into space. This is from the STEREO Ahead spacecraft at the 171 Angstroms wavelength in extreme ultraviolet light.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013A%26A...556A.105O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013A%26A...556A.105O"><span>A <span class="hlt">bright</span>-rimmed cloud sculpted by the H ii region Sh2-48</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ortega, M. E.; Paron, S.; Giacani, E.; Rubio, M.; Dubner, G.</p> <p>2013-08-01</p> <p>Aims: We characterize a <span class="hlt">bright</span>-rimmed cloud embedded in the H ii region Sh2-48 while searching for evidence of triggered star formation. Methods: We carried out <span class="hlt">observations</span> towards a region of 2' × 2' centered at RA = 18h22m11.39s, Dec = -14°35'24.81''(J2000) using the Atacama Submillimeter Telescope Experiment (ASTE; Chile) in the 12CO J = 3-2, 13CO J = 3-2, HCO+J = 4-3, and CS J = 7-6 lines with an angular resolution of about 22''. We also present radio continuum <span class="hlt">observations</span> at 5 GHz carried out with the Jansky Very Large Array (JVLA; EEUU) interferometer with a synthetized beam of 7'' × 5''. The molecular transitions were used to study the distribution and kinematics of the molecular gas of the <span class="hlt">bright</span>-rimmed cloud. The radio continuum data was used to characterize the ionized gas located on the illuminated border of this molecular condensation. Combining these <span class="hlt">observations</span> with infrared public data allowed us to build up a comprehensive picture of the current state of star formation within this cloud. Results: The analysis of our molecular <span class="hlt">observations</span> reveals a relatively dense clump with n(H2) ~ 3 × 103cm-3, located in projection onto the interior of the H ii region Sh2-48. The emission distribution of the four <span class="hlt">observed</span> molecular transitions has, at VLSR ~ 38 km s-1, morphological anticorrelation with the <span class="hlt">bright</span>-rimmed cloud as seen in the optical emission. From the new radio continuum <span class="hlt">observations</span>, we identify a thin layer of ionized gas located on the border of the clump that is facing the ionizing star. The ionized gas has an electron density of about 73 cm-3, which is a factor three higher than the typical critical density (nc ~ 25 cm-3), above which an ionized boundary layer can be formed and maintained. This supports the hypothesis that the clump is being photoionized by the nearby O9.5V star, BD-14 5014. From the evaluation of the pressure balance between the ionized and molecular gas, we conclude that the clump would be in a prepressure balance</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RAA....17...37R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RAA....17...37R"><span>Spatial Model of Sky <span class="hlt">Brightness</span> Magnitude in Langkawi Island, Malaysia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Redzuan Tahar, Mohammad; Kamarudin, Farahana; Umar, Roslan; Khairul Amri Kamarudin, Mohd; Sabri, Nor Hazmin; Ahmad, Karzaman; Rahim, Sobri Abdul; Sharul Aikal Baharim, Mohd</p> <p>2017-03-01</p> <p>Sky <span class="hlt">brightness</span> is an essential topic in the field of astronomy, especially for optical astronomical <span class="hlt">observations</span> that need very clear and dark sky conditions. This study presents the spatial model of sky <span class="hlt">brightness</span> magnitude in Langkawi Island, Malaysia. Two types of Sky Quality Meter (SQM) manufactured by Unihedron are used to measure the sky <span class="hlt">brightness</span> on a moonless night (or when the Moon is below the horizon), when the sky is cloudless and the locations are at least 100 m from the nearest light source. The selected locations are marked by their GPS coordinates. The sky <span class="hlt">brightness</span> data obtained in this study were interpolated and analyzed using a Geographic Information System (GIS), thus producing a spatial model of sky <span class="hlt">brightness</span> that clearly shows the dark and <span class="hlt">bright</span> sky areas in Langkawi Island. Surprisingly, our results show the existence of a few dark sites nearby areas of high human activity. The sky <span class="hlt">brightness</span> of 21.45 mag arcsec{}-2 in the Johnson-Cousins V-band, as the average of sky <span class="hlt">brightness</span> equivalent to 2.8 × {10}-4{cd} {{{m}}}-2 over the entire island, is an indication that the island is, overall, still relatively dark. However, the amount of development taking place might reduce the number in the near future as the island is famous as a holiday destination.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900056493&hterms=Einstein&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DEinstein','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900056493&hterms=Einstein&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DEinstein"><span>Einstein <span class="hlt">Observations</span> of Galactic supernova remnants</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Seward, Frederick D.</p> <p>1990-01-01</p> <p>This paper summarizes the <span class="hlt">observations</span> of Galactic supernova remnants with the imaging detectors of the Einstein Observatory. X-ray <span class="hlt">surface</span> <span class="hlt">brightness</span> contours of 47 remnants are shown together with gray-scale pictures. Count rates for these remnants have been derived and are listed for the HRI, IPC, and MPC detectors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007A%26A...476.1223C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007A%26A...476.1223C"><span>The HELLAS2XMM survey. XI. Unveiling the nature of X-ray <span class="hlt">bright</span> optically normal galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Civano, F.; Mignoli, M.; Comastri, A.; Vignali, C.; Fiore, F.; Pozzetti, L.; Brusa, M.; La Franca, F.; Matt, G.; Puccetti, S.; Cocchia, F.</p> <p>2007-12-01</p> <p>Aims:X-ray <span class="hlt">bright</span> optically normal galaxies (XBONGs) constitute a small but significant fraction of hard X-ray selected sources in recent Chandra and XMM-Newton surveys. Even though several possibilities were proposed to explain why a relatively luminous hard X-ray source does not leave any significant signature of its presence in terms of optical emission lines, the nature of XBONGs is still subject of debate. We aim to better understand their nature by means of a multiwavelength and morphological analysis of a small sample of these sources. Methods: Good-quality photometric near-infrared data (ISAAC/VLT) of four low-redshift (z = 0.1{-}0.3) XBONGs, selected from the HELLAS2XMM survey, have been used to search for the presence of the putative nucleus, applying the <span class="hlt">surface-brightness</span> decomposition technique through the least-squares fitting program GALFIT. Results: The <span class="hlt">surface</span> <span class="hlt">brightness</span> decomposition allows us to reveal a nuclear point-like source, likely to be responsible for the X-ray emission, in two out of the four sources. The results indicate that moderate amounts of gas and dust, covering a large solid angle (possibly 4π) at the nuclear source, combined with the low nuclear activity, may explain the lack of optical emission lines. The third XBONG is associated with an X-ray extended source and no nuclear excess is detected in the near infrared at the limits of our <span class="hlt">observations</span>. The last source is associated to a close (d≤ 1 arcsec) double system and the fitting procedure cannot achieve a firm conclusion. Based on <span class="hlt">observations</span> made at the European Southern Observatory, Paranal, Chile (ESO Programme ID 69.A-0554).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016WRR....52.4164G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016WRR....52.4164G"><span>Assimilation of gridded terrestrial water storage <span class="hlt">observations</span> from GRACE into a land <span class="hlt">surface</span> model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Girotto, Manuela; De Lannoy, Gabriëlle J. M.; Reichle, Rolf H.; Rodell, Matthew</p> <p>2016-05-01</p> <p><span class="hlt">Observations</span> of terrestrial water storage (TWS) from the Gravity Recovery and Climate Experiment (GRACE) satellite mission have a coarse resolution in time (monthly) and space (roughly 150,000 km2 at midlatitudes) and vertically integrate all water storage components over land, including soil moisture and groundwater. Data assimilation can be used to horizontally downscale and vertically partition GRACE-TWS <span class="hlt">observations</span>. This work proposes a variant of existing ensemble-based GRACE-TWS data assimilation schemes. The new algorithm differs in how the analysis increments are computed and applied. Existing schemes correlate the uncertainty in the modeled monthly TWS estimates with errors in the soil moisture profile state variables at a single instant in the month and then apply the increment either at the end of the month or gradually throughout the month. The proposed new scheme first computes increments for each day of the month and then applies the average of those increments at the beginning of the month. The new scheme therefore better reflects submonthly variations in TWS errors. The new and existing schemes are investigated here using gridded GRACE-TWS <span class="hlt">observations</span>. The assimilation results are validated at the monthly time scale, using in situ measurements of groundwater depth and soil moisture across the U.S. The new assimilation scheme yields improved (although not in a statistically significant sense) skill metrics for groundwater compared to the open-loop (no assimilation) simulations and compared to the existing assimilation schemes. A smaller impact is seen for <span class="hlt">surface</span> and root-zone soil moisture, which have a shorter memory and receive smaller increments from TWS assimilation than groundwater. These results motivate future efforts to combine GRACE-TWS <span class="hlt">observations</span> with <span class="hlt">observations</span> that are more sensitive to <span class="hlt">surface</span> soil moisture, such as L-band <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> from Soil Moisture Ocean Salinity (SMOS) or Soil Moisture Active Passive</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20160013297&hterms=storage&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dstorage','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20160013297&hterms=storage&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dstorage"><span>Assimilation of Gridded Terrestrial Water Storage <span class="hlt">Observations</span> from GRACE into a Land <span class="hlt">Surface</span> Model</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Girotto, Manuela; De Lannoy, Gabrielle J. M.; Reichle, Rolf H.; Rodell, Matthew</p> <p>2016-01-01</p> <p><span class="hlt">Observations</span> of terrestrial water storage (TWS) from the Gravity Recovery and Climate Experiment (GRACE) satellite mission have a coarse resolution in time (monthly) and space (roughly 150,000 km(sup 2) at midlatitudes) and vertically integrate all water storage components over land, including soil moisture and groundwater. Data assimilation can be used to horizontally downscale and vertically partition GRACE-TWS <span class="hlt">observations</span>. This work proposes a variant of existing ensemble-based GRACE-TWS data assimilation schemes. The new algorithm differs in how the analysis increments are computed and applied. Existing schemes correlate the uncertainty in the modeled monthly TWS estimates with errors in the soil moisture profile state variables at a single instant in the month and then apply the increment either at the end of the month or gradually throughout the month. The proposed new scheme first computes increments for each day of the month and then applies the average of those increments at the beginning of the month. The new scheme therefore better reflects submonthly variations in TWS errors. The new and existing schemes are investigated here using gridded GRACE-TWS <span class="hlt">observations</span>. The assimilation results are validated at the monthly time scale, using in situ measurements of groundwater depth and soil moisture across the U.S. The new assimilation scheme yields improved (although not in a statistically significant sense) skill metrics for groundwater compared to the open-loop (no assimilation) simulations and compared to the existing assimilation schemes. A smaller impact is seen for <span class="hlt">surface</span> and root-zone soil moisture, which have a shorter memory and receive smaller increments from TWS assimilation than groundwater. These results motivate future efforts to combine GRACE-TWS <span class="hlt">observations</span> with <span class="hlt">observations</span> that are more sensitive to <span class="hlt">surface</span> soil moisture, such as L-band <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> from Soil Moisture Ocean Salinity (SMOS) or Soil Moisture Active</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20090008678&hterms=by-product&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dby-product','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20090008678&hterms=by-product&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dby-product"><span>Titan <span class="hlt">Surface</span> Temperatures as Measured by Cassini CIRS</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jennings, Donald E.; Flasar, F.M.; Kunde, V.G.; Nixon, C.A.; Romani, P.N.; Samuelson, R.E.; Coustenis, A.; Courtin, R.</p> <p>2009-01-01</p> <p>Thermal radiation from the <span class="hlt">surface</span> of Titan reaches space through a spectral window of low opacity at 19-microns wavelength. This radiance gives a measure of the <span class="hlt">brightness</span> temperature of the <span class="hlt">surface</span>. Composite Infrared Spectrometer' (CIRS) <span class="hlt">observations</span> from Cassini during its first four years at Saturn have permitted latitude mapping of zonally averaged <span class="hlt">surface</span> temperatures. The measurements are corrected for atmospheric opacity using the dependence of radiance on emission angle. With the more complete latitude coverage and much larger dataset of CIRS we have improved upon the original results from Voyager IRIS. CIRS measures the equatorial <span class="hlt">surface</span> <span class="hlt">brightness</span> temperature to be 93.7+/-0.6 K, the same as the temperature measured at the Huygens landing site. The <span class="hlt">surface</span> <span class="hlt">brightness</span> temperature decreases by 2 K toward the south pole and by 3 K toward the north pole. The drop in <span class="hlt">surface</span> temperature between equator and north pole implies a 50% decrease in methane saturation vapor pressure and relative humidity; this may help explain the large northern lakes. The H2 mole fraction is derived as a by-product of our analysis and agrees with previous results. Evidence of seasonal variation in <span class="hlt">surface</span> and atmospheric temperatures is emerging from CIRS measurements over the Cassini mission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70032090','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70032090"><span>Fluvial channels on Titan: Initial Cassini RADAR <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lorenz, R.D.; Lopes, R.M.; Paganelli, F.; Lunine, J.I.; Kirk, R.L.; Mitchell, K.L.; Soderblom, L.A.; Stofan, E.R.; Ori, G.; Myers, M.; Miyamoto, H.; Radebaugh, J.; Stiles, B.; Wall, S.D.; Wood, C.A.</p> <p>2008-01-01</p> <p>Cassini radar images show a variety of fluvial channels on Titan's <span class="hlt">surface</span>, often several hundreds of kilometers in length. Some (predominantly at low- and mid-latitude) are radar-<span class="hlt">bright</span> and braided, resembling desert washes where fines have been removed by energetic <span class="hlt">surface</span> liquid flow, presumably from methane rainstorms. Others (predominantly at high latitudes) are radar-dark and meandering and drain into or connect polar lakes, suggesting slower-moving flow depositing fine-grained sediments. A third type, seen predominantly at mid- and high latitudes, have radar <span class="hlt">brightness</span> patterns indicating topographic incision, with valley widths of up to 3 km across and depth of several hundred meters. These <span class="hlt">observations</span> show that fluvial activity occurs at least occasionally at all latitudes, not only at the Huygens landing site, and can produce channels much larger in scale than those <span class="hlt">observed</span> there. The areas in which channels are prominent so far amount to about 1% of Titan's <span class="hlt">surface</span>, of which only a fraction is actually occupied by channels. The corresponding global sediment volume inferred is not enough to account for the extensive sand seas. Channels <span class="hlt">observed</span> so far have a consistent large-scale flow pattern, tending to flow polewards and eastwards. ?? 2008.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMAE33C0513Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMAE33C0513Y"><span>Long-delayed <span class="hlt">bright</span> dancing sprite with large horizontal displacement from its parent flash</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, J.; Lu, G.; Lee, L. J.; Feng, G.</p> <p>2015-12-01</p> <p>A long-delayed very <span class="hlt">bright</span> dancing sprite with large horizontal displacement from its parent flash was <span class="hlt">observed</span>. The dancing sprite lasted only 60 ms, and the morphology consisted of three fields with two slim dim sprite elements in the first two fields and a very <span class="hlt">bright</span> large sprite element in the third field, different from other <span class="hlt">observations</span>. The <span class="hlt">bright</span> sprite displaced at least 38 km from its parent flash and occurred over comparatively higher cloud top region. The parent flash was positive, with only one return stroke (~24 kA) and obvious continuing current process, and the charge moment change of the stroke was small (roughly the threshold for sprite production). All of the sprite elements occurred during the continuing current period, and the <span class="hlt">bright</span> sprite induced considerable current. The sprite dancing features may be linked to parent storm electrical structure, dynamics and microphysics, and the parent CG discharge process which was consistent with VHF <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AAS...22713716W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AAS...22713716W"><span>High Precision Photometry of <span class="hlt">Bright</span> Transiting Exoplanet Hosts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wilson, Maurice; Eastman, Jason; Johnson, John A.</p> <p>2016-01-01</p> <p>Within the past two decades, the successful search for exoplanets and the characterization of their physical properties have shown the immense progress that has been made towards finding planets with characteristics similar to Earth. For most exoplanets with a radius about the size of Earth, evaluating their physical properties, such as the mass, radius and equilibrium temperature, cannot be determined with satisfactory precision. The MINiature Exoplanet Radial Velocity Array (MINERVA) was recently built to obtain spectroscopic and photometric measurements to find, confirm, and characterize Earth-like exoplanets. MINERVA's spectroscopic survey targets the brightest, nearby stars which are well-suited to the array's capabilities, while its primary photometric goal is to search for transits around these <span class="hlt">bright</span> targets. Typically, it is difficult to find satisfactory comparison stars within a telescope's field of view when the primary target is very <span class="hlt">bright</span>. This issue is resolved by using one of MINERVA's telescopes to <span class="hlt">observe</span> the primary <span class="hlt">bright</span> star while the other telescopes <span class="hlt">observe</span> a distinct field of view that contains satisfactory <span class="hlt">bright</span> comparison stars. We describe the code used to identify nearby comparison stars, schedule the four telescopes, produce differential photometry from multiple telescopes, and show the first results from this effort.This work has been funded by the Ronald E. McNair Post-Baccalaureate Achievement Program, the ERAU Honors Program, the ERAU Undergraduate Research Spark Fund, and the Banneker Institute at the Harvard-Smithsonian Center for Astrophysics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...855L..21B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...855L..21B"><span>Evidence for Precursors of the Coronal Hole Jets in Solar <span class="hlt">Bright</span> Points</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bagashvili, Salome R.; Shergelashvili, Bidzina M.; Japaridze, Darejan R.; Kukhianidze, Vasil; Poedts, Stefaan; Zaqarashvili, Teimuraz V.; Khodachenko, Maxim L.; De Causmaecker, Patrick</p> <p>2018-03-01</p> <p>A set of 23 <span class="hlt">observations</span> of coronal jet events that occurred in coronal <span class="hlt">bright</span> points has been analyzed. The focus was on the temporal evolution of the mean <span class="hlt">brightness</span> before and during coronal jet events. In the absolute majority of the cases either single or recurrent coronal jets (CJs) were preceded by slight precursor disturbances <span class="hlt">observed</span> in the mean intensity curves. The key conclusion is that we were able to detect quasi-periodical oscillations with characteristic periods from sub-minute up to 3–4 minute values in the <span class="hlt">bright</span> point <span class="hlt">brightness</span> that precedes the jets. Our basic claim is that along with the conventionally accepted scenario of <span class="hlt">bright</span>-point evolution through new magnetic flux emergence and its reconnection with the initial structure of the <span class="hlt">bright</span> point and the coronal hole, certain magnetohydrodynamic (MHD) oscillatory and wavelike motions can be excited and these can take an important place in the <span class="hlt">observed</span> dynamics. These quasi-oscillatory phenomena might play the role of links between different epochs of the coronal jet ignition and evolution. They can be an indication of the MHD wave excitation processes due to the system entropy variations, density variations, or shear flows. It is very likely a sharp outflow velocity transverse gradients at the edges between the open and closed field line regions. We suppose that magnetic reconnections can be the source of MHD waves due to impulsive generation or rapid temperature variations, and shear flow driven nonmodel MHD wave evolution (self-heating and/or overreflection mechanisms).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24695177','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24695177"><span>Color constancy in a scene with <span class="hlt">bright</span> colors that do not have a fully natural <span class="hlt">surface</span> appearance.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fukuda, Kazuho; Uchikawa, Keiji</p> <p>2014-04-01</p> <p>Theoretical and experimental approaches have proposed that color constancy involves a correction related to some average of stimulation over the scene, and some of the studies showed that the average gives greater weight to surrounding <span class="hlt">bright</span> colors. However, in a natural scene, high-luminance elements do not necessarily carry information about the scene illuminant when the luminance is too high for it to appear as a natural object color. The question is how a surrounding color's appearance mode influences its contribution to the degree of color constancy. Here the stimuli were simple geometric patterns, and the luminance of surrounding colors was tested over the range beyond the luminosity threshold. <span class="hlt">Observers</span> performed perceptual achromatic setting on the test patch in order to measure the degree of color constancy and evaluated the surrounding <span class="hlt">bright</span> colors' appearance mode. Broadly, our results support the assumption that the visual system counts only the colors in the object-color appearance for color constancy. However, detailed analysis indicated that surrounding colors without a fully natural object-color appearance had some sort of influence on color constancy. Consideration of this contribution of unnatural object color might be important for precise modeling of human color constancy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMOS11A1778Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMOS11A1778Z"><span>Death of Darkness: Artificial Sky <span class="hlt">Brightness</span> in the Anthropocene</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zender, C. S.</p> <p>2016-12-01</p> <p>Many species (including ours) need darkness to survive and thrive yet light pollution in the anthropocene has received scant attention in Earth System Models (ESMs). Anthropogenic aerosols can brighten background sky <span class="hlt">brightness</span> and reduce the contrast between skylight and starlight. These are both aesthetic and health-related issues due to their accompanying disruption of circadian rhythms. We quantify aerosol contributions to light pollution using a single-column night sky model, NiteLite, suitable for implementation in ESMs. NiteLite accounts for physiologcal (photopic and scotopic vision, retinal diameter/age), anthropogenic (light and aerosol pollution properties), and natural (<span class="hlt">surface</span> albedo, trace gases) effects on background <span class="hlt">brightness</span> and threshold visibility. We find that stratospheric aerosol injection contemplated as a stop-gap measure to counter global warming would increase night-sky <span class="hlt">brightness</span> by about 25%, and thus eliminate last pristine dark sky areas on Earth. Our results suggest that ESMs incorporate light pollution so that associated societal impacts can be better quantified and included in policy deliberations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA10006&hterms=duck+hazard&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dduck%2Bhazard','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA10006&hterms=duck+hazard&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dduck%2Bhazard"><span>At <span class="hlt">Bright</span> Band Inside Victoria Crater</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2007-01-01</p> <p><p/> A layer of light-toned rock exposed inside Victoria Crater in the Meridiani Planum region of Mars appears to mark where the <span class="hlt">surface</span> was at the time, many millions of years ago, when an impact excavated the crater. NASA's Mars Exploration Rover Opportunity drove to this <span class="hlt">bright</span> band as the science team's first destination for the rover during investigations inside the crater. <p/> Opportunity's left front hazard-identification camera took this image just after the rover finished a drive of 2.25 meters (7 feet, 5 inches) during the rover's 1,305th Martian day, or sol, (Sept. 25, 2007). The rocks beneath the rover and its extended robotic arm are part of the <span class="hlt">bright</span> band. <p/> Victoria Crater has a scalloped shape of alternating alcoves and promontories around the crater's circumference. Opportunity descended into the crater two weeks earlier, within an alcove called 'Duck Bay.' Counterclockwise around the rim, just to the right of the arm in this image, is a promontory called 'Cabo Frio.'</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_14 --> <div id="page_15" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="281"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21230954','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21230954"><span>Stable vortex-<span class="hlt">bright</span>-soliton structures in two-component Bose-Einstein condensates.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Law, K J H; Kevrekidis, P G; Tuckerman, Laurette S</p> <p>2010-10-15</p> <p>We report the numerical realization of robust two-component structures in 2D and 3D Bose-Einstein condensates with nontrivial topological charge in one component. We identify a stable symbiotic state in which a higher-dimensional <span class="hlt">bright</span> soliton exists even in a homogeneous setting with defocusing interactions, due to the effective potential created by a stable vortex in the other component. The resulting vortex-<span class="hlt">bright</span>-solitons, generalizations of the recently experimentally <span class="hlt">observed</span> dark-<span class="hlt">bright</span> solitons, are found to be very robust both in the homogeneous medium and in the presence of external confinement.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110015365','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110015365"><span>Seasonal Changes in Titan's <span class="hlt">Surface</span> Temperatures</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jennins, Donald E.; Cottini, V.; Nixon, C. A.; Flasar, F. M.; Kunde, V. G.; Samuelson, R. E.; Romani, P. N.; Hesman, B. E.; Carlson, R. C.; Gorius, N. J. P.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20110015365'); toggleEditAbsImage('author_20110015365_show'); toggleEditAbsImage('author_20110015365_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20110015365_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20110015365_hide"></p> <p>2011-01-01</p> <p>Seasonal changes in Titan's <span class="hlt">surface</span> <span class="hlt">brightness</span> temperatures have been <span class="hlt">observed</span> by Cassini in the thermal infrared. The Composite Infrared Spectrometer (CIRS) measured <span class="hlt">surface</span> radiances at 19 micron in two time periods: one in late northern winter (Ls = 335d eg) and another centered on northern spring equinox (Ls = 0 deg). In both periods we constructed pole-to-pole maps of zonally averaged <span class="hlt">brightness</span> temperatures corrected for effects of the atmosphere. Between late northern winter and northern spring equinox a shift occurred in the temperature distribution, characterized by a warming of approximately 0.5 K in the north and a cooling by about the same amount in the south. At equinox the polar <span class="hlt">surface</span> temperatures were both near 91 K and the equator was 93.4 K. We measured a seasonal lag of delta Ls approximately 9 in the meridional <span class="hlt">surface</span> temperature distribution, consistent with the post-equinox results of Voyager 1 as well as with predictions from general circulation modeling. A slightly elevated temperature is <span class="hlt">observed</span> at 65 deg S in the relatively cloud-free zone between the mid-latitude and southern cloud regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19760013983','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19760013983"><span>Comet <span class="hlt">brightness</span> parameters: Definition, determination, and correlations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Meisel, D. D.; Morris, C. S.</p> <p>1976-01-01</p> <p>The power-law definition of comet <span class="hlt">brightness</span> is reviewed and possible systematic influences are discussed that can affect the derivation of m sub o and n values from visual magnitude estimates. A rationale for the Bobrovnikoff aperture correction method is given and it is demonstrated that the Beyer extrafocal method leads to large systematic effects which if uncorrected by an instrumental relationship result in values significantly higher than those derived according to the Bobrovnikoff guidelines. A series of visual <span class="hlt">brightness</span> parameter sets are presented which have been reduced to the same photometric system. Recommendations are given to insure that future <span class="hlt">observations</span> are reduced to the same system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E%26PSL.484..145W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E%26PSL.484..145W"><span>Spectroscopic <span class="hlt">observations</span> of the Moon at the lunar <span class="hlt">surface</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Yunzhao; Hapke, Bruce</p> <p>2018-02-01</p> <p>The Moon's reflectance spectrum records many of its important properties. However, prior to Chang'E-3 (CE-3), no spectra had previously been measured on the lunar <span class="hlt">surface</span>. Here we show the in situ reflectance spectra of the Moon acquired on the lunar <span class="hlt">surface</span> by the Visible-Near Infrared Spectrometer (VNIS) onboard the CE-3 rover. The VNIS detected thermal radiation from the lunar regolith, though with much shorter wavelength range than typical thermal radiometer. The measured temperatures are higher than expected from theoretical model, indicating low thermal inertia of the lunar soil and the effects of grain facet on soil temperature in submillimeter scale. The in situ spectra also reveal that 1) <span class="hlt">brightness</span> changes visible from orbit are related to the reduction in maturity due to the removal of the fine and weathered particles by the lander's rocket exhaust, not the smoothing of the <span class="hlt">surface</span> and 2) the spectra of the uppermost soil detected by remote sensing exhibit substantial differences with that immediately beneath, which has important implications for the remote compositional analysis. The reflectance spectra measured by VNIS not only reveal the thermal, compositional, and space-weathering properties of the Moon but also provide a means for the calibration of optical instruments that view the <span class="hlt">surface</span> remotely.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3283774','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3283774"><span>Excitation Spectra and <span class="hlt">Brightness</span> Optimization of Two-Photon Excited Probes</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Mütze, Jörg; Iyer, Vijay; Macklin, John J.; Colonell, Jennifer; Karsh, Bill; Petrášek, Zdeněk; Schwille, Petra; Looger, Loren L.; Lavis, Luke D.; Harris, Timothy D.</p> <p>2012-01-01</p> <p>Two-photon probe excitation data are commonly presented as absorption cross section or molecular <span class="hlt">brightness</span> (the detected fluorescence rate per molecule). We report two-photon molecular <span class="hlt">brightness</span> spectra for a diverse set of organic and genetically encoded probes with an automated spectroscopic system based on fluorescence correlation spectroscopy. The two-photon action cross section can be extracted from molecular <span class="hlt">brightness</span> measurements at low excitation intensities, while peak molecular <span class="hlt">brightness</span> (the maximum molecular <span class="hlt">brightness</span> with increasing excitation intensity) is measured at higher intensities at which probe photophysical effects become significant. The spectral shape of these two parameters was similar across all dye families tested. Peak molecular <span class="hlt">brightness</span> spectra, which can be obtained rapidly and with reduced experimental complexity, can thus serve as a first-order approximation to cross-section spectra in determining optimal wavelengths for two-photon excitation, while providing additional information pertaining to probe photostability. The data shown should assist in probe choice and experimental design for multiphoton microscopy studies. Further, we show that, by the addition of a passive pulse splitter, nonlinear bleaching can be reduced—resulting in an enhancement of the fluorescence signal in fluorescence correlation spectroscopy by a factor of two. This increase in fluorescence signal, together with the <span class="hlt">observed</span> resemblance of action cross section and peak <span class="hlt">brightness</span> spectra, suggests higher-order photobleaching pathways for two-photon excitation. PMID:22385865</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EPSC....8..346M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EPSC....8..346M"><span><span class="hlt">Bright</span> Perseids 2007-2012 statistics. Estimation of collision risks in circumterrestrial space</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Murtazov, A.</p> <p>2013-09-01</p> <p><span class="hlt">Bright</span> meteors are of serious hazard for space vehicles. In the Persieds shower these are the meteors brighter than 0m [2]. During 2007-2012 we conducted wide-angle CCD <span class="hlt">observations</span> of <span class="hlt">bright</span> Perseids [3-5]. <span class="hlt">Observations</span> were performed near Ryazan, Russia, (= 54.467 N, λ=39.750 E, H=200 m) using a Watec 902H camera and a Computar T2314FICS lens with the effective FOV of 140×100 arc degrees directed towards the local zenith. The sky control and meteor detection were provided using a Pinnacle Media Center EN or the Contrast as a grabber and an Intel Core.2 CPU processor, 1.83GHz, 500Mb RAM. Our results as compared to the visual meteors total number (IMO) are shown in Fig. 1. The averaged Perseids maximum lies within the solar longitudes 140.00-140.25 and here the average total shower spatial density is (80±6)·10-9km-3. The <span class="hlt">bright</span> Perseids average spatial density maximum is about (6±2) 10-9km-3. The <span class="hlt">bright</span> Perseids average percentage in the shower is calculated as the integrals ratio under the curves in Fig. 1 and is equal to 5% for the presented range of solar longitudes. It is natural to expect that the space densities of meteoroids decrease exponentially from the maximum [1]: D = D0exp{-B|λ-λ0|}, (1) where: D0 is the maximum meteor spatial density near the solar longitude λ0 and B - the factor determined empirically from <span class="hlt">observations</span>. The meteoroid flux F is equal to the number of particles passing through the elementary area per time unit: F(λ) = D(λ)·v, (2) where: v - is the meteor shower velocity. During the Perseids' maximum (D0=6·10-9km-3 and v=59km/s) the <span class="hlt">bright</span> meteoroid flux was equal to F (3.8±1.1)10-7km-2s-1, which corresponds to the hour rate HR15 for our camera FOV. The collision risk R here amounted to one collision per month on average with a 1 sq. km plane located normal to the meteor shower. An artificial space object rotating around the Earth constantly changes its orientation relative to the meteor shower, the Sun, and a ground</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20000073234&hterms=How+soil+form&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DHow%2Bsoil%2Bform','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20000073234&hterms=How+soil+form&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DHow%2Bsoil%2Bform"><span>Estimating Long Term <span class="hlt">Surface</span> Soil Moisture in the GCIP Area From Satellite Microwave <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Owe, Manfred; deJeu, Vrije; VandeGriend, Adriaan A.</p> <p>2000-01-01</p> <p>Soil moisture is an important component of the water and energy balances of the Earth's <span class="hlt">surface</span>. Furthermore, it has been identified as a parameter of significant potential for improving the accuracy of large-scale land <span class="hlt">surface</span>-atmosphere interaction models. However, accurate estimates of <span class="hlt">surface</span> soil moisture are often difficult to make, especially at large spatial scales. Soil moisture is a highly variable land <span class="hlt">surface</span> parameter, and while point measurements are usually accurate, they are representative only of the immediate site which was sampled. Simple averaging of point values to obtain spatial means often leads to substantial errors. Since remotely sensed <span class="hlt">observations</span> are already a spatially averaged or areally integrated value, they are ideally suited for measuring land <span class="hlt">surface</span> parameters, and as such, are a logical input to regional or larger scale land process models. A nine-year database of <span class="hlt">surface</span> soil moisture is being developed for the Central United States from satellite microwave <span class="hlt">observations</span>. This region forms much of the GCIP study area, and contains most of the Mississippi, Rio Grande, and Red River drainages. Daytime and nighttime microwave <span class="hlt">brightness</span> temperatures were <span class="hlt">observed</span> at a frequency of 6.6 GHz, by the Scanning Multichannel Microwave Radiometer (SMMR), onboard the Nimbus 7 satellite. The life of the SMMR instrument spanned from Nov. 1978 to Aug. 1987. At 6.6 GHz, the instrument provided a spatial resolution of approximately 150 km, and an orbital frequency over any pixel-sized area of about 2 daytime and 2 nighttime passes per week. Ground measurements of <span class="hlt">surface</span> soil moisture from various locations throughout the study area are used to calibrate the microwave <span class="hlt">observations</span>. Because ground measurements are usually only single point values, and since the time of satellite coverage does not always coincide with the ground measurements, the soil moisture data were used to calibrate a regional water balance for the top 1, 5, and 10 cm</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4216005','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4216005"><span>Intermittent Episodes of <span class="hlt">Bright</span> Light Suppress Myopia in the Chicken More than Continuous <span class="hlt">Bright</span> Light</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Lan, Weizhong; Feldkaemper, Marita; Schaeffel, Frank</p> <p>2014-01-01</p> <p>Purpose <span class="hlt">Bright</span> light has been shown a powerful inhibitor of myopia development in animal models. We studied which temporal patterns of <span class="hlt">bright</span> light are the most potent in suppressing deprivation myopia in chickens. Methods Eight-day-old chickens wore diffusers over one eye to induce deprivation myopia. A reference group (n = 8) was kept under office-like illuminance (500 lux) at a 10∶14 light∶dark cycle. Episodes of <span class="hlt">bright</span> light (15 000 lux) were super-imposed on this background as follows. Paradigm I: exposure to constant <span class="hlt">bright</span> light for either 1 hour (n = 5), 2 hours (n = 5), 5 hours (n = 4) or 10 hours (n = 4). Paradigm II: exposure to repeated cycles of <span class="hlt">bright</span> light with 50% duty cycle and either 60 minutes (n = 7), 30 minutes (n = 8), 15 minutes (n = 6), 7 minutes (n = 7) or 1 minute (n = 7) periods, provided for 10 hours. Refraction and axial length were measured prior to and immediately after the 5-day experiment. Relative changes were analyzed by paired t-tests, and differences among groups were tested by one-way ANOVA. Results Compared with the reference group, exposure to continuous <span class="hlt">bright</span> light for 1 or 2 hours every day had no significant protective effect against deprivation myopia. Inhibition of myopia became significant after 5 hours of <span class="hlt">bright</span> light exposure but extending the duration to 10 hours did not offer an additional benefit. In comparison, repeated cycles of 1∶1 or 7∶7 minutes of <span class="hlt">bright</span> light enhanced the protective effect against myopia and could fully suppress its development. Conclusions The protective effect of <span class="hlt">bright</span> light depends on the exposure duration and, to the intermittent form, the frequency cycle. Compared to the saturation effect of continuous <span class="hlt">bright</span> light, low frequency cycles of <span class="hlt">bright</span> light (1∶1 min) provided the strongest inhibition effect. However, our quantitative results probably might not be directly translated into humans, but rather need further amendments in clinical</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25360635','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25360635"><span>Intermittent episodes of <span class="hlt">bright</span> light suppress myopia in the chicken more than continuous <span class="hlt">bright</span> light.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lan, Weizhong; Feldkaemper, Marita; Schaeffel, Frank</p> <p>2014-01-01</p> <p><span class="hlt">Bright</span> light has been shown a powerful inhibitor of myopia development in animal models. We studied which temporal patterns of <span class="hlt">bright</span> light are the most potent in suppressing deprivation myopia in chickens. Eight-day-old chickens wore diffusers over one eye to induce deprivation myopia. A reference group (n = 8) was kept under office-like illuminance (500 lux) at a 10:14 light:dark cycle. Episodes of <span class="hlt">bright</span> light (15 000 lux) were super-imposed on this background as follows. Paradigm I: exposure to constant <span class="hlt">bright</span> light for either 1 hour (n = 5), 2 hours (n = 5), 5 hours (n = 4) or 10 hours (n = 4). Paradigm II: exposure to repeated cycles of <span class="hlt">bright</span> light with 50% duty cycle and either 60 minutes (n = 7), 30 minutes (n = 8), 15 minutes (n = 6), 7 minutes (n = 7) or 1 minute (n = 7) periods, provided for 10 hours. Refraction and axial length were measured prior to and immediately after the 5-day experiment. Relative changes were analyzed by paired t-tests, and differences among groups were tested by one-way ANOVA. Compared with the reference group, exposure to continuous <span class="hlt">bright</span> light for 1 or 2 hours every day had no significant protective effect against deprivation myopia. Inhibition of myopia became significant after 5 hours of <span class="hlt">bright</span> light exposure but extending the duration to 10 hours did not offer an additional benefit. In comparison, repeated cycles of 1:1 or 7:7 minutes of <span class="hlt">bright</span> light enhanced the protective effect against myopia and could fully suppress its development. The protective effect of <span class="hlt">bright</span> light depends on the exposure duration and, to the intermittent form, the frequency cycle. Compared to the saturation effect of continuous <span class="hlt">bright</span> light, low frequency cycles of <span class="hlt">bright</span> light (1:1 min) provided the strongest inhibition effect. However, our quantitative results probably might not be directly translated into humans, but rather need further amendments in clinical studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PASP..129c5003P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PASP..129c5003P"><span>Night Sky <span class="hlt">Brightness</span> at San Pedro Martir Observatory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Plauchu-Frayn, I.; Richer, M. G.; Colorado, E.; Herrera, J.; Córdova, A.; Ceseña, U.; Ávila, F.</p> <p>2017-03-01</p> <p>We present optical UBVRI zenith night sky <span class="hlt">brightness</span> measurements collected on 18 nights during 2013 to 2016 and SQM measurements obtained daily over 20 months during 2014 to 2016 at the Observatorio Astronómico Nacional on the Sierra San Pedro Mártir (OAN-SPM) in México. The UBVRI data is based upon CCD images obtained with the 0.84 m and 2.12 m telescopes, while the SQM data is obtained with a high-sensitivity, low-cost photometer. The typical moonless night sky <span class="hlt">brightness</span> at zenith averaged over the whole period is U = 22.68, B = 23.10, V = 21.84, R = 21.04, I = 19.36, and SQM = 21.88 {mag} {{arcsec}}-2, once corrected for zodiacal light. We find no seasonal variation of the night sky <span class="hlt">brightness</span> measured with the SQM. The typical night sky <span class="hlt">brightness</span> values found at OAN-SPM are similar to those reported for other astronomical dark sites at a similar phase of the solar cycle. We find a trend of decreasing night sky <span class="hlt">brightness</span> with decreasing solar activity during period of the <span class="hlt">observations</span>. This trend implies that the sky has become darker by Δ U = 0.7, Δ B = 0.5, Δ V = 0.3, Δ R=0.5 mag arcsec-2 since early 2014 due to the present solar cycle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130010955','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130010955"><span>Mineralogical Composition of the Different Types of <span class="hlt">Bright</span> Deposits on Vesta</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zambon, F.; Capaccioni, F.; DeSanctis, M. C.; Ammannito, E.; Li, J.-Y.; Longobardo, A.; Mittlefehldt, D. W.; Palomba, E.; Pieters, C. M.; Schroeder, S. E.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20130010955'); toggleEditAbsImage('author_20130010955_show'); toggleEditAbsImage('author_20130010955_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20130010955_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20130010955_hide"></p> <p>2013-01-01</p> <p>VIR-MS, Dawn's Visible and Infrared Mapping Spectrometer, obtained hyperspectral images of a wide part of Vesta's <span class="hlt">surface</span> at a variety of spatial resolutions [1]. Vesta spectra are similar to those of the howardite-eucrite-diogenite (HED) meteorites. Moreover, they are characterized by the two iron-bearing pyroxene bands at 0.9 (band I) and 1.9 microns (band II). Vesta <span class="hlt">surface</span>'s is dominated by eucrite/howardite with some diogenitic regions situated in the southern hemisphere near the Rheasilvia basin [2]. The <span class="hlt">surface</span> is heavily craterized and impacts can expose fresh material, thus generating the <span class="hlt">Bright</span> Material Deposits (BMD) <span class="hlt">observed</span> within and surrounding certain craters. BMD can be classified into six different types based on their morphological characteristics: Crater Wall/Scarp Material (CWM), Radial Material (RM), Slope Material (SM), Patchy Material (PM), Spot Material (SpM) and Diffuse Plains Material (DPM) [3]. The most widespread BMD are CWM, SM and RM. CWM, SM, RM originate from impacts. CWM is situated on the edge of the craters. Mass wasting from the crater walls and generates the SM, while RM is associated with the ejecta of the craters [4]. BMD are characterized by albedo greater than that of the vestan average, 0.38 [5]. Therefore the different types of deposits present distinct levels of reflectance respect to the Surrounding Regions (SR), in particular: the CWM and SM is approx.40% brighter, the RM is approx.30- 40% brighter; the SpM is about 20-25% brighter and the PM is about 20% brighter. Near the edge of the Rheasilvia basin it is possible to find some extremely <span class="hlt">bright</span> areas 80% brighter than the vestan average [6].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.C41D0757S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.C41D0757S"><span>Snowmelt and <span class="hlt">Surface</span> Freeze/Thaw Timings over Alaska derived from Passive Microwave <span class="hlt">Observations</span> using a Wavelet Classifier</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Steiner, N.; McDonald, K. C.; Dinardo, S. J.; Miller, C. E.</p> <p>2015-12-01</p> <p>Arctic permafrost soils contain a vast amount of organic carbon that will be released into the atmosphere as carbon dioxide or methane when thawed. <span class="hlt">Surface</span> to air greenhouse gas fluxes are largely dependent on such <span class="hlt">surface</span> controls as the frozen/thawed state of the snow and soil. Satellite remote sensing is an important means to create continuous mapping of <span class="hlt">surface</span> properties. Advances in the ability to determine soil and snow freeze/thaw timings from microwave frequency <span class="hlt">observations</span> improves upon our ability to predict the response of carbon gas emission to warming through synthesis with in-situ <span class="hlt">observation</span>, such as the 2012-2015 Carbon in Arctic Reservoir Vulnerability Experiment (CARVE). <span class="hlt">Surface</span> freeze/thaw or snowmelt timings are often derived using a constant or spatially/temporally variable threshold applied to time-series <span class="hlt">observations</span>. Alternately, time-series singularity classifiers aim to detect discontinuous changes, or "edges", in time-series data similar to those that occur from the large contrast in dielectric constant during the freezing or thaw of soil or snow. We use multi-scale analysis of continuous wavelet transform spectral gradient <span class="hlt">brightness</span> temperatures from various channel combinations of passive microwave radiometers, Advanced Microwave Scanning Radiometer (AMSR-E, AMSR2) and Special Sensor Microwave Imager (SSM/I F17) gridded at a 10 km posting with resolution proportional to the <span class="hlt">observational</span> footprint. Channel combinations presented here aim to illustrate and differentiate timings of "edges" from transitions in <span class="hlt">surface</span> water related to various landscape components (e.g. snow-melt, soil-thaw). To support an understanding of the physical basis of <span class="hlt">observed</span> "edges" we compare satellite measurements with simple radiative transfer microwave-emission modeling of the snow, soil and vegetation using in-situ <span class="hlt">observations</span> from the SNOw TELemetry (SNOTEL) automated weather stations. Results of freeze/thaw and snow-melt timings and trends are</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22654478-first-alma-observation-solar-plasmoid-ejection-from-ray-bright-point','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22654478-first-alma-observation-solar-plasmoid-ejection-from-ray-bright-point"><span>The First ALMA <span class="hlt">Observation</span> of a Solar Plasmoid Ejection from an X-Ray <span class="hlt">Bright</span> Point</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Shimojo, Masumi; Hudson, Hugh S.; White, Stephen M.</p> <p>2017-05-20</p> <p>Eruptive phenomena such as plasmoid ejections or jets are important features of solar activity and have the potential to improve our understanding of the dynamics of the solar atmosphere. Such ejections are often thought to be signatures of the outflows expected in regions of fast magnetic reconnection. The 304 Å EUV line of helium, formed at around 10{sup 5} K, is found to be a reliable tracer of such phenomena, but the determination of physical parameters from such <span class="hlt">observations</span> is not straightforward. We have <span class="hlt">observed</span> a plasmoid ejection from an X-ray <span class="hlt">bright</span> point simultaneously at millimeter wavelengths with ALMA, atmore » EUV wavelengths with SDO /AIA, and in soft X-rays with Hinode /XRT. This paper reports the physical parameters of the plasmoid obtained by combining the radio, EUV, and X-ray data. As a result, we conclude that the plasmoid can consist either of (approximately) isothermal ∼10{sup 5} K plasma that is optically thin at 100 GHz, or a ∼10{sup 4} K core with a hot envelope. The analysis demonstrates the value of the additional temperature and density constraints that ALMA provides, and future science <span class="hlt">observations</span> with ALMA will be able to match the spatial resolution of space-borne and other high-resolution telescopes.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20180001946','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20180001946"><span>Preliminary Evaluation of Influence of Aerosols on the Simulation of <span class="hlt">Brightness</span> Temperature in the NASA's Goddard Earth <span class="hlt">Observing</span> System Atmospheric Data Assimilation System</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kim, Jong; Akella, Santha; da Silva, Arlindo M.; Todling, Ricardo; McCarty, William</p> <p>2018-01-01</p> <p>This document reports on preliminary results obtained when studying the impact of aerosols on the calculation of <span class="hlt">brightness</span> temperature (BT) for satellite infrared (IR) instruments that are currently assimilated in a 3DVAR configuration of Goddard Earth <span class="hlt">Observing</span> System (GEOS)-atmospheric data assimilation system (ADAS). A set of fifteen aerosol species simulated by the Goddard Chemistry Aerosol Radiation and Transport (GOCART) model is used to evaluate the influence of the aerosol fields on the Community Radiative Transfer Model (CRTM) calculations taking place in the <span class="hlt">observation</span> operators of the Gridpoint Statistical Interpolation (GSI) analysis system of GEOSADAS. Results indicate that taking aerosols into account in the BT calculation improves the fit to <span class="hlt">observations</span> over regions with significant amounts of dust. The cooling effect obtained with the aerosol-affected BT leads to a slight warming of the analyzed <span class="hlt">surface</span> temperature (by about 0:5oK) in the tropical Atlantic ocean (off northwest Africa), whereas the effect on the air temperature aloft is negligible. In addition, this study identifies a few technical issues to be addressed in future work if aerosol-affected BT are to be implemented in reanalysis and operational settings. The computational cost of applying CRTM aerosol absorption and scattering options is too high to justify their use, given the size of the benefits obtained. Furthermore, the differentiation between clouds and aerosols in GSI cloud detection procedures needs satisfactory revision.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015csss...18..665S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015csss...18..665S"><span>The <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Contribution of II Peg: A Comparison of TiO Band Analysis and Doppler Imaging</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Senavci, H. V.; O'Neal, D.; Hussain, G. A. J.; Barnes, J. R.</p> <p>2015-01-01</p> <p>We investigate the <span class="hlt">surface</span> <span class="hlt">brightness</span> contribution of the very well known active SB1 binary II Pegasi , to determine the star spot filling factor and the spot temperature parameters. In this context, we analyze 54 spectra of the system taken over 6 nights in September - October of 1996, using the 2.1m Otto Struve Telescope equipped with SES at the McDonald Observatory. We measure the spot temperatures and spot filling factors by fitting TiO molecular bands in this spectroscopic dataset, with model atmosphere approximation using ATLAS9 and with proxy stars obtained with the same instrument. The same dataset is then used to also produce <span class="hlt">surface</span> spot maps using the Doppler imaging technique. We compare the spot filling factors obtained with the two independent techniques in order to better characterise the spot properties of the system and to better assess the limitations inherent to both techniques. The results obtained from both techniques show that the variation of spot filling factor as a function of phase agree well with each other, while the amount of TiO and DI spot</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20100017871','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20100017871"><span>Optical Photometric <span class="hlt">Observations</span> of GEO Debris</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Seitzer, Patrick; Rodriquez-Cowardin, Heather M.; Barker, Edwin S.; Abercromby, Kira J.; Kelecy, Thomas M.; Horstman, Matt</p> <p>2010-01-01</p> <p>We report on a continuing program of optical photometric measurements of faint orbital debris at geosynchronous Earth orbit (GEO). These <span class="hlt">observations</span> can be compared with laboratory studies of actual spacecraft materials in an effort to determine what the faint debris at GEO may be. We have optical <span class="hlt">observations</span> from Cerro Tololo Inter-American Observatory (CTIO) in Chile of two samples of debris: 1. GEO objects discovered in a survey with the University of Michigan's 0.6-m aperture Curtis-Schmidt telescope MODEST (for Michigan Orbital DEbris Survey Telescope), and then followed up in real-time with the CTIO/SMARTS 0.9-m for orbits and photometry. Our goal is to determine 6 parameter orbits and measure colors for all objects fainter than R = 15 t11 magnitude that are discovered in the MODEST survey. 2. A smaller sample of high area to mass ratio (AMR) objects discovered independently, and acquired using predictions from orbits derived from independent tracking data collected days prior to the <span class="hlt">observations</span>. Our optical <span class="hlt">observations</span> in standard astronomical BVRI filters are done with either telescope, and with the telescope tracking the debris object at the object's angular rate. <span class="hlt">Observations</span> in different filters are obtained sequentially. We have obtained 71 calibrated sequences of R-B-V-I-R magnitudes. A total of 66 of these sequences have 3 or more good measurements in all filters (not contaminated by star streaks or in Earth's shadow). Most of these sequences show <span class="hlt">brightness</span> variations, but a small subset has <span class="hlt">observed</span> <span class="hlt">brightness</span> variations consistent with that expected from <span class="hlt">observational</span> errors alone. The majority of these stable objects are redder than a solar color in both B-R and R-I. There is no dependence on color with <span class="hlt">brightness</span>. For a smaller sample of objects we have <span class="hlt">observed</span> with synchronized CCD cameras on the two telescopes. The CTIO 0.9-m <span class="hlt">observes</span> in B, and MODEST in R. The CCD cameras are electronically linked together so that the start time and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SPIE10621E..04L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SPIE10621E..04L"><span><span class="hlt">Brightness</span> checkerboard lattice method for the calibration of the coaxial reverse Hartmann test</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Xinji; Hui, Mei; Li, Ning; Hu, Shinan; Liu, Ming; Kong, Lingqin; Dong, Liquan; Zhao, Yuejin</p> <p>2018-01-01</p> <p>The coaxial reverse Hartmann test (RHT) is widely used in the measurement of large aspheric <span class="hlt">surfaces</span> as an auxiliary method for interference measurement, because of its large dynamic range, highly flexible test with low frequency of <span class="hlt">surface</span> errors, and low cost. And the accuracy of the coaxial RHT depends on the calibration. However, the calibration process remains inefficient, and the signal-to-noise ratio limits the accuracy of the calibration. In this paper, <span class="hlt">brightness</span> checkerboard lattices were used to replace the traditional dot matrix. The <span class="hlt">brightness</span> checkerboard method can reduce the number of dot matrix projections in the calibration process, thus improving efficiency. An LCD screen displayed a <span class="hlt">brightness</span> checkerboard lattice, in which the brighter checkerboard and the darker checkerboard alternately arranged. Based on the image on the detector, the relationship between the rays at certain angles and the photosensitive positions of the detector coordinates can be obtained. And a differential de-noising method can effectively reduce the impact of noise on the measurement results. Simulation and experimentation proved the feasibility of the method. Theoretical analysis and experimental results show that the efficiency of the <span class="hlt">brightness</span> checkerboard lattices is about four times that of the traditional dot matrix, and the signal-to-noise ratio of the calibration is significantly improved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890033162&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890033162&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour"><span>Coronal <span class="hlt">bright</span> points in microwaves</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kundu, M. R.; Nitta, N.</p> <p>1988-01-01</p> <p>An excellent map of the quiet sun showing coronal <span class="hlt">bright</span> points at 20-cm wavelength was produced using the VLA on February 13, 1987. The locations of <span class="hlt">bright</span> points (BPs) were studied relative to features on the photospheric magnetogram and Ca K spectroheliogram. Most <span class="hlt">bright</span> points appearing in the full 5-hour synthesized map are associated with small bipolar structures on the photospheric magnetogram; and the brightest part of a BP tends to lie on the boundary of a supergranulation network. The <span class="hlt">bright</span> points exhibit rapid variations in intensity superposed on an apparently slow variation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AN....330..425M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AN....330..425M"><span>Follow-up <span class="hlt">observations</span> of Comet 17P/Holmes after its extreme outburst in <span class="hlt">brightness</span> end of October 2007</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mugrauer, M.; Hohle, M. M.; Ginski, C.; Vanko, M.; Freistetter, F.</p> <p>2009-05-01</p> <p>We present follow-up <span class="hlt">observations</span> of comet 17/P Holmes after its extreme outburst in <span class="hlt">brightness</span>, which occurred end of October 2007. We obtained 58 V-band images of the comet between October 2007 and February 2008, using the Cassegrain-Teleskop-Kamera (CTK) at the University Observatory Jena. We present precise astrometry of the comet, which yields its most recent Keplerian orbital elements. Furthermore, we show that the comet's coma expands quite linearly with a velocity of about 1650 km/s between October and December 2007. The photometric monitoring of comet 17/P Holmes shows that its photometric activity level decreased by about 5.9 mag within 105 days after its outburst. Based on <span class="hlt">observations</span> obtained with telescopes of the University Observatory Jena, which is operated by the Astrophysical Institute of the Friedrich-Schiller-University.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA12525.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA12525.html"><span>Faint Ring, <span class="hlt">Bright</span> Arc</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2010-01-12</p> <p>In this image taken by NASA Cassini spacecraft, the <span class="hlt">bright</span> arc in Saturn faint G ring contains a little something special. Although it cant be seen here, the tiny moonlet Aegaeon orbits within the <span class="hlt">bright</span> arc.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_15 --> <div id="page_16" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="301"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.P32B..01M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.P32B..01M"><span>Viewing Mercury's <span class="hlt">Surface</span>-bound Exosphere from Orbit: Eighteen Months of <span class="hlt">Observations</span> by the Mercury Atmospheric and <span class="hlt">Surface</span> Composition Spectrometer aboard the MESSENGER Spacecraft</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McClintock, W. E.; Benna, M.; Burger, M. H.; Cassidy, T.; Killen, R. M.; Merkel, A. W.; Sarantos, M.; Solomon, S. C.; Sprague, A. L.; Vervack, R. J.</p> <p>2012-12-01</p> <p> of 180°. UVVS also <span class="hlt">observes</span> H. It is less well studied than Ca, Mg, and Na because signal from the exospheric H is often contaminated by emission from interplanetary hydrogen and sunlight reflected from the <span class="hlt">surface</span>. O has also been detected near the subsolar point, but its emission is too weak for routine study. UVVS <span class="hlt">observations</span> also include wavelength scans for neutral species that are known or are predicted to be present in the <span class="hlt">surface</span> materials (e.g., Si, Al, S, Mn, Fe, and OH), but emissions from these species are not sufficiently <span class="hlt">bright</span> for detection with current operational scenarios. The UVVS team uses a variety of techniques to relate exosphere composition and structure to source processes, including tomographic inversion and Monte Carlo modeling. Correlations of Mercury's neutral exosphere composition and structure with direct measurements of the space environment from MESSENGER's Magnetometer (MAG) and Energetic Particle and Plasma Spectrometer (EPPS) provide further insight into source processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130014383','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130014383"><span><span class="hlt">Surface</span> Temperature Measurements from a Stator Vane Doublet in a Turbine Engine Afterburner Flame using Ultra-<span class="hlt">Bright</span> Cr-Doped GdAlO3 Thermographic Phosphor</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Eldridge, Jeffrey I.; Jenkins, Thomas P.; Allison, Stephen W.; Wolfe, Douglas E.; Howard, Robert P.</p> <p>2013-01-01</p> <p>Luminescence-based <span class="hlt">surface</span> temperature measurements from an ultra-<span class="hlt">bright</span> Cr-doped GdAlO3 perovskite (GAP:Cr) coating were successfully conducted on an air-film-cooled stator vane doublet exposed to the afterburner flame of a J85 test engine at University of Tennessee Space Institute (UTSI). The objective of the testing at UTSI was to demonstrate that reliable thermal barrier coating (TBC) <span class="hlt">surface</span> temperatures based on luminescence decay of a thermographic phosphor could be obtained from the <span class="hlt">surface</span> of an actual engine component in an aggressive afterburner flame environment and to address the challenges of a highly radiant background and high velocity gases. A high-pressure turbine vane doublet from a Honeywell TECH7000 turbine engine was coated with a standard electron-beam physical vapor deposited (EB-PVD) 200-m-thick TBC composed of yttria-stabilized zirconia (YSZ) onto which a 25-m-thick GAP:Cr thermographic phosphor layer was deposited by EB-PVD. The ultra-<span class="hlt">bright</span> broadband luminescence from the GAP:Cr thermographic phosphor is shown to offer the advantage of over an order-of-magnitude greater emission intensity compared to rare-earth-doped phosphors in the engine test environment. This higher emission intensity was shown to be very desirable for overcoming the necessarily restricted probe light collection solid angle and for achieving high signal-to-background levels. Luminescence-decay-based <span class="hlt">surface</span> temperature measurements varied from 500 to over 1000C depending on engine operating conditions and level of air film cooling.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720021675','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720021675"><span>Microwave <span class="hlt">brightness</span> temperature of a windblown sea</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hall, F. G.</p> <p>1972-01-01</p> <p>A mathematical model is developed for the apparent temperature of the sea at all microwave frequencies. The model is a numerical model in which both the clear water structure and white water are accounted for as a function of wind speed. The model produces results similar to Stogryn's model at 19.35 GHz for wind speeds less than 8 m/sec; it can use radiosonde data to calculate atmospheric effects and can incorporate an empirically determined antenna gain pattern. The corresponding computer program is of modular design and the logic of the main program is capable of treating a horizontally inhomogeneous <span class="hlt">surface</span> or atmosphere. It is shown that a variation of microwave <span class="hlt">brightness</span> temperature with zenith angle is necessary to produce the wind sensitivity of the horizontally polarized <span class="hlt">brightness</span> temperature; the variation of sky temperature with frequency is sufficient to produce a frequency dependent wind sensitivity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DMP.B4004K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DMP.B4004K"><span>Experimental <span class="hlt">observation</span> of spatial quantum noise reduction below the standard quantum limit with <span class="hlt">bright</span> twin beams of light</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kumar, Ashok; Nunley, Hayden; Marino, Alberto</p> <p>2016-05-01</p> <p>Quantum noise reduction (QNR) below the standard quantum limit (SQL) has been a subject of interest for the past two to three decades due to its wide range of applications in quantum metrology and quantum information processing. To date, most of the attention has focused on the study of QNR in the temporal domain. However, many areas in quantum optics, specifically in quantum imaging, could benefit from QNR not only in the temporal domain but also in the spatial domain. With the use of a high quantum efficiency electron multiplier charge coupled device (EMCCD) camera, we have <span class="hlt">observed</span> spatial QNR below the SQL in <span class="hlt">bright</span> narrowband twin light beams generated through a four-wave mixing (FWM) process in hot rubidium atoms. Owing to momentum conservation in this process, the twin beams are momentum correlated. This leads to spatial quantum correlations and spatial QNR. Our preliminary results show a spatial QNR of over 2 dB with respect to the SQL. Unlike previous results on spatial QNR with faint and broadband photon pairs from parametric down conversion (PDC), we demonstrate spatial QNR with spectrally and spatially narrowband <span class="hlt">bright</span> light beams. The results obtained will be useful for atom light interaction based quantum protocols and quantum imaging. Work supported by the W.M. Keck Foundation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.5395L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.5395L"><span>Investigating the <span class="hlt">surface</span> <span class="hlt">brightness</span> profiles, ejected mass and speed from the outburst events of comet 67P/Churyumov-Gerasmenko</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lin, Zhong-Yi; Vincent, Jean-Baptiste; A'Hearn, Mike; Lara, Luisa; Knollenberg, Joerg; Ip, Wing-Huen; Osiris Team</p> <p>2016-04-01</p> <p>The OSIRIS (Optical, Spectroscopic, and Infrared Remote Imaging System) WAC and NAC camera onboard the ESA Rosetta spacecraft orbiting 67P/Churyumov-Gersimenko has captured a lot of outbursts since July, 2015. Most of their source regions were located at southern hemisphere of comet C-G. Including the March- and perihelion-outbursts, the detected events show a variety of morphological features (i.e. broad fan, collimated jet and so on). In this work, we investigate these events and characterize the physical properties, including the <span class="hlt">surface</span> <span class="hlt">brightness</span> profiles, ejected mass and speed if there were two or more images acquired by the same filter during the outburst timeframe.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24060566','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24060566"><span>Direct <span class="hlt">observation</span> of antisite defects in LiCoPO4 cathode materials by annular dark- and <span class="hlt">bright</span>-field electron microscopy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Truong, Quang Duc; Devaraju, Murukanahally Kempaiah; Tomai, Takaaki; Honma, Itaru</p> <p>2013-10-23</p> <p>LiCoPO4 cathode materials have been synthesized by a sol-gel route. X-ray diffraction analysis confirmed that LiCoPO4 was well-crystallized in an orthorhombic structure in the Pmna space group. From the high-resolution transmission electron microscopy (HR-TEM) image, the lattice fringes of {001} and {100} are well-resolved. The HR-TEM image and selected area electron diffraction pattern reveal the highly crystalline nature of LiCoPO4 having an ordered olivine structure. The atom-by-atom structure of LiCoPO4 olivine has been <span class="hlt">observed</span>, for the first time, using high-angle annular dark-field (HAADF) and annual <span class="hlt">bright</span>-field scanning transmission electron microscopy. We <span class="hlt">observed</span> the <span class="hlt">bright</span> contrast in Li columns in the HAADF images and strong contrast in the ABF images, directly indicating the antisite exchange defects in which Co atoms partly occupy the Li sites. The LiCoPO4 cathode materials delivered an initial discharge capacity of 117 mAh/g at a C/10 rate with moderate cyclic performance. The discharge profile of LiCoPO4 shows a plateau at 4.75 V, revealing its importance as a potentially high-voltage cathode. The direct visualization of atom-by-atom structure in this work represents important information for the understanding of the structure of the active cathode materials for Li-ion batteries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20070034151','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20070034151"><span>Thin Sea-Ice Thickness as Inferred from Passive Microwave and In Situ <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Naoki, Kazuhiro; Ukita, Jinro; Nishio, Fumihiko; Nakayama, Masashige; Comiso, Josefino C.; Gasiewski, Al</p> <p>2007-01-01</p> <p>Since microwave radiometric signals from sea-ice strongly reflect physical conditions of a layer near the ice <span class="hlt">surface</span>, a relationship of <span class="hlt">brightness</span> temperature with thickness is possible especially during the early stages of ice growth. Sea ice is most saline during formation stage and as the salinity decreases with time while at the same time the thickness of the sea ice increases, a corresponding change in the dielectric properties and hence the <span class="hlt">brightness</span> temperature may occur. This study examines the extent to which the relationships of thickness with <span class="hlt">brightness</span> temperature (and with emissivity) hold for thin sea-ice, approximately less than 0.2 -0.3 m, using near concurrent measurements of sea-ice thickness in the Sea of Okhotsk from a ship and passive microwave <span class="hlt">brightness</span> temperature data from an over-flying aircraft. The results show that the <span class="hlt">brightness</span> temperature and emissivity increase with ice thickness for the frequency range of 10-37 GHz. The relationship is more pronounced at lower frequencies and at the horizontal polarization. We also established an empirical relationship between ice thickness and salinity in the layer near the ice <span class="hlt">surface</span> from a field experiment, which qualitatively support the idea that changes in the near-<span class="hlt">surface</span> brine characteristics contribute to the <span class="hlt">observed</span> thickness-<span class="hlt">brightness</span> temperature/emissivity relationship. Our results suggest that for thin ice, passive microwave radiometric signals contain, ice thickness information which can be utilized in polar process studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999A%26A...344..342L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999A%26A...344..342L"><span>Looking at the <span class="hlt">bright</span> side of the rho Ophiuchi dark cloud. Far infrared spectrophotometric <span class="hlt">observations</span> of the rho Oph cloud with the ISO</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liseau, R.; White, G. J.; Larsson, B.; Sidher, S.; Olofsson, G.; Kaas, A.; Nordh, L.; Caux, E.; Lorenzetti, D.; Molinari, S.; Nisini, B.; Sibille, F.</p> <p>1999-04-01</p> <p>We present far infrared (45-195 mu m) spectrophotometric <span class="hlt">observations</span> with the Iso-Lws of the active star forming rho Oph main cloud (L 1688). The [C ii] 158 mu m and [O i] 63 mu m lines were detected at each of the 33 positions <span class="hlt">observed</span>, whereas the [O i] 145 mu m line was clearly seen toward twelve. The principal <span class="hlt">observational</span> result is that the [C ii] 158 mu m line fluxes exhibit a clear correlation with projected distance from the dominant stellar source in the field (HD 147889). We interpret this in terms of Pdr-type emission from the <span class="hlt">surface</span> layers of the rho Ophc. The <span class="hlt">observed</span> [C ii] 158 mu m/[O i] 63 mu m flux ratios are larger than unity everywhere. A comparison of the [C ii] 158 mu m line emission and the Fir dust continuum fluxes yields estimates of the efficiency at which the gas in the cloud converts stellar to [C ii] 158 mu m photons (chi_ {_C II},>_{ ~ },0.5%). We first develop an empirical model, which provides us with a three dimensional view of the far and <span class="hlt">bright</span> side of the dark rho Ophc, showing that the cloud <span class="hlt">surface</span> towards the putative energy source is concave. This model also yields quantitative estimates of the incident flux of ultraviolet radiation (G_0 ~ , \\powten{1} - \\powten{2}) and of the degree of clumpiness/texture of the cloud <span class="hlt">surface</span> (filling of the 80({') '} beam ~ ,0.2). Subsequently, we use theoretical models of Pdr s to derive the particle density, n(H), and the temperature structures, for T_gas and T_dust, in the <span class="hlt">surface</span> layers of the rho Ophc. T_gas is relatively low, ~ ,60 K, but higher than T_dust ( ~ ,30 K), and densities are generally found within the interval (1-3) \\powten{4} cm(-3) . These Pdr models are moderately successful in explaining the Lws <span class="hlt">observations</span>. They correctly predict the [O i] 63 mu m and [C ii] 158 mu m line intensities and the <span class="hlt">observed</span> absence of any molecular line emission. The models do fail, however, to reproduce the <span class="hlt">observed</span> small [O i] 63 mu m/[O i] 145 mu m ratios. We examine several possible</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28241664','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28241664"><span>Atmospheric correction for retrieving ground <span class="hlt">brightness</span> temperature at commonly-used passive microwave frequencies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Han, Xiao-Jing; Duan, Si-Bo; Li, Zhao-Liang</p> <p>2017-02-20</p> <p>An analysis of the atmospheric impact on ground <span class="hlt">brightness</span> temperature (Tg) is performed for numerous land <span class="hlt">surface</span> types at commonly-used frequencies (i.e., 1.4 GHz, 6.93 GHz, 10.65 GHz, 18.7 GHz, 23.8 GHz, 36.5 GHz and 89.0 GHz). The results indicate that the atmosphere has a negligible impact on Tg at 1.4 GHz for land <span class="hlt">surfaces</span> with emissivities greater than 0.7, at 6.93 GHz for land <span class="hlt">surfaces</span> with emissivities greater than 0.8, and at 10.65 GHz for land <span class="hlt">surfaces</span> with emissivities greater than 0.9 if a root mean square error (RMSE) less than 1 K is desired. To remove the atmospheric effect on Tg, a generalized atmospheric correction method is proposed by parameterizing the atmospheric transmittance τ and upwelling atmospheric <span class="hlt">brightness</span> temperature Tba↑. Better accuracies with Tg RMSEs less than 1 K are achieved at 1.4 GHz, 6.93 GHz, 10.65 GHz, 18.7 GHz and 36.5 GHz, and worse accuracies with RMSEs of 1.34 K and 4.35 K are obtained at 23.8 GHz and 89.0 GHz, respectively. Additionally, a simplified atmospheric correction method is developed when lacking sufficient input data to perform the generalized atmospheric correction method, and an emissivity-based atmospheric correction method is presented when the emissivity is known. Consequently, an appropriate atmospheric correction method can be selected based on the available data, frequency and required accuracy. Furthermore, this study provides a method to estimate τ and Tba↑ of different frequencies using the atmospheric parameters (total water vapor content in <span class="hlt">observation</span> direction Lwv, total cloud liquid water content Lclw and mean temperature of cloud Tclw), which is important for simultaneously determining the land <span class="hlt">surface</span> parameters using multi-frequency passive microwave satellite data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=257635','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=257635"><span>L band <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> over a corn canopy during the entire growth cycle</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>During a field campaign covering the 2002 corn growing season, a dual polarized tower mounted L-band (1.4 GHz) radiometer (LRAD) provided <span class="hlt">brightness</span> temperature (T¬B) measurements at preset intervals, incidence and azimuth angles. These radiometer measurements were supported by an extensive characte...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01058&hterms=Dark+web&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDark%2Bweb','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01058&hterms=Dark+web&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDark%2Bweb"><span><span class="hlt">Bright</span> and Dark Slopes on Ganymede</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1997-01-01</p> <p>Ridges on the edge of Ganymede's north polar cap show <span class="hlt">bright</span> east-facing slopes and dark west-facing slopes with troughs of darker material below the larger ridges. North is to the top. The <span class="hlt">bright</span> slopes may be due to grain size differences, differences in composition between the original <span class="hlt">surface</span> and the underlying material, frost deposition, or illumination effects. The large 2.4 kilometer (1.5 mile) diameter crater in this image shows frost deposits located on the north-facing rim slope, away from the sun. A smaller 675 meter (2200 foot) diameter crater in the center of the image is surrounded by a <span class="hlt">bright</span> deposit which may be ejecta from the impact. Ejecta deposits such as this are uncommon for small craters on Ganymede. This image measures 18 by 19 kilometers (11 by 12 miles) and has a resolution of 45 meters (148 feet) per pixel. NASA's Galileo spacecraft obtained this image on September 6, 1996 during its second orbit around Jupiter.<p/>The Jet Propulsion Laboratory, Pasadena, CA manages the Galileo mission for NASA's Office of Space Science, Washington, DC. JPL is an operating division of California Institute of Technology (Caltech).<p/>This image and other images and data received from Galileo are posted on the World Wide Web, on the Galileo mission home page at URL http://galileo.jpl.nasa.gov. Background information and educational context for the images can be found at URL http://www.jpl.nasa.gov/galileo/sepo</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AAS...21349111M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AAS...21349111M"><span>Light and Velocity Variability in Seven <span class="hlt">Bright</span> Proto-Planetary Nebulae</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McGuire, Ryan</p> <p>2009-01-01</p> <p>Light and Velocity Variability in Seven <span class="hlt">Bright</span> Proto-Planetary Nebulae R.B. McGuire, C.M. Steele, B.J. Hrivnak, W. Lu, D. Bohlender, C.D. Scarfe We present new contemporaneous light and velocity <span class="hlt">observations</span> of seven proto-planetary nebulae obtained over the past two years. Proto-planetary nebulae are objects evolving between the AGB and planetary nebula phases. In these seven objects, the central star is <span class="hlt">bright</span> (V= 7-10), surrounded by a faint nebula. We knew from past monitoring that the light from each of these varied by a few tenths of a magnitude over intervals of 30-150 days and that the velocity varied by 10 km/s. These appear to be due to pulsation. With these new contemporaneous <span class="hlt">observations</span>, we are able to measure the correlation between the <span class="hlt">brightness</span>, color, and velocity, which will constrain the pulsation models. This is an ongoing project with the light monitoring being carried out with the Valparaiso University 0.4 m telescope and CCD camera and the radial velocity <span class="hlt">observations</span> being carried out with the Dominion Astrophysical Observatory 1.8 m telescope and spectrograph. This research is partially supported by NSF grant 0407087 and the Indiana Space Grant Consortium.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017hst..prop15186T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017hst..prop15186T"><span>Enabling HST UV Exploration of the Low <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Universe: A Pilot Study with the WFC3 X Filter Set</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thilker, David</p> <p>2017-08-01</p> <p>We request 17 orbits to conduct a pilot study to examine the effectiveness of the WFC3/UVIS F300X filter for studying fundamental problems in star formation in the low density regime. In principle, the broader bandpass and higher throughput of F300X can halve the required <span class="hlt">observing</span> time relative to F275W, the filter of choice for studying young stellar populations in nearby galaxies. Together with F475W and F600LP, this X filter set may be as effective as standard UVIS broadband filters for characterizing the physical properties of such populations. We will <span class="hlt">observe</span> 5 low <span class="hlt">surface</span> <span class="hlt">brightness</span> targets with a range of properties to test potential issues with F300X: the red tail to 4000A and a red leak beyond, ghosts, and the wider bandpass. Masses and ages of massive stars, young star clusters, and clumps derived from photometry from the X filter set will be compared with corresponding measurements from standard filters. Beyond testing, our program will provide the first sample spanning a range of LSB galaxy properties for which HST UV imaging will be obtained, and a glimpse into the ensemble properties of the quanta of star formation in these strange environments. The increased <span class="hlt">observing</span> efficiency would make more tractable programs which require several tens to hundreds of orbits to aggregate sufficient numbers of massive stars, young star clusters, and clumps to build statistical samples. We are hopeful that our pilot <span class="hlt">observations</span> will broadly enable high-resolution UV imaging exploration of the low density frontier of star formation while HST is still in good health.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JASTP.129....1Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JASTP.129....1Y"><span>Long-delayed <span class="hlt">bright</span> dancing sprite with large Horizontal displacement from its parent flash</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Jing; Lu, Gaopeng; Lee, Li-Jou; Feng, Guili</p> <p>2015-07-01</p> <p>We reported in this paper the <span class="hlt">observation</span> of a very <span class="hlt">bright</span> long-delayed dancing sprite with distinct horizontal displacement from its parent stroke. The dancing sprite lasted only 60 ms, and the morphology consisted of three fields with two slim dim sprite elements in the first two fields and a very <span class="hlt">bright</span> large element in the third field, different from other <span class="hlt">observations</span> where the dancing sprites usually contained multiple elements over a longer time interval, and the sprite shape and <span class="hlt">brightness</span> in the video field are often similar to the previous fields. The <span class="hlt">bright</span> sprite was displaced at least 38 km from its parent cloud-to-ground (CG) stroke and occurred over comparatively higher cloud top region. The parent flash of this compact dancing sprite was of positive polarity, with only one return stroke (approximately +24 kA) and obvious continuing current process, and the charge moment change of stroke was small (barely above the threshold for sprite production). All the sprite elements occurred during the continuing current stage, and the <span class="hlt">bright</span> long-delayed sprite element induced a considerable current pulse. The dancing feature of this sprite may be linked to the electrical charge structure, dynamics and microphysics of parent storm, and the inferred development of parent CG flash was consistent with previous very high-frequency (VHF) <span class="hlt">observations</span> of lightning in the same region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApJ...820L..10J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApJ...820L..10J"><span>Extreme <span class="hlt">Brightness</span> Temperatures and Refractive Substructure in 3C273 with RadioAstron</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Johnson, Michael D.; Kovalev, Yuri Y.; Gwinn, Carl R.; Gurvits, Leonid I.; Narayan, Ramesh; Macquart, Jean-Pierre; Jauncey, David L.; Voitsik, Peter A.; Anderson, James M.; Sokolovsky, Kirill V.; Lisakov, Mikhail M.</p> <p>2016-03-01</p> <p>Earth-space interferometry with RadioAstron provides the highest direct angular resolution ever achieved in astronomy at any wavelength. RadioAstron detections of the classic quasar 3C 273 on interferometric baselines up to 171,000 km suggest <span class="hlt">brightness</span> temperatures exceeding expected limits from the “inverse-Compton catastrophe” by two orders of magnitude. We show that at 18 cm, these estimates most likely arise from refractive substructure introduced by scattering in the interstellar medium. We use the scattering properties to estimate an intrinsic <span class="hlt">brightness</span> temperature of 7× {10}12 {{K}}, which is consistent with expected theoretical limits, but which is ˜15 times lower than estimates that neglect substructure. At 6.2 cm, the substructure influences the measured values appreciably but gives an estimated <span class="hlt">brightness</span> temperature that is comparable to models that do not account for the substructure. At 1.35 {{cm}}, the substructure does not affect the extremely high inferred <span class="hlt">brightness</span> temperatures, in excess of {10}13 {{K}}. We also demonstrate that for a source having a Gaussian <span class="hlt">surface</span> <span class="hlt">brightness</span> profile, a single long-baseline estimate of refractive substructure determines an absolute minimum <span class="hlt">brightness</span> temperature, if the scattering properties along a given line of sight are known, and that this minimum accurately approximates the apparent <span class="hlt">brightness</span> temperature over a wide range of total flux densities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19910004789','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19910004789"><span>A comparison of UV <span class="hlt">surface</span> <span class="hlt">brightness</span> and HI <span class="hlt">surface</span> densities for spiral galaxies</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Federman, S. R.; Strom, C.</p> <p>1990-01-01</p> <p>Shaya and Federman (1987) suggested that the ambient ultraviolet flux at 1000 A permeating a spiral galaxy controls the neutral hydrogen (HI) <span class="hlt">surface</span> density in the galaxy. They found that the atomic envelopes surrounding small molecular clouds, because of their great number, provide the major contribution to the HI <span class="hlt">surface</span> density over the stellar disk. The increase in HI <span class="hlt">surface</span> density with later Hubble types was ascribed to the stronger UV fields from more high-mass stars in later Hubble types. These hypotheses are based on the <span class="hlt">observations</span> of nearby diffuse interstellar clouds, which show a sharp atomic-to-molecular transition (Savage et al. 1977), and on the theoretical framework introduced by Federman, Glassgold, and Kwan (1979). Atomic envelopes around interstellar clouds in the solar neighborhood arise when a steady state is reached between photodissociation of H2 and the formation of H2 on grains. The photodissociation process involves photons with wavelengths between 912 A and 1108 A. Shaya and Federman used H-alpha flux as an approximate measure for the far UV flux and made their comparisons based on averages over Hubble type. Here, researchers compare, on an individual basis, UV data obtained with space-borne and balloon-borne instruments for galaxies with measurements of HI <span class="hlt">surface</span> density (Warmels 1988a, b). The comparisons substantiate the conclusion of Shaya and Federman that the far UV field controls the HI content of spiral galaxies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22340258-andromeda-m31-optical-infrared-disk-survey-insights-wide-field-near-ir-surface-photometry','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22340258-andromeda-m31-optical-infrared-disk-survey-insights-wide-field-near-ir-surface-photometry"><span>Andromeda (M31) optical and infrared disk survey. I. Insights in wide-field near-IR <span class="hlt">surface</span> photometry</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Sick, Jonathan; Courteau, Stéphane; Cuillandre, Jean-Charles</p> <p></p> <p>We present wide-field near-infrared J and K{sub s} images of the Andromeda Galaxy (M31) taken with WIRCam at the Canada-France-Hawaii Telescope as part of the Andromeda Optical and Infrared Disk Survey. This data set allows simultaneous <span class="hlt">observations</span> of resolved stars and near-infrared (NIR) <span class="hlt">surface</span> <span class="hlt">brightness</span> across M31's entire bulge and disk (within R = 22 kpc), permitting a direct test of the stellar composition of near-infrared light in a nearby galaxy. Here we develop NIR <span class="hlt">observation</span> and reduction methods to recover a uniform <span class="hlt">surface</span> <span class="hlt">brightness</span> map across the 3° × 1° disk of M31 with 27 WIRCam fields. Two sky-targetmore » nodding strategies are tested, and we find that strictly minimizing sky sampling latency cannot improve background subtraction accuracy to better than 2% of the background level due to spatio-temporal variations in the NIR skyglow. We fully describe our WIRCam reduction pipeline and advocate using flats built from night-sky images over a single night, rather than dome flats that do not capture the WIRCam illumination field. Contamination from scattered light and thermal background in sky flats has a negligible effect on the <span class="hlt">surface</span> <span class="hlt">brightness</span> shape compared to the stochastic differences in background shape between sky and galaxy disk fields, which are ∼0.3% of the background level. The most dramatic calibration step is the introduction of scalar sky offsets to each image that optimizes <span class="hlt">surface</span> <span class="hlt">brightness</span> continuity. Sky offsets reduce the mean <span class="hlt">surface</span> <span class="hlt">brightness</span> difference between <span class="hlt">observation</span> blocks from 1% to <0.1% of the background level, though the absolute background level remains statistically uncertain to 0.15% of the background level. We present our WIRCam reduction pipeline and performance analysis to give specific recommendations for the improvement of NIR wide-field imaging methods.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21574786-response-granulation-small-scale-bright-features-quiet-sun','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21574786-response-granulation-small-scale-bright-features-quiet-sun"><span>RESPONSE OF GRANULATION TO SMALL-SCALE <span class="hlt">BRIGHT</span> FEATURES IN THE QUIET SUN</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Andic, A.; Chae, J.; Goode, P. R.</p> <p>2011-04-10</p> <p>We detected 2.8 <span class="hlt">bright</span> points (BPs) per Mm{sup 2} in the quiet Sun with the New Solar Telescope at Big Bear Solar Observatory, using the TiO 705.68 nm spectral line at an angular resolution {approx}0.''1 to obtain a 30 minute data sequence. Some BPs formed knots that were stable in time and influenced the properties of the granulation pattern around them. The <span class="hlt">observed</span> granulation pattern within {approx}3'' of knots presents smaller granules than those <span class="hlt">observed</span> in a normal granulation pattern, i.e., around the knots a suppressed convection is detected. <span class="hlt">Observed</span> BPs covered {approx}5% of the solar <span class="hlt">surface</span> and were notmore » homogeneously distributed. BPs had an average size of 0.''22, they were detectable for 4.28 minutes on average, and had an averaged contrast of 0.1% in the deep red TiO spectral line.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMSA23A4051F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMSA23A4051F"><span>Geomagnetically conjugate <span class="hlt">observations</span> of ionospheric and thermospheric variations accompanied with a midnight <span class="hlt">brightness</span> wave at low latitudes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fukushima, D.; Shiokawa, K.; Otsuka, Y.; Kubota, M.; Yokoyama, T.; Nishioka, M.; Komonjinda, S.; Yatini, C. Y.</p> <p>2014-12-01</p> <p>A midnight <span class="hlt">brightness</span> wave (MBW) is the phenomenon that the OI (630-nm) airglow enhancement propagates poleward once at local midnight. In this study, we first conducted geomagnetically conjugate <span class="hlt">observations</span> of 630nm airglow for an MBW at conjugate stations. An airglow enhancement which is considered to be an MBW was <span class="hlt">observed</span> in the 630-nm airglow images at Kototabang, Indonesia (geomagnetic latitude (MLAT): 10.0S) at around local midnight from 1540 to 1730 UT (from 2240 to 2430 LT) on 7 February 2011. This MBW was propagating south-southwestward, which is geomagnetically poleward, with a velocity of 290 m/s. However, similar wave was not <span class="hlt">observed</span> in the 630-nm airglow images at Chiang Mai, Thailand (MLAT: 8.9N), which is close to being conjugate point of Kototabang. This result indicates that the MBW does not have geomagnetic conjugacy. We simultaneously <span class="hlt">observed</span> thermospheric neutral winds <span class="hlt">observed</span> by a co-located Fabry-Perot interferometer at Kototabang. The <span class="hlt">observed</span> meridional winds turned from northward (geomagnetically equatorward) to southward (geomagnetically poleward) just before the MBW was <span class="hlt">observed</span>. The bottomside ionospheric heights <span class="hlt">observed</span> by ionosondes rapidly decreased at Kototabang and slightly increased at Chiang Mai simultaneously with the MBW passage. In the presentation, we discuss the MBW generation by the <span class="hlt">observed</span> poleward neutral winds at Kototabang, and the cause of the coinciding small height increase at Chiang Mai by the polarization electric field inside the <span class="hlt">observed</span> MBW at Kototabang.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFM.P24A..05T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFM.P24A..05T"><span><span class="hlt">Bright</span> Fans in Mars Cryptic Region Caused by Adiabatic Cooling of CO2 Gas Jets.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Titus, T. N.; Kieffer, H. H.; Langevin, Y.; Murchie, S.; Seelos, F.; Vincendon, M.</p> <p>2007-12-01</p> <p>Over the last decade, <span class="hlt">observations</span> of the retreat of the southern seasonal cap of Mars have revealed the presence of exotic processes within an area now informally referred to as the cryptic region. The appearance of dark spots, fans, blotches, and halos have been a "hot" topic of scientific discussion since they were first <span class="hlt">observed</span> by the Mars Global Surveyor (MGS) Mars Orbiter Camera (MOC) [Malin et al., 1998]. Further <span class="hlt">observations</span> by the Mars Odyssey (ODY) Thermal Emission Imaging System (THEMIS) showed that the dark features remained cold throughout the early-to-mid spring, suggesting that these features were either CO2 ice or were in thermal contact with CO2 ice [Kieffer et al., 2006]. In this paper, we present <span class="hlt">observations</span> in the near-infrared at spatial resolutions that have previously been unavailable. We present further evidence that many of these features in the cryptic region are the result of cold jets, as first described by Kieffer [2000, 2007]. The adiabatic cooling of gas spewing downwind from the jets produces CO2 frost, thus forming the <span class="hlt">bright</span> fans. The <span class="hlt">bright</span> fans appear to be devoid of H2O ice, thus further supporting the hypothesis that they are formed from the downwind settling of CO2 frost. In some areas, the <span class="hlt">bright</span> fans are adjacent to dark fans and appear to start from common vertices, while in other areas, <span class="hlt">bright</span> fan-like deposits occur without the strong presence of dark fans. References: Kieffer, H.H. (2000) Annual Punctuated CO2 Slab-Ice and Jets on Mars, International Conference on Mars Polar Science and Exploration, p. 93. Kieffer, H.H. et al. (2006) Nature, 442,793-796. Kieffer, H.H. (2007) JGR, in press. Malin, M.C., M.H. Carr, G.E. Danielson, M.E. Davies, W.K. Hartmann, A.P. Ingersoll, P.B. James, H. Masursky, A.S. McEwen, L.A. Soderblom, P. Thomas, J. Veverka, M.A. Caplinger, M.A. Ravine, and T.A. Soulanille (1998) Early views of the Martian <span class="hlt">surface</span> from the Mars orbiter camera of Mars global surveyor, Science, 279, 1681-1685.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_16 --> <div id="page_17" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="321"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080013171&hterms=Transformation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DTransformation','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080013171&hterms=Transformation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DTransformation"><span>A New Bond Albedo for Performing Orbital Debris <span class="hlt">Brightness</span> to Size Transformations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mulrooney, Mark K.; Matney, Mark J.</p> <p>2008-01-01</p> <p>We have developed a technique for estimating the intrinsic size distribution of orbital debris objects via optical measurements alone. The process is predicated on the empirically <span class="hlt">observed</span> power-law size distribution of debris (as indicated by radar RCS measurements) and the log-normal probability distribution of optical albedos as ascertained from phase (Lambertian) and range-corrected telescopic <span class="hlt">brightness</span> measurements. Since the <span class="hlt">observed</span> distribution of optical <span class="hlt">brightness</span> is the product integral of the size distribution of the parent [debris] population with the albedo probability distribution, it is a straightforward matter to transform a given distribution of optical <span class="hlt">brightness</span> back to a size distribution by the appropriate choice of a single albedo value. This is true because the integration of a powerlaw with a log-normal distribution (Fredholm Integral of the First Kind) yields a Gaussian-blurred power-law distribution with identical power-law exponent. Application of a single albedo to this distribution recovers a simple power-law [in size] which is linearly offset from the original distribution by a constant whose value depends on the choice of the albedo. Significantly, there exists a unique Bond albedo which, when applied to an <span class="hlt">observed</span> <span class="hlt">brightness</span> distribution, yields zero offset and therefore recovers the original size distribution. For physically realistic powerlaws of negative slope, the proper choice of albedo recovers the parent size distribution by compensating for the <span class="hlt">observational</span> bias caused by the large number of small objects that appear anomalously large (<span class="hlt">bright</span>) - and thereby skew the small population upward by rising above the detection threshold - and the lower number of large objects that appear anomalously small (dim). Based on this comprehensive analysis, a value of 0.13 should be applied to all orbital debris albedo-based <span class="hlt">brightness</span>-to-size transformations regardless of data source. Its prima fascia genesis, derived and constructed</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870027099&hterms=microwaves+water+structure&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dmicrowaves%2Bwater%2Bstructure','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870027099&hterms=microwaves+water+structure&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dmicrowaves%2Bwater%2Bstructure"><span>Satellite microwave and in situ <span class="hlt">observations</span> of the Weddell Sea ice cover and its marginal ice zone</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Comiso, J. C.; Sullivan, C. W.</p> <p>1986-01-01</p> <p>The radiative and physical characteristics of the Weddell Sea ice cover and its marginal ice zone are analyzed using multichannel satellite passive microwave data and ship and helicopter <span class="hlt">observations</span> obtained during the 1983 Antarctic Marine Ecosystem Research. Winter and spring <span class="hlt">brightness</span> temperatures are examined; spatial variability in the <span class="hlt">brightness</span> temperatures of consolidated ice in winter and spring cyclic increases and decrease in <span class="hlt">brightness</span> temperatures of consolidated ice with an amplitude of 50 K at 37 GHz and 20 K at 18 GHz are <span class="hlt">observed</span>. The roles of variations in air temperature and <span class="hlt">surface</span> characteristics in the variability of spring <span class="hlt">brightness</span> temperatures are investigated. Ice concentrations are derived using the frequency and polarization techniques, and the data are compared with the helicopter and ship <span class="hlt">observations</span>. Temporal changes in the ice margin structure and the mass balance of fresh water and of biological features of the marginal ice zone are studied.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27333609','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27333609"><span>Automated Adaptive <span class="hlt">Brightness</span> in Wireless Capsule Endoscopy Using Image Segmentation and Sigmoid Function.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shrestha, Ravi; Mohammed, Shahed K; Hasan, Md Mehedi; Zhang, Xuechao; Wahid, Khan A</p> <p>2016-08-01</p> <p>Wireless capsule endoscopy (WCE) plays an important role in the diagnosis of gastrointestinal (GI) diseases by capturing images of human small intestine. Accurate diagnosis of endoscopic images depends heavily on the quality of captured images. Along with image and frame rate, <span class="hlt">brightness</span> of the image is an important parameter that influences the image quality which leads to the design of an efficient illumination system. Such design involves the choice and placement of proper light source and its ability to illuminate GI <span class="hlt">surface</span> with proper <span class="hlt">brightness</span>. Light emitting diodes (LEDs) are normally used as sources where modulated pulses are used to control LED's <span class="hlt">brightness</span>. In practice, instances like under- and over-illumination are very common in WCE, where the former provides dark images and the later provides <span class="hlt">bright</span> images with high power consumption. In this paper, we propose a low-power and efficient illumination system that is based on an automated <span class="hlt">brightness</span> algorithm. The scheme is adaptive in nature, i.e., the <span class="hlt">brightness</span> level is controlled automatically in real-time while the images are being captured. The captured images are segmented into four equal regions and the <span class="hlt">brightness</span> level of each region is calculated. Then an adaptive sigmoid function is used to find the optimized <span class="hlt">brightness</span> level and accordingly a new value of duty cycle of the modulated pulse is generated to capture future images. The algorithm is fully implemented in a capsule prototype and tested with endoscopic images. Commercial capsules like Pillcam and Mirocam were also used in the experiment. The results show that the proposed algorithm works well in controlling the <span class="hlt">brightness</span> level accordingly to the environmental condition, and as a result, good quality images are captured with an average of 40% <span class="hlt">brightness</span> level that saves power consumption of the capsule.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9910E..1AY','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9910E..1AY"><span>An optical to IR sky <span class="hlt">brightness</span> model for the LSST</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yoachim, Peter; Coughlin, Michael; Angeli, George Z.; Claver, Charles F.; Connolly, Andrew J.; Cook, Kem; Daniel, Scott; Ivezić, Željko; Jones, R. Lynne; Petry, Catherine; Reuter, Michael; Stubbs, Christopher; Xin, Bo</p> <p>2016-07-01</p> <p>To optimize the <span class="hlt">observing</span> strategy of a large survey such as the LSST, one needs an accurate model of the night sky emission spectrum across a range of atmospheric conditions and from the near-UV to the near-IR. We have used the ESO SkyCalc Sky Model Calculator1, 2 to construct a library of template spectra for the Chilean night sky. The ESO model includes emission from the upper and lower atmosphere, scattered starlight, scattered moonlight, and zodiacal light. We have then extended the ESO templates with an empirical fit to the twilight sky emission as measured by a Canon all-sky camera installed at the LSST site. With the ESO templates and our twilight model we can quickly interpolate to any arbitrary sky position and date and return the full sky spectrum or <span class="hlt">surface</span> <span class="hlt">brightness</span> magnitudes in the LSST filter system. Comparing our model to all-sky <span class="hlt">observations</span>, we find typical residual RMS values of +/-0.2-0.3 magnitudes per square arcsecond.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980169266','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980169266"><span><span class="hlt">Surface</span> <span class="hlt">Brightness</span> Profiles and Energetics of Intracluster Gas in Cool Galaxy Clusters and ROSAT <span class="hlt">Observations</span> of <span class="hlt">Bright</span>, Early-Type Galaxies</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>White, Raymond E., III</p> <p>1994-01-01</p> <p>Preliminary results on the elliptical galaxy NGC 1407 were published in the proceedings of the first ROSAT symposium. NGC 1407 is embedded in diffuse X-ray-emitting gas which is extensive enough that it is likely to be related to the surrounding group of galaxies, rather than just NGC 1407. Spectral data for NGC 1407 (AO2) and IC 1459 (AO3) are also included in a complete sample of elliptical galaxies I compiled in collaboration with David Davis. This allowed us to construct the first complete X-ray sample of optically-selected elliptical galaxies. The complete sample allows us to apply Malmquist bias corrections to the <span class="hlt">observed</span> correlation between X-ray and optical luminosities. I continue to work on the implications of this first complete X-ray sample of elliptical galaxies. Paul Eskridge Dave Davis and I also analyzed three long ROSAT PSPC <span class="hlt">observations</span> of the small (but not dwarf) elliptical galaxy M32. We found the X-ray spectra and variability to be consistent with either a Low Mass X-Ray Binary (LMXRB) or a putative 'micro"-AGN.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19790037025&hterms=bouguer&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dbouguer','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19790037025&hterms=bouguer&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dbouguer"><span>The effect of <span class="hlt">surface</span> anisotropy on the accuracy of total ozone estimates from satellite <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fraser, R. S.; Ahmad, Z.</p> <p>1978-01-01</p> <p>The total amount of ozone in a vertical column of the earth's atmosphere is being derived from satellite measurements of the intensity of ultraviolet sunlight scattered by the earth-atmosphere system. The algorithm for deriving the ozone amount utilizes the assumption that the earth's <span class="hlt">surface</span> reflects the incident light isotropically according to Lambert's law. Natural <span class="hlt">surface</span> reflection deviates more or less from this law. Two extreme examples of anisotropic reflection from dark ocean and from <span class="hlt">bright</span> snow are analyzed by means of models for their effects on the derived values of ozone.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=STS055-151-120&hterms=popcorn&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dpopcorn','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=STS055-151-120&hterms=popcorn&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dpopcorn"><span>STS-55 Earth <span class="hlt">observation</span> of the Timor Sea</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1993-01-01</p> <p>STS-55 Earth <span class="hlt">observation</span> taken from Columbia, Orbiter Vehicle (OV) 102, shows the Timor Sea along the south coast of Timor. The sunglint pattern shows a sharp boundary in sea <span class="hlt">surface</span> temperature, with cooler water along the coast and warmer water offshore. The sunglint <span class="hlt">brightness</span> reveals water <span class="hlt">surface</span> roughness with <span class="hlt">bright</span> indicating smooth water and dark representing rough water. Cooler water is smoother because it acts to stabilize the atmospheric boundary layer, while the warm water acts to destabilize the atmosphere. Another indication of water temperature is the cloud pattern. Advection within the atmosphere as a result of warming at the sea <span class="hlt">surface</span> forms low-level clouds with the small, popcorn-like appearance seen in upper right corner of the photograph. The cool water, on the other hand, is relatively free of the popcorn-like clouds. The distribution of the clouds indicates that the wind is blowing toward the upper right corner of the photograph. Also note the line of low-level</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19880015722','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19880015722"><span>Automation of <span class="hlt">surface</span> <span class="hlt">observations</span> program</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Short, Steve E.</p> <p>1988-01-01</p> <p>At present, <span class="hlt">surface</span> weather <span class="hlt">observing</span> methods are still largely manual and labor intensive. Through the nationwide implementation of Automated <span class="hlt">Surface</span> <span class="hlt">Observing</span> Systems (ASOS), this situation can be improved. Two ASOS capability levels are planned. The first is a basic-level system which will automatically <span class="hlt">observe</span> the weather parameters essential for aviation operations and will operate either with or without supplemental contributions by an <span class="hlt">observer</span>. The second is a more fully automated, stand-alone system which will <span class="hlt">observe</span> and report the full range of weather parameters and will operate primarily in the unattended mode. Approximately 250 systems are planned by the end of the decade. When deployed, these systems will generate the standard hourly and special long-line transmitted weather <span class="hlt">observations</span>, as well as provide continuous weather information direct to airport users. Specific ASOS configurations will vary depending upon whether the operation is unattended, minimally attended, or fully attended. The major functions of ASOS are data collection, data processing, product distribution, and system control. The program phases of development, demonstration, production system acquisition, and operational implementation are described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.901a2013H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.901a2013H"><span>Auroral <span class="hlt">bright</span> spot in Jupiter’s active region in corresponding to solar wind dynamic</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Haewsantati, K.; Wannawichian, S.; Clarke, J. T.; Nichols, J. D.</p> <p>2017-09-01</p> <p>Jupiter’s polar emission has <span class="hlt">brightness</span> whose behavior appears to be unstable. This work focuses on the <span class="hlt">bright</span> spot in active region which is a section of Jupiter’s polar emission. Images of the aurora were taken by Advanced Camera for Surveys (ACS) onboard the Hubble Space Telescope (HST). Previously, two <span class="hlt">bright</span> spots, which were found on 13 th May 2007, were suggested to be fixed on locations described by system III longitude. The <span class="hlt">bright</span> spot’s origin in equatorial plane was proposed to be at distance 80-90 Jovian radii and probably associated with the solar wind properties. This study analyzes additional data on May 2007 to study long-term variation of <span class="hlt">brightness</span> and locations of <span class="hlt">bright</span> spots. The newly modified magnetosphere-ionosphere mapping based on VIP4 and VIPAL model is used to locate the origin of <span class="hlt">bright</span> spot in magnetosphere. Furthermore, the Michigan Solar Wind Model or mSWiM is also used to study the variation of solar wind dynamic pressure during the time of <span class="hlt">bright</span> spot’s <span class="hlt">observation</span>. We found that the <span class="hlt">bright</span> spots appear in similar locations which correspond to similar origins in magnetosphere. In addition, the solar wind dynamic pressure should probably affect the <span class="hlt">bright</span> spot’s variation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3674266','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3674266"><span>Controlled formation and reflection of a <span class="hlt">bright</span> solitary matter-wave</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Marchant, A. L.; Billam, T. P.; Wiles, T. P.; Yu, M. M. H.; Gardiner, S. A.; Cornish, S. L.</p> <p>2013-01-01</p> <p><span class="hlt">Bright</span> solitons are non-dispersive wave solutions, arising in a diverse range of nonlinear, one-dimensional systems, including atomic Bose–Einstein condensates with attractive interactions. In reality, cold-atom experiments can only approach the idealized one-dimensional limit necessary for the realization of true solitons. Nevertheless, it remains possible to create <span class="hlt">bright</span> solitary waves, the three-dimensional analogue of solitons, which maintain many of the key properties of their one-dimensional counterparts. Such solitary waves offer many potential applications and provide a rich testing ground for theoretical treatments of many-body quantum systems. Here we report the controlled formation of a <span class="hlt">bright</span> solitary matter-wave from a Bose–Einstein condensate of 85Rb, which is <span class="hlt">observed</span> to propagate over a distance of ∼1.1 mm in 150 ms with no <span class="hlt">observable</span> dispersion. We demonstrate the reflection of a solitary wave from a repulsive Gaussian barrier and contrast this to the case of a repulsive condensate, in both cases finding excellent agreement with theoretical simulations using the three-dimensional Gross–Pitaevskii equation. PMID:23673650</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016cosp...41E.152B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016cosp...41E.152B"><span><span class="hlt">Bright</span> ice spots on the nucleus of comet 67P/Churyumov-Gerasimenko as <span class="hlt">observed</span> by Rosetta OSIRIS and VIRTIS instruments</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barucci, Maria Antonietta; Fulchignoni, Marcello; Pommerol, Antoine; Erard, Stéphane; Oklay, Nilda; Tosi, Federico; Capaccioni, Fabrizio; Sierks, Holger; Filacchione, Gianrico; Bockelee-Morvan, Dominique; Guettler, Carsten; Fornasier, Sonia; Raponi, Andrea; Deshapriya, J. D. P.; Feller, Clement; Ciarniello, Mauro; Leyrat, Cedric</p> <p>2016-07-01</p> <p>Since the Rosetta mission arrived at the comet 67P/Churyumov-Gerasimenko (67/P C-G) on August 2014, the comet nucleus has been mapped by both OSIRIS (Optical, Spectroscopic, and Infrared Remote Imaging System), and VIRTIS (Visible Infrared Thermal Imaging Spectrometer) acquiring a huge quantity of <span class="hlt">surface</span>'s images and spectra, producing the most detailed maps at the highest spatial resolution of a cometary nucleus. The OSIRIS imaging system (NAC & WAC) has a set of filters at different wavelengths from the ultraviolet (269 nm) to the near-infrared (989 nm). The OSIRIS imaging system has been the first instrument with the capability to map a comet <span class="hlt">surface</span> at a high resolution reaching a maximum resolution of 11cm/px during the closest fly-by on February 14, 2015 at a distance of about 6 km from the nucleus <span class="hlt">surface</span> while the VIRTIS spectro-imager (with two channels M and H) operates from 0.25 to 5m with medium and high spectral resolution. The spectral analysis on global scale from the VIRTIS data indicates that the nucleus presents different terrains covered by a very dark and dehydrated organic-rich material [1]. OSIRIS images indicate a morphologically complex and dark <span class="hlt">surface</span> with a variety of terrain types and several intricate features [2]. The <span class="hlt">surface</span> shows albedo variation and from the spectrophotometric analysis a large heterogeneity on the <span class="hlt">surface</span> properties [3, 4, 5]. Limited evidences of exposed H2O ice have been found on the <span class="hlt">surface</span> of 67/P C-G up to now [6, 7, 8], even though ices are considered to be a major constituent of cometary nuclei. The aim of this work is, taking advantage of the high resolution of the OSIRIS images, i) to detect the <span class="hlt">bright</span> spots at all dimensions by albedo and spectral slope analyses, ii) to select those spots which could be resolved by VIRTIS and iii ) to deeply analyse the corresponding spectra. The OSIRIS analysis has been carried out on the colours and spectrophotometry of the whole 67/P C-G nucleus from images acquired</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvA..97a3629G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvA..97a3629G"><span>Three-dimensional vortex-<span class="hlt">bright</span> solitons in a spin-orbit-coupled spin-1 condensate</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gautam, Sandeep; Adhikari, S. K.</p> <p>2018-01-01</p> <p>We demonstrate stable and metastable vortex-<span class="hlt">bright</span> solitons in a three-dimensional spin-orbit-coupled three-component hyperfine spin-1 Bose-Einstein condensate (BEC) using numerical solution and variational approximation of a mean-field model. The spin-orbit coupling provides attraction to form vortex-<span class="hlt">bright</span> solitons in both attractive and repulsive spinor BECs. The ground state of these vortex-<span class="hlt">bright</span> solitons is axially symmetric for weak polar interaction. For a sufficiently strong ferromagnetic interaction, we <span class="hlt">observe</span> the emergence of a fully asymmetric vortex-<span class="hlt">bright</span> soliton as the ground state. We also numerically investigate moving solitons. The present mean-field model is not Galilean invariant, and we use a Galilean-transformed mean-field model for generating the moving solitons.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA21914.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA21914.html"><span>Map of Ceres' <span class="hlt">Bright</span> Spots</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-12</p> <p>This map from NASA's Dawn mission shows locations of <span class="hlt">bright</span> material on dwarf planet Ceres. There are more than 300 <span class="hlt">bright</span> areas, called "faculae," on Ceres. Scientists have divided them into four categories: <span class="hlt">bright</span> areas on the floors of crater (red), on the rims or walls of craters (green), in the ejecta blankets of craters (blue), and on the flanks of the mountain Ahuna Mons (yellow). https://photojournal.jpl.nasa.gov/catalog/PIA21914</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29108645','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29108645"><span>The bidirectional congruency effect of <span class="hlt">brightness</span>-valence metaphoric association in the Stroop-like and priming paradigms.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huang, Yanli; Tse, Chi-Shing; Xie, Jiushu</p> <p>2017-11-04</p> <p>The conceptual metaphor theory (Lakoff & Johnson, 1980, 1999) postulates a unidirectional metaphoric association between abstract and concrete concepts: sensorimotor experience activated by concrete concepts facilitates the processing of abstract concepts, but not the other way around. However, this unidirectional view has been challenged by studies that reported a bidirectional metaphoric association. In three experiments, we tested the directionality of the <span class="hlt">brightness</span>-valence metaphoric association, using Stroop-like paradigm, priming paradigm, and Stroop-like paradigm with a go/no-go manipulation. Both mean and vincentile analyses of reaction time data were performed. We showed that the directionality of <span class="hlt">brightness</span>-valence metaphoric congruency effect could be modulated by the activation level of the <span class="hlt">brightness</span>/valence information. Both <span class="hlt">brightness</span>-to-valence and valence-to-<span class="hlt">brightness</span> metaphoric congruency effects occurred in the priming paradigm, which could be attributed to the presentation of prime that pre-activated the <span class="hlt">brightness</span> or valence information. However, in the Stroop-like paradigm the metaphoric congruency effect was only <span class="hlt">observed</span> in the <span class="hlt">brightness</span>-to-valence direction, but not in the valence-to-<span class="hlt">brightness</span> direction. When the go/no-go manipulation was used to boost the activation of word meaning in the Stroop-like paradigm, the valence-to-<span class="hlt">brightness</span> metaphoric congruency effect was <span class="hlt">observed</span>. Vincentile analyses further revealed that valence-to-<span class="hlt">brightness</span> metaphoric congruency effect approached significance in the Stroop-like paradigm when participants' reaction times were slower (at around 490ms). The implications of the current findings on the conceptual metaphor theory and embodied cognition are discussed. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001321.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001321.html"><span><span class="hlt">Bright</span> Comet ISON</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2013-11-22</p> <p>Comet ISON shines <span class="hlt">brightly</span> in this image taken on the morning of 19 Nov. 2013. This is a 10-second exposure taken with the Marshall Space Flight Center 20" telescope in New Mexico. The camera there is black and white, but the smaller field of view allows for a better "zoom in" on the comet's coma, which is essentially the head of the comet. Credit: NASA/MSFC/MEO/Cameron McCarty -------- More details on Comet ISON: Comet ISON began its trip from the Oort cloud region of our solar system and is now travelling toward the sun. The comet will reach its closest approach to the sun on Thanksgiving Day -- 28 Nov 2013 -- skimming just 730,000 miles above the sun's <span class="hlt">surface</span>. If it comes around the sun without breaking up, the comet will be visible in the Northern Hemisphere with the naked eye, and from what we see now, ISON is predicted to be a particularly <span class="hlt">bright</span> and beautiful comet. Catalogued as C/2012 S1, Comet ISON was first spotted 585 million miles away in September 2012. This is ISON's very first trip around the sun, which means it is still made of pristine matter from the earliest days of the solar system’s formation, its top layers never having been lost by a trip near the sun. Comet ISON is, like all comets, a dirty snowball made up of dust and frozen gases like water, ammonia, methane and carbon dioxide -- some of the fundamental building blocks that scientists believe led to the formation of the planets 4.5 billion years ago. NASA has been using a vast fleet of spacecraft, instruments, and space- and Earth-based telescope, in order to learn more about this time capsule from when the solar system first formed. The journey along the way for such a sun-grazing comet can be dangerous. A giant ejection of solar material from the sun could rip its tail off. Before it reaches Mars -- at some 230 million miles away from the sun -- the radiation of the sun begins to boil its water, the first step toward breaking apart. And, if it survives all this, the intense radiation</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22454802','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22454802"><span><span class="hlt">Bright</span> light treatment as add-on therapy for depression in 28 adolescents: a randomized trial.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Niederhofer, Helmut; von Klitzing, Kai</p> <p>2011-01-01</p> <p>In the last decade, a significant incidence of depression in the younger population has been <span class="hlt">observed</span>. <span class="hlt">Bright</span> light therapy, an effective therapeutic option for depressed adults, could also provide safe, economical, and effective rapid recovery in adolescents. The randomized trial included 28 inpatients (18 females and 10 males) between 14 and 17 years old with depressive complaints. The study was conducted between February and December of 2010 in Rodewisch, Germany. Half of the patients (n = 14) first received placebo (50 lux) 1 hour a day in the morning from 9:00 am to 10:00 am for 1 week and then received <span class="hlt">bright</span> light therapy (2,500 lux) for 1 week in the morning from 9:00 am to 10:00 am. The other half (n = 14) first received <span class="hlt">bright</span> light therapy and then received placebo. Patients were encouraged to continue ongoing treatment (fluoxetine 20 mg/day and 2 sessions of psychotherapy/week) because there were no changes in medication/dosage and psychotherapy since 1 month before the 4-week study period. For assessment of depressive symptoms, the Beck Depression Inventory (BDI) was administered 1 week before and 1 day before placebo treatment, on the day between placebo and <span class="hlt">bright</span> light treatment, and on the day after and 1 week after <span class="hlt">bright</span> light treatment. Saliva samples of melatonin and cortisol were collected at 8:00 am and 8:00 pm 1 week before and 1 day before placebo treatment, on the day between placebo and <span class="hlt">bright</span> light treatment, on the day after <span class="hlt">bright</span> light treatment, and 1 week after <span class="hlt">bright</span> light treatment and were assayed for melatonin and cortisol to <span class="hlt">observe</span> any change in circadian timing. The BDI scores improved significantly (P = .015). The assays of saliva showed significant differences between treatment and placebo for evening melatonin (P = .040). No significant adverse reactions were <span class="hlt">observed</span>. Antidepressant response to <span class="hlt">bright</span> light treatment in this age group was statistically superior to placebo. World Health Organization International Clinical</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12191006','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12191006"><span>Morphologic examination of CD3-CD4(<span class="hlt">bright</span>) cells in rat liver.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yamamoto, Satoshi; Sato, Yosinobu; Abo, Toru; Hatakeyama, Katsuyosi</p> <p>2002-01-01</p> <p>Recently, we found CD3-CD4(<span class="hlt">bright</span>) cells with comparative specificity for normal rat liver. In the current study, we investigated the type and form of both CD3-CD4(<span class="hlt">bright</span>) cells and CD3-CD4(dull) cells in the rat liver. The <span class="hlt">surface</span> phenotype of hepatic mononuclear cells in Lewis rats was identified by using monoclonal antibodies including anti-CD4, anti-CD3, and antimacrophage in conjunction with two- or three-color immunofluorescence analysis. CD3-CD4(<span class="hlt">bright</span>) cells and CD3-CD4(dull) cells were examined morphologically using May-Giemsa staining and scanning electron microscopy. The distribution of CD3-CD4(<span class="hlt">bright</span>) cells and CD3-CD4(dull) cells 48 hours after intravenous administration of liposome-encapsulated dichloromethylene diphosphate was also investigated. In comparison to CD3-CD4(dull) cells, CD3-CD4(<span class="hlt">bright</span>) cells were slightly larger macrophages with abundant cytoplasmic granules, being present with comparative specificity for normal rat liver and showing negligible effects by intravenous liposome-encapsulated dichloromethylene diphosphate administration. These data suggest that in normal young rat liver these CD3-CD4(dull) and CD3-CD4(<span class="hlt">bright</span>) cells may be dendritic cells and Kupffer cells that shift from the liver to the spleen or vice versa. These cells may also be able to locally proliferate in liver or spleen due to changes in the developing liver.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21587316-where-fossils-first-galaxies-ii-true-fossils-ghost-halos-missing-bright-satellites','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21587316-where-fossils-first-galaxies-ii-true-fossils-ghost-halos-missing-bright-satellites"><span>WHERE ARE THE FOSSILS OF THE FIRST GALAXIES? II. TRUE FOSSILS, GHOST HALOS, AND THE MISSING <span class="hlt">BRIGHT</span> SATELLITES</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Bovill, Mia S.; Ricotti, Massimo, E-mail: msbovill@astro.umd.edu</p> <p></p> <p>We use a new set of cold dark matter simulations of the local universe to investigate the distribution of fossils of primordial dwarf galaxies within and around the Milky Way. Throughout, we build upon previous results showing agreement between the <span class="hlt">observed</span> stellar properties of a subset of the ultra-faint dwarfs and our simulated fossils. Here, we show that fossils of the first galaxies have galactocentric distributions and cumulative luminosity functions consistent with <span class="hlt">observations</span>. In our model, we predict {approx}300 luminous satellites orbiting the Milky Way, 50%-70% of which are well-preserved fossils. Within the Milky Way virial radius, the majority ofmore » these fossils have luminosities L{sub V} < 10{sup 6} L{sub sun}. Despite our multidimensional agreement with <span class="hlt">observations</span> at low masses and luminosities, the primordial model produces an overabundance of <span class="hlt">bright</span> dwarf satellites (L{sub V} > 10{sup 4} L{sub sun}) with respect to <span class="hlt">observations</span> where <span class="hlt">observations</span> are nearly complete. The '<span class="hlt">bright</span> satellite problem' is most evident in the outer parts of the Milky Way. We estimate that, although relatively <span class="hlt">bright</span>, the primordial stellar populations are very diffuse, producing a population with <span class="hlt">surface</span> <span class="hlt">brightnesses</span> below surveys' detection limits, and are easily stripped by tidal forces. Although we cannot yet present unmistakable evidence for the existence of the fossils of first galaxies in the Local Group, the results of our studies suggest <span class="hlt">observational</span> strategies that may demonstrate their existence: (1) the detection of 'ghost halos' of primordial stars around isolated dwarfs would prove that stars formed in minihalos (M < 10{sup 8} M{sub sun}) before reionization and strongly suggest that at least a fraction of the ultra-faint dwarfs are fossils of the first galaxies; and (2) the existence of a yet unknown population of {approx}150 Milky Way ultra-faints with half-light radii r{sub hl} {approx} 100-1000 pc and luminosities L{sub V} < 10{sup 4} L{sub sun</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20000111076&hterms=hydra&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dhydra','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20000111076&hterms=hydra&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dhydra"><span>Chandra <span class="hlt">Observations</span> of Hydra A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>McNamara, Brian; Lavoie, Anthony R. (Technical Monitor)</p> <p>2000-01-01</p> <p>We present Chandra X-ray <span class="hlt">Observations</span> of the Hydra A cluster of galaxies, and we report the discovery of structure in the central 80 kpc of the cluster's X-ray-emitting gas. The most remarkable structures are depressions in the X-ray <span class="hlt">surface</span> <span class="hlt">brightness</span>, approx. 25 - 35 kpc diameter, that are coincident with Hydra A's radio lobes. The depressions are nearly devoid of X-ray-emitting gas, and there is no evidence for shock-heated gas surrounding the radio lobes. We suggest the gas within the <span class="hlt">surface</span> <span class="hlt">brightness</span> depressions was displaced as the radio lobes expanded subsonically, leaving cavities in the hot atmosphere. The gas temperature declines from 4 keV at 70 kpc to 3 keV in the inner 20 kpc of the brightest cluster galaxy (BCG), and the cooling time of the gas is approx. 600 Myr in the inner 10 kpc. These properties are consistent with the presence of a approx. 34 solar mass/yr cooling flow within a 70 kpc radius. <span class="hlt">Bright</span> X-ray emission is present in the BCG surrounding a recently-accreted disk of nebular emission and young stars. The star formation rate is commensurate with the cooling rate of the hot gas within the volume of the disk, although the sink for the material that may be cooling at larger radii remains elusive.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA615469','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA615469"><span>Simultaneous Position, Velocity, Attitude, Angular Rates, and <span class="hlt">Surface</span> Parameter Estimation Using Astrometric and Photometric <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2013-07-01</p> <p>Additionally, a physically consistent BRDF and radiation pressure model is utilized thus enabling an accurate physical link between the <span class="hlt">observed</span>... BRDF and radiation pressure model is utilized thus enabling an accurate physical link between the <span class="hlt">observed</span> photometric <span class="hlt">brightness</span> and the attitudinal...source and the <span class="hlt">observer</span> is ( ) VLVLH ˆˆˆˆˆ ++= (2) with angles α and β from N̂ and is used in many analytic BRDF models . There are many</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_17 --> <div id="page_18" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="341"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..12210811M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..12210811M"><span>The Variability of Atmospheric Deuterium <span class="hlt">Brightness</span> at Mars: Evidence for Seasonal Dependence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mayyasi, Majd; Clarke, John; Bhattacharyya, Dolon; Deighan, Justin; Jain, Sonal; Chaffin, Michael; Thiemann, Edward; Schneider, Nick; Jakosky, Bruce</p> <p>2017-10-01</p> <p>The enhanced ratio of deuterium to hydrogen on Mars has been widely interpreted as indicating the loss of a large column of water into space, and the hydrogen content of the upper atmosphere is now known to be highly variable. The variation in the properties of both deuterium and hydrogen in the upper atmosphere of Mars is indicative of the dynamical processes that produce these species and propagate them to altitudes where they can escape the planet. Understanding the seasonal variability of D is key to understanding the variability of the escape rate of water from Mars. Data from a 15 month <span class="hlt">observing</span> campaign, made by the Mars Atmosphere and Volatile Evolution Imaging Ultraviolet Spectrograph high-resolution echelle channel, are used to determine the <span class="hlt">brightness</span> of deuterium as <span class="hlt">observed</span> at the limb of Mars. The D emission is highly variable, with a peak in <span class="hlt">brightness</span> just after southern summer solstice. The trends of D <span class="hlt">brightness</span> are examined against extrinsic as well as intrinsic sources. It is found that the fluctuations in deuterium <span class="hlt">brightness</span> in the upper atmosphere of Mars (up to 400 km), corrected for periodic solar variations, vary on timescales that are similar to those of water vapor fluctuations lower in the atmosphere (20-80 km). The <span class="hlt">observed</span> variability in deuterium may be attributed to seasonal factors such as regional dust storm activity and subsequent circulation lower in the atmosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013DPS....4511208B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013DPS....4511208B"><span><span class="hlt">Observations</span> of NEA 1998 QE2 with the SMA and VLA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Butler, Bryan J.; Gurwell, M. A.; Moullet, A.</p> <p>2013-10-01</p> <p>Long wavelength (submm to cm) <span class="hlt">observations</span> of NEAs are an important tool in their physical characterization. Such <span class="hlt">observations</span> offer a unique probe into the subsurfaces of these bodies, to depths of 10's of cm, and reveal the <span class="hlt">surface</span> and near-<span class="hlt">surface</span> temperatures. These temperatures are critical in constraining the magnitude of the Yarkovsky effect, which is important for these small bodies in the inner solar system as it forces orbital drift [1]. Such <span class="hlt">observations</span> also probe the physical state of the material in the upper layer of the NEAs; notably the thermal inertia. This is a strong indicator of bulk <span class="hlt">surface</span> properties and can be utilized to distinguish rocky from porous <span class="hlt">surfaces</span>. Very low thermal inertia may also indicate "rubble pile" type internal structure in NEAs [2]. Asteroid 1998 QE2 approached to within 0.04 AU on June 1, 2013; its closest approach in two centuries. With a diameter of ~2.5 km [3], it was a relatively <span class="hlt">bright</span> target for radio wavelength <span class="hlt">observations</span>, even given the weakness of the emission at those wavelengths (10's of microJy in the cm to 100's of milliJy in the submm). We used the SubMillimeter Array (SMA) and Very Large Array (VLA) to <span class="hlt">observe</span> the asteroid from wavelengths of 1 mm to 7 cm. We will present the <span class="hlt">observed</span> <span class="hlt">brightness</span> temperatures as a function of wavelength, and implications for temperature and physical state of the <span class="hlt">surface</span> and near-<span class="hlt">surface</span>. [1] Delbo et al. 2007, Icarus, 190, 236. [2] Muller et al. 2007, IAUS 236, 261. [3] Trilling et al. 2010, AJ, 140, 770.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012OptEn..51f4302M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012OptEn..51f4302M"><span><span class="hlt">Bright</span>-dark soliton pairs in a self-mode locking fiber laser</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Meng, Yichang; Zhang, Shumin; Li, Hongfei; Du, Juan; Hao, Yanping; Li, Xingliang</p> <p>2012-06-01</p> <p>We have experimentally <span class="hlt">observed</span> <span class="hlt">bright</span>-dark soliton pairs in an erbium-doped fiber ring laser for the first time. This approach is different from the vector dark domain wall solitons which separate the two orthogonal linear polarization eigenstates of the laser emission. In our laser, the <span class="hlt">bright</span>-dark soliton pairs can co-exist in any one polarization state. Numerical simulations based on the coupled complex Ginzburg-Landau equations have confirmed the experimental results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980219468','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980219468"><span><span class="hlt">Bright</span> Points and Subflares in UV Lines and in X-Rays</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rovira, M.; Schmieder, B.; Demoulin, P.; Simnett, G. M.; Hagyard, M. J.; Reichmann, E.; Tandberg-Hanssen, E.</p> <p>1998-01-01</p> <p>We have analysed an active region which was <span class="hlt">observed</span> in Halpha (MSDP), UV lines (SMM/UVSP), and in X rays (SMM/HXIS). In this active region there were only a few subflares and many small <span class="hlt">bright</span> points visible in UV and in X rays. Using an extrapolation based on the Fourier transform we have computed magnetic field lines connecting different photospheric magnetic polarities from ground-based magnetograms. Along the magnetic inversion lines we find 2 different zones: 1. a high shear region (less than 70 degrees) where subflares occur 2. a low shear region along the magnetic inversion line where UV <span class="hlt">bright</span> points are <span class="hlt">observed</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...856...99P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...856...99P"><span>The <span class="hlt">Bright</span> γ-ray Flare of 3C 279 in 2015 June: AGILE Detection and Multifrequency Follow-up <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pittori, C.; Lucarelli, F.; Verrecchia, F.; Raiteri, C. M.; Villata, M.; Vittorini, V.; Tavani, M.; Puccetti, S.; Perri, M.; Donnarumma, I.; Vercellone, S.; Acosta-Pulido, J. A.; Bachev, R.; Benítez, E.; Borman, G. A.; Carnerero, M. I.; Carosati, D.; Chen, W. P.; Ehgamberdiev, Sh. A.; Goded, A.; Grishina, T. S.; Hiriart, D.; Hsiao, H. Y.; Jorstad, S. G.; Kimeridze, G. N.; Kopatskaya, E. N.; Kurtanidze, O. M.; Kurtanidze, S. O.; Larionov, V. M.; Larionova, L. V.; Marscher, A. P.; Mirzaqulov, D. O.; Morozova, D. A.; Nilsson, K.; Samal, M. R.; Sigua, L. A.; Spassov, B.; Strigachev, A.; Takalo, L. O.; Antonelli, L. A.; Bulgarelli, A.; Cattaneo, P.; Colafrancesco, S.; Giommi, P.; Longo, F.; Morselli, A.; Paoletti, F.</p> <p>2018-04-01</p> <p>We report the AGILE detection and the results of the multifrequency follow-up <span class="hlt">observations</span> of a <span class="hlt">bright</span> γ-ray flare of the blazar 3C 279 in 2015 June. We use AGILE and Fermi gamma-ray data, together with Swift X-ray andoptical-ultraviolet data, and ground-based GASP-WEBT optical <span class="hlt">observations</span>, including polarization information, to study the source variability and the overall spectral energy distribution during the γ-ray flare. The γ-ray flaring data, compared with as yet unpublished simultaneous optical data that will allow constraints on the big blue bump disk luminosity, show very high Compton dominance values of ∼100, with the ratio of γ-ray to optical emission rising by a factor of three in a few hours. The multiwavelength behavior of the source during the flare challenges one-zone leptonic theoretical models. The new <span class="hlt">observations</span> during the 2015 June flare are also compared with already published data and nonsimultaneous historical 3C 279 archival data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017A%26A...608A.142V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...608A.142V"><span>The Fornax Deep Survey with VST. III. Low <span class="hlt">surface</span> <span class="hlt">brightness</span> dwarfs and ultra diffuse galaxies in the center of the Fornax cluster</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Venhola, Aku; Peletier, Reynier; Laurikainen, Eija; Salo, Heikki; Lisker, Thorsten; Iodice, Enrichetta; Capaccioli, Massimo; Kleijn, Gijs Verdoes; Valentijn, Edwin; Mieske, Steffen; Hilker, Michael; Wittmann, Carolin; van de Ven, Glenn; Grado, Aniello; Spavone, Marilena; Cantiello, Michele; Napolitano, Nicola; Paolillo, Maurizio; Falcón-Barroso, Jesús</p> <p>2017-12-01</p> <p>Context. Studies of low <span class="hlt">surface</span> <span class="hlt">brightness</span> (LSB) galaxies in nearby clusters have revealed a sub-population of extremely diffuse galaxies with central <span class="hlt">surface</span> <span class="hlt">brightness</span> of μ0,g' > 24 mag arcsec-2, total luminosity Mg' fainter than -16 mag and effective radius between 1.5 kpc <Re < 10 kpc. The origin of these ultra diffuse galaxies (UDGs) is still unclear, although several theories have been suggested. As the UDGs overlap with the dwarf-sized galaxies in their luminosities, it is important to compare their properties in the same environment. If a continuum is found between the properties of UDGs and the rest of the LSB population, it would be consistent with the idea that they have a common origin. Aims: Our aim is to exploit the deep g', r' and i'-band images of the Fornax Deep Survey (FDS), in order to identify LSB galaxies in an area of 4 deg2 in the center of the Fornax cluster. The identified galaxies are divided into UDGs and dwarf-sized LSB galaxies, and their properties are compared. Methods: We identified visually all extended structures having r'-band central <span class="hlt">surface</span> <span class="hlt">brightness</span> of μ0,r' > 23 mag arcsec-2. We classified the objects based on their appearance into galaxies and tidal structures, and perform 2D Sérsic model fitting with GALFIT to measure the properties of those classified as galaxies. We analyzed their radial distribution and orientations with respect of the cluster center, and with respect to the other galaxies in our sample. We also studied their colors and compare the LSB galaxies in Fornax with those in other environments. Results: Our final sample complete in the parameter space of the previously known UDGs, consists of 205 galaxies of which 196 are LSB dwarfs (with Re < 1.5 kpc) and nine are UDGs (Re > 1.5 kpc). We show that the UDGs have (1) g'-r' colors similar to those of LSB dwarfs of the same luminosity; (2) the largest UDGs (Re > 3 kpc) in our sample appear different from the other LSB galaxies, in that they are significantly</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25620199','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25620199"><span>Phase advancing human circadian rhythms with morning <span class="hlt">bright</span> light, afternoon melatonin, and gradually shifted sleep: can we reduce morning <span class="hlt">bright</span>-light duration?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Crowley, Stephanie J; Eastman, Charmane I</p> <p>2015-02-01</p> <p>Efficient treatments to phase-advance human circadian rhythms are needed to attenuate circadian misalignment and the associated negative health outcomes that accompany early-morning shift work, early school start times, jet lag, and delayed sleep phase disorder. This study compared three morning <span class="hlt">bright</span>-light exposure patterns from a single light box (to mimic home treatment) in combination with afternoon melatonin. Fifty adults (27 males) aged 25.9 ± 5.1 years participated. Sleep/dark was advanced 1 h/day for three treatment days. Participants took 0.5 mg of melatonin 5 h before the baseline bedtime on treatment day 1, and an hour earlier each treatment day. They were exposed to one of three <span class="hlt">bright</span>-light (~5000 lux) patterns upon waking each morning: four 30-min exposures separated by 30 min of room light (2-h group), four 15-min exposures separated by 45 min of room light (1-h group), and one 30-min exposure (0.5-h group). Dim-light melatonin onsets (DLMOs) before and after treatment determined the phase advance. Compared to the 2-h group (phase shift = 2.4 ± 0.8 h), smaller phase-advance shifts were seen in the 1-h (1.7 ± 0.7 h) and 0.5-h (1.8 ± 0.8 h) groups. The 2-h pattern produced the largest phase advance; however, the single 30-min <span class="hlt">bright</span>-light exposure was as effective as 1 h of <span class="hlt">bright</span> light spread over 3.25 h, and it produced 75% of the phase shift <span class="hlt">observed</span> with 2 h of <span class="hlt">bright</span> light. A 30-min morning <span class="hlt">bright</span>-light exposure with afternoon melatonin is an efficient treatment to phase-advance human circadian rhythms. Copyright © 2014 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4344919','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4344919"><span>Phase advancing human circadian rhythms with morning <span class="hlt">bright</span> light, afternoon melatonin, and gradually shifted sleep: can we reduce morning <span class="hlt">bright</span> light duration?</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Crowley, Stephanie J.; Eastman, Charmane I.</p> <p>2015-01-01</p> <p>OBJECTIVE Efficient treatments to phase advance human circadian rhythms are needed to attenuate circadian misalignment and the associated negative health outcomes that accompany early morning shift work, early school start times, jet lag, and delayed sleep phase disorder. This study compared three morning <span class="hlt">bright</span> light exposure patterns from a single light box (to mimic home treatment) in combination with afternoon melatonin. METHODS Fifty adults (27 males) aged 25.9±5.1 years participated. Sleep/dark was advanced 1 hour/day for 3 treatment days. Participants took 0.5 mg melatonin 5 hours before baseline bedtime on treatment day 1, and an hour earlier each treatment day. They were exposed to one of three <span class="hlt">bright</span> light (~5000 lux) patterns upon waking each morning: four 30-minute exposures separated by 30 minutes of room light (2 h group); four 15-minute exposures separated by 45 minutes of room light (1 h group), and one 30-minute exposure (0.5 h group). Dim light melatonin onsets (DLMOs) before and after treatment determined the phase advance. RESULTS Compared to the 2 h group (phase shift=2.4±0.8 h), smaller phase advance shifts were seen in the 1 h (1.7±0.7 h) and 0.5 h (1.8±0.8 h) groups. The 2-hour pattern produced the largest phase advance; however, the single 30-minute <span class="hlt">bright</span> light exposure was as effective as 1 hour of <span class="hlt">bright</span> light spread over 3.25 h, and produced 75% of the phase shift <span class="hlt">observed</span> with 2 hours of <span class="hlt">bright</span> light. CONCLUSIONS A 30-minute morning <span class="hlt">bright</span> light exposure with afternoon melatonin is an efficient treatment to phase advance human circadian rhythms. PMID:25620199</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15524938','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15524938"><span>Generation of dark and <span class="hlt">bright</span> spin wave envelope soliton trains through self-modulational instability in magnetic films.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Mingzhong; Kalinikos, Boris A; Patton, Carl E</p> <p>2004-10-08</p> <p>The generation of dark spin wave envelope soliton trains from a continuous wave input signal due to spontaneous modulational instability has been <span class="hlt">observed</span> for the first time. The dark soliton trains were formed from high dispersion dipole-exchange spin waves propagated in a thin yttrium iron garnet film with pinned <span class="hlt">surface</span> spins at frequencies situated near the dipole gaps in the dipole-exchange spin wave spectrum. Dark and <span class="hlt">bright</span> soliton trains were generated for one and the same film through placement of the input carrier frequency in regions of negative and positive dispersion, respectively. Two unreported effects in soliton dynamics, hysteresis and period doubling, were also <span class="hlt">observed</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870014851','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870014851"><span>IRAS <span class="hlt">observations</span> of the Pleiades</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cox, P.; he ultraviolet.</p> <p>1987-01-01</p> <p>The Infrared Astronomy Satellite (IRAS) <span class="hlt">observations</span> of the Pleiades region are reported. The data show large flux densities at 12 and 25 microns, extended over the optical nebulosity. This strong excess emission, implying temperatures of a few hundred degrees Kelvin, indicates a population of very small grains in the Pleiades. It is suggested that these grains are similar to the small grains needed to explain the <span class="hlt">surface</span> <span class="hlt">brightness</span> measurements made in the ultraviolet.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28988026','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28988026"><span>Quantitative Image Restoration in <span class="hlt">Bright</span> Field Optical Microscopy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gutiérrez-Medina, Braulio; Sánchez Miranda, Manuel de Jesús</p> <p>2017-11-07</p> <p><span class="hlt">Bright</span> field (BF) optical microscopy is regarded as a poor method to <span class="hlt">observe</span> unstained biological samples due to intrinsic low image contrast. We introduce quantitative image restoration in <span class="hlt">bright</span> field (QRBF), a digital image processing method that restores out-of-focus BF images of unstained cells. Our procedure is based on deconvolution, using a point spread function modeled from theory. By comparing with reference images of bacteria <span class="hlt">observed</span> in fluorescence, we show that QRBF faithfully recovers shape and enables quantify size of individual cells, even from a single input image. We applied QRBF in a high-throughput image cytometer to assess shape changes in Escherichia coli during hyperosmotic shock, finding size heterogeneity. We demonstrate that QRBF is also applicable to eukaryotic cells (yeast). Altogether, digital restoration emerges as a straightforward alternative to methods designed to generate contrast in BF imaging for quantitative analysis. Copyright © 2017 Biophysical Society. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A24A..01K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A24A..01K"><span>Polarimetric <span class="hlt">Observations</span> of the Lunar <span class="hlt">Surface</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, S.</p> <p>2017-12-01</p> <p>Polarimetric images contain valuable information on the lunar <span class="hlt">surface</span> such as grain size and porosity of the regolith, from which one can estimate the space weathering environment on the lunar <span class="hlt">surface</span>. Surprisingly, polarimetric <span class="hlt">observation</span> has never been conducted from the lunar orbit before. A Wide-Angle Polarimetric Camera (PolCam) has been recently selected as one of three Korean science instruments onboard the Korea Pathfinder Lunar Orbiter (KPLO), which is aimed to be launched in 2019/2020 as the first Korean lunar mission. PolCam will obtain 80 m-resolution polarimetric images of the whole lunar <span class="hlt">surface</span> between -70º and +70º latitudes at 320, 430 and 750 nm bands for phase angles up to 115º. I will also discuss previous polarimetric studies on the lunar <span class="hlt">surface</span> based on our ground-based <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999AAS...194.1503D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999AAS...194.1503D"><span>Groundbased <span class="hlt">Observations</span> of [C I] 9850A Emission from Comet Hale-Bopp</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Doane, N. E.; Oliversen, R. J.; Scherb, F.; Morgenthaler, J. P.; Roesler, F. L.; Woodward, R. C.; Harris, W. M.; Hilton, G. M.</p> <p>1999-05-01</p> <p>High spectral resolution <span class="hlt">observations</span> of Comet Hale-Bopp [C I] 9850A emission were obtained at the NSO McMath-Pierce main telescope on 13 nights during 1997 March 9 to 10 and April 7 to 19. Spectra with good signal-to-noise were obtained using a dual- etalon 50mm Fabry-Perot spectrometer (R 40,000) with a 6 arcmin field of view. The comet was <span class="hlt">observed</span> over a 0.92-1.00 AU range of heliocentric distances. Most <span class="hlt">observations</span> were centered on the comet nucleus where the <span class="hlt">surface</span> <span class="hlt">brightness</span> ranged from about 70 to 170 Rayleighs. Several <span class="hlt">observations</span> were also centered approximately 5 arcmin sunward and tailward of the comet nucleus. The sunward [C I] emission was fainter than the tailward emission. Assuming that CO photodissociation is the source of cometary C(1D) (and neglecting quenching), for a <span class="hlt">surface</span> <span class="hlt">brightness</span> of 120 Rayleighs, we estimate a (lower limit) CO production rate of about 2x10(30) per sec. These [C I] observationsare the first extensive set reported for this cometary emission line.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EP%26S...69..112F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EP%26S...69..112F"><span>Geomagnetically conjugate <span class="hlt">observations</span> of ionospheric and thermospheric variations accompanied by a midnight <span class="hlt">brightness</span> wave at low latitudes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fukushima, D.; Shiokawa, K.; Otsuka, Y.; Kubota, M.; Yokoyama, T.; Nishioka, M.; Komonjinda, S.; Yatini, C. Y.</p> <p>2017-08-01</p> <p>We conducted geomagnetically conjugate <span class="hlt">observations</span> of 630-nm airglow for a midnight <span class="hlt">brightness</span> wave (MBW) at Kototabang, Indonesia [geomagnetic latitude (MLAT): 10.0°S], and Chiang Mai, Thailand (MLAT: 8.9°N), which are geomagnetically conjugate points at low latitudes. An airglow enhancement that was considered to be an MBW was <span class="hlt">observed</span> in OI (630-nm) airglow images at Kototabang around local midnight from 2240 to 2430 LT on February 7, 2011. This MBW propagated south-southwestward, which is geomagnetically poleward, at a velocity of 290 m/s. However, a similar wave was not <span class="hlt">observed</span> in the 630-nm airglow images at Chiang Mai. This is the first evidence of an MBW that does not have geomagnetic conjugacy, which also implies generation of MBW only in one side of the hemisphere from the equator. We simultaneously <span class="hlt">observed</span> thermospheric neutral winds <span class="hlt">observed</span> by a co-located Fabry-Perot interferometer at Kototabang. The <span class="hlt">observed</span> meridional winds turned from northward (geomagnetically equatorward) to southward (geomagnetically poleward) just before the wave was <span class="hlt">observed</span>. This indicates that the <span class="hlt">observed</span> MBW was generated by the poleward winds which push ionospheric plasma down along geomagnetic field lines, thereby increasing the 630-nm airglow intensity. The bottomside ionospheric heights <span class="hlt">observed</span> by ionosondes rapidly decreased at Kototabang and slightly increased at Chiang Mai. We suggest that the polarization electric field inside the <span class="hlt">observed</span> MBW is projected to the northern hemisphere, causing the small height increase <span class="hlt">observed</span> at Chiang Mai. This implies that electromagnetic coupling between hemispheres can occur even though the original disturbance is caused purely by the neutral wind.[Figure not available: see fulltext.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950034213&hterms=Carl+Rogers&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DCarl%2BRogers','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950034213&hterms=Carl+Rogers&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DCarl%2BRogers"><span>Extreme Ultraviolet Explorer <span class="hlt">Bright</span> Source List</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Malina, Roger F.; Marshall, Herman L.; Antia, Behram; Christian, Carol A.; Dobson, Carl A.; Finley, David S.; Fruscione, Antonella; Girouard, Forrest R.; Hawkins, Isabel; Jelinsky, Patrick</p> <p>1994-01-01</p> <p>Initial results from the analysis of the Extreme Ultraviolet Explorer (EUVE) all-sky survey (58-740 A) and deep survey (67-364 A) are presented through the EUVE <span class="hlt">Bright</span> Source List (BSL). The BSL contains 356 confirmed extreme ultraviolet (EUV) point sources with supporting information, including positions, <span class="hlt">observed</span> EUV count rates, and the identification of possible optical counterparts. One-hundred twenty-six sources have been detected longward of 200 A.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19830006291','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19830006291"><span>Adjusting the tasseled cap <span class="hlt">brightness</span> and greenness factors for atmospheric path radiance and absorption on a pixel by pixel basis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jackson, R. D.; Slater, P. N.; Pinter, P. J. (Principal Investigator)</p> <p>1982-01-01</p> <p>A radiative transfer model was used to convert ground measured reflectances into the radiance at the top of the atmosphere, for several levels of atmospheric path radiance. The radiance in MSS7 (0.8 to 1.1 m) was multiplied by the transmission fraction for atmospheres having different levels of precipitable water. The radiance values were converted to simulated LANDSAT digital counts for four path radiance levels and four levels of precipitable water. These values were used to calculate the Kauth-Thomas <span class="hlt">brightness</span>, greenness, yellowness, and nonsuch factors. <span class="hlt">Brightness</span> was affected by <span class="hlt">surface</span> conditions and path radiance. Greenness was affected by <span class="hlt">surface</span> conditions, path radiance, and precipitable water. Yellowness was affected by path radiance and nonsuch by precipitable water, and both factors changed only slightly with <span class="hlt">surface</span> conditions. Yellowness and nonsuch were used to adjust <span class="hlt">brightness</span> and greenness to produce factors that were affected only by <span class="hlt">surface</span> conditions such as soils and vegetation, and not by path radiance and precipitable water.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016nova.pres..886K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016nova.pres..886K"><span>How <span class="hlt">Bright</span> Can Supernovae Get?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kohler, Susanna</p> <p>2016-04-01</p> <p>Supernovae enormous explosions associated with the end of a stars life come in a variety of types with different origins. A new study has examined how the brightest supernovae in the Universe are produced, and what limits might be set on their <span class="hlt">brightness</span>.Ultra-Luminous <span class="hlt">Observations</span>Recent <span class="hlt">observations</span> have revealed many ultra-luminous supernovae, which haveenergies that challenge our abilities to explain them usingcurrent supernova models. An especially extreme example is the 2015 discovery of the supernova ASASSN-15lh, which shone with a peak luminosity of ~2*1045 erg/s, nearly a trillion times brighter than the Sun. ASASSN-15lh radiated a whopping ~2*1052 erg in the first four months after its detection.How could a supernova that <span class="hlt">bright</span> be produced? To explore the answer to that question, Tuguldur Sukhbold and Stan Woosley at University of California, Santa Cruz, have examined the different sources that could produce supernovae and calculated upper limits on the potential luminosities ofeach of these supernova varieties.Explosive ModelsSukhbold and Woosley explore multiple different models for core-collapse supernova explosions, including:Prompt explosionA stars core collapses and immediately explodes.Pair instabilityElectron/positron pair production at a massive stars center leads to core collapse. For high masses, radioactivity can contribute to delayed energy output.Colliding shellsPreviously expelled shells of material around a star collide after the initial explosion, providing additional energy release.MagnetarThe collapsing star forms a magnetar a rapidly rotating neutron star with an incredibly strong magnetic field at its core, which then dumps energy into the supernova ejecta, further brightening the explosion.They then apply these models to different types of stars.Setting the LimitThe authors show that the light curve of ASASSN-15lh (plotted in orange) can be described by a model (black curve) in which a magnetar with an initial spin period of 0.7 ms</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018sptz.prop14130B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018sptz.prop14130B"><span><span class="hlt">Bright</span> galaxies at z=9-11 from pure-parallel HST <span class="hlt">observations</span>: Building a unique sample for JWST with Spitzer/IRAC</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bouwens, Rychard; Morashita, Takahiro; Stefanon, Mauro; Magee, Dan</p> <p>2018-05-01</p> <p>The combination of <span class="hlt">observations</span> taken by Hubble and Spitzer revealed the unexpected presence of sources as <span class="hlt">bright</span> as our own Milky Way as early as 400 Myr after the Big Bang, potentially highlighting a new highly efficient regime for star formation in L>L* galaxies at very early times. Yet, the sample of high-quality z>8 galaxies with both HST and Spitzer/IRAC imaging is still small, particularly at the highest luminosities. We propose here to remedy this situation and use Spitzer/IRAC to efficiently follow up the most promising z>8 sources from our Hubble Brightest of Reionizing Galaxies (BoRG) survey, which covers a footprint on the sky similar to CANDELS, provides a deeper search than ground-based surveys like UltraVISTA, and is robust against cosmic variance because of its 210 independent lines of sight. The proposed new 3.6 micron <span class="hlt">observations</span> will continue our Spitzer cycle 12 and 13 BORG911 programs, targeting 15 additional fields, leveraging over 200 new HST orbits to identify a final sample of about 8 <span class="hlt">bright</span> galaxies at z >= 8.5. For optimal time use (just 20 hours), our goal is to readily discriminate between z>8 sources (undetected or marginally detected in IRAC) and z 2 interlopers (strongly detected in IRAC) with just 1-2 hours per pointing. The high-quality candidates that we will identify with IRAC will be ideal targets for further studies investigating the ionization state of the distant universe through near-IR Keck/VLT spectroscopy. They will also be uniquely suited to measurement of the redshift and stellar population properties through JWST/NIRSPEC <span class="hlt">observations</span>, with the potential to elucidate how the first generations of stars are assembled in the earliest stages of the epoch of reionization.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21137206-observational-difference-between-gamma-ray-properties-optically-dark-bright-grbs','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21137206-observational-difference-between-gamma-ray-properties-optically-dark-bright-grbs"><span><span class="hlt">Observational</span> difference between gamma and X-ray properties of optically dark and <span class="hlt">bright</span> GRBs</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Balazs, L. G.; Horvath, I.; Bagoly, Zs.</p> <p>2008-05-22</p> <p>Using the discriminant analysis of the multivariate statistical analysis we compared the distribution of the physical quantities of the optically dark and <span class="hlt">bright</span> GRBs, detected by the BAT and XRT on board of the Swift Satellite. We found that the GRBs having detected optical transients (OT) have systematically higher peak fluxes and lower HI column densities than those without OT.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...836L..20Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...836L..20Z"><span>H2O Megamasers toward Radio-<span class="hlt">bright</span> Seyfert 2 Nuclei</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, J. S.; Liu, Z. W.; Henkel, C.; Wang, J. Z.; Coldwell, G. V.</p> <p>2017-02-01</p> <p>Using the Effelsberg-100 m telescope, we perform a successful pilot survey on H2O maser emission toward a small sample of radio-<span class="hlt">bright</span> Seyfert 2 galaxies with a redshift larger than 0.04. The targets were selected from a large Seyfert 2 sample derived from the spectroscopic Sloan Digital Sky Survey Data Release 7 (SDSS-DR7). One source, SDSS J102802.9+104630.4 (z ˜ 0.0448), was detected four times during our <span class="hlt">observations</span>, with a typical maser flux density of ˜30 mJy and a corresponding (very large) luminosity of ˜1135 L ⊙. The successful detection of this radio-<span class="hlt">bright</span> Seyfert 2 and an additional tentative detection support our previous statistical results that H2O megamasers tend to arise from Seyfert 2 galaxies with large radio luminosity. The finding provides further motivation for an upcoming larger H2O megamaser survey toward Seyfert 2s with particularly radio-<span class="hlt">bright</span> nuclei with the basic goal to improve our understanding of the nuclear environment of active megamaser host galaxies. Based on <span class="hlt">observations</span> with the 100 m telescope of the MPIfR (Max-Planck-Institut für Radioastronomie) at Effelsberg.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999ASSL..239..231B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999ASSL..239..231B"><span><span class="hlt">Brightness</span> Variations in the Solar Atmosphere as Seen by SOHO</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brkovic, A.; Rüedi, I.; Solanki, S. K.; Huber, M. C. E.; Stenflo, J. O.; Stucki, K.; Harrison, R.; Fludra, A.</p> <p></p> <p>We present preliminary results of a statistical analysis of the <span class="hlt">brightness</span> variations of solar features at different levels in the solar atmosphere. We <span class="hlt">observed</span> quiet Sun regions at disc centre using the Coronal Diagnostic Spectrometer (CDS) onboard the Solar and Heliospheric Observatory (SOHO). We find significant variability at all time scales in all parts of the quiet Sun, from darkest intranetwork to brightest network. Such variations are <span class="hlt">observed</span> simultaneously in the chromospheric He I 584.33 Angstroms (2 \\cdot 10^4 K) line, the transition region O V 629.74 Angstroms (2.5 \\cdot 10^5 K) and coronal Mg IX 368.06 Angstroms (10^6 K) line. The relative variability is independent of <span class="hlt">brightness</span> and most of the variability appears to take place on time scales longer than 5 minutes for all 3 spectral lines. No significant differences are <span class="hlt">observed</span> between the different data sets.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018TCry...12..921R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018TCry...12..921R"><span>Arctic sea ice signatures: L-band <span class="hlt">brightness</span> temperature sensitivity comparison using two radiation transfer models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Richter, Friedrich; Drusch, Matthias; Kaleschke, Lars; Maaß, Nina; Tian-Kunze, Xiangshan; Mecklenburg, Susanne</p> <p>2018-03-01</p> <p>Sea ice is a crucial component for short-, medium- and long-term numerical weather predictions. Most importantly, changes of sea ice coverage and areas covered by thin sea ice have a large impact on heat fluxes between the ocean and the atmosphere. L-band <span class="hlt">brightness</span> temperatures from ESA's Earth Explorer SMOS (Soil Moisture and Ocean Salinity) have been proven to be a valuable tool to derive thin sea ice thickness. These retrieved estimates were already successfully assimilated in forecasting models to constrain the ice analysis, leading to more accurate initial conditions and subsequently more accurate forecasts. However, the <span class="hlt">brightness</span> temperature measurements can potentially be assimilated directly in forecasting systems, reducing the data latency and providing a more consistent first guess. As a first step towards such a data assimilation system we studied the forward operator that translates geophysical parameters provided by a model into <span class="hlt">brightness</span> temperatures. We use two different radiative transfer models to generate top of atmosphere <span class="hlt">brightness</span> temperatures based on ORAP5 model output for the 2012/2013 winter season. The simulations are then compared against actual SMOS measurements. The results indicate that both models are able to capture the general variability of measured <span class="hlt">brightness</span> temperatures over sea ice. The simulated <span class="hlt">brightness</span> temperatures are dominated by sea ice coverage and thickness changes are most pronounced in the marginal ice zone where new sea ice is formed. There we <span class="hlt">observe</span> the largest differences of more than 20 K over sea ice between simulated and <span class="hlt">observed</span> <span class="hlt">brightness</span> temperatures. We conclude that the assimilation of SMOS <span class="hlt">brightness</span> temperatures yields high potential for forecasting models to correct for uncertainties in thin sea ice areas and suggest that information on sea ice fractional coverage from higher-frequency <span class="hlt">brightness</span> temperatures should be used simultaneously.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001ApJS..132..129M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001ApJS..132..129M"><span>An Ultraviolet/Optical Atlas of <span class="hlt">Bright</span> Galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marcum, Pamela M.; O'Connell, Robert W.; Fanelli, Michael N.; Cornett, Robert H.; Waller, William H.; Bohlin, Ralph C.; Neff, Susan G.; Roberts, Morton S.; Smith, Andrew M.; Cheng, K.-P.; Collins, Nicholas R.; Hennessy, Gregory S.; Hill, Jesse K.; Hill, Robert S.; Hintzen, Paul; Landsman, Wayne B.; Ohl, Raymond G.; Parise, Ronald A.; Smith, Eric P.; Freedman, Wendy L.; Kuchinski, Leslie E.; Madore, Barry; Angione, Ronald; Palma, Christopher; Talbert, Freddie; Stecher, Theodore P.</p> <p>2001-02-01</p> <p>We present wide-field imagery and photometry of 43 selected nearby galaxies of all morphological types at ultraviolet and optical wavelengths. The ultraviolet (UV) images, in two broad bands at 1500 and 2500 Å, were obtained using the Ultraviolet Imaging Telescope (UIT) during the Astro-1 Spacelab mission. The UV images have ~3" resolution, and the comparison sets of ground-based CCD images (in one or more of B, V, R, and Hα) have pixel scales and fields of view closely matching the UV frames. The atlas consists of multiband images and plots of UV/optical <span class="hlt">surface</span> <span class="hlt">brightness</span> and color profiles. Other associated parameters, such as integrated photometry and half-light radii, are tabulated. In an appendix, we discuss the sensitivity of different wavebands to a galaxy's star formation history in the form of ``history weighting functions'' and emphasize the importance of UV <span class="hlt">observations</span> as probes of evolution during the past 10-1000 Myr. We find that UV galaxy morphologies are usually significantly different from visible band morphologies as a consequence of spatially inhomogeneous stellar populations. Differences are quite pronounced for systems in the middle range of Hubble types, Sa through Sc, but less so for ellipticals or late-type disks. Normal ellipticals and large spiral bulges are fainter and more compact in the UV. However, they typically exhibit smooth UV profiles with far-UV/optical color gradients which are larger than any at optical/IR wavelengths. The far-UV light in these cases is probably produced by extreme horizontal branch stars and their descendants in the dominant, low-mass, metal-rich population. The cool stars in the large bulges of Sa and Sb spirals fade in the UV while hot OB stars in their disks brighten, such that their Hubble classifications become significantly later. In the far-UV, early-type spirals often appear as peculiar, ringlike systems. In some spiral disks, UV-<span class="hlt">bright</span> structures closely outline the spiral pattern; in others, the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010028709','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010028709"><span>A Methodology for <span class="hlt">Surface</span> Soil Moisture and Vegetation Optical Depth Retrieval Using the Microwave Polarization Difference Index</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Owe, Manfred; deJeu, Richard; Walker, Jeffrey; Zukor, Dorothy J. (Technical Monitor)</p> <p>2001-01-01</p> <p>A methodology for retrieving <span class="hlt">surface</span> soil moisture and vegetation optical depth from satellite microwave radiometer data is presented. The procedure is tested with historical 6.6 GHz <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> from the Scanning Multichannel Microwave Radiometer over several test sites in Illinois. Results using only nighttime data are presented at this time, due to the greater stability of nighttime <span class="hlt">surface</span> temperature estimation. The methodology uses a radiative transfer model to solve for <span class="hlt">surface</span> soil moisture and vegetation optical depth simultaneously using a non-linear iterative optimization procedure. It assumes known constant values for the scattering albedo and roughness. <span class="hlt">Surface</span> temperature is derived by a procedure using high frequency vertically polarized <span class="hlt">brightness</span> temperatures. The methodology does not require any field <span class="hlt">observations</span> of soil moisture or canopy biophysical properties for calibration purposes and is totally independent of wavelength. Results compare well with field <span class="hlt">observations</span> of soil moisture and satellite-derived vegetation index data from optical sensors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMED11D0146H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMED11D0146H"><span>The Magnetic Evolution of Coronal Hole <span class="hlt">Bright</span> Points</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>He, Y.; Muglach, K.</p> <p>2017-12-01</p> <p>Space weather refers to the state of the heliosphere and the geospace environment that are caused primarily by solar activity. Coronal mass ejections and flares originate in active regions and filaments close to the solar <span class="hlt">surface</span> and can cause geomagnetic storms and solar energetic particles events, which can damage both spacecraft and ground-based systems that are critical for society's well-being. Coronal <span class="hlt">bright</span> points are small-scale magnetic regions on the sun that seem to be similar to active regions, but are about an order of magnitude smaller. Due to their shorter lifetime, the complete evolutionary cycle of these mini active regions can be studied, from the time they appear in extreme-ultraviolet (EUV) images to the time they fade. We are using data from the Solar Dynamics Observatory (SDO) to study both the coronal EUV flux and the photospheric magnetic field and compare them to activities of the coronal <span class="hlt">bright</span> point.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010SPIE.7840E..25L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010SPIE.7840E..25L"><span>Study of the model of calibrating differences of <span class="hlt">brightness</span> temperature from geostationary satellite generated by time zone differences</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Weidong; Shan, Xinjian; Qu, Chunyan</p> <p>2010-11-01</p> <p>In comparison with polar-orbiting satellites, geostationary satellites have a higher time resolution and wider field of visions, which can cover eleven time zones (an image covers about one third of the Earth's <span class="hlt">surface</span>). For a geostationary satellite panorama graph at a point of time, the <span class="hlt">brightness</span> temperature of different zones is unable to represent the thermal radiation information of the <span class="hlt">surface</span> at the same point of time because of the effect of different sun solar radiation. So it is necessary to calibrate <span class="hlt">brightness</span> temperature of different zones with respect to the same point of time. A model of calibrating the differences of the <span class="hlt">brightness</span> temperature of geostationary satellite generated by time zone differences is suggested in this study. A total of 16 curves of four positions in four different stages are given through sample statistics of <span class="hlt">brightness</span> temperature of every 5 days synthetic data which are from four different time zones (time zones 4, 6, 8, and 9). The above four stages span January -March (winter), April-June (spring), July-September (summer), and October-December (autumn). Three kinds of correct situations and correct formulas based on curves changes are able to better eliminate <span class="hlt">brightness</span> temperature rising or dropping caused by time zone differences.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...859....5L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...859....5L"><span>Difference in Dwarf Galaxy <span class="hlt">Surface</span> <span class="hlt">Brightness</span> Profiles as a Function of Environment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Youngdae; Park, Hong Soo; Kim, Sang Chul; Moon, Dae-Sik; Lee, Jae-Joon; Kim, Dong-Jin; Cha, Sang-Mok</p> <p>2018-05-01</p> <p>We investigate <span class="hlt">surface</span> <span class="hlt">brightness</span> profiles (SBPs) of dwarf galaxies in field, group, and cluster environments. With deep BV I images from the Korea Microlensing Telescope Network Supernova Program, SBPs of 38 dwarfs in the NGC 2784 group are fitted by a single-exponential or double-exponential model. We find that 53% of the dwarfs are fitted with single-exponential profiles (“Type I”), while 47% of the dwarfs show double-exponential profiles; 37% of all dwarfs have smaller sizes for the outer part than the inner part (“Type II”), while 10% have a larger outer than inner part (“Type III”). We compare these results with those in the field and in the Virgo cluster, where the SBP types of 102 field dwarfs are compiled from a previous study and the SBP types of 375 cluster dwarfs are measured using SDSS r-band images. As a result, the distributions of SBP types are different in the three environments. Common SBP types for the field, the NGC 2784 group, and the Virgo cluster are Type II, Type I and II, and Type I and III profiles, respectively. After comparing the sizes of dwarfs in different environments, we suggest that since the sizes of some dwarfs are changed due to environmental effects, SBP types are capable of being transformed and the distributions of SBP types in the three environments are different. We discuss possible environmental mechanisms for the transformation of SBP types. Based on data collected at KMTNet Telescopes and SDSS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009A%26A...497..287D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009A%26A...497..287D"><span>The plasma filling factor of coronal <span class="hlt">bright</span> points. II. Combined EIS and TRACE results</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dere, K. P.</p> <p>2009-04-01</p> <p>Aims: In a previous paper, the volumetric plasma filling factor of coronal <span class="hlt">bright</span> points was determined from spectra obtained with the Extreme ultraviolet Imaging Spectrometer (EIS). The analysis of these data showed that the median plasma filling factor was 0.015. One interpretation of this result was that the small filling factor was consistent with a single coronal loop with a width of 1-2´´, somewhat below the apparent width. In this paper, higher spatial resolution <span class="hlt">observations</span> with the Transition Region and Corona Explorer (TRACE) are used to test this interpretation. Methods: Rastered spectra of regions of the quiet Sun were recorded by the EIS during operations with the Hinode satellite. Many of these regions were simultaneously <span class="hlt">observed</span> with TRACE. Calibrated intensities of Fe xii lines were obtained and images of the quiet corona were constructed from the EIS measurements. Emission measures were determined from the EIS spectra and geometrical widths of coronal <span class="hlt">bright</span> points were obtained from the TRACE images. Electron densities were determined from density-sensitive line ratios measured with EIS. A comparison of the emission measure and <span class="hlt">bright</span> point widths with the electron densities yielded the plasma filling factor. Results: The median electron density of coronal <span class="hlt">bright</span> points is 3 × 109 cm-3 at a temperature of 1.6 × 106 K. The volumetric plasma filling factor of coronal <span class="hlt">bright</span> points was found to vary from 3 × 10-3 to 0.3 with a median value of 0.04. Conclusions: The current set of EIS and TRACE coronal <span class="hlt">bright</span>-point <span class="hlt">observations</span> indicate the median value of their plasma filling factor is 0.04. This can be interpreted as evidence of a considerable subresolution structure in coronal <span class="hlt">bright</span> points or as the result of a single completely filled plasma loop with widths on the order of 0.2-1.5´´ that has not been spatially resolved in these measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18653259','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18653259"><span><span class="hlt">Bright</span>Stat.com: free statistics online.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Stricker, Daniel</p> <p>2008-10-01</p> <p>Powerful software for statistical analysis is expensive. Here I present <span class="hlt">Bright</span>Stat, a statistical software running on the Internet which is free of charge. <span class="hlt">Bright</span>Stat's goals, its main capabilities and functionalities are outlined. Three different sample runs, a Friedman test, a chi-square test, and a step-wise multiple regression are presented. The results obtained by <span class="hlt">Bright</span>Stat are compared with results computed by SPSS, one of the global leader in providing statistical software, and VassarStats, a collection of scripts for data analysis running on the Internet. Elementary statistics is an inherent part of academic education and <span class="hlt">Bright</span>Stat is an alternative to commercial products.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999A%26AS..137..101M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999A%26AS..137..101M"><span>Near-infrared <span class="hlt">observations</span> of galaxies in Pisces-Perseus. I. vec H-band <span class="hlt">surface</span> photometry of 174 spiral</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Moriondo, G.; Baffa, C.; Casertano, S.; Chincarini, G.; Gavazzi, G.; Giovanardi, C.; Hunt, L. K.; Pierini, D.; Sperandio, M.; Trinchieri, G.</p> <p>1999-05-01</p> <p>We present near-infrared, H-band (1.65 $() μm), <span class="hlt">surface</span> photometry of 174 spiral galaxies in the area of the Pisces-Perseus supercluster. The images, acquired with the ARNICA camera mounted on various telescopes, are used to derive radial profiles of <span class="hlt">surface</span> <span class="hlt">brightness</span>, ellipticities, and position angles, together with global parameters such as H-band magnitudes and diameters Radial profiles in tabular form and images FITS files are also available upon request from gmorio@arcetri.astro.it.}. The mean relation between H-band isophotal diameter D_{21.5} and the B-band D25 implies a B-H color of the outer disk bluer than 3.5; moreover, D_{21.5}/D25 depends on (global) color and absolute luminosity. The correlations among the various photometric parameters suggest a ratio between isophotal radius D_{21.5}/2 and disk scale length of ~ m3.5 and a mean disk central <span class="hlt">brightness</span> ~ meq 17.5 H-mag arcsec^{-2}. We confirm the trend of the concentration index C31$ with absolute luminosity and, to a lesser degree, with morphological type. We also assess the influence of non-axisymmetric structures on the radial profiles and on the derived parameters. Based on <span class="hlt">observations</span> at the TIRGO, NOT, and VATT telescopes. TIRGO (Gornergrat, CH) is operated by CAISMI-CNR, Arcetri, Firenze. NOT (La Palma, Canary Islands) is operated by NOTSA, the Nordic Observatory Scientific Association. VATT (Mt. Graham, Az) is operated by VORG, the Vatican Observatory Research Group Table 3 and Fig. 4 are only available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsweb.u-strasbg.fr/Abstract.html.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21044607','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21044607"><span>Characterization of <span class="hlt">brightness</span> and stoichiometry of <span class="hlt">bright</span> particles by flow-fluorescence fluctuation spectroscopy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Johnson, Jolene; Chen, Yan; Mueller, Joachim D</p> <p>2010-11-03</p> <p>Characterization of <span class="hlt">bright</span> particles at low concentrations by fluorescence fluctuation spectroscopy (FFS) is challenging, because the event rate of particle detection is low and fluorescence background contributes significantly to the measured signal. It is straightforward to increase the event rate by flow, but the high background continues to be problematic for fluorescence correlation spectroscopy. Here, we characterize the use of photon-counting histogram analysis in the presence of flow. We demonstrate that a photon-counting histogram efficiently separates the particle signal from the background and faithfully determines the <span class="hlt">brightness</span> and concentration of particles independent of flow speed, as long as undersampling is avoided. <span class="hlt">Brightness</span> provides a measure of the number of fluorescently labeled proteins within a complex and has been used to determine stoichiometry of protein complexes in vivo and in vitro. We apply flow-FFS to determine the stoichiometry of the group specific antigen protein within viral-like particles of the human immunodeficiency virus type-1 from the <span class="hlt">brightness</span>. Our results demonstrate that flow-FFS is a sensitive method for the characterization of complex macromolecular particles at low concentrations. Copyright © 2010 Biophysical Society. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.U11B0026M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.U11B0026M"><span>Sublimation of Exposed Snow Queen <span class="hlt">Surface</span> Water Ice as <span class="hlt">Observed</span> by the Phoenix Mars Lander</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Markiewicz, W. J.; Keller, H. U.; Kossacki, K. J.; Mellon, M. T.; Stubbe, H. F.; Bos, B. J.; Woida, R.; Drube, L.; Leer, K.; Madsen, M. B.; Goetz, W.; El Maarry, M. R.; Smith, P.</p> <p>2008-12-01</p> <p>One of the first images obtained by the Robotic Arm Camera on the Mars Phoenix Lander was that of the <span class="hlt">surface</span> beneath the spacecraft. This image, taken on sol 4 (Martian day) of the mission, was intended to check the stability of the footpads of the lander and to document the effect the retro-rockets had on the Martian <span class="hlt">surface</span>. Not completely unexpected the image revealed an oval shaped, relatively <span class="hlt">bright</span> and apparently smooth object, later named Snow Queen, surrounded by the regolith similar to that already seen throughout the landscape of the landing site. The object was suspected to be the <span class="hlt">surface</span> of the ice table uncovered by the blast of the retro-rockets during touchdown. High resolution HiRISE images of the landing site from orbit, show a roughly circular dark region of about 40 m diameter with the lander in the center. A plausible explanation for this region being darker than the rest of the visible Martian Northern Planes (here polygonal patterns) is that a thin layer of the material ejected by the retro-rockets covered the original <span class="hlt">surface</span>. Alternatively the thrusters may have removed the fine <span class="hlt">surface</span> dust during the last stages of the descent. A simple estimate requires that about 10 cm of the <span class="hlt">surface</span> material underneath the lander is needed to be ejected and redistributed to create the <span class="hlt">observed</span> dark circular region. 10 cm is comparable to 4-5 cm predicted depth at which the ice table was expected to be found at the latitude of the Phoenix landing site. The models also predicted that exposed water ice should sublimate at a rate not faster but probably close to 1 mm per sol. Snow Queen was further documented on sols 5, 6 and 21 with no obvious changes detected. The following time it was imaged was on sol 45, 24 sols after the previous <span class="hlt">observation</span>. This time some clear changes were obvious. Several small cracks, most likely due to thermal cycling and sublimation of water ice appeared. Nevertheless, the bulk of Snow Queen <span class="hlt">surface</span> remained smooth. The next</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA12753.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA12753.html"><span><span class="hlt">Bright</span> Enceladus</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-02-14</p> <p>Saturn moon Enceladus reflects sunlight <span class="hlt">brightly</span> while the planet and its rings fill the background in this view from NASA Cassini spacecraft. Enceladus is one of the most reflective bodies in the solar system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008cosp...37.3090S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008cosp...37.3090S"><span>Calculation of gyrosynchrotron radiation <span class="hlt">brightness</span> temperature for outer <span class="hlt">bright</span> loop of ICME</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sun, Weiying; Wu, Ji; Wang, C. B.; Wang, S.</p> <p></p> <p>:Solar polar orbit radio telescope (SPORT) is proposed to detect the high density plasma clouds of outer <span class="hlt">bright</span> loop of ICMEs from solar orbit with large inclination. Of particular interest is following the propagation of the plasma clouds with remote sensor in radio wavelength band. Gyrosynchrotron emission is a main radio radiation mechanism of the plasma clouds and can provide information of interplanetary magnetic field. In this paper, we statistically analyze the electron density, electron temperature and magnetic field of background solar wind in time of quiet sun and ICMEs propagation. We also estimate the fluctuation range of the electron density, electron temperature and magnetic field of outer <span class="hlt">bright</span> loop of ICMEs. Moreover, we calculate and analyze the emission <span class="hlt">brightness</span> temperature and degree of polarization on the basis of the study of gyrosynchrotron emission, absorption and polarization characteristics as the optical depth is less than or equal to 1.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AAS...23115223O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AAS...23115223O"><span>Investigating the <span class="hlt">Bright</span> End of LSST Photometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ojala, Elle; Pepper, Joshua; LSST Collaboration</p> <p>2018-01-01</p> <p>The Large Synoptic Survey Telescope (LSST) will begin operations in 2022, conducting a wide-field, synoptic multiband survey of the southern sky. Some fraction of objects at the <span class="hlt">bright</span> end of the magnitude regime <span class="hlt">observed</span> by LSST will overlap with other wide-sky surveys, allowing for calibration and cross-checking between surveys. The LSST is optimized for <span class="hlt">observations</span> of very faint objects, so much of this data overlap will be comprised of saturated images. This project provides the first in-depth analysis of saturation in LSST images. Using the PhoSim package to create simulated LSST images, we evaluate saturation properties of several types of stars to determine the <span class="hlt">brightness</span> limitations of LSST. We also collect metadata from many wide-field photometric surveys to provide cross-survey accounting and comparison. Additionally, we evaluate the accuracy of the PhoSim modeling parameters to determine the reliability of the software. These efforts will allow us to determine the expected useable data overlap between <span class="hlt">bright</span>-end LSST images and faint-end images in other wide-sky surveys. Our next steps are developing methods to extract photometry from saturated images.This material is based upon work supported in part by the National Science Foundation through Cooperative Agreement 1258333 managed by the Association of Universities for Research in Astronomy (AURA), and the Department of Energy under Contract No. DE-AC02-76SF00515 with the SLAC National Accelerator Laboratory. Additional LSST funding comes from private donations, grants to universities, and in-kind support from LSSTC Institutional Members.Thanks to NSF grant PHY-135195 and the 2017 LSSTC Grant Award #2017-UG06 for making this project possible.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ARBl...24..109S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ARBl...24..109S"><span>Research of the relationships between light dispersion and contrast of the registered image at different background <span class="hlt">brightness</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stoyanov, Stiliyan; Mardirossian, Garo</p> <p>2012-10-01</p> <p>The light diffraction is for telescope apparatuses an especially important characteristic which has an influence on the record image contrast from the eye <span class="hlt">observer</span>. The task of the investigation is to determine to what degree the coefficient of light diffraction influences the record image <span class="hlt">brightness</span>. The object of the theoretical research are experimental results provided from a telescope system experiment in the process of <span class="hlt">observation</span> of remote objects with different <span class="hlt">brightness</span> of the background in the fixed light diffraction coefficients and permanent contrast of the background in respect to the object. The received values and the ratio of the image contrast to the light diffraction coefficient is shown in a graphic view. It's settled that with increasing of the value of background <span class="hlt">brightness</span> in permanent background contrast in respect to the object, the image contrast sharply decrease. The relationship between the increase of the light diffraction coefficient and the decrease of the <span class="hlt">brightness</span> of the project image from telescope apparatuses can be <span class="hlt">observed</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H21D1487F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H21D1487F"><span>Predicting Near Real-Time Inundation Occurrence from Complimentary Satellite Microwave <span class="hlt">Brightness</span> Temperature <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fisher, C. K.; Pan, M.; Wood, E. F.</p> <p>2017-12-01</p> <p>Throughout the world, there is an increasing need for new methods and data that can aid decision makers, emergency responders and scientists in the monitoring of flood events as they happen. In many regions, it is possible to examine the extent of historical and real-time inundation occurrence from visible and infrared imagery provided by sensors such as MODIS or the Landsat TM; however, this is not possible in regions that are densely vegetated or are under persistent cloud cover. In addition, there is often a temporal mismatch between the sampling of a particular sensor and a given flood event, leading to limited <span class="hlt">observations</span> in near real-time. As a result, there is a need for alternative methods that take full advantage of complimentary remotely sensed data sources, such as available microwave <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> (e.g., SMAP, SMOS, AMSR2, AMSR-E, and GMI), to aid in the estimation of global flooding. The objective of this work was to develop a high-resolution mapping of inundated areas derived from multiple satellite microwave sensor <span class="hlt">observations</span> with a daily temporal resolution. This system consists of first retrieving water fractions from complimentary microwave sensors (AMSR-2 and SMAP) which may spatially and temporally overlap in the region of interest. Using additional information in a Random Forest classifier, including high resolution topography and multiple datasets of inundated area (both historical and empirical), the resulting retrievals are spatially downscaled to derive estimates of the extent of inundation at a scale relevant to management and flood response activities ( 90m or better) instead of the relatively coarse resolution water fractions, which are limited by the microwave sensor footprints ( 5-50km). Here we present the training and validation of this method for the 2015 floods that occurred in Houston, Texas. Comparing the predicted inundation against historical occurrence maps derived from the Landsat TM record and MODIS</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014DPS....4641604B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014DPS....4641604B"><span><span class="hlt">Observations</span> of Venus at 1-meter wavelength</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Butler, Bryan J.</p> <p>2014-11-01</p> <p>Radio wavelength <span class="hlt">observations</span> of Venus (including from the Magellan spacecraft) have been a powerful method of probing its <span class="hlt">surface</span> and atmosphere since the 1950's. The emission is generally understood to come from a combination of emission and absorption in the subsurface, <span class="hlt">surface</span>, and atmosphere at cm and shorter wavelengths [1]. There is, however, a long-standing mystery regarding the long wavelength emission from Venus. First discovered at wavelengths of 50 cm and greater [2], the effect was later confirmed to extend to wavelengths as short as 13 cm [1,3]. The <span class="hlt">brightness</span> temperatures are depressed significantly 50 K around 10-20 cm, increasing to as much as 200 K around 1 m) from what one would expect from a "normal" <span class="hlt">surface</span> (e.g., similar to the Moon or Earth) [1-3].No simple <span class="hlt">surface</span> and subsurface model of Venus can reproduce these large depressions in the long wavelength emission [1-3]. Simple atmospheric and ionospheric models fail similarly. In an attempt to constrain the <span class="hlt">brightness</span> temperature spectrum more fully, new <span class="hlt">observations</span> have been made at wavelengths that cover the range 60 cm to 1.3 m at the Very Large Array, using the newly available low-band receiving systems there [4]. The new <span class="hlt">observations</span> were made over a very wide wavelength range and at several Venus phases, with that wide parameter space coverage potentially allowing us to pinpoint the cause of the phenomenon. The <span class="hlt">observations</span> and potential interpretations will be presented and discussed.[1] Butler et al. 2001, Icarus, 154, 226. [2] Schloerb et al. 1976, Icarus, 29, 329; Muhleman et al. 1973, ApJ, 183, 1081; Condon et al. 1973, ApJ, 183, 1075; Kuzmin 1965, Radiophysics. [3] Butler & Sault 2003, IAUSS, 1E, 17B. [4] Intema et al. 2014, BASI, 1.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JQSRT.205..278H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JQSRT.205..278H"><span>Measuring night sky <span class="hlt">brightness</span>: methods and challenges</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hänel, Andreas; Posch, Thomas; Ribas, Salvador J.; Aubé, Martin; Duriscoe, Dan; Jechow, Andreas; Kollath, Zoltán; Lolkema, Dorien E.; Moore, Chadwick; Schmidt, Norbert; Spoelstra, Henk; Wuchterl, Günther; Kyba, Christopher C. M.</p> <p>2018-01-01</p> <p>Measuring the <span class="hlt">brightness</span> of the night sky has become an increasingly important topic in recent years, as artificial lights and their scattering by the Earth's atmosphere continue spreading around the globe. Several instruments and techniques have been developed for this task. We give an overview of these, and discuss their strengths and limitations. The different quantities that can and should be derived when measuring the night sky <span class="hlt">brightness</span> are discussed, as well as the procedures that have been and still need to be defined in this context. We conclude that in many situations, calibrated consumer digital cameras with fisheye lenses provide the best relation between ease-of-use and wealth of obtainable information on the night sky. While they do not obtain full spectral information, they are able to sample the complete sky in a period of minutes, with colour information in three bands. This is important, as given the current global changes in lamp spectra, changes in sky radiance <span class="hlt">observed</span> only with single band devices may lead to incorrect conclusions regarding long term changes in sky <span class="hlt">brightness</span>. The acquisition of all-sky information is desirable, as zenith-only information does not provide an adequate characterization of a site. Nevertheless, zenith-only single-band one-channel devices such as the "Sky Quality Meter" continue to be a viable option for long-term studies of night sky <span class="hlt">brightness</span> and for studies conducted from a moving platform. Accurate interpretation of such data requires some understanding of the colour composition of the sky light. We recommend supplementing long-term time series derived with such devices with periodic all-sky sampling by a calibrated camera system and calibrated luxmeters or luminance meters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013IAUS..289..414M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013IAUS..289..414M"><span>Improving distance estimates to nearby <span class="hlt">bright</span> stars: Combining astrometric data from Hipparcos, Nano-JASMINE and Gaia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Michalik, Daniel; Lindegren, Lennart; Hobbs, David; Lammers, Uwe; Yamada, Yoshiyuki</p> <p>2013-02-01</p> <p>Starting in 2013, Gaia will deliver highly accurate astrometric data, which eventually will supersede most other stellar catalogues in accuracy and completeness. It is, however, limited to <span class="hlt">observations</span> from magnitude 6 to 20 and will therefore not include the brightest stars. Nano-JASMINE, an ultrasmall Japanese astrometry satellite, will <span class="hlt">observe</span> these <span class="hlt">bright</span> stars, but with much lower accuracy. Hence, the Hipparcos catalogue from 1997 will likely remain the main source of accurate distances to <span class="hlt">bright</span> nearby stars. We are investigating how this might be improved by optimally combining data from all three missions through a joint astrometric solution. This would take advantage of the unique features of each mission: the historic <span class="hlt">bright</span>-star measurements of Hipparcos, the updated <span class="hlt">bright</span>-star <span class="hlt">observations</span> of Nano-JASMINE, and the very accurate reference frame of Gaia. The long temporal baseline between the missions provides additional benefits for the determination of proper motions and binary detection, which indirectly improve the parallax determination further. We present a quantitative analysis of the expected gains based on simulated data for all three missions.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005PhDT.........3U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005PhDT.........3U"><span><span class="hlt">Brightness</span> and magnetic evolution of solar coronal <span class="hlt">bright</span> points</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ugarte Urra, Ignacio</p> <p></p> <p>This thesis presents a study of the <span class="hlt">brightness</span> and magnetic evolution of several Extreme ultraviolet (EUV) coronal <span class="hlt">bright</span> points (hereafter BPs). The study was carried out using several instruments on board the Solar and Heliospheric Observatory, supported by the high resolution imaging from the Transition Region And Coronal Explorer. The results confirm that, down to 1" resolution, BPs are made of small loops with lengths of [approximate]6 Mm and cross-sections of ≈2 Mm. The loops are very dynamic, evolving in time scales as short as 1 - 2 minutes. This is reflected in a highly variable EUV response with fluctuations highly correlated in spectral lines at transition region temperatures, but not always at coronal temperatures. A wavelet analysis of the intensity variations reveals the existence of quasi-periodic oscillations with periods ranging 400--1000s, in the range of periods characteristic of the chromospheric network. The link between BPs and network <span class="hlt">bright</span> points is discussed, as well as the interpretation of the oscillations in terms of global acoustic modes of closed magnetic structures. A comparison of the magnetic flux evolution of the magnetic polarities to the EUV flux changes is also presented. Throughout their lifetime, the intrinsic EUV emission of BPs is found to be dependent on the total magnetic flux of the polarities. In short time scales, co-spatial and co-temporal coronal images and magnetograms, reveal the signature of heating events that produce sudden EUV brightenings simultaneous to magnetic flux cancellations. This is interpreted in terms of magnetic reconnection events. Finally, a electron density study of six coronal <span class="hlt">bright</span> points produces values of ≈1.6×10 9 cm -3 , closer to active region plasma than to quiet Sun. The analysis of a large coronal loop (half length of 72 Mm) introduces the discussion on the prospects of future plasma diagnostics of BPs with forthcoming solar missions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9730E..0CH','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9730E..0CH"><span>Teradiode's high <span class="hlt">brightness</span> semiconductor lasers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huang, Robin K.; Chann, Bien; Burgess, James; Lochman, Bryan; Zhou, Wang; Cruz, Mike; Cook, Rob; Dugmore, Dan; Shattuck, Jeff; Tayebati, Parviz</p> <p>2016-03-01</p> <p>TeraDiode is manufacturing multi-kW-class ultra-high <span class="hlt">brightness</span> fiber-coupled direct diode lasers for industrial applications. A fiber-coupled direct diode laser with a power level of 4,680 W from a 100 μm core diameter, <0.08 numerical aperture (NA) output fiber at a single center wavelength was demonstrated. Our TeraBlade industrial platform achieves world-record <span class="hlt">brightness</span> levels for direct diode lasers. The fiber-coupled output corresponds to a Beam Parameter Product (BPP) of 3.5 mm-mrad and is the lowest BPP multi-kW-class direct diode laser yet reported. This laser is suitable for industrial materials processing applications, including sheet metal cutting and welding. This 4-kW fiber-coupled direct diode laser has comparable <span class="hlt">brightness</span> to that of industrial fiber lasers and CO2 lasers, and is over 10x brighter than state-of-the-art direct diode lasers. We have also demonstrated novel high peak power lasers and high <span class="hlt">brightness</span> Mid-Infrared Lasers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22663934-observed-asteroid-surface-area-thermal-infrared','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22663934-observed-asteroid-surface-area-thermal-infrared"><span><span class="hlt">OBSERVED</span> ASTEROID <span class="hlt">SURFACE</span> AREA IN THE THERMAL INFRARED</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Nugent, C. R.; Mainzer, A.; Masiero, J.</p> <p></p> <p>The rapid accumulation of thermal infrared <span class="hlt">observations</span> and shape models of asteroids has led to increased interest in thermophysical modeling. Most of these infrared <span class="hlt">observations</span> are unresolved. We consider what fraction of an asteroid’s <span class="hlt">surface</span> area contributes the bulk of the emitted thermal flux for two model asteroids of different shapes over a range of thermal parameters. The resulting <span class="hlt">observed</span> <span class="hlt">surface</span> in the infrared is generally more fragmented than the area <span class="hlt">observed</span> in visible wavelengths, indicating high sensitivity to shape. For objects with low values of the thermal parameter, small fractions of the <span class="hlt">surface</span> contribute the majority of thermally emittedmore » flux. Calculating <span class="hlt">observed</span> areas could enable the production of spatially resolved thermal inertia maps from non-resolved <span class="hlt">observations</span> of asteroids.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ATel11448....1D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ATel11448....1D"><span>Very <span class="hlt">bright</span> optical transient near the Trifid and Lagoon Nebulae</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dunsby, Peter</p> <p>2018-03-01</p> <p>Peter Dunsby (University of Cape Town) reports the detection of a very <span class="hlt">bright</span> optical transient in the region between the Lagoon and Trifid Nebulae based on <span class="hlt">observations</span> obtained from Cape Town on 20 March 2018, between 01:00 and 03:45 UT. The object was visible throughout the full duration of the <span class="hlt">observations</span> and not seen when this field was <span class="hlt">observed</span> previously (08 March 2018).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA00569&hterms=sponge&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dsponge','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA00569&hterms=sponge&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dsponge"><span><span class="hlt">Bright</span> Summer Afternoon on the Mars Utopian Planitia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1976-01-01</p> <p>A UTOPIAN <span class="hlt">BRIGHT</span> SUMMER AFTERNOON ON MARS--Looking south from Viking 2 on September 6, the orange-red <span class="hlt">surface</span> of the nearly level plain upon which the spacecraft sits is seen strewn with rocks as large as three feet across. Many of these rocks are porous and sponge-like, similar to some of Earth's volcanic rocks. Other rocks are coarse-grained such as the large rock at lower left. Between the rocks, the <span class="hlt">surface</span> is blanketed with fine-grained material that, in places, is piled into small drifts and banked against some of the larger blocks. The cylindrical mast with the orange cable is the low-gain antenna used to receive commands from Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19459701','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19459701"><span>Ferri<span class="hlt">BRIGHT</span>: a rationally designed fluorescent probe for redox active metals.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kennedy, Daniel P; Kormos, Chad M; Burdette, Shawn C</p> <p>2009-06-24</p> <p>The novel catechol-BODIPY dyad, 8-(3,4-dihydroxyphenyl)-2,6-bis(ethoxycarbonyl)-1,3,5,7-tetramethyl-4,4-difluoro-4-bora-3a,4a-diaza-s-indacene (Ferri<span class="hlt">BRIGHT</span>) was rationally designed with the aid of computational methods. Ferri<span class="hlt">BRIGHT</span> could be prepared by standard one-pot synthesis of BODIPY fluorophores from 3,4-bis(benzyloxy)benzaldehyde (1) and 3,5-dimethyl-4-(ethoxycarbonyl)pyrrole (3); however, isolating the dipyrrin intermediate 8-[3,4-bis(benzyloxy)phenyl]-2,6-bis(ethoxycarbonyl)-1,3,5,7-tetramethyl-4,4-diaza-s-indacene (7) prior to reaction with excess BF(3).OEt(2) led to marked improvements in the isolated overall yield of the desired compound. In addition to these improvements in fluorophore synthesis, microwave-assisted palladium-catalyzed hydrogenolysis of benzyl ethers was used to reduce reaction times and catalyst loading in preparation of the desired compound. When Ferri<span class="hlt">BRIGHT</span> is exposed to excess FeCl(3), CuCl(2), [Co(NH(3))(5)Cl]Cl(2), 2,3-dichloro-5,6-dicyanobenzoquinone, or ceric ammonium nitrate in methanol, a significant enhancement of fluorescence is <span class="hlt">observed</span>. Ferri<span class="hlt">BRIGHT</span>-Q, the product resulting from the oxidation of the pendant catechol to the corresponding quinone, was found to be the emissive species. Ferri<span class="hlt">BRIGHT</span>-Q was synthesized independently, isolated, and fully characterized to allow for direct comparison with the spectroscopic data acquired in solution. Biologically relevant reactive oxygen species, such as H(2)O(2), (*)OH, (1)O(2), O(2)(*-), and bleach (NaOCl), failed to cause any changes in the emission intensity of Ferri<span class="hlt">BRIGHT</span>. In accordance with the quantum mechanical calculations, the quantum yield of fluorescence for Ferri<span class="hlt">BRIGHT</span> (Phi(fl) approximately 0) and Ferri<span class="hlt">BRIGHT</span>-Q (Phi(fl) = 0.026, lambda(ex)/lambda(em) = 490 nm/510 nm) suggests that photoinduced electron transfer between the catechol and the BODIPY dye is attenuated upon oxidation, which results in fluorescence enhancement. Binding studies of Ferri<span class="hlt">BRIGHT</span> with Ga(NO(3</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004ApJ...616..439S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004ApJ...616..439S"><span>Giant Pulses from PSR B1937+21 with Widths <=15 Nanoseconds and Tb>=5×1039 K, the Highest <span class="hlt">Brightness</span> Temperature <span class="hlt">Observed</span> in the Universe</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Soglasnov, V. A.; Popov, M. V.; Bartel, N.; Cannon, W.; Novikov, A. Yu.; Kondratiev, V. I.; Altunin, V. I.</p> <p>2004-11-01</p> <p>Giant radio pulses of the millisecond pulsar B1937+21 were recorded with the S2 VLBI system at 1.65 GHz with NASA/JPL's 70 m radio telescope at Tidbinbilla, Australia. These pulses have been <span class="hlt">observed</span> as strong as 65,000 Jy with widths <=15 ns, corresponding to a <span class="hlt">brightness</span> temperature of Tb>=5×1039 K, the highest <span class="hlt">observed</span> in the universe. The vast majority of these pulses occur in 5.8 and 8.2 μs windows at the very trailing edges of the regular main pulse and interpulse profiles, respectively. Giant pulses occur, in general, with a single spike. Only in one case of 309 was the structure clearly more complex. The cumulative distribution is fitted by a power law with index -1.40+/-0.01 with a low-energy but no high-energy cutoff. We estimate that giant pulses occur frequently but are only rarely detected. When corrected for the directivity factor, 25 giant pulses are estimated to be generated in one neutron star revolution alone. The intensities of the giant pulses of the main pulses and interpulses are not correlated with each other nor with the intensities or energies of the main pulses and interpulses themselves. Their radiation energy density can exceed 300 times the plasma energy density at the <span class="hlt">surface</span> of the neutron star and can even exceed the magnetic field energy density at that <span class="hlt">surface</span>. We therefore do not think that the generation of giant pulses is linked to the plasma mechanisms in the magnetosphere. Instead we suggest that it is directly related to discharges in the polar cap region of the pulsar.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17797099','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17797099"><span>Temperatures of the martian <span class="hlt">surface</span> and atmosphere: viking <span class="hlt">observation</span> of diurnal and geometric variations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kieffer, H H; Christensen, P R; Martin, T Z; Miner, E D; Palluconi, F D</p> <p>1976-12-11</p> <p>Selected <span class="hlt">observations</span> made with the Viking infrared thermal mapper after the first landing are reported. Atmospheric temperatures measured at the latitude of the Viking 2 landing site (48 degrees N) over most of a martian day reveal a diurnal variation of at least 15 K, with peak temperatures occurring near 2.2 hours after noon, implying significant absorption of sunlight in the lower 30 km of the atmosphere by entrained dust. The summit temperature of Arsia Mons varies by a factor of nearly two each day; large diurnal temperature variation is characteristic of the south Tharsis upland and implies the presence of low thermal inertia material. The thermal inertia of material on the floors of several typical large craters is found to be higher than for the surrounding terrain; this suggests that craters are somehow effective in sorting aeolian material. <span class="hlt">Brightness</span> temperatures of the Viking 1 landing area decrease at large emission angles; the intensity of reflected sunlight shows a more complex dependence on geometry than expected, implying atmospheric as well as <span class="hlt">surface</span> scattering.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8795756','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8795756"><span><span class="hlt">Brightness</span> discrimination test is not useful in screening for open angle glaucoma.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Peter, E; Thomas, R; Muliyil, J</p> <p>1996-06-01</p> <p><span class="hlt">Brightness</span> discrimination test (BDT) is routinely employed to assess asymmetrical optic nerve dysfunction and has been suggested as a screening test for primary open angle glaucoma (POAG). We tested the reliability and validity of BDT in the diagnosis of POAG. The study groups included 34 patients with established primary open angle glaucoma, 20 glaucoma suspects, and 33 age-sex matched controls. Cataract was not an exclusion criterion in these groups. The normal <span class="hlt">brightness</span> score was determined to be 88% (mean score, 94%-2 SD) in a pilot study. <span class="hlt">Brightness</span> discrimination test was performed in all subjects by two <span class="hlt">observers</span> independently. BDT showed an excellent interobserver agreement (weighted Kappa 0.84). The presence of a cataract alone increased the risk of <span class="hlt">brightness</span> impairment twofold, glaucoma alone increased the risk eightfold, and the presence of both conditions by 17 times compared to those with neither condition. BDT was not a useful test in the diagnosis of POAG (sensitivity 67% and specificity 93%); the ability to detect a significant field defect was also poor (sensitivity 53% and specificity 76%). There was poor association between decreased <span class="hlt">brightness</span> scores and asymmetrical field defects as determined by the Humphrey's field analyzer (HFA).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19760042316&hterms=sparrow&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsparrow','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19760042316&hterms=sparrow&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsparrow"><span>The Skylab ten color photoelectric polarimeter. [sky <span class="hlt">brightness</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Weinberg, J. L.; Hahn, R. C.; Sparrow, J. G.</p> <p>1975-01-01</p> <p>A 10-color photoelectric polarimeter was used during Skylab missions SL-2 and SL-3 to measure sky <span class="hlt">brightness</span> and polarization associated with zodiacal light, background starlight, and the spacecraft corona. A description is given of the instrument and <span class="hlt">observing</span> routines together with initial results on the spacecraft corona and polarization of the zodiacal light.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/869440','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/869440"><span>High <span class="hlt">brightness</span> electron accelerator</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Sheffield, Richard L.; Carlsten, Bruce E.; Young, Lloyd M.</p> <p>1994-01-01</p> <p>A compact high <span class="hlt">brightness</span> linear accelerator is provided for use, e.g., in a free electron laser. The accelerator has a first plurality of acclerating cavities having end walls with four coupling slots for accelerating electrons to high velocities in the absence of quadrupole fields. A second plurality of cavities receives the high velocity electrons for further acceleration, where each of the second cavities has end walls with two coupling slots for acceleration in the absence of dipole fields. The accelerator also includes a first cavity with an extended length to provide for phase matching the electron beam along the accelerating cavities. A solenoid is provided about the photocathode that emits the electons, where the solenoid is configured to provide a substantially uniform magnetic field over the photocathode <span class="hlt">surface</span> to minimize emittance of the electons as the electrons enter the first cavity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.P41C..08T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.P41C..08T"><span>ALMA <span class="hlt">observation</span> of Ceres' <span class="hlt">Surface</span> Temperature.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Titus, T. N.; Li, J. Y.; Sykes, M. V.; Ip, W. H.; Lai, I.; Moullet, A.</p> <p>2016-12-01</p> <p>Ceres, the largest object in the main asteroid belt, has been mapped by the Dawn spacecraft. The mapping includes measuring <span class="hlt">surface</span> temperatures using the Visible and Infrared (VIR) spectrometer at high spatial resolution. However, the VIR instrument has a long wavelength cutoff at 5 μm, which prevents the accurate measurement of <span class="hlt">surface</span> temperatures below 180 K. This restricts temperature determinations to low and mid-latitudes at mid-day. <span class="hlt">Observations</span> from the Atacama Large Millimeter/submillimeter Array (ALMA) [1], while having lower spatial resolution, are sensitive to the full range of <span class="hlt">surface</span> temperatures that are expected at Ceres. Forty reconstructed images at 75 km/beam resolution were acquired of Ceres that were consistent with a low thermal inertia <span class="hlt">surface</span>. The diurnal temperature profiles were compared to the KRC thermal model [2, 3], which has been extensively used for Mars [e.g. 4, 5]. Variations in temperature as a function of local time are <span class="hlt">observed</span> and are compared to predictions from the KRC model. The model temperatures are converted to radiance (Jy/Steradian) and are corrected for near-<span class="hlt">surface</span> thermal gradients and limb effects for comparison to <span class="hlt">observations</span>. Initial analysis is consistent with the presence of near-<span class="hlt">surface</span> water ice in the north polar region. The edge of the ice table is between 50° and 70° North Latitude, consistent with the enhanced detection of hydrogen by the Dawn GRaND instrument [6]. Further analysis will be presented. This work is supported by the NASA Solar System <span class="hlt">Observations</span> Program. References: [1] Wootten A. et al. (2015) IAU General Assembly, Meeting #29, #2237199 [2] Kieffer, H. H., et al. (1977) JGR, 82, 4249-4291. [3] Kieffer, Hugh H., (2013) Journal of Geophysical Research: Planets, 118(3), 451-470. [4] Titus, T. N., H. H. Kieffer, and P. N. Christensen (2003) Science, 299, 1048-1051. [5] Fergason, R. L. et al. (2012) Space Sci. Rev, 170, 739-773[6] Prettyman, T. et al. (2016) LPSC 47, #2228.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01179&hterms=Dark+web&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDark%2Bweb','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01179&hterms=Dark+web&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDark%2Bweb"><span>Dark and <span class="hlt">Bright</span> Ridges on Europa</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1998-01-01</p> <p>This high-resolution image of Jupiter's moon Europa, taken by NASA's Galileo spacecraft camera, shows dark, relatively smooth region at the lower right hand corner of the image which may be a place where warm ice has welled up from below. The region is approximately 30 square kilometers in area. An isolated <span class="hlt">bright</span> hill stands within it. The image also shows two prominent ridges which have different characteristics; youngest ridge runs from left to top right and is about 5 kilometers in width (about 3.1 miles). The ridge has two <span class="hlt">bright</span>, raised rims and a central valley. The rims of the ridge are rough in texture. The inner and outer walls show <span class="hlt">bright</span> and dark debris streaming downslope, some of it forming broad fans. This ridge overlies and therefore must be younger than a second ridge running from top to bottom on the left side of the image. This dark 2 km wide ridge is relatively flat, and has smaller-scale ridges and troughs along its length.<p/>North is to the top of the picture, and the sun illuminates the <span class="hlt">surface</span> from the upper left. This image, centered at approximately 14 degrees south latitude and 194 degrees west longitude, covers an area approximately 15 kilometers by 20 kilometers (9 miles by 12 miles). The resolution is 26 meters (85 feet) per picture element. This image was taken on December 16, 1997 at a range of 1300 kilometers (800 miles) by Galileo's solid state imaging system.<p/>The Jet Propulsion Laboratory, Pasadena, CA manages the Galileo mission for NASA's Office of Space Science, Washington, DC. JPL is an operating division of California Institute of Technology (Caltech).<p/>This image and other images and data received from Galileo are posted on the World Wide Web, on the Galileo mission home page at URL http://www.jpl.nasa.gov/ galileo.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JASTP.169...83D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JASTP.169...83D"><span>Response of noctilucent cloud <span class="hlt">brightness</span> to daily solar variations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dalin, P.; Pertsev, N.; Perminov, V.; Dubietis, A.; Zadorozhny, A.; Zalcik, M.; McEachran, I.; McEwan, T.; Černis, K.; Grønne, J.; Taustrup, T.; Hansen, O.; Andersen, H.; Melnikov, D.; Manevich, A.; Romejko, V.; Lifatova, D.</p> <p>2018-04-01</p> <p>For the first time, long-term data sets of ground-based <span class="hlt">observations</span> of noctilucent clouds (NLC) around the globe have been analyzed in order to investigate a response of NLC to solar UV irradiance variability on a day-to-day scale. NLC <span class="hlt">brightness</span> has been considered versus variations of solar Lyman-alpha flux. We have found that day-to-day solar variability, whose effect is generally masked in the natural NLC variability, has a statistically significant effect when considering large statistics for more than ten years. Average increase in day-to-day solar Lyman-α flux results in average decrease in day-to-day NLC <span class="hlt">brightness</span> that can be explained by robust physical mechanisms taking place in the summer mesosphere. Average time lags between variations of Lyman-α flux and NLC <span class="hlt">brightness</span> are short (0-3 days), suggesting a dominant role of direct solar heating and of the dynamical mechanism compared to photodissociation of water vapor by solar Lyman-α flux. All found regularities are consistent between various ground-based NLC data sets collected at different locations around the globe and for various time intervals. Signatures of a 27-day periodicity seem to be present in the NLC <span class="hlt">brightness</span> for individual summertime intervals; however, this oscillation cannot be unambiguously retrieved due to inevitable periods of tropospheric cloudiness.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA21398.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA21398.html"><span>Occator <span class="hlt">Bright</span> Spots in 3-D</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-03-09</p> <p>This 3-D image, or anaglyph, shows the center of Occator Crater, the brightest area on dwarf planet Ceres, using data from NASA's Dawn mission. The <span class="hlt">bright</span> central area, including a dome that is 0.25 miles (400 meters) high, is called Cerealia Facula. The secondary, scattered <span class="hlt">bright</span> areas are called Vinalia Faculae. A 2017 study suggests that the central <span class="hlt">bright</span> area is significantly younger than Occator Crater. Estimates put Cerealia Facula at 4 million years old, while Occator Crater is approximately 34 million years old. The reflective material that appears so <span class="hlt">bright</span> in this image is made of carbonate salts, according to Dawn researchers. The Vinalia Faculae seem to be composed of carbonates mixed with dark material. http://photojournal.jpl.nasa.gov/catalog/PIA21398</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvS..21c2802T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvS..21c2802T"><span>Time-resolved <span class="hlt">brightness</span> measurements by streaking</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Torrance, Joshua S.; Speirs, Rory W.; McCulloch, Andrew J.; Scholten, Robert E.</p> <p>2018-03-01</p> <p><span class="hlt">Brightness</span> is a key figure of merit for charged particle beams, and time-resolved <span class="hlt">brightness</span> measurements can elucidate the processes involved in beam creation and manipulation. Here we report on a simple, robust, and widely applicable method for the measurement of beam <span class="hlt">brightness</span> with temporal resolution by streaking one-dimensional pepperpots, and demonstrate the technique to characterize electron bunches produced from a cold-atom electron source. We demonstrate <span class="hlt">brightness</span> measurements with 145 ps temporal resolution and a minimum resolvable emittance of 40 nm rad. This technique provides an efficient method of exploring source parameters and will prove useful for examining the efficacy of techniques to counter space-charge expansion, a critical hurdle to achieving single-shot imaging of atomic scale targets.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1989SPIE.1010..193C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1989SPIE.1010..193C"><span><span class="hlt">Surface</span> Relief of Mapping</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Costa, Manuel F.; Almeida, Jose B.</p> <p>1989-02-01</p> <p>We will describe in this communication a noncont act method of measuring <span class="hlt">surface</span> profile, it does not require any <span class="hlt">surface</span> preparation, and it can be used with a very large range of <span class="hlt">surfaces</span> from highly reflecting to non reflecting ones and as complex as textile <span class="hlt">surfaces</span>. This method is reasonably immune to dispersion and diffraction, which usually make very difficult the application of non contact profilometry methods to a wide range of materials and situations, namely on quality control systems in industrial production lines. The method is based on the horizontal shift of the <span class="hlt">bright</span> spot on a horizontal <span class="hlt">surface</span> when this is illuminated with an oblique beam and moved vertically. in order to make the profilometry the sample is swept by an oblique light beam and the <span class="hlt">bright</span> spot position is compared with a reference position. The <span class="hlt">bright</span> spot must be as small as possible, particularly in very irregular <span class="hlt">surfaces</span>; so the light beam diameter must be as small as possible and the incidence angle must not be too small. The sensivity of a system based on this method will be given, mostly, by the reception optical system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70028728','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70028728"><span>Titan: Preliminary results on <span class="hlt">surface</span> properties and photometry from VIMS <span class="hlt">observations</span> of the early flybys</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Buratti, B.J.; Sotin, Christophe; Brown, R.H.; Hicks, M.D.; Clark, R.N.; Mosher, J.A.; McCord, T.B.; Jaumann, R.; Baines, K.H.; Nicholson, P.D.; Momary, T.; Simonelli, D.P.; Sicardy, B.</p> <p>2006-01-01</p> <p>Cassini <span class="hlt">observations</span> of the <span class="hlt">surface</span> of Titan offer unprecedented views of its <span class="hlt">surface</span> through atmospheric windows in the 1-5 ??m region. Images obtained in windows for which the haze opacity is low can be used to derive quantitative photometric parameters such as albedo and albedo distribution, and physical properties such as roughness and particle characteristics. Images from the early Titan flybys, particularly T0, Ta, and T5 have been analyzed to create albedo maps in the 2.01 and 2.73 ??m windows. We find the average normal reflectance at these two wavelengths to be 0.15??0.02 and 0.035??0.003, respectively. Titan's <span class="hlt">surface</span> is bifurcated into two albedo regimes, particularly at 2.01 ??m. Analysis of these two regimes to understand the physical character of the <span class="hlt">surface</span> was accomplished with a macroscopic roughness model. We find that the two types of <span class="hlt">surface</span> have substantially different roughness, with the low-albedo <span class="hlt">surface</span> exhibiting mean slope angles of ???18??, and the high-albedo terrain having a much more substantial roughness with a mean slope angle of ???34??. A single-scattering phase function approximated by a one-term Henyey-Greenstein equation was also fit to each unit. Titan's <span class="hlt">surface</span> is back-scattering (g???0.3-0.4), and does not exhibit substantially different backscattering behavior between the two terrains. Our results suggest that two distinct geophysical domains exist on Titan: a <span class="hlt">bright</span> region cut by deep drainage channels and a relatively smooth <span class="hlt">surface</span>. The two terrains are covered by a film or a coating of particles perhaps precipitated from the satellite's haze layer and transported by eolian processes. Our results are preliminary: more accurate values for the <span class="hlt">surface</span> albedo and physical parameters will be derived as more data is gathered by the Cassini spacecraft and as a more complete radiative transfer model is developed from both Cassini orbiter and Huygens Lander measurements. ?? 2006 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017OptEn..56k4103K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017OptEn..56k4103K"><span>On correct evaluation techniques of <span class="hlt">brightness</span> enhancement effect measurement data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kukačka, Leoš; Dupuis, Pascal; Motomura, Hideki; Rozkovec, Jiří; Kolář, Milan; Zissis, Georges; Jinno, Masafumi</p> <p>2017-11-01</p> <p>This paper aims to establish confidence intervals of the quantification of <span class="hlt">brightness</span> enhancement effects resulting from the use of pulsing <span class="hlt">bright</span> light. It is found that the methods used so far may yield significant bias in the published results, overestimating or underestimating the enhancement effect. The authors propose to use a linear algebra method called the total least squares. Upon an example dataset, it is shown that this method does not yield biased results. The statistical significance of the results is also computed. It is concluded over an <span class="hlt">observation</span> set that the currently used linear algebra methods present many patterns of noise sensitivity. Changing algorithm details leads to inconsistent results. It is thus recommended to use the method with the lowest noise sensitivity. Moreover, it is shown that this method also permits one to obtain an estimate of the confidence interval. This paper neither aims to publish results about a particular experiment nor to draw any particular conclusion about existence or nonexistence of the <span class="hlt">brightness</span> enhancement effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.tmp.1339V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.tmp.1339V"><span>HD 89345: a <span class="hlt">bright</span> oscillating star hosting a transiting warm Saturn-sized planet <span class="hlt">observed</span> by K2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Van Eylen, V.; Dai, F.; Mathur, S.; Gandolfi, D.; Albrecht, S.; Fridlund, M.; García, R. A.; Guenther, E.; Hjorth, M.; Justesen, A. B.; Livingston, J.; Lund, M. N.; Pérez Hernández, F.; Prieto-Arranz, J.; Regulo, C.; Bugnet, L.; Everett, M. E.; Hirano, T.; Nespral, D.; Nowak, G.; Palle, E.; Silva Aguirre, V.; Trifonov, T.; Winn, J. N.; Barragán, O.; Beck, P. G.; Chaplin, W. J.; Cochran, W. D.; Csizmadia, S.; Deeg, H.; Endl, M.; Heeren, P.; Grziwa, S.; Hatzes, A. P.; Hidalgo, D.; Korth, J.; Mathis, S.; Montañes Rodriguez, P.; Narita, N.; Patzold, M.; Persson, C. M.; Rodler, F.; Smith, A. M. S.</p> <p>2018-05-01</p> <p>We report the discovery and characterization of HD 89345b (K2-234b; EPIC 248777106b), a Saturn-sized planet orbiting a slightly evolved star. HD 89345 is a <span class="hlt">bright</span> star (V = 9.3 mag) <span class="hlt">observed</span> by the K2 mission with one-minute time sampling. It exhibits solar-like oscillations. We conducted asteroseismology to determine the parameters of the star, finding the mass and radius to be 1.12^{+0.04}_{-0.01} M_⊙ and 1.657^{+0.020}_{-0.004} R_⊙, respectively. The star appears to have recently left the main sequence, based on the inferred age, 9.4^{+0.4}_{-1.3} Gyr, and the non-detection of mixed modes. The star hosts a "warm Saturn" (P = 11.8 days, Rp = 6.86 ± 0.14 R⊕). Radial-velocity follow-up <span class="hlt">observations</span> performed with the FIES, HARPS, and HARPS-N spectrographs show that the planet has a mass of 35.7 ± 3.3 M⊕. The data also show that the planet's orbit is eccentric (e ≈ 0.2). An investigation of the rotational splitting of the oscillation frequencies of the star yields no conclusive evidence on the stellar inclination angle. We further obtained Rossiter-McLaughlin <span class="hlt">observations</span>, which result in a broad posterior of the stellar obliquity. The planet seems to conform to the same patterns that have been <span class="hlt">observed</span> for other sub-Saturns regarding planet mass and multiplicity, orbital eccentricity, and stellar metallicity.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.468.2372N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.468.2372N"><span><span class="hlt">Brightness</span> temperature - obtaining the physical properties of a non-equipartition plasma</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nokhrina, E. E.</p> <p>2017-06-01</p> <p>The limit on the intrinsic <span class="hlt">brightness</span> temperature, attributed to `Compton catastrophe', has been established being 1012 K. Somewhat lower limit of the order of 1011.5 K is implied if we assume that the radiating plasma is in equipartition with the magnetic field - the idea that explained why the <span class="hlt">observed</span> cores of active galactic nuclei (AGNs) sustained the limit lower than the `Compton catastrophe'. Recent <span class="hlt">observations</span> with unprecedented high resolution by the RadioAstron have revealed systematic exceed in the <span class="hlt">observed</span> <span class="hlt">brightness</span> temperature. We propose means of estimating the degree of the non-equipartition regime in AGN cores. Coupled with the core-shift measurements, the method allows us to independently estimate the magnetic field strength and the particle number density at the core. We show that the ratio of magnetic energy to radiating plasma energy is of the order of 10-5, which means the flow in the core is dominated by the particle energy. We show that the magnetic field obtained by the <span class="hlt">brightness</span> temperature measurements may be underestimated. We propose for the relativistic jets with small viewing angles the non-uniform magnetohydrodynamic model and obtain the expression for the magnetic field amplitude about two orders higher than that for the uniform model. These magnetic field amplitudes are consistent with the limiting magnetic field suggested by the `magnetically arrested disc' model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017xru..conf..116K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017xru..conf..116K"><span>CXB <span class="hlt">surface</span> <span class="hlt">brightness</span> fluctuations: A new frontier of ICM structure and outskirts studies of (un)resolved galaxy clusters?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kolodzig, A.; Gilfanov, M.; Hutsi, G.; Sunyaev, R.</p> <p>2017-10-01</p> <p><span class="hlt">Surface</span> <span class="hlt">brightness</span> fluctuations of the cosmic X-ray background (CXB) carry unique information about the intracluster-medium (ICM) structure of galaxy clusters and groups up to the virial radius, which is inaccessible by conventional <span class="hlt">observations</span> of selected nearby resolved clusters. We present results of our CXB fluctuation analysis of the ˜5ks-deep, ˜9deg^2-large Chandra survey XBOOTES. We find that our fluctuation signal of resolved clusters is dominated by nearby, high-luminosity sources. The shape of its power spectrum suggests that for the brightest cluster we are sensitive to the ICM structure up to ˜2× R_{500};(˜2 Mpc/h). The energy spectrum of the fluctuation signal from resolved and unresolved clusters follows a typical ICM spectrum, where redshifts and temperatures are consistent with expectations. It also demonstrates that fluctuations of our unresolved CXB are dominated by unresolved clusters with an average z˜0.4 and T˜1.3keV, suggesting an average L_{0.5-2keV}˜3×10^{42} erg/s and M_{500}˜4×10^{13} M_{Sun}/h. Comparison with modeling suggests, that our fluctuation signal can be described with the one-halo-term of clusters and that it might be sensitive to the presence of substructures. Discrepancies between model and measurement could be utilized to improve our understanding of the ICM structure in a statistical manner. We briefly discuss the potential of larger surveys (e.g. Stripe82, XXL, SRG/eRosita).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26016658','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26016658"><span>Afternoon nap and <span class="hlt">bright</span> light exposure improve cognitive flexibility post lunch.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Slama, Hichem; Deliens, Gaétane; Schmitz, Rémy; Peigneux, Philippe; Leproult, Rachel</p> <p>2015-01-01</p> <p>Beneficial effects of napping or <span class="hlt">bright</span> light exposure on cognitive performance have been reported in participants exposed to sleep loss. Nonetheless, few studies investigated the effect of these potential countermeasures against the temporary drop in performance <span class="hlt">observed</span> in mid-afternoon, and even less so on cognitive flexibility, a crucial component of executive functions. This study investigated the impact of either an afternoon nap or <span class="hlt">bright</span> light exposure on post-prandial alterations in task switching performance in well-rested participants. Twenty-five healthy adults participated in two randomized experimental conditions, either wake versus nap (n=15), or <span class="hlt">bright</span> light versus placebo (n=10). Participants were tested on a switching task three times (morning, post-lunch and late afternoon sessions). The interventions occurred prior to the post-lunch session. In the nap/wake condition, participants either stayed awake watching a 30-minute documentary or had the opportunity to take a nap for 30 minutes. In the <span class="hlt">bright</span> light/placebo condition, participants watched a documentary under either <span class="hlt">bright</span> blue light or dim orange light (placebo) for 30 minutes. The switch cost estimates cognitive flexibility and measures task-switching efficiency. Increased switch cost scores indicate higher difficulties to switch between tasks. In both control conditions (wake or placebo), accuracy switch-cost score increased post lunch. Both interventions (nap or <span class="hlt">bright</span> light) elicited a decrease in accuracy switch-cost score post lunch, which was associated with diminished fatigue and decreased variability in vigilance. Additionally, there was a trend for a post-lunch benefit of <span class="hlt">bright</span> light with a decreased latency switch-cost score. In the nap group, improvements in accuracy switch-cost score were associated with more NREM sleep stage N1. Thus, exposure to <span class="hlt">bright</span> light during the post-lunch dip, a countermeasure easily applicable in daily life, results in similar beneficial effects as</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4446306','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4446306"><span>Afternoon Nap and <span class="hlt">Bright</span> Light Exposure Improve Cognitive Flexibility Post Lunch</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Schmitz, Rémy; Peigneux, Philippe; Leproult, Rachel</p> <p>2015-01-01</p> <p>Beneficial effects of napping or <span class="hlt">bright</span> light exposure on cognitive performance have been reported in participants exposed to sleep loss. Nonetheless, few studies investigated the effect of these potential countermeasures against the temporary drop in performance <span class="hlt">observed</span> in mid-afternoon, and even less so on cognitive flexibility, a crucial component of executive functions. This study investigated the impact of either an afternoon nap or <span class="hlt">bright</span> light exposure on post-prandial alterations in task switching performance in well-rested participants. Twenty-five healthy adults participated in two randomized experimental conditions, either wake versus nap (n=15), or <span class="hlt">bright</span> light versus placebo (n=10). Participants were tested on a switching task three times (morning, post-lunch and late afternoon sessions). The interventions occurred prior to the post-lunch session. In the nap/wake condition, participants either stayed awake watching a 30-minute documentary or had the opportunity to take a nap for 30 minutes. In the <span class="hlt">bright</span> light/placebo condition, participants watched a documentary under either <span class="hlt">bright</span> blue light or dim orange light (placebo) for 30 minutes. The switch cost estimates cognitive flexibility and measures task-switching efficiency. Increased switch cost scores indicate higher difficulties to switch between tasks. In both control conditions (wake or placebo), accuracy switch-cost score increased post lunch. Both interventions (nap or <span class="hlt">bright</span> light) elicited a decrease in accuracy switch-cost score post lunch, which was associated with diminished fatigue and decreased variability in vigilance. Additionally, there was a trend for a post-lunch benefit of <span class="hlt">bright</span> light with a decreased latency switch-cost score. In the nap group, improvements in accuracy switch-cost score were associated with more NREM sleep stage N1. Thus, exposure to <span class="hlt">bright</span> light during the post-lunch dip, a countermeasure easily applicable in daily life, results in similar beneficial effects as</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150016525','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150016525"><span>A <span class="hlt">Bright</span> Lunar Impact Flash Linked to the Virginid Meteor Complex</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Moser, D. E.; Suggs, R. M.; Suggs, R. J.</p> <p>2015-01-01</p> <p>On 17 March 2013 at 03:50:54 UTC, NASA detected a <span class="hlt">bright</span> impact flash on the Moon caused by a meteoroid impacting the lunar <span class="hlt">surface</span>. There was meteor activity in Earth's atmosphere the same night from the Virginid Meteor Complex. The impact crater associated with the impact flash was found and imaged by Lunar Reconnaissance Orbiter (LRO). Goal: Monitor the Moon for impact flashes produced by meteoroids striking the lunar <span class="hlt">surface</span>. Determine meteoroid flux in the 10's gram to kilogram size range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C11F..03M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C11F..03M"><span>Effect of Different Tree canopies on the <span class="hlt">Brightness</span> Temperature of Snowpack</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mousavi, S.; De Roo, R. D.; Brucker, L.</p> <p>2017-12-01</p> <p>Snow stores the water we drink and is essential to grow food that we eat. But changes in snow quantities such as snow water equivalent (SWE) are underway and have serious consequences. So, effective management of the freshwater reservoir requires to monitor frequently (weekly or better) the spatial distribution of SWE and snowpack wetness. Both microwave radar and radiometer systems have long been considered as relevant remote sensing tools in retrieving globally snow physical parameters of interest thanks to their all-weather operation capability. However, their <span class="hlt">observations</span> are sensitive to the presence of tree canopies, which in turns impacts SWE estimation. To address this long-lasting challenge, we parked a truck-mounted microwave radiometer system for an entire winter in a rare area where it exists different tree types in the different cardinal directions. We used dual-polarization microwave radiometers at three different frequencies (1.4, 19, and 37 GHz), mounted on a boom truck to <span class="hlt">observe</span> continuously the snowpack surrounding the truck. <span class="hlt">Observations</span> were recorded at different incidence angles. These measurements have been collected in Grand Mesa National Forest, Colorado as part of the NASA SnowEx 2016-17. In this presentation, the effect of Engelmann Spruce and Aspen trees on the measured <span class="hlt">brightness</span> temperature of snow is discussed. It is shown that Engelmann Spruce trees increases the <span class="hlt">brightness</span> temperature of the snowpack more than Aspen trees do. Moreover, the elevation angular dependence of the measured <span class="hlt">brightness</span> temperatures of snowpack with and without tree canopies is investigated in the context of SWE retrievals. A time-lapse camera was monitoring a snow post installed in the sensors' field of view to characterize the <span class="hlt">brightness</span> temperature change as snow depth evolved. Also, our study takes advantage of the snowpit measurements that were collected near the radiometers' field of view.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97o5403Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97o5403Y"><span>Electromagnetically induced transparency control in terahertz metasurfaces based on <span class="hlt">bright-bright</span> mode coupling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yahiaoui, R.; Burrow, J. A.; Mekonen, S. M.; Sarangan, A.; Mathews, J.; Agha, I.; Searles, T. A.</p> <p>2018-04-01</p> <p>We demonstrate a classical analog of electromagnetically induced transparency (EIT) in a highly flexible planar terahertz metamaterial (MM) comprised of three-gap split-ring resonators. The keys to achieve EIT in this system are the frequency detuning and hybridization processes between two <span class="hlt">bright</span> modes coexisting in the same unit cell as opposed to <span class="hlt">bright</span>-dark modes. We present experimental verification of two <span class="hlt">bright</span> modes coupling for a terahertz EIT-MM in the context of numerical results and theoretical analysis based on a coupled Lorentz oscillator model. In addition, a hybrid variation of the EIT-MM is proposed and implemented numerically to dynamically tune the EIT window by incorporating photosensitive silicon pads in the split gap region of the resonators. As a result, this hybrid MM enables the active optical control of a transition from the on state (EIT mode) to the off state (dipole mode).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19840003488','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19840003488"><span>Passive microwave sensing of soil moisture content: Soil bulk density and <span class="hlt">surface</span> roughness</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wang, J. R.</p> <p>1982-01-01</p> <p>Microwave radiometric measurements over bare fields of different <span class="hlt">surface</span> roughnesses were made at the frequencies of 1.4 GHz, 5 GHz, and 10.7 GHz to study the frequency dependence as well as the possible time variation of <span class="hlt">surface</span> roughness. The presence of <span class="hlt">surface</span> roughness was found to increase the <span class="hlt">brightness</span> temperature of soils and reduce the slope of regression between <span class="hlt">brightness</span> temperature and soil moisture content. The frequency dependence of the <span class="hlt">surface</span> roughness effect was relatively weak when compared with that of the vegetation effect. Radiometric time series <span class="hlt">observation</span> over a given field indicated that field <span class="hlt">surface</span> roughness might gradually diminish with time, especially after a rainfall or irrigation. This time variation of <span class="hlt">surface</span> roughness served to enhance the uncertainty in remote soil moisture estimate by microwave radiometry. Three years of radiometric measurements over a test site revealed a possible inconsistency in the soil bulk density determination, which turned out to be an important factor in the interpretation of radiometric data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015Icar..251..211B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015Icar..251..211B"><span>Transient <span class="hlt">bright</span> "halos" on the South Polar Residual Cap of Mars: Implications for mass-balance</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Becerra, Patricio; Byrne, Shane; Brown, Adrian J.</p> <p>2015-05-01</p> <p>Spacecraft imaging of Mars' south polar region during mid-southern summer of Mars year 28 (2007) <span class="hlt">observed</span> <span class="hlt">bright</span> halo-like features surrounding many of the pits, scarps and slopes of the heavily eroded carbon dioxide ice of the South Polar Residual Cap (SPRC). These features had not been <span class="hlt">observed</span> before, and have not been <span class="hlt">observed</span> since. We report on the results of an <span class="hlt">observational</span> study of these halos, and spectral modeling of the SPRC <span class="hlt">surface</span> at the time of their appearance. Image analysis was performed using data from MRO's Context Camera (CTX), and High Resolution Imaging Science Experiment (HiRISE), as well as images from Mars Global Surveyor's (MGS) Mars Orbiter Camera (MOC). Data from MRO's Compact Reconnaissance Imaging Spectrometer for Mars (CRISM) were used for the spectral analysis of the SPRC ice at the time of the halos. These data were compared with a Hapke reflectance model of the <span class="hlt">surface</span> to constrain their formation mechanism. We find that the unique appearance of the halos is intimately linked to a near-perihelion global dust storm that occurred shortly before they were <span class="hlt">observed</span>. The combination of vigorous summertime sublimation of carbon dioxide ice from sloped <span class="hlt">surfaces</span> on the SPRC and simultaneous settling of dust from the global storm, resulted in a sublimation wind that deflected settling dust particles away from the edges of these slopes, keeping these areas relatively free of dust compared to the rest of the cap. The fact that the halos were not exhumed in subsequent years indicates a positive mass-balance for flat portions of the SPRC in those years. A net accumulation mass-balance on flat <span class="hlt">surfaces</span> of the SPRC is required to preserve the cap, as it is constantly being eroded by the expansion of the pits and scarps that populate its <span class="hlt">surface</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170007932','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170007932"><span>SMAP Level 4 <span class="hlt">Surface</span> and Root Zone Soil Moisture</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Reichle, R.; De Lannoy, G.; Liu, Q.; Ardizzone, J.; Kimball, J.; Koster, R.</p> <p>2017-01-01</p> <p>The SMAP Level 4 soil moisture (L4_SM) product provides global estimates of <span class="hlt">surface</span> and root zone soil moisture, along with other land <span class="hlt">surface</span> variables and their error estimates. These estimates are obtained through assimilation of SMAP <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> into the Goddard Earth <span class="hlt">Observing</span> System (GEOS-5) land <span class="hlt">surface</span> model. The L4_SM product is provided at 9 km spatial and 3-hourly temporal resolution and with about 2.5 day latency. The soil moisture and temperature estimates in the L4_SM product are validated against in situ <span class="hlt">observations</span>. The L4_SM product meets the required target uncertainty of 0.04 m(exp. 3)m(exp. -3), measured in terms of unbiased root-mean-square-error, for both <span class="hlt">surface</span> and root zone soil moisture.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006APS..MARU12008M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006APS..MARU12008M"><span><span class="hlt">Observation</span> of <span class="hlt">surface</span> layering in a nonmetallic liquid</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mo, Haiding; Evmenenko, Guennadi; Kewalramani, Sumit; Kim, Kyungil; Dutta, Pulak; Ehrlich, Steven</p> <p>2006-03-01</p> <p>Non-monotonic density profiles (layers) have previously been <span class="hlt">observed</span> at the free <span class="hlt">surfaces</span> of many metallic liquids, but not in isotropic dielectric liquids. Whether the presence of an electron gas is necessary for <span class="hlt">surface</span> layering has been the subject of debate. Until recently, MD simulations have suggested that layering at free liquid interface may be a generic phenomenon and is not limited to the metallic liquids^1. The theories predict that if normal liquids can be cooled down to temperatures low enough, layering structure should be <span class="hlt">observed</span> experimentally. However, this is difficult for most molecular liquids because these liquids freeze well above the temperature necessary for <span class="hlt">observing</span> the layering structure. By studying the <span class="hlt">surface</span> structure of liquid TEHOS (tetrakis(2-ethylhexoxy)silane), which combines relatively low freezing point and high boiling point compared to that of most molecular liquids, we have <span class="hlt">observed</span> the evidence of layering at the free interface of liquid TEHOS using x-ray reflectivity. When cooled to T/Tc 0.25 (well above the bulk freezing point, Tc is the critical temperature of TEHOS), the <span class="hlt">surface</span> roughness drops sharply and density oscillations appear near the <span class="hlt">surface</span>. Lateral ordering of the <span class="hlt">surface</span> layers is liquid-like, just as at liquid metal <span class="hlt">surfaces</span>. 1. E. Chac'on and P. Tarazona, Phys. Rev. Lett. 91 166103-1 (2003)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.884a2045R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.884a2045R"><span>Tolerance of image enhancement <span class="hlt">brightness</span> and contrast in lateral cephalometric digital radiography for Steiner analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rianti, R. A.; Priaminiarti, M.; Syahraini, S. I.</p> <p>2017-08-01</p> <p>Image enhancement <span class="hlt">brightness</span> and contrast can be adjusted on lateral cephalometric digital radiographs to improve image quality and anatomic landmarks for measurement by Steiner analysis. To determine the limit value for adjustments of image enhancement <span class="hlt">brightness</span> and contrast in lateral cephalometric digital radiography for Steiner analysis. Image enhancement <span class="hlt">brightness</span> and contrast were adjusted on 100 lateral cephalometric radiography in 10-point increments (-30, -20, -10, 0, +10, +20, +30). Steiner analysis measurements were then performed by two <span class="hlt">observers</span>. Reliabilities were tested by the Interclass Correlation Coefficient (ICC) and significance tested by ANOVA or the Kruskal Wallis test. No significant differences were detected in lateral cephalometric analysis measurements following adjustment of the image enhancement <span class="hlt">brightness</span> and contrast. The limit value of adjustments of the image enhancement <span class="hlt">brightness</span> and contrast associated with incremental 10-point changes (-30, -20, -10, 0, +10, +20, +30) does not affect the results of Steiner analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28842470','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28842470"><span>IL15 Infusion of Cancer Patients Expands the Subpopulation of Cytotoxic CD56<span class="hlt">bright</span> NK Cells and Increases NK-Cell Cytokine Release Capabilities.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dubois, Sigrid; Conlon, Kevin C; Müller, Jürgen R; Hsu-Albert, Jennifer; Beltran, Nancy; Bryant, Bonita R; Waldmann, Thomas A</p> <p>2017-10-01</p> <p>The cytokine IL15 is required for survival and activation of natural killer (NK) cells as well as expansion of NK-cell populations. Here, we compare the effects of continuous IL15 infusions on NK-cell subpopulations in cancer patients. Infusions affected the CD56 <span class="hlt">bright</span> NK-cell subpopulation in that the expansion rates exceeded those of CD56 dim NK-cell populations with a 350-fold increase in their total cell numbers compared with 20-fold expansion for the CD56 dim subset. CD56 <span class="hlt">bright</span> NK cells responded with increased cytokine release to various stimuli, as expected given their immunoregulatory functions. Moreover, CD56 <span class="hlt">bright</span> NK cells gained the ability to kill various target cells at levels that are typical for CD56 dim NK cells. Some increased cytotoxic activities were also <span class="hlt">observed</span> for CD56 dim NK cells. IL15 infusions induced expression changes on the <span class="hlt">surface</span> of both NK-cell subsets, resulting in a previously undescribed and similar phenotype. These data suggest that IL15 infusions expand and arm CD56 <span class="hlt">bright</span> NK cells that alone or in combination with tumor-targeting antibodies may be useful in the treatment of cancer. Cancer Immunol Res; 5(10); 929-38. ©2017 AACR . ©2017 American Association for Cancer Research.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA10140&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA10140&hterms=bright+hour&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dbright%2Bhour"><span>Active Processes: <span class="hlt">Bright</span> Streaks and Dark Fans</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2007-01-01</p> <p><p/> [figure removed for brevity, see original site] [figure removed for brevity, see original site] Figure 1Figure 2 <p/> In a region of the south pole known informally as 'Ithaca' numerous fans of dark frost form every spring. HiRISE collected a time lapse series of these images, starting at L<sub>s</sub> = 185 and culminating at L<sub>s</sub> = 294. 'L<sub>s</sub>' is the way we measure time on Mars: at L<sub>s</sub> = 180 the sun passes the equator on its way south; at L<sub>s</sub> = 270 it reaches its maximum subsolar latitude and summer begins. <p/> In the earliest image (figure 1) fans are dark, but small narrow <span class="hlt">bright</span> streaks can be detected. In the next image (figure 2), acquired at L<sub>s</sub> = 187, just 106 hours later, dramatic differences are apparent. The dark fans are larger and the <span class="hlt">bright</span> fans are more pronounced and easily detectable. The third image in the sequence shows no <span class="hlt">bright</span> fans at all. <p/> We believe that the <span class="hlt">bright</span> streaks are fine frost condensed from the gas exiting the vent. The conditions must be just right for the <span class="hlt">bright</span> frost to condense. <p/> <span class="hlt">Observation</span> Geometry Image PSP_002622_0945 was taken by the High Resolution Imaging Science Experiment (HiRISE) camera onboard the Mars Reconnaissance Orbiter spacecraft on 16-Feb-2007. The complete image is centered at -85.2 degrees latitude, 181.5 degrees East longitude. The range to the target site was 246.9 km (154.3 miles). At this distance the image scale is 49.4 cm/pixel (with 2 x 2 binning) so objects 148 cm across are resolved. The image shown here has been map-projected to 50 cm/pixel . The image was taken at a local Mars time of 05:46 PM and the scene is illuminated from the west with a solar incidence angle of 88 degrees, thus the sun was about 2 degrees above the horizon. At a solar longitude of 185.1 degrees, the season on Mars is Northern Autumn.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=S81-40833&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dphoto%2Bimage','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=S81-40833&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dphoto%2Bimage"><span>Earth <span class="hlt">observation</span> photo taken by JPL with the Shuttle Imaging Radar-A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1981-01-01</p> <p>Earth <span class="hlt">observation</span> photo taken by the Jet Propulsion Laboratory (JPL) with the Shuttle Imaging Radar-A (SIR-A). This image shows a 50 by 120 kilometer (30 by 75 mile) area of the Mediterranean Sea and the eastern coast of Central Sardinia (left). The city of Arbatose is seen as a <span class="hlt">bright</span> area along the coast in the lower part of the image, and the star-like spot off the coast is a ship's reflection. The Gulf of Orsei is near the top of the image. <span class="hlt">Bright</span>, mottled features in the sea (right) represent <span class="hlt">surface</span> choppiness.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ACP....12.4143D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ACP....12.4143D"><span>Lidar and radar measurements of the melting layer: <span class="hlt">observations</span> of dark and <span class="hlt">bright</span> band phenomena</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Di Girolamo, P.; Summa, D.; Cacciani, M.; Norton, E. G.; Peters, G.; Dufournet, Y.</p> <p>2012-05-01</p> <p>Multi-wavelength lidar measurements in the melting layer revealing the presence of dark and <span class="hlt">bright</span> bands have been performed by the University of BASILicata Raman lidar system (BASIL) during a stratiform rain event. Simultaneously radar measurements have been also performed from the same site by the University of Hamburg cloud radar MIRA 36 (35.5 GHz), the University of Hamburg dual-polarization micro rain radar (24.15 GHz) and the University of Manchester UHF wind profiler (1.29 GHz). Measurements from BASIL and the radars are illustrated and discussed in this paper for a specific case study on 23 July 2007 during the Convective and Orographically-induced Precipitation Study (COPS). Simulations of the lidar dark and <span class="hlt">bright</span> band based on the application of concentric/eccentric sphere Lorentz-Mie codes and a melting layer model are also provided. Lidar and radar measurements and model results are also compared with measurements from a disdrometer on ground and a two-dimensional cloud (2DC) probe on-board the ATR42 SAFIRE. Measurements and model results are found to confirm and support the conceptual microphysical/scattering model elaborated by Sassen et al. (2005).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ATel11586....1S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ATel11586....1S"><span>Optical confirmation of Gaia18ayp <span class="hlt">brightness</span> increase</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Spano, M.; Blanco-Cuaresma, S.; Roelens, M.; Mowlavi, N.; Eyer, L.</p> <p>2018-04-01</p> <p>We report confirmation of Gaia_Science_Alerts, <span class="hlt">brightness</span> increase of the QSO [VV2006] J233633.0-411547, Gaia18ayp . Images were obtained through modified Gunn R and V band filter of the ECAM instrument installed on the Swiss 1.2m Euler telescope at La Silla, on 2018 April 21- 22. Magnitudes according to the MJD of <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017A%26A...606A..46G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...606A..46G"><span>Magnetic topological analysis of coronal <span class="hlt">bright</span> points</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Galsgaard, K.; Madjarska, M. S.; Moreno-Insertis, F.; Huang, Z.; Wiegelmann, T.</p> <p>2017-10-01</p> <p>Context. We report on the first of a series of studies on coronal <span class="hlt">bright</span> points which investigate the physical mechanism that generates these phenomena. Aims: The aim of this paper is to understand the magnetic-field structure that hosts the <span class="hlt">bright</span> points. Methods: We use longitudinal magnetograms taken by the Solar Optical Telescope with the Narrowband Filter Imager. For a single case, magnetograms from the Helioseismic and Magnetic Imager were added to the analysis. The longitudinal magnetic field component is used to derive the potential magnetic fields of the large regions around the <span class="hlt">bright</span> points. A magneto-static field extrapolation method is tested to verify the accuracy of the potential field modelling. The three dimensional magnetic fields are investigated for the presence of magnetic null points and their influence on the local magnetic domain. Results: In nine out of ten cases the <span class="hlt">bright</span> point resides in areas where the coronal magnetic field contains an opposite polarity intrusion defining a magnetic null point above it. We find that X-ray <span class="hlt">bright</span> points reside, in these nine cases, in a limited part of the projected fan-dome area, either fully inside the dome or expanding over a limited area below which typically a dominant flux concentration resides. The tenth <span class="hlt">bright</span> point is located in a bipolar loop system without an overlying null point. Conclusions: All <span class="hlt">bright</span> points in coronal holes and two out of three <span class="hlt">bright</span> points in quiet Sun regions are seen to reside in regions containing a magnetic null point. An as yet unidentified process(es) generates the brigh points in specific regions of the fan-dome structure. The movies are available at http://www.aanda.org</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA148882','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA148882"><span>Individual Differences in Chromatic <span class="hlt">Brightness</span> Matching.</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1984-10-03</p> <p>very unreliable..." More recently, Boynton (15) has written, "Consider... a 555-nm green light on one side of a bi-partite field with a 4 6 5-nm blue...field immediately adjacent to it... We ask an <span class="hlt">observer</span> to adjust the intensity of the blue field until it looks ’equally <span class="hlt">bright</span>’ as the green one. This...clearly being blue, blue- green , green , yellow- green , yellow, and red. Their spectral transmittance curves are shown in Fig. 2. All were broad-band filters</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120015232','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120015232"><span>Inter-Calibration of EIS, XRT and AIA using Active Region and <span class="hlt">Bright</span> Point Data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mulu-Moore, Fana M.; Winebarger, Amy R.; Winebarger, Amy R.; Farid, Samaiyah I.</p> <p>2012-01-01</p> <p>Certain limitations in our solar instruments have created the need to use several instruments together for long term and/or large field of view studies. We will, therefore, present an intercalibration study of the EIS, XRT and AIA instruments using active region and <span class="hlt">bright</span> point data. We will use the DEMs calculated from EIS <span class="hlt">bright</span> point <span class="hlt">observations</span> to determine the expected AIA and XRT intensities. We will them compare to the <span class="hlt">observed</span> intensities and calculate a correction factor. We will consider data taken over a year to see if there is a time dependence to the correction factor. We will then determine if the correction factors are valid for active region <span class="hlt">observations</span>.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.471.2882W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.471.2882W"><span>Beyond the Kepler/K2 <span class="hlt">bright</span> limit: variability in the seven brightest members of the Pleiades</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>White, T. R.; Pope, B. J. S.; Antoci, V.; Pápics, P. I.; Aerts, C.; Gies, D. R.; Gordon, K.; Huber, D.; Schaefer, G. H.; Aigrain, S.; Albrecht, S.; Barclay, T.; Barentsen, G.; Beck, P. G.; Bedding, T. R.; Fredslund Andersen, M.; Grundahl, F.; Howell, S. B.; Ireland, M. J.; Murphy, S. J.; Nielsen, M. B.; Silva Aguirre, V.; Tuthill, P. G.</p> <p>2017-11-01</p> <p>The most powerful tests of stellar models come from the brightest stars in the sky, for which complementary techniques, such as astrometry, asteroseismology, spectroscopy and interferometry, can be combined. The K2 mission is providing a unique opportunity to obtain high-precision photometric time series for <span class="hlt">bright</span> stars along the ecliptic. However, <span class="hlt">bright</span> targets require a large number of pixels to capture the entirety of the stellar flux, and CCD saturation, as well as restrictions on data storage and bandwidth, limit the number and <span class="hlt">brightness</span> of stars that can be <span class="hlt">observed</span>. To overcome this, we have developed a new photometric technique, which we call halo photometry, to <span class="hlt">observe</span> very <span class="hlt">bright</span> stars using a limited number of pixels. Halo photometry is simple, fast and does not require extensive pixel allocation, and will allow us to use K2 and other photometric missions, such as TESS, to <span class="hlt">observe</span> very <span class="hlt">bright</span> stars for asteroseismology and to search for transiting exoplanets. We apply this method to the seven brightest stars in the Pleiades open cluster. Each star exhibits variability; six of the stars show what are most likely slowly pulsating B-star pulsations, with amplitudes ranging from 20 to 2000 ppm. For the star Maia, we demonstrate the utility of combining K2 photometry with spectroscopy and interferometry to show that it is not a `Maia variable', and to establish that its variability is caused by rotational modulation of a large chemical spot on a 10 d time-scale.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AAS...21330108M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AAS...21330108M"><span>Network based sky <span class="hlt">Brightness</span> Monitor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McKenna, Dan; Pulvermacher, R.; Davis, D. R.</p> <p>2009-01-01</p> <p>We have developed and are currently testing an autonomous 2 channel photometer designed to measure the night sky <span class="hlt">brightness</span> in the visual wavelengths over a multi-year campaign. The photometer uses a robust silicon sensor filtered with Hoya CM500 glass. The Sky <span class="hlt">brightness</span> is measured every minute at two elevation angles typically zenith and 20 degrees to monitor <span class="hlt">brightness</span> and transparency. The Sky <span class="hlt">Brightness</span> monitor consists of two units, the remote photometer and a network interface. Currently these devices use 2.4 Ghz transceivers with a free space range of 100 meters. The remote unit is battery powered with day time recharging using a solar panel. Data received by the network interface transmits data via standard POP Email protocol. A second version is under development for radio sensitive areas using an optical fiber for data transmission. We will present the current comparison with the National Park Service sky monitoring camera. We will also discuss the calibration methods used for standardization and temperature compensation. This system is expected to be deployed in the next year and be operated by the International Dark Sky Association SKYMONITOR project.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24076544','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24076544"><span><span class="hlt">Bright</span> light induces choroidal thickening in chickens.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lan, Weizhong; Feldkaemper, Marita; Schaeffel, Frank</p> <p>2013-11-01</p> <p><span class="hlt">Bright</span> light is a potent inhibitor of myopia development in animal models. Because development of refractive errors has been linked to changes in choroidal thickness, we have studied in chickens whether <span class="hlt">bright</span> light may exert its effects on myopia also through changes in choroidal thickness. Three-day-old chickens were exposed to "<span class="hlt">bright</span> light" (15,000 lux; n = 14) from 10 AM to 4 PM but kept under "normal light" (500 lux) during the remaining time of the light phase for 5 days (total duration of light phase 8 AM to 6 PM). A control group (n = 14) was kept under normal light during the entire light phase. Choroidal thickness was measured in alert, hand-held animals with optical coherence tomography at 10 AM, 4 PM, and 8 PM every day. Complete data sets were available for 12 chicks in <span class="hlt">bright</span> light group and nine in normal light group. The striking inter-individual variability in choroidal thickness (coefficient of variance: 23%) made it necessary to normalize changes to the individual baseline thickness of the choroid. During the 6 hours of exposure to <span class="hlt">bright</span> light, choroidal thickness decreased by -5.2 ± 4.0% (mean ± SEM). By contrast, in the group kept under normal light, choroidal thickness increased by +15.4 ± 4.7% (difference between both groups p = 0.003). After an additional 4 hours, choroidal thickness increased also in the "<span class="hlt">bright</span> light group" by +17.8 ± 3.5%, while there was little further change (+0.6 ± 4.0%) in the "normal light group" (difference p = 0.004). Finally, the choroid was thicker in the "<span class="hlt">bright</span> light group" (+7.6 ± 26.0%) than in the "normal light group" (day 5: -18.6 ± 26.9%; difference p = 0.036). <span class="hlt">Bright</span> light stimulates choroidal thickening in chickens, although the response is smaller than with experimentally imposed myopic defocus, and it occurs with some time delay. It nevertheless suggests that choroidal thickening is also involved in myopia inhibition by <span class="hlt">bright</span> light.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20818480','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20818480"><span>The effect of <span class="hlt">bright</span> light on sleepiness among rapid-rotating 12-hour shift workers.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sadeghniiat-Haghighi, Khosro; Yazdi, Zohreh; Jahanihashemi, Hassan; Aminian, Omid</p> <p>2011-01-01</p> <p>About 20% of workers in industrialized countries are shift workers and more than half of them work on night or rotating shifts. Most night workers complain of sleepiness due to lack of adjustment of the circadian rhythm. In simulated night-work experiments, scheduled exposure to <span class="hlt">bright</span> light has been shown to reduce these complaints. Our study assessed the effects of <span class="hlt">bright</span> light exposure on sleepiness during night work in an industrial setting. In a cross-over design, 94 workers at a ceramic factory were exposed to either <span class="hlt">bright</span> (2500 lux) or normal light (300 lux) during breaks on night shifts. We initiated 20-minute breaks between 24.00 and 02.00 hours. Sleepiness ratings were determined using the Stanford Sleepiness Scale at 22.00, 24.00, 02.00 and 04.00 hours. Under normal light conditions, sleepiness peaked at 02:00 hours. A significant reduction (22% compared to normal light conditions) in sleepiness was <span class="hlt">observed</span> after workers were exposed to <span class="hlt">bright</span> light. Exposure to <span class="hlt">bright</span> light may be effective in reducing sleepiness among night workers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AAS...22314832Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AAS...22314832Y"><span>Winter sky <span class="hlt">brightness</span> & cloud cover over Dome A</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Yi; Moore, A. M.; Fu, J.; Ashley, M.; Cui, X.; Feng, L.; Gong, X.; Hu, Z.; Laurence, J.; LuongVan, D.; Riddle, R. L.; Shang, Z.; Sims, G.; Storey, J.; Tothill, N.; Travouillon, T.; Wang, L.; Yang, H.; Yang, J.; Zhou, X.; Zhu, Z.; Burton, M. G.</p> <p>2014-01-01</p> <p>At the summit of the Antarctic plateau, Dome A offers an intriguing location for future large scale optical astronomical Observatories. The Gattini DomeA project was created to measure the optical sky <span class="hlt">brightness</span> and large area cloud cover of the winter-time sky above this high altitude Antarctic site. The wide field camera and multi-filter system was installed on the PLATO instrument module as part of the Chinese-led traverse to Dome A in January 2008. This automated wide field camera consists of an Apogee U4000 interline CCD coupled to a Nikon fish-eye lens enclosed in a heated container with glass window. The system contains a filter mechanism providing a suite of standard astronomical photometric filters (Bessell B, V, R), however, the absence of tracking systems, together with the ultra large field of view 85 degrees) and strong distortion have driven us to seek a unique way to build our data reduction pipeline. We present here the first measurements of sky <span class="hlt">brightness</span> in the photometric B, V, and R band, cloud cover statistics measured during the 2009 winter season and an estimate of the transparency. In addition, we present example light curves for <span class="hlt">bright</span> targets to emphasize the unprecedented <span class="hlt">observational</span> window function available from this ground-based location. A ~0.2 magnitude agreement of our simultaneous test at Palomar Observatory with NSBM(National Sky <span class="hlt">Brightness</span> Monitor), as well as an 0.04 magnitude photometric accuracy for typical 6th magnitude stars limited by the instrument design, indicating we obtained reasonable results based on our ~7mm effective aperture fish-eye lens.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013RScI...84h3703N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013RScI...84h3703N"><span>A <span class="hlt">brightness</span> exceeding simulated Langmuir limit</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nakasuji, Mamoru</p> <p>2013-08-01</p> <p>When an excitation of the first lens determines a beam is parallel beam, a <span class="hlt">brightness</span> that is 100 times higher than Langmuir limit is measured experimentally, where Langmuir limits are estimated using a simulated axial cathode current density which is simulated based on a measured emission current. The measured <span class="hlt">brightness</span> is comparable to Langmuir limit, when the lens excitation is such that an image position is slightly shorter than a lens position. Previously measured values of <span class="hlt">brightness</span> for cathode apical radii of curvature 20, 60, 120, 240, and 480 μm were 8.7, 5.3, 3.3, 2.4, and 3.9 times higher than their corresponding Langmuir limits, respectively, in this experiment, the lens excitation was such that the lens and the image positions were 180 mm and 400 mm, respectively. From these measured <span class="hlt">brightness</span> for three different lens excitation conditions, it is concluded that the <span class="hlt">brightness</span> depends on the first lens excitation. For the electron gun operated in a space charge limited condition, some of the electrons emitted from the cathode are returned to the cathode without having crossed a virtual cathode. Therefore, method that assumes a Langmuir limit defining method using a Maxwellian distribution of electron velocities may need to be revised. For the condition in which the values of the exceeding the Langmuir limit are measured, the simulated trajectories of electrons that are emitted from the cathode do not cross the optical axis at the crossover, thus the law of sines may not be valid for high <span class="hlt">brightness</span> electron beam systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=plasmon&pg=2&id=EJ231003','ERIC'); return false;" href="https://eric.ed.gov/?q=plasmon&pg=2&id=EJ231003"><span>Plasmon <span class="hlt">Surface</span> Polariton Dispersion by Direct Optical <span class="hlt">Observation</span>.</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Swalen, J. D.; And Others</p> <p>1980-01-01</p> <p>Describes several simple experiments that can be used to <span class="hlt">observe</span> directly the dispersion curve of plasmon <span class="hlt">surface</span> polaritons (PSP) on flat metal <span class="hlt">surfaces</span>. A method is described of <span class="hlt">observing</span> the increonental change in the wave vector of the PSP due to coatings that differ in thickness by a few nanometers. (Author/CS)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.478....2C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.478....2C"><span>Investigating a population of infrared-<span class="hlt">bright</span> gamma-ray burst host galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chrimes, Ashley A.; Stanway, Elizabeth R.; Levan, Andrew J.; Davies, Luke J. M.; Angus, Charlotte R.; Greis, Stephanie M. L.</p> <p>2018-07-01</p> <p>We identify and explore the properties of an infrared-<span class="hlt">bright</span> gamma-ray burst (GRB) host population. Candidate hosts are selected by coincidence with sources in WISE, with matching to random coordinates and a false alarm probability analysis showing that the contamination fraction is ˜0.5. This methodology has already identified the host galaxy of GRB 080517. We combine survey photometry from Pan-STARRS, SDSS, APASS, 2MASS, GALEX, and WISE with our own WHT/ACAM and VLT/X-shooter <span class="hlt">observations</span> to classify the candidates and identify interlopers. Galaxy SED fitting is performed using MAGPHYS, in addition to stellar template fitting, yielding 13 possible IR-<span class="hlt">bright</span> hosts. A further seven candidates are identified from the previously published work. We report a candidate host for GRB 061002, previously unidentified as such. The remainder of the galaxies have already been noted as potential hosts. Comparing the IR-<span class="hlt">bright</span> population properties including redshift z, stellar mass M⋆, star formation rate SFR, and V-band attenuation AV to GRB host catalogues in the literature, we find that the infrared-<span class="hlt">bright</span> population is biased towards low z, high M⋆, and high AV. This naturally arises from their initial selection - local and dusty galaxies are more likely to have the required IR flux to be detected in WISE. We conclude that while IR-<span class="hlt">bright</span> GRB hosts are not a physically distinct class, they are useful for constraining existing GRB host populations, particularly for long GRBs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.tmp..989C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.tmp..989C"><span>Investigating a population of infrared-<span class="hlt">bright</span> gamma-ray burst host galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chrimes, Ashley A.; Stanway, Elizabeth R.; Levan, Andrew J.; Davies, Luke J. M.; Angus, Charlotte R.; Greis, Stephanie M. L.</p> <p>2018-04-01</p> <p>We identify and explore the properties of an infrared-<span class="hlt">bright</span> gamma-ray burst (GRB) host population. Candidate hosts are selected by coincidence with sources in WISE, with matching to random coordinates and a false alarm probability analysis showing that the contamination fraction is ˜ 0.5. This methodology has already identified the host galaxy of GRB 080517. We combine survey photometry from Pan-STARRS, SDSS, APASS, 2MASS, GALEX and WISE with our own WHT/ACAM and VLT/X-shooter <span class="hlt">observations</span> to classify the candidates and identify interlopers. Galaxy SED fitting is performed using MAGPHYS, in addition to stellar template fitting, yielding 13 possible IR-<span class="hlt">bright</span> hosts. A further 7 candidates are identified from previously published work. We report a candidate host for GRB 061002, previously unidentified as such. The remainder of the galaxies have already been noted as potential hosts. Comparing the IR-<span class="hlt">bright</span> population properties including redshift z, stellar mass M⋆, star formation rate SFR and V-band attenuation AV to GRB host catalogues in the literature, we find that the infrared-<span class="hlt">bright</span> population is biased toward low z, high M⋆ and high AV. This naturally arises from their initial selection - local and dusty galaxies are more likely to have the required IR flux to be detected in WISE. We conclude that while IR-<span class="hlt">bright</span> GRB hosts are not a physically distinct class, they are useful for constraining existing GRB host populations, particularly for long GRBs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://rosap.ntl.bts.gov/view/dot/20882','DOTNTL'); return false;" href="https://rosap.ntl.bts.gov/view/dot/20882"><span>Assessment of the broca-sulzer phenomenon via inter- and intra-modality matching procedures : studies of signal-light <span class="hlt">brightness</span>.</span></a></p> <p><a target="_blank" href="http://ntlsearch.bts.gov/tris/index.do">DOT National Transportation Integrated Search</a></p> <p></p> <p>1968-10-01</p> <p>Signal lights are presented to an <span class="hlt">observer</span> as flashes with finite duration; thus, the effect of flash duration on the apparent <span class="hlt">brightness</span> of the signal is important. The relation of effective signal <span class="hlt">brightness</span> to flash duration and luminance finds ex...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11837952','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11837952"><span><span class="hlt">Bright</span>-light mask treatment of delayed sleep phase syndrome.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cole, Roger J; Smith, Julian S; Alcalá, Yvonne C; Elliott, Jeffrey A; Kripke, Daniel F</p> <p>2002-02-01</p> <p>We treated delayed sleep phase syndrome (DSPS) with an illuminated mask that provides light through closed eyelids during sleep. Volunteers received either <span class="hlt">bright</span> white light (2,700 lux, n = 28) or dim red light placebo (0.1 lux, n = 26) for 26 days at home. Mask lights were turned on (< 0.01 lux) 4 h before arising, ramped up for 1 h, and remained on at full <span class="hlt">brightness</span> until arising. Volunteers also attempted to systematically advance sleep time, avoid naps, and avoid evening <span class="hlt">bright</span> light. The light mask was well tolerated and produced little sleep disturbance. The acrophase of urinary 6-sulphatoxymelatonin (6-SMT) excretion advanced significantly from baseline in the <span class="hlt">bright</span> group (p < 0.0006) and not in the dim group, but final phases were not significantly earlier in the <span class="hlt">bright</span> group (ANCOVA ns). <span class="hlt">Bright</span> treatment did produce significantly earlier phases, however, among volunteers whose baseline 6-SMT acrophase was later than the median of 0602 h (<span class="hlt">bright</span> shift: 0732-0554 h, p < 0.0009; dim shift: 0746-0717 h, ns; ANCOVA p = 0.03). In this subgroup, sleep onset advanced significantly only with <span class="hlt">bright</span> but not dim treatment (sleep onset shift: <span class="hlt">bright</span> 0306-0145 h, p < 0.0002; dim 0229-0211 h, ns; ANCOVA p < .05). Despite equal expectations at baseline, participants rated <span class="hlt">bright</span> treatment as more effective than dim treatment (p < 0.04). We conclude that <span class="hlt">bright</span>-light mask treatment advances circadian phase and provides clinical benefit in DSPS individuals whose initial circadian delay is relatively severe.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2000A%26A...353.1083B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000A%26A...353.1083B"><span>EUV <span class="hlt">brightness</span> variations in the quiet Sun</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brković, A.; Rüedi, I.; Solanki, S. K.; Fludra, A.; Harrison, R. A.; Huber, M. C. E.; Stenflo, J. O.; Stucki, K.</p> <p>2000-01-01</p> <p>The Coronal Diagnostic Spectrometer (CDS) onboard the SOHO satellite has been used to obtain movies of quiet Sun regions at disc centre. These movies were used to study <span class="hlt">brightness</span> variations of solar features at three different temperatures sampled simultaneously in the chromospheric He I 584.3 Ä (2 * 104 K), the transition region O V 629.7 Ä (2.5 * 105 K) and coronal Mg IX 368.1 Ä (106 K) lines. In all parts of the quiet Sun, from darkest intranetwork to brightest network, we find significant variability in the He I and O V line, while the variability in the Mg IX line is more marginal. The relative variability, defined by rms of intensity normalised to the local intensity, is independent of <span class="hlt">brightness</span> and strongest in the transition region line. Thus the relative variability is the same in the network and the intranetwork. More than half of the points on the solar <span class="hlt">surface</span> show a relative variability, determined over a period of 4 hours, greater than 15.5% for the O V line, but only 5% of the points exhibit a variability above 25%. Most of the variability appears to take place on time-scales between 5 and 80 minutes for the He I and O V lines. Clear signs of ``high variability'' events are found. For these events the variability as a function of time seen in the different lines shows a good correlation. The correlation is higher for more variable events. These events coincide with the (time averaged) brightest points on the solar <span class="hlt">surface</span>, i.e. they occur in the network. The spatial positions of the most variable points are identical in all the lines.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20110007159&hterms=jupiter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Djupiter','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20110007159&hterms=jupiter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Djupiter"><span>On the Long-Term Variability of Jupiter's Winds and <span class="hlt">Brightness</span> as <span class="hlt">Observed</span> from Hubble</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Simon-Miller, Amy A.; Gierasch, Peter J.</p> <p>2010-01-01</p> <p>Hubble Space Telescope Wide Field Planetary Camera 2 imaging data of Jupiter were combined with wind profiles from Voyager and Cassini data to study long-term variability in Jupiter's winds and cloud <span class="hlt">brightness</span>. Searches for evidence of wind velocity periodicity yielded a few latitudes with potential variability; the most significant periods were found nearly symmetrically about the equator at 0 deg., 10-12 deg. N, and 14-18 deg. S planetographic latitude. The low to mid-latitude signals have components consistent with the measured stratospheric temperature Quasi-Quadrennial Oscillation (QQO) period of-5 years, while the equatorial signal is approximately seasonal and could be tied to mesoscale wave formation, robustness tests indicate that a constant or continuously varying periodic signal near 4.5 years would appear with high significance in the data periodograms as long as uncertainties or noise in the data are not of greater magnitude. However, the lack of a consistent signal over many latitudes makes it difficult to interpret as a QQO-related change. In addition, further analyses of calibrated 410-nm and 953-nm <span class="hlt">brightness</span> scans found few corresponding changes in troposphere haze and cloud structure on QQO timescales. However, stratospheric haze reflectance at 255-nm did appear to vary on seasonal timescales, though the data do not have enough temporal coverage or photometric accuracy to be conclusive. Sufficient temporal coverage and spacing, as well as data quality, are critical to this type of search.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19780030814&hterms=bright+hour&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dbright%2Bhour','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19780030814&hterms=bright+hour&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dbright%2Bhour"><span><span class="hlt">Bright</span> X-ray arcs and the emergence of solar magnetic flux</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chapman, G. A.; Broussard, R. M.</p> <p>1977-01-01</p> <p>The Skylab S-056 and S-082A experiments and ground-based magnetograms have been used to study the role of <span class="hlt">bright</span> X-ray arcs and the emergence of solar magnetic flux in the McMath region 12476. The S-056 X-ray images show a system of one or sometimes two <span class="hlt">bright</span> arcs within a diffuse emitting region. The arcs seem to directly connect regions of opposite magnetic polarity in the photosphere. Magnetograms suggest the possible emergence of a magnetic flux. The width of the main arc is approximately 6 arcsec when most clearly defined, and the length is approximately 30-50 arcsec. Although the arc system is <span class="hlt">observed</span> to vary in <span class="hlt">brightness</span> over a period exceeding 24 hours, it remains fixed in orientation. The temperature of the main arc is approximately 3 x 10 to the 6th K. It is suggested that merging magnetic fields may provide the primary energy source, perhaps accompanied by resistive heating from a force-free current.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20551588','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20551588"><span>Effect of evening exposure to <span class="hlt">bright</span> or dim light after daytime <span class="hlt">bright</span> light on absorption of dietary carbohydrates the following morning.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hirota, Naoko; Sone, Yoshiaki; Tokura, Hiromi</p> <p>2010-01-01</p> <p>We had previously reported on the effect of exposure to light on the human digestive system: daytime <span class="hlt">bright</span> light exposure has a positive effect, whereas, evening <span class="hlt">bright</span> light exposure has a negative effect on the efficiency of dietary carbohydrate absorption from the evening meal. These results prompted us to examine whether the light intensity to which subjects are exposed in the evening affects the efficiency of dietary carbohydrate absorption the following morning. In this study, subjects were exposed to either 50 lux (dim light conditions) or 2,000 lux (<span class="hlt">bright</span> light conditions) in the evening for 9 h (from 15:00 to 24:00) after staying under <span class="hlt">bright</span> light in the daytime (under 2,000 lux from 07:00 to 15:00). We measured unabsorbed dietary carbohydrates using the breath-hydrogen test the morning after exposure to either <span class="hlt">bright</span> light or dim light the previous evening. Results showed that there was no significant difference between the two conditions in the amount of breath hydrogen. This indicates that evening exposure to <span class="hlt">bright</span> or dim light after <span class="hlt">bright</span> light exposure in the daytime has no varying effect on digestion or absorption of dietary carbohydrates in the following morning's breakfast.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMGC51D0442G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMGC51D0442G"><span>Inter-annual variation of the <span class="hlt">surface</span> temperature of tropical forests from SSM/I <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gao, H.; Fu, R.; Li, W.; Zhang, S.; Dickinson, R. E.</p> <p>2014-12-01</p> <p>Land <span class="hlt">surface</span> temperatures (LST) within tropical rain forests contribute to climate variation, but <span class="hlt">observational</span> data are very limited in these regions. In this study, all weather canopy sky temperatures were retrieved using the passive microwave remote sensing data from the Special Sensor Microwave/Imager (SSM/I) and the Special Sensor Microwave Imager/Sounder (SSMIS) over the Amazon and Congo rainforests. The remote sensing data used were collected from 1996 to 2012 using two separate satellites—F13 (1996-2009) and F17 (2007-2012). An inter-sensor calibration between the <span class="hlt">brightness</span> temperatures collected by the two satellites was conducted in order to ensure consistency amongst the instruments. The interannual changes of LST associated with the dry and wet anomalies were investigated in both regions. The dominant spatial and temporal patterns for inter-seasonal variations of the LST over the tropical rainforest were analyzed, and the impacts of droughts and El Niños (on LST) were also investigated. The remote sensing results suggest that the morning LST is mainly controlled by atmospheric humidity (which controls longwave radiation) whereas the late afternoon LST is controlled by solar radiation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004PhDT.........1U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004PhDT.........1U"><span><span class="hlt">Brightness</span> and magnetic evolution of solar coronal <span class="hlt">bright</span> points</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ugarte-Urra, I.</p> <p>2004-12-01</p> <p>This thesis presents a study of the <span class="hlt">brightness</span> and magnetic evolution of several Extreme ultraviolet (EUV) coronal <span class="hlt">bright</span> points (hereafter BPs). BPs are loop-like features of enhanced emission in the coronal EUV and X-ray images of the Sun, that are associated to the interaction of opposite photospheric magnetic polarities with magnetic fluxes of ≈1018 - 1019 Mx. The study was carried out using several instruments on board the Solar and Heliospheric Observatory (SOHO): the Extreme Ultraviolet Imager (EIT), the Coronal Diagnostic Spectrometer (CDS) and the Michelson Doppler Imager (MDI), supported by the high resolution imaging from the Transition Region And Coronal Explorer (TRACE). The results confirm that, down to 1'' (i.e. ~715 km) resolution, BPs are made of small loops with lengths of ~6 Mm and cross-sections of ~2 Mm. The loops are very dynamic, evolving in time scales as short as 1 - 2 minutes. This is reflected in a highly variable EUV response with fluctuations highly correlated in spectral lines at transition region temperatures (in the range 3.2x10^4 - 3.5x10^5 K), but not always at coronal temperatures. A wavelet analysis of the intensity variations reveals, for the first time, the existence of quasi-periodic oscillations with periods ranging 400 -- 1000 s, in the range of periods characteristic of the chromospheric network. The link between BPs and network <span class="hlt">bright</span> points is discussed, as well as the interpretation of the oscillations in terms of global acoustic modes of closed magnetic structures. A comparison of the magnetic flux evolution of the magnetic polarities to the EUV flux changes is also presented. Throughout their lifetime, the intrinsic EUV emission of BPs is found to be dependent on the total magnetic flux of the polarities. In short time scales, co-spatial and co-temporal TRACE and MDI images, reveal the signature of heating events that produce sudden EUV brightenings simultaneous to magnetic flux cancellations. This is interpreted in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=345521','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=345521"><span>Assessment of the SMAP Level-4 <span class="hlt">Surface</span> and Root-Zone Soil Moisture Product Using In Situ Measurements</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>The Soil Moisture Active Passive (SMAP) mission Level-4 <span class="hlt">Surface</span> and Root-Zone Soil Moisture (L4_SM) data product is generated by assimilating SMAP L-band <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> into the NASA Catchment land <span class="hlt">surface</span> model. The L4_SM product is available from 31 March 2015 to present (with...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SPIE10514E..0GR','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SPIE10514E..0GR"><span>Next generation diode lasers with enhanced <span class="hlt">brightness</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ried, S.; Rauch, S.; Irmler, L.; Rikels, J.; Killi, A.; Papastathopoulos, E.; Sarailou, E.; Zimer, H.</p> <p>2018-02-01</p> <p>High-power diode lasers are nowadays well established manufacturing tools in high power materials processing, mainly for tactile welding, <span class="hlt">surface</span> treatment and cladding applications. Typical beam parameter products (BPP) of such lasers range from 30 to 50 mm·mrad at several kilowatts of output power. TRUMPF offers a product line of diode lasers to its customers ranging from 150 W up to 6 kW of output power. These diode lasers combine high reliability with small footprint and high efficiency. However, up to now these lasers are limited in <span class="hlt">brightness</span> due to the commonly used spatial and coarse spectral beam combining techniques. Recently diode lasers with enhanced <span class="hlt">brightness</span> have been presented by use of dense wavelength multiplexing (DWM). In this paper we report on TRUMPF's diode lasers utilizing DWM. We demonstrate a 2 kW and a 4 kW system ideally suited for fine welding and scanner welding applications. The typical laser efficiency is in the range of 50%. The system offers plug and play exchange of the fiber beam delivery cable, multiple optical outputs and integrated cooling in a very compact package. An advanced control system offers flexible integration in any customer's shop floor environment and includes industry 4.0 capabilities (e.g. condition monitoring and predictive maintenance).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720021659','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720021659"><span>Microwave emission measurements of sea <span class="hlt">surface</span> roughness, soil moisture, and sea ice structure</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gloersen, P.; Wilheit, T. T.; Schmugge, T. J.</p> <p>1972-01-01</p> <p>In order to demonstrate the feasibility of the microwave radiometers to be carried aboard the Nimbus 5 and 6 satellites and proposed for one of the earth observatory satellites, remote measurements of microwave radiation at wavelengths ranging from 0.8 to 21 cm have been made of a variety of the earth's <span class="hlt">surfaces</span> from the NASA CV-990 A/C. <span class="hlt">Brightness</span> temperatures of sea water <span class="hlt">surfaces</span> of varying roughness, of terrain with varying soil moisture, and of sea ice of varying structure were <span class="hlt">observed</span>. In each case, around truth information was available for correlation with the microwave <span class="hlt">brightness</span> temperature. The utility of passive microwave radiometry in determining ocean <span class="hlt">surface</span> wind speeds, at least for values higher than 7 meters/second has been demonstrated. In addition, it was shown that radiometric signatures can be used to determine soil moisture in unvegetated terrain to within five percentage points by weight. Finally, it was demonstrated that first year thick, multi-year, and first year thin sea ice can be distinguished by <span class="hlt">observing</span> their differing microwave emissivities at various wavelengths.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JKPS...61.1046H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JKPS...61.1046H"><span>Precursor state of oxygen molecules on the Si(001) <span class="hlt">surface</span> during the initial room-temperature adsorption</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hwang, Eunkyung; Chang, Yun Hee; Kim, Yong-Sung; Koo, Ja-Yong; Kim, Hanchul</p> <p>2012-10-01</p> <p>The initial adsorption of oxygen molecules on Si(001) is investigated at room temperature. The scanning tunneling microscopy images reveal a unique <span class="hlt">bright</span> O2-induced feature. The very initial sticking coefficient of O2 below 0.04 Langmuir is measured to be ˜0.16. Upon thermal annealing at 250-600 °C, the <span class="hlt">bright</span> O2-induced feature is destroyed, and the Si(001) <span class="hlt">surface</span> is covered with dark depressions that seem to be oxidized structures with -Si-O-Si- bonds. This suggests that the <span class="hlt">observed</span> <span class="hlt">bright</span> O2-induced feature is an intermediate precursor state that may be either a silanone species or a molecular adsorption structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.P41D2863K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.P41D2863K"><span>Coronagraphic <span class="hlt">Observations</span> of the Lunar Sodium Exosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Killen, R. M.; Johnson, J. D.; Morgan, T. H.; Potter, A. E.</p> <p>2017-12-01</p> <p>We have designed, built and installed a small robotic coronagraph at the Winer Observatory in Sonoita, Arizona, in order to <span class="hlt">observe</span> the sodium exosphere out to one-half degree around the Moon. <span class="hlt">Observations</span> are obtained remotely every available clear night from our home base at Goddard Space Flight Center. Our data encompass lunations in 2015, 2016, and 2017, thus we have a long baseline of sodium exospheric calibrated images. We employ an Andover temperature-controlled 1.5 Å wide narrow-band filter centered on the sodium D2 line, and a similar 1.5 Å filter centered blueward of the D2 line by 5 Å. Exposures of 10 minutes are required to image the sodium corona at good signal to noise. Autoguiding is performed locking onto a small <span class="hlt">bright</span> crater each night. Following each onband-offband exposure pair, on- and off-band images of the lunar <span class="hlt">surface</span> are collected by taking a 0.1- 0.5 second exposures with the open filter. The sodium is calibrated using the counts in the open Moon images and the Hapke function. We use both dark and <span class="hlt">bright</span> Hapke parameters for comparison check using Mare and highlands, respectively. In order to obtain the sodium profile around the entire limb, the images are transformed using a polar transform and the profiles are extracted automatically. We have derived zenith column abundances and <span class="hlt">surface</span> abundances around the lunar limb for each <span class="hlt">observation</span> and we fit these <span class="hlt">observations</span> with a 3-dimensional model. We compare our lunar model derived from these <span class="hlt">observations</span> with the data from the spectrograph onboard the LADEE spacecraft.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910000198&hterms=outer+space&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Douter%2Bspace','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910000198&hterms=outer+space&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Douter%2Bspace"><span>Microwave <span class="hlt">Brightness</span> Of Land <span class="hlt">Surfaces</span> From Outer Space</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kerr, Yann H.; Njoku, Eni G.</p> <p>1991-01-01</p> <p>Mathematical model approximates microwave radiation emitted by land <span class="hlt">surfaces</span> traveling to microwave radiometer in outer space. Applied to measurements made by Scanning Multichannel Microwave Radiometer (SMMR). Developed for interpretation of microwave imagery of Earth to obtain distributions of various chemical, physical, and biological characteristics across its <span class="hlt">surface</span>. Intended primarily for use in mapping moisture content of soil and fraction of Earth covered by vegetation. Advanced Very-High-Resolution Radiometer (AVHRR), provides additional information on vegetative cover, thereby making possible retrieval of soil-moisture values from SMMR measurements. Possible to monitor changes of land <span class="hlt">surface</span> during intervals of 5 to 10 years, providing significant data for mathematical models of evolution of climate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1993SPIE.1951...32L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1993SPIE.1951...32L"><span>Determination of debris albedo from visible and infrared <span class="hlt">brightnesses</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lambert, John V.; Osteen, Thomas J.; Kraszewski, Butch</p> <p>1993-09-01</p> <p>The Air Force Phillips Laboratory is conducting measurements to characterize the orbital debris environment using wide-field optical systems located at the Air Force's Maui, Hawaii, Space Surveillance Site. Conversion of the <span class="hlt">observed</span> visible <span class="hlt">brightnesses</span> of detected debris objects to physical sizes require knowledge of the albedo (reflectivity). A thermal model for small debris objects has been developed and is used to calculate albedos from simultaneous visible and thermal infrared <span class="hlt">observations</span> of catalogued debris objects. The model and initial results will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss023e029061.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss023e029061.html"><span>Earth <span class="hlt">Observations</span> taken by the Expedition 23 Crew</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2010-04-28</p> <p>ISS023-E-029061 (28 April 2010) --- City lights at night along the France-Italy border, Europe are featured in this image photographed by an Expedition 23 crew member on the International Space Station (ISS). The <span class="hlt">brightly</span> lit metropolitan areas of Torino (Italy), Lyon, and Marseille (both in France) stand out amidst numerous smaller urban areas in this dramatic photograph. The image captures the night time appearance of the France-Italy border area between the mountainous Alps to the north (not shown) and the island of Corsica in the Ligurian Sea to the south (top). The full moon reflects <span class="hlt">brightly</span> on the water <span class="hlt">surface</span> and also illuminates the tops of low patchy clouds over the border (center). This image was taken by an ISS crew member at approximately 11:55 p.m. local time when the station was located over the France-Belgium border near Luxembourg. Crew members orbiting Earth frequently collect images that include sunglint, or sunlight that reflects off a water <span class="hlt">surface</span> at such an angle that it travels directly back towards the <span class="hlt">observer</span>. Sunglint typically lends a mirror-like appearance to the water <span class="hlt">surface</span>. During clear sky conditions reflected light from the moon can produce the same effect (moon glint) as illustrated in this view. The <span class="hlt">observer</span> was looking towards the southeast at an oblique viewing angle at the time the image was taken; in other words, looking outwards from the ISS, not straight down towards Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.A33F0305B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.A33F0305B"><span>Sensitivity analysis of <span class="hlt">observed</span> reflectivity to ice particle <span class="hlt">surface</span> roughness using MISR satellite <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bell, A.; Hioki, S.; Wang, Y.; Yang, P.; Di Girolamo, L.</p> <p>2016-12-01</p> <p>Previous studies found that including ice particle <span class="hlt">surface</span> roughness in forward light scattering calculations significantly reduces the differences between <span class="hlt">observed</span> and simulated polarimetric and radiometric <span class="hlt">observations</span>. While it is suggested that some degree of roughness is desirable, the appropriate degree of <span class="hlt">surface</span> roughness to be assumed in operational cloud property retrievals and the sensitivity of retrieval products to this assumption remains uncertain. In an effort to extricate this ambiguity, we will present a sensitivity analysis of space-borne multi-angle <span class="hlt">observations</span> of reflectivity, to varying degrees of <span class="hlt">surface</span> roughness. This process is two fold. First, sampling information and statistics of Multi-angle Imaging SpectroRadiometer (MISR) sensor data aboard the Terra platform, will be used to define the most coming viewing <span class="hlt">observation</span> geometries. Using these defined geometries, reflectivity will be simulated for multiple degrees of roughness using results from adding-doubling radiative transfer simulations. Sensitivity of simulated reflectivity to <span class="hlt">surface</span> roughness can then be quantified, thus yielding a more robust retrieval system. Secondly, sensitivity of the inverse problem will be analyzed. Spherical albedo values will be computed by feeding blocks of MISR data comprising cloudy pixels over ocean into the retrieval system, with assumed values of <span class="hlt">surface</span> roughness. The sensitivity of spherical albedo to the inclusion of <span class="hlt">surface</span> roughness can then be quantified, and the accuracy of retrieved parameters can be determined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015NatSR...512653F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015NatSR...512653F"><span>Development of <span class="hlt">bright</span> fluorescent quadracyclic adenine analogues: TDDFT-calculation supported rational design</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Foller Larsen, Anders; Dumat, Blaise; Wranne, Moa S.; Lawson, Christopher P.; Preus, Søren; Bood, Mattias; Gradén, Henrik; Marcus Wilhelmsson, L.; Grøtli, Morten</p> <p>2015-07-01</p> <p>Fluorescent base analogues (FBAs) comprise a family of increasingly important molecules for the investigation of nucleic acid structure and dynamics. We recently reported the quantum chemical calculation supported development of four microenvironment sensitive analogues of the quadracyclic adenine (qA) scaffold, the qANs, with highly promising absorptive and fluorescence properties that were very well predicted by TDDFT calculations. Herein, we report on the efficient synthesis, experimental and theoretical characterization of nine novel quadracyclic adenine derivatives. The brightest derivative, 2-CNqA, displays a 13-fold increased <span class="hlt">brightness</span> (ɛΦF = 4500) compared with the parent compound qA and has the additional benefit of being a virtually microenvironment-insensitive fluorophore, making it a suitable candidate for nucleic acid incorporation and use in quantitative FRET and anisotropy experiments. TDDFT calculations, conducted on the nine novel qAs a posteriori, successfully describe the relative fluorescence quantum yield and <span class="hlt">brightness</span> of all qA derivatives. This <span class="hlt">observation</span> suggests that the TDDFT-based rational design strategy may be employed for the development of <span class="hlt">bright</span> fluorophores built up from a common scaffold to reduce the otherwise costly and time-consuming screening process usually required to obtain useful and <span class="hlt">bright</span> FBAs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29809046','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29809046"><span>Exposure to <span class="hlt">bright</span> light biases effort-based decisions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bijleveld, Erik; Knufinke, Melanie</p> <p>2018-06-01</p> <p>Secreted in the evening and the night, melatonin suppresses activity of the mesolimbic dopamine pathway, a brain pathway involved in reward processing. However, exposure to <span class="hlt">bright</span> light diminishes-or even prevents-melatonin secretion. Thus, we hypothesized that reward processing, in the evening, is more pronounced in <span class="hlt">bright</span> light (vs. dim light). Healthy human participants carried out three tasks that tapped into various aspects of reward processing (effort expenditure for rewards task [EEfRT]; two-armed bandit task [2ABT]; balloon analogue risk task [BART). <span class="hlt">Brightness</span> was manipulated within-subjects (<span class="hlt">bright</span> vs. dim light), in separate evening sessions. During the EEfRT, participants used reward-value information more strongly when they were exposed to <span class="hlt">bright</span> light (vs. dim light). This finding supported our hypothesis. However, exposure to <span class="hlt">bright</span> light did not significantly affect task behavior on the 2ABT and the BART. While future research is necessary (e.g., to zoom in on working mechanisms), these findings have potential implications for the design of physical work environments. (PsycINFO Database Record (c) 2018 APA, all rights reserved).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29890335','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29890335"><span>Effects of social anxiety on metaphorical associations between emotional valence and clothing <span class="hlt">brightness</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ishikawa, Kenta; Suzuki, Hikaru; Okubo, Matia</p> <p>2018-06-05</p> <p>Individuals with social anxiety have various types of deficiencies in emotional processing. Diversity of deficiencies may imply that socially anxious individuals have malfunctions in fundamental parts of emotional processing. Therefore, we hypothesized that social anxiety contributes to deficiencies in building on the metaphorical relationship between emotional experience and <span class="hlt">brightness</span>. We conducted a judgment task of valences of faces with manipulated clothing <span class="hlt">brightness</span> (<span class="hlt">bright</span> or dark). A congruency effect between the emotional valence and clothing <span class="hlt">brightness</span> was <span class="hlt">observed</span> in participants with low social anxiety. However, this pattern was not found in participants with high social anxiety. The results suggested that a deficiency in metaphorical associations leads to maladaptive emotional processing in individuals with social anxiety. Our findings cannot be directly generalized to clinical populations. Such populations should be tested in the future studies. We may expand Lakoff and Johnson's (1999) conceptual metaphor theory by showing the relationships between social anxiety and malfunction in metaphorical processing. Malfunctions in metaphorical processing could lead to various types of psychological disorders which have deficiencies in emotional processing. Copyright © 2018 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A11F1936C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A11F1936C"><span>CYGNSS <span class="hlt">Surface</span> Wind <span class="hlt">Observations</span> and <span class="hlt">Surface</span> Flux Estimates within Low-Latitude Extratropical Cyclones</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Crespo, J.; Posselt, D. J.</p> <p>2017-12-01</p> <p>The Cyclone Global Navigation Satellite System (CYGNSS), launched in December 2016, aims to improve estimates of <span class="hlt">surface</span> wind speeds over the tropical oceans. While CYGNSS's core mission is to provide better estimates of <span class="hlt">surface</span> winds within the core of tropical cyclones, previous research has shown that the constellation, with its orbital inclination of 35°, also has the ability to <span class="hlt">observe</span> numerous extratropical cyclones that form in the lower latitudes. Along with its high spatial and temporal resolution, CYGNSS can provide new insights into how extratropical cyclones develop and evolve, especially in the presence of thick clouds and precipitation. We will demonstrate this by presenting case studies of multiple extratropical cyclones <span class="hlt">observed</span> by CYGNSS early on in its mission in both Northern and Southern Hemispheres. By using the improved estimates of <span class="hlt">surface</span> wind speeds from CYGNSS, we can obtain better estimates of <span class="hlt">surface</span> latent and sensible heat fluxes within and around extratropical cyclones. <span class="hlt">Surface</span> heat fluxes, driven by <span class="hlt">surface</span> winds and strong vertical gradients of water vapor and temperature, play a key role in marine cyclogenesis as they increase instability within the boundary layer and may contribute to extreme marine cyclogenesis. In the past, it has been difficult to estimate <span class="hlt">surface</span> heat fluxes from space borne instruments, as these fluxes cannot be <span class="hlt">observed</span> directly from space, and deficiencies in spatial coverage and attenuation from clouds and precipitation lead to inaccurate estimates of <span class="hlt">surface</span> flux components, such as <span class="hlt">surface</span> wind speeds. While CYGNSS only contributes estimates of <span class="hlt">surface</span> wind speeds, we can combine this data with other reanalysis and satellite data to provide improved estimates of <span class="hlt">surface</span> sensible and latent heat fluxes within and around extratropical cyclones and throughout the entire CYGNSS mission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ATel11688....1F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ATel11688....1F"><span>The ZTF <span class="hlt">Bright</span> Transient Survey</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fremling, C.; Sharma, Y.; Kulkarni, S. R.; Miller, A. A.; Taggart, K.; Perley, D. A.; Gooba, A.</p> <p>2018-06-01</p> <p>As a supplement to the Zwicky Transient Facility (ZTF; ATel #11266) public alerts (ATel #11685) we plan to report (following ATel #11615) <span class="hlt">bright</span> probable supernovae identified in the raw alert stream from the ZTF Northern Sky Survey ("Celestial Cinematography"; see Bellm & Kulkarni, 2017, Nature Astronomy 1, 71) to the Transient Name Server (https://wis-tns.weizmann.ac.il) on a daily basis; the ZTF <span class="hlt">Bright</span> Transient Survey (BTS; see Kulkarni et al., 2018; arXiv:1710.04223).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.tmp.1251M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.tmp.1251M"><span>A lower occurrence rate of <span class="hlt">bright</span> X-ray flares in SN-GRBs than z < 1 GRBs: evidence of energy partitions?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mu, Hui-Jun; Gu, Wei-Min; Mao, Jirong; Liu, Tong; Hou, Shu-Jin; Lin, Da-Bin; Wang, Junfeng; Fang, Taotao; Liang, En-Wei</p> <p>2018-05-01</p> <p>The occurrence rates of <span class="hlt">bright</span> X-ray flares in z < 1 gamma-ray bursts (GRBs) with or without <span class="hlt">observed</span> supernovae (SNe) association were compared. Our Sample I: the z < 1 long GRBs (LGRBs) with SNe association (SN-GRBs) and with early Swift/X-Ray Telescope (XRT) <span class="hlt">observations</span>, consists of 18 GRBs, among which only two GRBs have <span class="hlt">bright</span> X-ray flares. Our Sample II: for comparison, all the z < 1 LGRBs without <span class="hlt">observed</span> SNe association and with early Swift/XRT <span class="hlt">observations</span>, consists of 45 GRBs, among which 16 GRBs present <span class="hlt">bright</span> X-ray flares. Thus, the study indicates a lower occurrence rate of <span class="hlt">bright</span> X-ray flares in Sample I (11.1%) than in Sample II (35.6%). In addition, if dim X-ray fluctuations are included as flares, then 16.7% of Sample I and 55.6% of Sample II are found to have flares, again showing the discrepancy between these two samples. We examined the physical origin of these <span class="hlt">bright</span> X-ray flares and found that most of them are probably related to the central engine reactivity. To understand the discrepancy, we propose that such a lower occurrence rate of flares in the SN-GRB sample may hint at an energy partition among the GRB, SNe, and X-ray flares under a saturated energy budget of massive star explosion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e002110.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e002110.html"><span><span class="hlt">Bright</span> Solar Flare</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>A <span class="hlt">bright</span> solar flare is captured by the EIT 195Å instrument on 1998 May 2. A solar flare (a sudden, rapid, and intense variation in <span class="hlt">brightness</span>) occurs when magnetic energy that has built up in the solar atmosphere is suddenly released, launching material outward at millions of km per hour. The Sun’s magnetic fields tend to restrain each other and force the buildup of tremendous energy, like twisting rubber bands, so much that they eventually break. At some point, the magnetic lines of force merge and cancel in a process known as magnetic reconnection, causing plasma to forcefully escape from the Sun. Credit: NASA/GSFC/SOHO/ESA To learn more go to the SOHO website: sohowww.nascom.nasa.gov/home.html To learn more about NASA's Sun Earth Day go here: sunearthday.nasa.gov/2010/index.php</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PASJ...70S...7C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PASJ...70S...7C"><span>The <span class="hlt">bright</span>-star masks for the HSC-SSP survey</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Coupon, Jean; Czakon, Nicole; Bosch, James; Komiyama, Yutaka; Medezinski, Elinor; Miyazaki, Satoshi; Oguri, Masamune</p> <p>2018-01-01</p> <p>We present the procedure to build and validate the <span class="hlt">bright</span>-star masks for the Hyper-Suprime-Cam Strategic Subaru Proposal (HSC-SSP) survey. To identify and mask the saturated stars in the full HSC-SSP footprint, we rely on the Gaia and Tycho-2 star catalogues. We first assemble a pure star catalogue down to GGaia < 18 after removing ˜1.5% of sources that appear extended in the Sloan Digital Sky Survey (SDSS). We perform visual inspection on the early data from the S16A internal release of HSC-SSP, finding that our star catalogue is 99.2% pure down to GGaia < 18. Second, we build the mask regions in an automated way using stacked detected source measurements around <span class="hlt">bright</span> stars binned per GGaia magnitude. Finally, we validate those masks by visual inspection and comparison with the literature of galaxy number counts and angular two-point correlation functions. This version (Arcturus) supersedes the previous version (Sirius) used in the S16A internal and DR1 public releases. We publicly release the full masks and tools to flag objects in the entire footprint of the planned HSC-SSP <span class="hlt">observations</span> at "ftp://obsftp.unige.ch/pub/coupon/<span class="hlt">bright</span>StarMasks/HSC-SSP/".</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830066004&hterms=surface+density&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dsurface%2Bdensity','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830066004&hterms=surface+density&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dsurface%2Bdensity"><span>Passive microwave sensing of soil moisture content - The effects of soil bulk density and <span class="hlt">surface</span> roughness</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wang, J. R.</p> <p>1983-01-01</p> <p>Microwave radiometric measurements over bare fields of different <span class="hlt">surface</span> roughness were made at frequencies of 1.4 GHz, 5 GHz, and 10.7 GHz to study the frequency dependence, as well as the possible time variation, of <span class="hlt">surface</span> roughness. An increase in <span class="hlt">surface</span> roughness was found to increase the <span class="hlt">brightness</span> temperature of soils and reduce the slope of regression between <span class="hlt">brightness</span> temperature and soil moisture content. The frequency dependence of the <span class="hlt">surface</span> roughness effect was relatively weak when compared with that of the vegetation effect. Radiometric time-series <span class="hlt">observations</span> over a given field indicate that field <span class="hlt">surface</span> roughness might gradually diminish with time, especially after a rainfall or irrigation. The variation of <span class="hlt">surface</span> roughness increases the uncertainty of remote soil moisture estimates by microwave radiometry. Three years of radiometric measurements over a test site revealed a possible inconsistency in the soil bulk density determination, which is an important factor in the interpretation of radiometric data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.469S.475R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.469S.475R"><span>Spatial variations of <span class="hlt">brightness</span>, colour and polarization of dust in comet 67P/Churyumov-Gerasimenko</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rosenbush, Vera K.; Ivanova, Oleksandra V.; Kiselev, Nikolai N.; Kolokolova, Ludmilla O.; Afanasiev, Viktor L.</p> <p>2017-07-01</p> <p>We present post-perihelion photometric and polarimetric <span class="hlt">observations</span> of comet 67P/Churyumov-Gerasimenko performed at the 6-m telescope of the SAO RAS in the g-sdss (465/65 nm), r-sdss (620/60 nm) and R filters. <span class="hlt">Observations</span> in November and December 2015 and April 2016 covered the range of heliocentric distance 1.62-2.72 au and phase angle 33.2°-10.4°. The comet was very active. Two persistent jets and long dust tail were <span class="hlt">observed</span> during the whole <span class="hlt">observing</span> period; one more jet was detected only in December. The radial profiles of <span class="hlt">surface</span> <span class="hlt">brightness</span>, colour and polarization significantly differed for the coma, jets and tail, and changed with increasing heliocentric distance. The dust production Afρ decreased from 162 cm at r = 1.62 au to 51 cm at r = 2.72 au. The dust colour (g-r) gradually changed from 0.8 mag in the innermost coma to about 0.4 mag in the outer coma. The spectral slope was 8.2 ± 1.7 per cent/100 nm in the 465 to 620 nm wavelength domain. In November and December, the polarization in the near-nucleus area was about 8 per cent, dropped sharply to 2 per cent at the distance above 5000 km and then gradually increased with distance from the nucleus, reaching ˜8 per cent at 40 000 km. In April, at a phase angle 10.4°, the polarization varied between -0.6 per cent in the near-nucleus area and -4 per cent in the outer coma. Circular polarization was not detected in the comet. The spatial variations of <span class="hlt">brightness</span>, colour and polarization in different structural features suggest some evolution of particle properties, most likely decreasing the size of dust particles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900048604&hterms=relationship+form&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Drelationship%2Bform','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900048604&hterms=relationship+form&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Drelationship%2Bform"><span>Relationship of magnetic field strength and <span class="hlt">brightness</span> of fine-structure elements in the solar temperature minimum region</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cook, J. W.; Ewing, J. A.</p> <p>1990-01-01</p> <p>A quantitative relationship was determined between magnetic field strength (or magnetic flux) from photospheric magnetograph <span class="hlt">observations</span> and the <span class="hlt">brightness</span> temperature of solar fine-structure elements <span class="hlt">observed</span> at 1600 A, where the predominant flux source is continuum emission from the solar temperature minimum region. A Kitt Peak magnetogram and spectroheliograph <span class="hlt">observations</span> at 1600 A taken during a sounding rocket flight of the High Resolution Telescope and Spectrograph from December 11, 1987 were used. The statistical distributions of <span class="hlt">brightness</span> temperature in the quiet sun at 1600 A, and absolute value of magnetic field strength in the same area were determined from these <span class="hlt">observations</span>. Using a technique which obtains the best-fit relationship of a given functional form between these two histogram distributions, a quantitative relationship was determined between absolute value of magnetic field strength B and <span class="hlt">brightness</span> temperature which is essentially linear from 10 to 150 G. An interpretation is suggested, in which a basal heating occurs generally, while brighter elements are produced in magnetic regions with temperature enhancements proportional to B.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120013516','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120013516"><span>Advances in Assimilation of Satellite-Based Passive Microwave <span class="hlt">Observations</span> for Soil-Moisture Estimation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>De Lannoy, Gabrielle J. M.; Pauwels, Valentijn; Reichle, Rolf H.; Draper, Clara; Koster, Randy; Liu, Qing</p> <p>2012-01-01</p> <p>Satellite-based microwave measurements have long shown potential to provide global information about soil moisture. The European Space Agency (ESA) Soil Moisture and Ocean Salinity (SMOS, [1]) mission as well as the future National Aeronautics and Space Administration (NASA) Soil Moisture Active and Passive (SMAP, [2]) mission measure passive microwave emission at L-band frequencies, at a relatively coarse (40 km) spatial resolution. In addition, SMAP will measure active microwave signals at a higher spatial resolution (3 km). These new L-band missions have a greater sensing depth (of -5cm) compared with past and present C- and X-band microwave sensors. ESA currently also disseminates retrievals of SMOS <span class="hlt">surface</span> soil moisture that are derived from SMOS <span class="hlt">brightness</span> temperature <span class="hlt">observations</span> and ancillary data. In this research, we address two major challenges with the assimilation of recent/future satellite-based microwave measurements: (i) assimilation of soil moisture retrievals versus <span class="hlt">brightness</span> temperatures for <span class="hlt">surface</span> and root-zone soil moisture estimation and (ii) scale-mismatches between satellite <span class="hlt">observations</span>, models and in situ validation data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170008275&hterms=software&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dsoftware','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170008275&hterms=software&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dsoftware"><span>Extended <span class="hlt">Bright</span> Bodies - Flight and Ground Software Challenges on the Cassini Mission at Saturn</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sung, Tina S.; Burk, Thomas A.</p> <p>2016-01-01</p> <p>Extended <span class="hlt">bright</span> bodies in the Saturn environment such as Saturn's rings, the planet itself, and Saturn's satellites near the Cassini spacecraft may interfere with the star tracker's ability to find stars. These interferences can create faulty spacecraft attitude knowledge, which would decrease the pointing accuracy or even trip a fault protection response on board the spacecraft. The effects of the extended <span class="hlt">bright</span> body interference were <span class="hlt">observed</span> in December of 2000 when Cassini flew by Jupiter. Based on this flight experience and expected star tracker behavior at Saturn, the Cassini AACS operations team defined flight rules to suspend the star tracker during predicted interference windows. The flight rules are also implemented in the existing ground software called Kinematic Predictor Tool to create star identification suspend commands to be uplinked to the spacecraft for future predicted interferences. This paper discusses the details of how extended <span class="hlt">bright</span> bodies impact Cassini's acquisition of attitude knowledge, how the <span class="hlt">observed</span> data helped the ground engineers in developing flight rules, and how automated methods are used in the flight and ground software to ensure the spacecraft is continuously operated within these flight rules. This paper also discusses how these established procedures will continue to be used to overcome new <span class="hlt">bright</span> body challenges that Cassini will encounter during its dips inside the rings of Saturn for its final orbits of a remarkable 20-year mission at Saturn.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvA..97d3623K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvA..97d3623K"><span>Dark-<span class="hlt">bright</span> soliton pairs: Bifurcations and collisions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Katsimiga, G. C.; Kevrekidis, P. G.; Prinari, B.; Biondini, G.; Schmelcher, P.</p> <p>2018-04-01</p> <p>The statics, stability, and dynamical properties of dark-<span class="hlt">bright</span> soliton pairs are investigated here, motivated by applications in a homogeneous two-component repulsively interacting Bose-Einstein condensate. One of the intraspecies interaction coefficients is used as the relevant parameter controlling the deviation from the integrable Manakov limit. Two different families of stationary states are identified consisting of dark-<span class="hlt">bright</span> solitons that are either antisymmetric (out-of-phase) or asymmetric (mass imbalanced) with respect to their <span class="hlt">bright</span> soliton. Both of the above dark-<span class="hlt">bright</span> configurations coexist at the integrable limit of equal intra and interspecies repulsions and are degenerate in that limit. However, they are found to bifurcate from it in a transcritical bifurcation. This bifurcation interchanges the stability properties of the bound dark-<span class="hlt">bright</span> pairs rendering the antisymmetric states unstable and the asymmetric ones stable past the associated critical point (and vice versa before it). Finally, on the dynamical side, it is found that large kinetic energies and thus rapid soliton collisions are essentially unaffected by the intraspecies variation, while cases involving near equilibrium states or breathing dynamics are significantly modified under such a variation.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_23 --> <div id="page_24" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="461"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170002529','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170002529"><span>Si III OV <span class="hlt">Bright</span> Line of Scattering Polarized Light That Has Been <span class="hlt">Observed</span> in the CLASP and Its Center-to-Limb Variation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Katsukawa, Yukio; Ishikawa, Ryoko; Kano, Ryohei; Kubo, Masahito; Noriyuki, Narukage; Kisei, Bando; Hara, Hirohisa; Yoshiho, Suematsu; Goto, Motouji; Ishikawa, Shinnosuke; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20170002529'); toggleEditAbsImage('author_20170002529_show'); toggleEditAbsImage('author_20170002529_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20170002529_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20170002529_hide"></p> <p>2017-01-01</p> <p>The CLASP (Chromospheric Lyman-Alpha Spectro- Polarimeter) rocket experiment, in addition to the ultraviolet region of the Ly alpha emission line (121.57 nm), emission lines of Si III (120.65 nm) and OV (121.83 nm) is can be <span class="hlt">observed</span>. These are optically thin line compared to a Ly alpha line, if Rarere captured its polarization, there is a possibility that dripping even a new physical diagnosis chromosphere-transition layer. In particular, OV <span class="hlt">bright</span> light is a release from the transition layer, further, three P one to one S(sub 0) is a forbidden line (cross-triplet transition between lines), it was not quite know whether to polarization.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...854...75S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...854...75S"><span>Following the Cosmic Evolution of Pristine Gas. II. The Search for Pop III–<span class="hlt">bright</span> Galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sarmento, Richard; Scannapieco, Evan; Cohen, Seth</p> <p>2018-02-01</p> <p>Direct <span class="hlt">observational</span> searches for Population III (Pop III) stars at high redshift are faced with the question of how to select the most promising targets for spectroscopic follow-up. To help answer this, we use a large-scale cosmological simulation, augmented with a new subgrid model that tracks the fraction of pristine gas, to follow the evolution of high-redshift galaxies and the Pop III stars they contain. We generate rest-frame ultraviolet (UV) luminosity functions for our galaxies and find that they are consistent with current z≥slant 7 <span class="hlt">observations</span>. Throughout the redshift range 7≤slant z≤slant 15, we identify “Pop III–bright” galaxies as those with at least 75% of their flux coming from Pop III stars. While less than 1% of galaxies brighter than {m}UV,{AB}}=31.4 mag are Pop III–<span class="hlt">bright</span> in the range 7≤slant z≤slant 8, roughly 17% of such galaxies are Pop III–<span class="hlt">bright</span> at z = 9, immediately before reionization occurs in our simulation. Moving to z = 10, {m}UV,{AB}}=31.4 mag corresponds to larger, more luminous galaxies, and the Pop III–<span class="hlt">bright</span> fraction falls off to 5%. Finally, at the highest redshifts, a large fraction (29% at z = 14 and 41% at z = 15) of all galaxies are Pop III–<span class="hlt">bright</span> regardless of magnitude. While {m}UV,{AB}}=31.4 mag galaxies are extremely rare during this epoch, we find that 13% of galaxies at z = 14 are Pop III–<span class="hlt">bright</span> with {m}UV,{AB}}≤slant 33 mag, a intrinsic magnitude within reach of the James Webb Space Telescope using lensing. Thus, we predict that the best redshift to search for luminous Pop III–<span class="hlt">bright</span> galaxies is just before reionization, while lensing surveys for fainter galaxies should push to the highest redshifts possible.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JARS...12a6032E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JARS...12a6032E"><span>Simulation of the <span class="hlt">brightness</span> temperatures <span class="hlt">observed</span> by the visible infrared imaging radiometer suite instrument</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Evrard, Rebecca L.; Ding, Yifeng</p> <p>2018-01-01</p> <p>Clouds play a large role in the Earth's global energy budget, but the impact of cirrus clouds is still widely questioned and researched. Cirrus clouds reside high in the atmosphere and due to cold temperatures are comprised of ice crystals. Gaining a better understanding of ice cloud optical properties and the distribution of cirrus clouds provides an explanation for the contribution of cirrus clouds to the global energy budget. Using radiative transfer models (RTMs), accurate simulations of cirrus clouds can enhance the understanding of the global energy budget as well as improve the use of global climate models. A newer, faster RTM such as the visible infrared imaging radiometer suite (VIIRS) fast radiative transfer model (VFRTM) is compared to a rigorous RTM such as the line-by-line radiative transfer model plus the discrete ordinates radiative transfer program. By comparing <span class="hlt">brightness</span> temperature (BT) simulations from both models, the accuracy of the VFRTM can be obtained. This study shows root-mean-square error <0.2 K for BT difference using reanalysis data for atmospheric profiles and updated ice particle habit information from the moderate-resolution imaging spectroradiometer collection 6. At a higher resolution, the simulated results of the VFRTM are compared to the <span class="hlt">observations</span> of VIIRS resulting in a <1.5 % error from the VFRTM for all cases. The VFRTM is validated and is an appropriate RTM to use for global cloud retrievals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19770020110','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19770020110"><span>An interpretation of photometric parameters of <span class="hlt">bright</span> desert regions of Mars and their dependence on wave length</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Weaver, W. R.; Meador, W. E.</p> <p>1977-01-01</p> <p>Photometric data from the <span class="hlt">bright</span> desert areas of Mars were used to determine the dependence of the three photometric parameters of the photometric function on wavelength and to provide qualitative predictions about the physical properties of the <span class="hlt">surface</span>. Knowledge of the parameters allowed the <span class="hlt">brightness</span> of these areas of Mars to be determined for any scattering geometry in the wavelength range of 0.45 to 0.70 micron. The changes that occur in the photometric parameters due to changes in wavelength were shown to be consistent with their physical interpretations, and the predictions of <span class="hlt">surface</span> properties were shown to be consistent with conditions expected to exist in these regions of Mars. The photometric function was shown to have potential as a diagnostic tool for the qualitative determination of <span class="hlt">surface</span> properties, and the consistency of the behavior of the photometric parameters was considered to be support for the validity of the photometric function.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22520034-herschel-spectroscopic-observations-little-things-dwarf-galaxies','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22520034-herschel-spectroscopic-observations-little-things-dwarf-galaxies"><span>HERSCHEL SPECTROSCOPIC <span class="hlt">OBSERVATIONS</span> OF LITTLE THINGS DWARF GALAXIES</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Cigan, Phil; Young, Lisa; Cormier, Diane</p> <p></p> <p>We present far-infrared (FIR) spectral line <span class="hlt">observations</span> of five galaxies from the Little Things sample: DDO 69, DDO 70, DDO 75, DDO 155, and WLM. While most studies of dwarfs focus on <span class="hlt">bright</span> systems or starbursts due to <span class="hlt">observational</span> constraints, our data extend the <span class="hlt">observed</span> parameter space into the regime of low <span class="hlt">surface</span> <span class="hlt">brightness</span> dwarf galaxies with low metallicities and moderate star formation rates. Our targets were <span class="hlt">observed</span> with Herschel at the [C ii] 158 μm, [O i] 63 μm, [O iii] 88 μm, and [N ii] 122 μm emission lines using the PACS Spectrometer. These high-resolution maps allow usmore » for the first time to study the FIR properties of these systems on the scales of larger star-forming complexes. The spatial resolution in our maps, in combination with star formation tracers, allows us to identify separate photodissociation regions (PDRs) in some of the regions we <span class="hlt">observed</span>. Our systems have widespread [C ii] emission that is <span class="hlt">bright</span> relative to continuum, averaging near 0.5% of the total infrared (TIR) budget—higher than in solar-metallicity galaxies of other types. [N ii] is weak, suggesting that the [C ii] emission in our galaxies comes mostly from PDRs instead of the diffuse ionized interstellar medium (ISM). These systems exhibit efficient cooling at low dust temperatures, as shown by ([O i]+[C ii])/TIR in relation to 60 μm/100 μm, and low [O i]/[C ii] ratios which indicate that [C ii] is the dominant coolant of the ISM. We <span class="hlt">observe</span> [O iii]/[C ii] ratios in our galaxies that are lower than those published for other dwarfs, but similar to levels noted in spirals.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19800008728','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19800008728"><span>UBVR <span class="hlt">observation</span> of V1357 Cyg = Cyg X-1. Search of the optical radiation of the accretion disk</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Shevchenko, V. S.</p> <p>1979-01-01</p> <p>Data from 30 nights of V 1357 Cyg <span class="hlt">observations</span> in July, August, and September of 1977 are presented. The contribution of the disk to the optic <span class="hlt">brightness</span> of the system is computed with regard for the heating of its <span class="hlt">surface</span> by ultraviolet radiation from V 1357 Cyg and X-ray radiation from Cyg X-1. The disk radiation explains the irregular variability in the system <span class="hlt">brightness</span>. The possibility of the eclipse of the star by the disk and the disk by the star is discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFM.A33E1653H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFM.A33E1653H"><span>Development of Yellow Sand Image Products Using Infrared <span class="hlt">Brightness</span> Temperature Difference Method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ha, J.; Kim, J.; Kwak, M.; Ha, K.</p> <p>2007-12-01</p> <p>A technique for detection of airborne yellow sand dust using meteorological satellite has been developed from various bands from ultraviolet to infrared channels. Among them, Infrared (IR) channels have an advantage of detecting aerosols over high reflecting <span class="hlt">surface</span> as well as during nighttime. There had been suggestion of using <span class="hlt">brightness</span> temperature difference (BTD) between 11 and 12¥ìm. We have found that the technique is highly depends on <span class="hlt">surface</span> temperature, emissivity, and zenith angle, which results in changing the threshold of BTD. In order to overcome these problems, we have constructed the background <span class="hlt">brightness</span> temperature threshold of BTD and then aerosol index (AI) has been determined from subtracting the background threshold from BTD of our interested scene. Along with this, we utilized high temporal coverage of geostationary satellite, MTSAT, to improve the reliability of the determined AI signal. The products have been evaluated by comparing the forecasted wind field with the movement fiend of AI. The statistical score test illustrates that this newly developed algorithm produces a promising result for detecting mineral dust by reducing the errors with respect to the current BTD method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJS..234...17C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJS..234...17C"><span>The Most Compact <span class="hlt">Bright</span> Radio-loud AGNs. II. VLBA <span class="hlt">Observations</span> of 10 Sources at 43 and 86 GHz</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cheng, X.-P.; An, T.; Hong, X.-Y.; Yang, J.; Mohan, P.; Kellermann, K. I.; Lister, M. L.; Frey, S.; Zhao, W.; Zhang, Z.-L.; Wu, X.-C.; Li, X.-F.; Zhang, Y.-K.</p> <p>2018-01-01</p> <p>Radio-loud active galactic nuclei (AGNs), hosting powerful relativistic jet outflows, provide an excellent laboratory for studying jet physics. Very long baseline interferometry (VLBI) enables high-resolution imaging on milli-arcsecond (mas) and sub-mas scales, making it a powerful tool to explore the inner jet structure, shedding light on the formation, acceleration, and collimation of AGN jets. In this paper, we present Very Long Baseline Array <span class="hlt">observations</span> of 10 radio-loud AGNs at 43 and 86 GHz that were selected from the Planck catalog of compact sources and are among the brightest in published VLBI images at and below 15 GHz. The image noise levels in our <span class="hlt">observations</span> are typically 0.3 and 1.5 mJy beam‑1 at 43 and 86 GHz, respectively. Compared with the VLBI data <span class="hlt">observed</span> at lower frequencies from the literature, our <span class="hlt">observations</span> with higher resolutions (with the highest resolution being up to 0.07 mas at 86 GHz and 0.18 mas at 43 GHz) and at higher frequencies detected new jet components at sub-parsec scales, offering valuable data for studies of the physical properties of the innermost jets. These include the compactness factor of the radio structure (the ratio of core flux density to total flux density), and core <span class="hlt">brightness</span> temperature ({T}{{b}}). In all these sources, the compact core accounts for a significant fraction (> 60 % ) of the total flux density. Their correlated flux density at the longest baselines is higher than 0.16 Jy. The compactness of these sources make them good phase calibrators of millimeter-wavelength ground-based and space VLBI.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11242037','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11242037"><span>Flooding of Ganymede's <span class="hlt">bright</span> terrains by low-viscosity water-ice lavas.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schenk, P M; McKinnon, W B; Gwynn, D; Moore, J M</p> <p>2001-03-01</p> <p>Large regions of the jovian moon Ganymede have been resurfaced, but the means has been unclear. Suggestions have ranged from volcanic eruptions of liquid water or solid ice to tectonic deformation, but definitive high-resolution morphological evidence has been lacking. Here we report digital elevation models of parts of the <span class="hlt">surface</span> of Ganymede, derived from stereo pairs combining data from the Voyager and Galileo spacecraft, which reveal <span class="hlt">bright</span>, smooth terrains that lie at roughly constant elevations 100 to 1,000 metres below the surrounding rougher terrains. These topographic data, together with new images that show fine-scale embayment and burial of older features, indicate that the smooth terrains were formed by flooding of shallow structural troughs by low-viscosity water-ice lavas. The oldest and most deformed areas (the 'reticulate' terrains) in general have the highest relative elevations, whereas units of the most common resurfaced type--the grooved terrain--lie at elevations between those of the smooth and reticulate terrains. <span class="hlt">Bright</span> terrain, which accounts for some two-thirds of the <span class="hlt">surface</span>, probably results from a continuum of processes, including crustal rifting, shallow flooding and groove formation. Volcanism plays an integral role in these processes, and is consistent with partial melting of Ganymede's interior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013RAA....13.1255Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013RAA....13.1255Y"><span>Moon night sky <span class="hlt">brightness</span> simulation for the Xinglong station</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yao, Song; Zhang, Hao-Tong; Yuan, Hai-Long; Zhao, Yong-Heng; Dong, Yi-Qiao; Bai, Zhong-Rui; Deng, Li-Cai; Lei, Ya-Juan</p> <p>2013-10-01</p> <p>Using a sky <span class="hlt">brightness</span> monitor at the Xinglong station of National Astronomical Observatories, Chinese Academy of Sciences, we collected data from 22 dark clear nights and 90 moon nights. We first measured the sky <span class="hlt">brightness</span> variation with time for dark nights and found a clear correlation between sky <span class="hlt">brightness</span> and human activity. Then with a modified sky <span class="hlt">brightness</span> model of moon nights and data from these nights, we derived the typical value for several important parameters in the model. With these results, we calculated the sky <span class="hlt">brightness</span> distribution under a given moon condition for the Xinglong station. Furthermore, we simulated the sky <span class="hlt">brightness</span> distribution of a moon night for a telescope with a 5° field of view (such as LAMOST). These simulations will be helpful for determining the limiting magnitude and exposure time, as well as planning the survey for LAMOST during moon nights.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008ApSS..255.1606S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008ApSS..255.1606S"><span>A high <span class="hlt">brightness</span> source for nano-probe secondary ion mass spectrometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Smith, N. S.; Tesch, P. P.; Martin, N. P.; Kinion, D. E.</p> <p>2008-12-01</p> <p>The two most prevalent ion source technologies in the field of <span class="hlt">surface</span> analysis and <span class="hlt">surface</span> machining are the Duoplasmatron and the liquid metal ion source (LMIS). There have been many efforts in this area of research to develop an alternative source [ S.K. Guharay, J. Orloff, M. Wada, IEEE Trans. Plasma Sci. 33 (6) (2005) 1911; N.S. Smith, W.P. Skoczylas, S.M. Kellogg, D.E. Kinion, P.P. Tesch, O. Sutherland, A. Aanesland, R.W. Boswell, J. Vac. Sci. Technol. B 24 (6) (2006) 2902-2906] with the <span class="hlt">brightness</span> of a LMIS and yet the ability to produce secondary ion yield enhancing species such as oxygen. However, to date a viable alternative has not been realized. The high <span class="hlt">brightness</span> and small virtual source size of the LMIS are advantageous for forming high resolution probes but a significant disadvantage when beam currents in excess of 100 nA are required, due to the effects of spherical aberration from the optical column. At these higher currents a source with a high angular intensity is optimal and in fact the relatively moderate <span class="hlt">brightness</span> of today's plasma ion sources prevail in this operating regime. Both the LMIS and Duoplasmatron suffer from a large axial energy spread resulting in further limitations when forming focused beams at the chromatic limit where the figure-of-merit is inversely proportional to the square of the energy spread. Also, both of these ion sources operate with a very limited range of ion species. This article reviews some of the latest developments and some future potential in this area of instrument development. Here we present an approach to source development that could lead to oxygen ion beam SIMS imaging with 10 nm resolution, using a 'broad area' RF gas phase ion source.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16283926','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16283926"><span><span class="hlt">Bright</span> green light treatment of depression for older adults [ISRCTN69400161].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Loving, Richard T; Kripke, Daniel F; Knickerbocker, Nancy C; Grandner, Michael A</p> <p>2005-11-09</p> <p><span class="hlt">Bright</span> white light has been successfully used for the treatment of depression. There is interest in identifying which spectral colors of light are the most efficient in the treatment of depression. It is theorized that green light could decrease the intensity duration of exposure needed. Late Wake Treatment (LWT), sleep deprivation for the last half of one night, is associated with rapid mood improvement which has been sustained by light treatment. Because spectral responsiveness may differ by age, we examined whether green light would provide efficient antidepressant treatment in an elder age group. We contrasted one hour of <span class="hlt">bright</span> green light (1,200 Lux) and one hour of dim red light placebo (<10 Lux) in a randomized treatment trial with depressed elders. Participants were <span class="hlt">observed</span> in their homes with mood scales, wrist actigraphy and light monitoring. On the day prior to beginning treatment, the participants self-administered LWT. The protocol was completed by 33 subjects who were 59 to 80 years old. Mood improved on average 23% for all subjects, but there were no significant statistical differences between treatment and placebo groups. There were negligible adverse reactions to the <span class="hlt">bright</span> green light, which was well tolerated. <span class="hlt">Bright</span> green light was not shown to have an antidepressant effect in the age group of this study, but a larger trial with brighter green light might be of value.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70033298','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70033298"><span>Titan's <span class="hlt">surface</span> from the Cassini RADAR radiometry data during SAR mode</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Paganelli, F.; Janssen, M.A.; Lopes, R.M.; Stofan, E.; Wall, S.D.; Lorenz, R.D.; Lunine, J.I.; Kirk, R.L.; Roth, L.; Elachi, C.</p> <p>2008-01-01</p> <p>We present initial results on the calibration and interpretation of the high-resolution radiometry data acquired during the Synthetic Aperture Radar (SAR) mode (SAR-radiometry) of the Cassini Radar Mapper during its first five flybys of Saturn's moon Titan. We construct maps of the <span class="hlt">brightness</span> temperature at the 2-cm wavelength coincident with SAR swath imaging. A preliminary radiometry calibration shows that <span class="hlt">brightness</span> temperature in these maps varies from 64 to 89 K. <span class="hlt">Surface</span> features and physical properties derived from the SAR-radiometry maps and SAR imaging are strongly correlated; in general, we find that <span class="hlt">surface</span> features with high radar reflectivity are associated with radiometrically cold regions, while <span class="hlt">surface</span> features with low radar reflectivity correlate with radiometrically warm regions. We examined scatterplots of the normalized radar cross-section ??0 versus <span class="hlt">brightness</span> temperature, outlining signatures that characterize various terrains and <span class="hlt">surface</span> features. The results indicate that volume scattering is important in many areas of Titan's <span class="hlt">surface</span>, particularly Xanadu, while other areas exhibit complex <span class="hlt">brightness</span> temperature variations consistent with variable slopes or <span class="hlt">surface</span> material and compositional properties. ?? 2007.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA259620','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA259620"><span><span class="hlt">Surface</span> <span class="hlt">Observation</span> Climatic Summaries for Nellis AFB, Nevada</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1992-05-01</p> <p>DISTRIBUTION OF THIS DOMWI! TO THE PUBLIC AT LARGE, OR BY THE DEFENSE TECHNICAL IMKNMTI1M CENTER (DTIC) TO THE NATIOAL T•ECICRL INFO TION SERVICE (NTS). JOSEPH...DOCUMENTS FORMERLY KNOW AS THE REVISED UNIFON4 StlMMRRY OF <span class="hlt">SURFACE</span> <span class="hlt">OBSERVATIONS</span> (RUSSW) AND THE LIMITED <span class="hlt">SURFACE</span> <span class="hlt">OBSERVATIONS</span> CLIMATIC SWSU.R (LISOCS...RECORD (POR). -SUMMARY OF DAY- (SOD) INFOEATIOR IS SUMMARIZED )FRO ALL AVAILABLE DATA IN THE OL-A, USARETJC CLIMATIC DATABASE. 14. SUBJECT TOM</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JEI....23b3011W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JEI....23b3011W"><span>Color constancy using <span class="hlt">bright</span>-neutral pixels</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Yanfang; Luo, Yupin</p> <p>2014-03-01</p> <p>An effective illuminant-estimation approach for color constancy is proposed. <span class="hlt">Bright</span> and near-neutral pixels are selected to jointly represent the illuminant color and utilized for illuminant estimation. To assess the representing capability of pixels, <span class="hlt">bright</span>-neutral strength (BNS) is proposed by combining pixel chroma and <span class="hlt">brightness</span>. Accordingly, a certain percentage of pixels with the largest BNS is selected to be the representative set. For every input image, a proper percentage value is determined via an iterative strategy by seeking the optimal color-corrected image. To compare various color-corrected images of an input image, image color-cast degree (ICCD) is devised using means and standard deviations of RGB channels. Experimental evaluation on standard real-world datasets validates the effectiveness of the proposed approach.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20070019869&hterms=ocean+salinity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Docean%2Bsalinity','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20070019869&hterms=ocean+salinity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Docean%2Bsalinity"><span>Microwave Remote Sensing Modeling of Ocean <span class="hlt">Surface</span> Salinity and Winds Using an Empirical Sea <span class="hlt">Surface</span> Spectrum</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yueh, Simon H.</p> <p>2004-01-01</p> <p>Active and passive microwave remote sensing techniques have been investigated for the remote sensing of ocean <span class="hlt">surface</span> wind and salinity. We revised an ocean <span class="hlt">surface</span> spectrum using the CMOD-5 geophysical model function (GMF) for the European Remote Sensing (ERS) C-band scatterometer and the Ku-band GMF for the NASA SeaWinds scatterometer. The predictions of microwave <span class="hlt">brightness</span> temperatures from this model agree well with satellite, aircraft and tower-based microwave radiometer data. This suggests that the impact of <span class="hlt">surface</span> roughness on microwave <span class="hlt">brightness</span> temperatures and radar scattering coefficients of sea <span class="hlt">surfaces</span> can be consistently characterized by a roughness spectrum, providing physical basis for using combined active and passive remote sensing techniques for ocean <span class="hlt">surface</span> wind and salinity remote sensing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21713499','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21713499"><span>Night-sky <span class="hlt">brightness</span> monitoring in Hong Kong: a city-wide light pollution assessment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pun, Chun Shing Jason; So, Chu Wing</p> <p>2012-04-01</p> <p>Results of the first comprehensive light pollution survey in Hong Kong are presented. The night-sky <span class="hlt">brightness</span> was measured and monitored around the city using a portable light-sensing device called the Sky Quality Meter over a 15-month period beginning in March 2008. A total of 1,957 data sets were taken at 199 distinct locations, including urban and rural sites covering all 18 Administrative Districts of Hong Kong. The survey shows that the environmental light pollution problem in Hong Kong is severe-the urban night skies (sky <span class="hlt">brightness</span> at 15.0 mag arcsec(- 2)) are on average ~ 100 times brighter than at the darkest rural sites (20.1 mag arcsec(- 2)), indicating that the high lighting densities in the densely populated residential and commercial areas lead to light pollution. In the worst polluted urban location studied, the night-sky at 13.2 mag arcsec(- 2) can be over 500 times brighter than the darkest sites in Hong Kong. The <span class="hlt">observed</span> night-sky <span class="hlt">brightness</span> is found to be affected by human factors such as land utilization and population density of the <span class="hlt">observation</span> sites, together with meteorological and/or environmental factors. Moreover, earlier night skies (at 9:30 p.m. local time) are generally brighter than later time (at 11:30 p.m.), which can be attributed to some public and commercial lightings being turned off later at night. On the other hand, no concrete relationship between the <span class="hlt">observed</span> sky <span class="hlt">brightness</span> and air pollutant concentrations could be established with the limited survey sampling. Results from this survey will serve as an important database for the public to assess whether new rules and regulations are necessary to control the use of outdoor lightings in Hong Kong.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040081177','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040081177"><span>The Gamma-Ray <span class="hlt">Bright</span> BL Lac Object RX J1211+2242</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Beckmann, V.; Favre, P.; Tavecchio, F.; Bussien, T.; Fliri, J.; Wolter, A.</p> <p>2004-01-01</p> <p>RX J1211+2242 is an optically faint (B approximately equal to 19.2mag) but X-ray <span class="hlt">bright</span> (f2-10kev = 5 x l0(exp -12)erg per square centimeter per second) AGN, which has been shown to be a BL Lac object at redshift z = 0.455. The ROSAT X-ray, Calar Alto optical, and NVSS radio data suggest that the peak of the synchrotron emission of this object is at energies as high as several keV. BeppoSAX <span class="hlt">observations</span> have been carried out simultaneously with optical <span class="hlt">observations</span> in order to extend the coverage to higher energies. The new data indeed indicate a turn-over in the 2 - 10keV energy region. We propose that RX J1211+2242 is the counterpart of the unidentified EGRET source 3EG J1212+2304, making it a gamma-ray emitter with properties similar to, for example, Markarian 501 in its <span class="hlt">bright</span> state, though being at a much larger distance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20070009994','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20070009994"><span>Calcium Sulfate in Atacama Desert Basalt: A Possible Analog for <span class="hlt">Bright</span> Material in Adirondack Basalt, Gusev Crater</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sutter, B.; Golden, D. C.; Amundson, R.; Chong-Diaz, G.; Ming, D. W.</p> <p>2007-01-01</p> <p>The Atacama Desert in northern Chile is one of the driest deserts on Earth (< 2mm/y). The hyper-arid conditions allow extraordinary accumulations of sulfates, chlorides, and nitrates in Atacama soils. Examining salt accumulations in the Atacama may assist understanding salt accumulations on Mars. Recent work examining sulfate soils on basalt parent material <span class="hlt">observed</span> white material in the interior vesicles of <span class="hlt">surface</span> basalt. This is strikingly similar to the <span class="hlt">bright</span>-white material present in veins and vesicles of the Adirondack basalt rocks at Gusev Crater which are presumed to consist of S, Cl, and/or Br. The abundance of soil gypsum/anhydrite in the area of the Atacama basalt suggested that the white material consisted of calcium sulfate (Ca-SO4) which was later confirmed by SEM/EDS analysis. This work examines the Ca-SO4 of Atacama basalt in an effort to provide insight into the possible nature of the <span class="hlt">bright</span> material in the Adirondack basalt of Gusev Crater. The objectives of this work are to (i) discuss variations in Ca-SO4 crystal morphology in the vesicles and (ii) examine the Ca-SO4 interaction(s) with the basalt interior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040112701&hterms=pacemaker&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dpacemaker','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040112701&hterms=pacemaker&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dpacemaker"><span>Dynamic resetting of the human circadian pacemaker by intermittent <span class="hlt">bright</span> light</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rimmer, D. W.; Boivin, D. B.; Shanahan, T. L.; Kronauer, R. E.; Duffy, J. F.; Czeisler, C. A.</p> <p>2000-01-01</p> <p>In humans, experimental studies of circadian resetting typically have been limited to lengthy episodes of exposure to continuous <span class="hlt">bright</span> light. To evaluate the time course of the human endogenous circadian pacemaker's resetting response to brief episodes of intermittent <span class="hlt">bright</span> light, we studied 16 subjects assigned to one of two intermittent lighting conditions in which the subjects were presented with intermittent episodes of <span class="hlt">bright</span>-light exposure at 25- or 90-min intervals. The effective duration of <span class="hlt">bright</span>-light exposure was 31% or 63% compared with a continuous 5-h <span class="hlt">bright</span>-light stimulus. Exposure to intermittent <span class="hlt">bright</span> light elicited almost as great a resetting response compared with 5 h of continuous <span class="hlt">bright</span> light. We conclude that exposure to intermittent <span class="hlt">bright</span> light produces robust phase shifts of the endogenous circadian pacemaker. Furthermore, these results demonstrate that humans, like other species, exhibit an enhanced sensitivity to the initial minutes of <span class="hlt">bright</span>-light exposure.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19850026410','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850026410"><span>Quasiperiodic oscillations in <span class="hlt">bright</span> galactic-bulge X-ray sources</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lamb, F. K.; Shibazaki, N.; Alpar, M. A.; Shaham, J.</p> <p>1985-01-01</p> <p>Quasiperiodic oscillations with frequencies in the range 5-50 Hz have recently been discovered in X-rays from two <span class="hlt">bright</span> galactic-bulge sources and Sco X-1. These sources are weakly magnetic neutron stars accreting from disks which the plasma is clumped. The interaction of the magnetosphere with clumps in the inner disk causes oscillations in the X-ray flux with many of the properties <span class="hlt">observed</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017DPS....4911031G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017DPS....4911031G"><span>Exploring Extreme Retro-reflection by Asteroids Using Las Cumbres Observatory Robotic Telescope <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Goguen, Jay D.; Bauer, James M.</p> <p>2017-10-01</p> <p>The reflectivity of solar system <span class="hlt">surfaces</span> ‘spikes’ sharply when the Sun is less than 1 degree from directly behind the <span class="hlt">observer</span>. The Galileo spacecraft measured the reflectivity of part of Europa’s <span class="hlt">surface</span> to increase by as much as a factor of 8 as the <span class="hlt">observer</span> moves from 5 degrees to the exact backscattering direction! One mechanism explains this spike as coherent light scattering that occurs only close to this unique retro-reflection geometry. Due to the tight linear alignment of the target, <span class="hlt">observer</span> and Sun required to measure the peak <span class="hlt">brightness</span> of the spike, accurate and complete measurements of the amplitude and decay of the spike exist for only a few targets. We used the unique capabilities of the automated Las Cumbres Observatory global telescope network (LCO) to systematically measure this extreme opposition surge for 60+ asteroids sampling a variety of taxonomic classes in the Bus/DeMeo taxonomy.Each asteroid was <span class="hlt">observed</span> in the SDSS r’ and g’ filters during the ~8 hour interval when it passes within ~0.1 deg of the point opposite the Sun on the sky. Supporting <span class="hlt">observations</span> of each asteroid with LCO collected over ~50 days measure asteroid rotation and phase angle <span class="hlt">brightness</span> changes to enable accurate characterization of the retro-reflection spike. This data set vastly increases the number and variety of the <span class="hlt">surfaces</span> characterized at such small phase angles compared to existing asteroid data. We examine how the spike characteristics vary with <span class="hlt">surface</span> composition, albedo, and wavelength providing new constraints on physical models of this ubiquitous yet poorly understood phenomenon.Analysis and modeling of these measurements will advance our understanding of the physical mechanism responsible for this enhanced retro-reflection thereby improving our ability to characterize these <span class="hlt">surfaces</span> from remote <span class="hlt">observations</span>. The ability to infer <span class="hlt">surface</span> physical properties from remote sensing data is a key capability for future asteroid missions, manned</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017yCat..18400021T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017yCat..18400021T"><span>VizieR Online Data Catalog: FIR data of IR-<span class="hlt">bright</span> dust-obscured galaxies (DOGs) (Toba+, 2017)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Toba, Y.; Nagao, T.; Wang, W.-H.; Matsuhara, H.; Akiyama, M.; Goto, T.; Koyama, Y.; Ohyama, Y.; Yamamura, I.</p> <p>2017-11-01</p> <p>We investigate the star-forming activity of a sample of infrared (IR)-<span class="hlt">bright</span> dust-obscured galaxies (DOGs) that show an extreme red color in the optical and IR regime, (i-[22])AB>7.0. Combining an IR-<span class="hlt">bright</span> DOG sample with the flux at 22μm>3.8mJy discovered by Toba & Nagao (2016ApJ...820...46T) with the IRAS faint source catalog version 2 and AKARI far-IR (FIR) all-sky survey <span class="hlt">bright</span> source catalog version 2, we selected 109 DOGs with FIR data. For a subsample of seven IR-<span class="hlt">bright</span> DOGs with spectroscopic redshifts (0.07<z<1.0) that were obtained from the literature, we estimated their IR luminosity, star formation rate (SFR), and stellar mass based on the spectral energy distribution fitting. We found that (1) the WISE 22μm luminosity at the <span class="hlt">observed</span> frame is a good indicator of IR luminosity for IR-<span class="hlt">bright</span> DOGs and (2) the contribution of the active galactic nucleus to IR luminosity increases with IR luminosity. By comparing the stellar mass and SFR relation for our DOG sample and the literature, we found that most of the IR-<span class="hlt">bright</span> DOGs lie significantly above the main sequence of star-forming galaxies at similar redshift, indicating that the majority of IRAS- or AKARI-detected IR-<span class="hlt">bright</span> DOGs are starburst galaxies. (1 data file).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001SPIE.4283..586B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001SPIE.4283..586B"><span>Numerical simulations of novel high-power high-<span class="hlt">brightness</span> diode laser structures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Boucke, Konstantin; Rogg, Joseph; Kelemen, Marc T.; Poprawe, Reinhart; Weimann, Guenter</p> <p>2001-07-01</p> <p>One of the key topics in today's semiconductor laser development activities is to increase the <span class="hlt">brightness</span> of high-power diode lasers. Although structures showing an increased <span class="hlt">brightness</span> have been developed specific draw-backs of these structures lead to a still strong demand for investigation of alternative concepts. Especially for the investigation of basically novel structures easy-to-use and fast simulation tools are essential to avoid unnecessary, cost and time consuming experiments. A diode laser simulation tool based on finite difference representations of the Helmholtz equation in 'wide-angle' approximation and the carrier diffusion equation has been developed. An optimized numerical algorithm leads to short execution times of a few seconds per resonator round-trip on a standard PC. After each round-trip characteristics like optical output power, beam profile and beam parameters are calculated. A graphical user interface allows online monitoring of the simulation results. The simulation tool is used to investigate a novel high-power, high-<span class="hlt">brightness</span> diode laser structure, the so-called 'Z-Structure'. In this structure an increased <span class="hlt">brightness</span> is achieved by reducing the divergency angle of the beam by angular filtering: The round trip path of the beam is two times folded using internal total reflection at <span class="hlt">surfaces</span> defined by a small index step in the semiconductor material, forming a stretched 'Z'. The sharp decrease of the reflectivity for angles of incidence above the angle of total reflection leads to a narrowing of the angular spectrum of the beam. The simulations of the 'Z-Structure' indicate an increase of the beam quality by a factor of five to ten compared to standard broad-area lasers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19840016444','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19840016444"><span>An X-ray survey of variable radio <span class="hlt">bright</span> quasars</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Henriksen, M. J.; Marshall, F. E.; Mushotzky, R. F.</p> <p>1984-01-01</p> <p>A sample consisting primarily of radio <span class="hlt">bright</span> quasars was <span class="hlt">observed</span> in X-rays with the Einstein Observatory for times ranging from 1500 to 5000 seconds. Detected sources had luminosities ranging from 0.2 to 41.0 x 10 to the 45th power ergs/sec in the 0.5 to 4.5 keV band. Three of the fourteen objects which were reobserved showed flux increases greater than a factor of two on a time scale greater than six months. No variability was detected during the individual <span class="hlt">observations</span>. The optical and X-ray luminosities are correlated, which suggests a common origin. However, the relationship (L sub x is approximately L sub op to the (.89 + or - .15)) found for historic radio variables may be significantly different than that reported for other radio <span class="hlt">bright</span> sources. Some of the <span class="hlt">observed</span> X-ray fluxes were substantially below the predicted self-Compton flux, assuming incoherent synchrotron emission and using VLBI results to constrain the size of the emission region, which suggests relativistic expansion in these sources. Normal CIV emission in two of the sources with an overpredicted Compton component suggests that although they, like BL Lac objects, have highly relativistic material apparently moving at small angle to the line of sight, they have a smaller fraction of the continuum component in the beam.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20050243590','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20050243590"><span>Chandra X-ray <span class="hlt">Observation</span> of a Mature Cloud-Shock Interaction in the <span class="hlt">Bright</span> Eastern Knot of Puppis A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hwang, Una; Flanagan, Kathryn A.; Petre, Robert</p> <p>2005-01-01</p> <p>We present Chandra X-ray images and spectra of the most prominent cloud-shock interaction region in the Puppis A supernova remnant. The <span class="hlt">Bright</span> Eastern Knot (BEK) has two main morphological components: (1) a <span class="hlt">bright</span> compact knot that lies directly behind the apex of an indentation in the eastern X-ray boundary and (2) lying 1 westward behind the shock, a curved vertical structure (bar) that is separated from a smaller <span class="hlt">bright</span> cloud (cap) by faint diffuse emission. Based on hardness images and spectra, we identify the bar and cap as a single shocked interstellar cloud. Its morphology strongly resembles the "voided sphere" structures seen at late times in Klein et al. experimental simulat.ions of cloud-shock interactions, when the crushing of the cloud by shear instabilities is well underway. We infer an intera.ction time of roughly cloud-crushing timescales, which translates to 2000-4000 years, based on the X-ray temperature, physical size, and estimated expansion of the shocked cloud. This is the first X-ray identified example of a cloud-shock interaction in this advanced phase. Closer t o the shock front, the X-ray emission of the compact knot in the eastern part of the BEK region implies a recent interaction with relatively denser gas, some of which lies in front of the remnant. The complex spatial relationship of the X-ray emission of the compact knot to optical [O III] emission suggests that there are multiple cloud interactions occurring along the line of sight.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014MNRAS.441..910R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014MNRAS.441..910R"><span>MOST detects corotating <span class="hlt">bright</span> spots on the mid-O-type giant ξ Persei</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ramiaramanantsoa, Tahina; Moffat, Anthony F. J.; Chené, André-Nicolas; Richardson, Noel D.; Henrichs, Huib F.; Desforges, Sébastien; Antoci, Victoria; Rowe, Jason F.; Matthews, Jaymie M.; Kuschnig, Rainer; Weiss, Werner W.; Sasselov, Dimitar; Rucinski, Slavek M.; Guenther, David B.</p> <p>2014-06-01</p> <p>We have used the MOST (Microvariability and Oscillations of STars) microsatellite to obtain four weeks of contiguous high-precision broad-band visual photometry of the O7.5III(n)((f)) star ξ Persei in 2011 November. This star is well known from previous work to show prominent DACs (discrete absorption components) on time-scales of about 2 d from UV spectroscopy and non-radial pulsation with one (l = 3) p-mode oscillation with a period of 3.5 h from optical spectroscopy. Our MOST-orbit (101.4 min) binned photometry fails to reveal any periodic light variations above the 0.1 mmag 3σ noise level for periods of a few hours, while several prominent Fourier peaks emerge at the 1 mmag level in the two-day period range. These longer period variations are unlikely due to pulsations, including gravity modes. From our simulations based upon a simple spot model, we deduce that we are seeing the photometric modulation of several corotating <span class="hlt">bright</span> spots on the stellar <span class="hlt">surface</span>. In our model, the starting times (random) and lifetimes (up to several rotations) vary from one spot to another yet all spots rotate at the same period of 4.18 d, the best-estimated rotation period of the star. This is the first convincing reported case of corotating <span class="hlt">bright</span> spots on an O star, with important implications for drivers of the DACs (resulting from corotating interaction regions) with possible <span class="hlt">bright</span>-spot generation via a breakout at the <span class="hlt">surface</span> of a global magnetic field generated by a subsurface convection zone.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AMT....10.3947D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AMT....10.3947D"><span>Long-term <span class="hlt">observations</span> minus background monitoring of ground-based <span class="hlt">brightness</span> temperatures from a microwave radiometer network</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>De Angelis, Francesco; Cimini, Domenico; Löhnert, Ulrich; Caumont, Olivier; Haefele, Alexander; Pospichal, Bernhard; Martinet, Pauline; Navas-Guzmán, Francisco; Klein-Baltink, Henk; Dupont, Jean-Charles; Hocking, James</p> <p>2017-10-01</p> <p>Ground-based microwave radiometers (MWRs) offer the capability to provide continuous, high-temporal-resolution <span class="hlt">observations</span> of the atmospheric thermodynamic state in the planetary boundary layer (PBL) with low maintenance. This makes MWR an ideal instrument to supplement radiosonde and satellite <span class="hlt">observations</span> when initializing numerical weather prediction (NWP) models through data assimilation. State-of-the-art data assimilation systems (e.g. variational schemes) require an accurate representation of the differences between model (background) and <span class="hlt">observations</span>, which are then weighted by their respective errors to provide the best analysis of the true atmospheric state. In this perspective, one source of information is contained in the statistics of the differences between <span class="hlt">observations</span> and their background counterparts (O-B). Monitoring of O-B statistics is crucial to detect and remove systematic errors coming from the measurements, the <span class="hlt">observation</span> operator, and/or the NWP model. This work illustrates a 1-year O-B analysis for MWR <span class="hlt">observations</span> in clear-sky conditions for an European-wide network of six MWRs. <span class="hlt">Observations</span> include MWR <span class="hlt">brightness</span> temperatures (TB) measured by the two most common types of MWR instruments. Background profiles are extracted from the French convective-scale model AROME-France before being converted into TB. The <span class="hlt">observation</span> operator used to map atmospheric profiles into TB is the fast radiative transfer model RTTOV-gb. It is shown that O-B monitoring can effectively detect instrument malfunctions. O-B statistics (bias, standard deviation, and root mean square) for water vapour channels (22.24-30.0 GHz) are quite consistent for all the instrumental sites, decreasing from the 22.24 GHz line centre ( ˜ 2-2.5 K) towards the high-frequency wing ( ˜ 0.8-1.3 K). Statistics for zenith and lower-elevation <span class="hlt">observations</span> show a similar trend, though values increase with increasing air mass. O-B statistics for temperature channels show different</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ATel11615....1F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ATel11615....1F"><span><span class="hlt">Bright</span> ZTF transients</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fremling, C.; Kulkarni, S. R.; Taggart, K.; Perley, D.</p> <p>2018-05-01</p> <p>As a part of ongoing commissioning of the Zwicky Transient Facility (ZTF; ATel #11266) Alert Infrastructure, here we report <span class="hlt">bright</span> probable supernovae identified in the raw alert stream resulting from the public ZTF Northern Sky Survey ("Celestial Cinematagrophy"; see Bellm & Kulkarni, Nature Astronomy 1, 71, 2017).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA31A2571S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA31A2571S"><span>On the relation between GNSS phase scintillation and auroral <span class="hlt">brightness</span> around satellite's IPP</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Spanswick, E.; Mushini, S. C.; Skone, S.; Donovan, E.</p> <p>2017-12-01</p> <p>Aurora occurs in different well-known morphologies, or types, including arcs and patchy-pulsating aurora (PPA). Previous <span class="hlt">observational</span> studies have demonstrated that global navigation satellite system (GNSS) signals transiting the ionosphere in regions of aurora can contain varying levels of scintillation. These scintillations are often attributed to the ionospheric disturbances associated with auroral precipitation, which in extreme cases can affect the accuracy of these systems. One question that remains unanswered is whether a satellite's line of sight transmission through the aurora is a sufficient condition for signal scintillation. Previous studies have used "level" or "strength" of auroral emission as a proxy indicator for scintillation using limited datasets. In general, these results are mixed and inconclusive. In this study, we use a large data set (700 Auroral arc events) to statistically study the relationship between aurora and scintillation of GPS signals. This is one of the largest datasets used in this type of studies. We utilize the THEMIS (Time History of Events and Macroscale Interactions during Substorms) All-Sky Imagers (ASIs) located at Fort Smith (59.9 N, 248.1 E geog.) and Gillam (56.5 N, 265.4 E geog.), Canada. Corresponding GPS data were obtained from CHAIN (Canadian High Arctic Ionospheric Network) GPS receivers collocated with the ASIs. These GPS receivers are custom made receivers capable of providing high rate GPS signal power and phase <span class="hlt">observations</span> as well as scintillation indices. To obtain information how aurora is affecting the signal, <span class="hlt">brightness</span> around satellite's Ionospheric Pierce Point (IPP) was calculated and correlated with sigma phi from the satellite's signal. A very low correlation of 0.003 was <span class="hlt">observed</span> between them. Correlation between the rate of change of <span class="hlt">brightness</span> around the satellite's IPP and sigma phi was also calculated and a correlation coefficient of 0.7 was <span class="hlt">observed</span> between them. These results indicate that GPS</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22061121','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22061121"><span>A Zn-porphyrin complex contributes to <span class="hlt">bright</span> red color in Parma ham.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wakamatsu, J; Nishimura, T; Hattori, A</p> <p>2004-05-01</p> <p>The Italian traditional dry-cured ham (Parma ham) shows a stable <span class="hlt">bright</span> red color that is achieved without the use of nitrite and/or nitrate. In this study we examined the pigment spectroscopically, fluoroscopically and by using HPLC and ESI-HR-MASS analysis. Porphyrin derivative other than acid hematin were contained in the HCl-containing acetone extract from Parma ham. A strong fluorescence peak at 588 nm and a weak fluorescence peak at 641 nm were <span class="hlt">observed</span>. By HPLC analysis the acetone extract of Parma ham was <span class="hlt">observed</span> at the single peak, which eluted at the same time as Zn-protoporphyrin IX and emitted fluorescence. The results of ESI-HR-MS analysis showed both agreement with the molecular weight of Zn-protoporphyrin IX and the characteristic isotope pattern caused by Zn isotopes. These results suggest that the <span class="hlt">bright</span> red color in Parma ham is caused by Zn-protoporphyrin IX.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130012041','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130012041"><span>Hubble Space Telescope <span class="hlt">Observations</span> of the HD 202628 Debris Disk</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Krist, John E.; Stapelfeldt, Karl R.; Bryden, Geoffrey; Plavchan, Peter</p> <p>2012-01-01</p> <p>A ring-shaped debris disk around the G2V star HD 202628 (d = 24.4 pc) was imaged in scattered light at visible wavelengths using the coronagraphic mode of the Space Telescope Imaging Spectrograph on the Hubble Space Telescope. The ring is inclined by approx.64deg from face-on, based on the apparent major/minor axis ratio, with the major axis aligned along PA = 130deg. It has inner and outer radii (> 50% maximum <span class="hlt">surface</span> <span class="hlt">brightness</span>) of 139 AU and 193 AU in the northwest ansae and 161 AU and 223 AU in the southeast ((Delta)r/r approx. = 0.4). The maximum visible radial extent is approx. 254 AU. With a mean <span class="hlt">surface</span> <span class="hlt">brightnesses</span> of V approx. = 24 mag arcsec.(sup -2), this is the faintest debris disk <span class="hlt">observed</span> to date in reflected light. The center of the ring appears offset from the star by approx.28 AU (deprojected). An ellipse fit to the inner edge has an eccentricity of 0.18 and a = 158 AU. This offset, along with the relatively sharp inner edge of the ring, suggests the influence of a planetary-mass companion. There is a strong similarity with the debris ring around Fomalhaut, though HD 202628 is a more mature star with an estimated age of about 2 Gyr. We also provide <span class="hlt">surface</span> <span class="hlt">brightness</span> limits for nine other stars in our study with strong Spitzer excesses around which no debris disks were detected in scattered light (HD 377, HD 7590, HD 38858, HD 45184, HD 73350, HD 135599, HD 145229, HD 187897, and HD 201219).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AJ....144...45K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AJ....144...45K"><span>Hubble Space Telescope <span class="hlt">Observations</span> of the HD 202628 Debris Disk</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krist, John E.; Stapelfeldt, Karl R.; Bryden, Geoffrey; Plavchan, Peter</p> <p>2012-08-01</p> <p>A ring-shaped debris disk around the G2V star HD 202628 (d = 24.4 pc) was imaged in scattered light at visible wavelengths using the coronagraphic mode of the Space Telescope Imaging Spectrograph on the Hubble Space Telescope. The ring is inclined by ~64° from face-on, based on the apparent major/minor axis ratio, with the major axis aligned along P.A. = 130°. It has inner and outer radii (>50% maximum <span class="hlt">surface</span> <span class="hlt">brightness</span>) of 139 AU and 193 AU in the northwest ansae and 161 AU and 223 AU in the southeast (Δr/r ≈ 0.4). The maximum visible radial extent is ~254 AU. With mean <span class="hlt">surface</span> <span class="hlt">brightness</span> of V ≈ 24 mag arcsec-2, this is the faintest debris disk <span class="hlt">observed</span> to date in reflected light. The center of the ring appears offset from the star by ~28 AU (deprojected). An ellipse fit to the inner edge has an eccentricity of 0.18 and a = 158 AU. This offset, along with the relatively sharp inner edge of the ring, suggests the influence of a planetary-mass companion. There is a strong similarity with the debris ring around Fomalhaut, though HD 202628 is a more mature star with an estimated age of about 2 Gyr. We also provide <span class="hlt">surface</span> <span class="hlt">brightness</span> limits for nine other stars in our study with strong Spitzer excesses around which no debris disks were detected in scattered light (HD 377, HD 7590, HD 38858, HD 45184, HD 73350, HD 135599, HD 145229, HD 187897, and HD 201219).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20150005800&hterms=Ripple+labs&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DRipple%2Blabs','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20150005800&hterms=Ripple+labs&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DRipple%2Blabs"><span>Sea <span class="hlt">Surface</span> Salinity and Wind Retrieval Algorithm Using Combined Passive-Active L-Band Microwave Data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yueh, Simon H.; Chaubell, Mario J.</p> <p>2011-01-01</p> <p>Aquarius is a combined passive/active L-band microwave instrument developed to map the salinity field at the <span class="hlt">surface</span> of the ocean from space. The data will support studies of the coupling between ocean circulation, the global water cycle, and climate. The primary science objective of this mission is to monitor the seasonal and interannual variation of the large scale features of the <span class="hlt">surface</span> salinity field in the open ocean with a spatial resolution of 150 kilometers and a retrieval accuracy of 0.2 practical salinity units globally on a monthly basis. The measurement principle is based on the response of the L-band (1.413 gigahertz) sea <span class="hlt">surface</span> <span class="hlt">brightness</span> temperatures (T (sub B)) to sea <span class="hlt">surface</span> salinity. To achieve the required 0.2 practical salinity units accuracy, the impact of sea <span class="hlt">surface</span> roughness (e.g. wind-generated ripples and waves) along with several factors on the <span class="hlt">observed</span> <span class="hlt">brightness</span> temperature has to be corrected to better than a few tenths of a degree Kelvin. To the end, Aquarius includes a scatterometer to help correct for this <span class="hlt">surface</span> roughness effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10228E..0PK','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10228E..0PK"><span><span class="hlt">Bright</span>-dark rogue wave in mode-locked fibre laser (Conference Presentation)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kbashi, Hani; Kolpakov, Stanislav; Martinez, Amós; Mou, Chengbo; Sergeyev, Sergey V.</p> <p>2017-05-01</p> <p><span class="hlt">Bright</span>-Dark Rogue Wave in Mode-Locked Fibre Laser Hani Kbashi1*, Amos Martinez1, S. A. Kolpakov1, Chengbo Mou, Alex Rozhin1, Sergey V. Sergeyev1 1Aston Institute of Photonic Technologies, School of Engineering and Applied Science Aston University, Birmingham, B4 7ET, UK kbashihj@aston.ac.uk , 0044 755 3534 388 Keywords: Optical rogue wave, <span class="hlt">Bright</span>-Dark rogue wave, rogue wave, mode-locked fiber laser, polarization instability. Abstract: Rogue waves (RWs) are statistically rare localized waves with high amplitude that suddenly appear and disappear in oceans, water tanks, and optical systems [1]. The investigation of these events in optics, optical rogue waves, is of interest for both fundamental research and applied science. Recently, we have shown that the adjustment of the in-cavity birefringence and pump polarization leads to emerge optical RW events [2-4]. Here, we report the first experimental <span class="hlt">observation</span> of vector <span class="hlt">bright</span>-dark RWs in an erbium-doped stretched pulse mode-locked fiber laser. The change of induced in-cavity birefringence provides an opportunity to <span class="hlt">observe</span> RW events at pump power is a little higher than the lasing threshold. Polarization instabilities in the laser cavity result in the coupling between two orthogonal linearly polarized components leading to the emergence of <span class="hlt">bright</span>-dark RWs. The <span class="hlt">observed</span> clusters belongs to the class of slow optical RWs because their lifetime is of order of a thousand of laser cavity roundtrip periods. References: 1. D. R. Solli, C. Ropers, P. Koonath,and B. Jalali, Optical rogue waves," Nature, 450, 1054-1057, 2007. 2. S. V. Sergeyev, S. A. Kolpakov, C. Mou, G. Jacobsen, S. Popov, and V. Kalashnikov, "Slow deterministic vector rogue waves," Proc. SPIE 9732, 97320K (2016). 3. S. A. Kolpakov, H. Kbashi, and S. V. Sergeyev, "Dynamics of vector rogue waves in a fiber laser with a ring cavity," Optica, 3, 8, 870, (2016). 5. S. Kolpakov, H. Kbashi, and S. Sergeyev, "Slow optical rogue waves in a unidirectional fiber laser</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.2015H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.2015H"><span>Simultaneous Assimilation of AMSR-E <span class="hlt">Brightness</span> Temperature and MODIS LST to Improve Soil Moisture with Dual Ensemble Kalman Smoother</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huang, Chunlin; Chen, Weijin; Wang, Weizhen; Gu, Juan</p> <p>2017-04-01</p> <p>Uncertainties in model parameters can easily cause systematic differences between model states and <span class="hlt">observations</span> from ground or satellites, which significantly affect the accuracy of soil moisture estimation in data assimilation systems. In this paper, a novel soil moisture assimilation scheme is developed to simultaneously assimilate AMSR-E <span class="hlt">brightness</span> temperature (TB) and MODIS Land <span class="hlt">Surface</span> Temperature (LST), which can correct model bias by simultaneously updating model states and parameters with dual ensemble Kalman filter (DEnKS). The Common Land Model (CoLM) and a Q-h Radiative Transfer Model (RTM) are adopted as model operator and <span class="hlt">observation</span> operator, respectively. The assimilation experiment is conducted in Naqu, Tibet Plateau, from May 31 to September 27, 2011. Compared with in-situ measurements, the accuracy of soil moisture estimation is tremendously improved in terms of a variety of scales. The updated soil temperature by assimilating MODIS LST as input of RTM can reduce the differences between the simulated and <span class="hlt">observed</span> <span class="hlt">brightness</span> temperatures to a certain degree, which helps to improve the estimation of soil moisture and model parameters. The updated parameters show large discrepancy with the default ones and the former effectively reduces the states bias of CoLM. Results demonstrate the potential of assimilating both microwave TB and MODIS LST to improve the estimation of soil moisture and related parameters. Furthermore, this study also indicates that the developed scheme is an effective soil moisture downscaling approach for coarse-scale microwave TB.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016Icar..268...50L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016Icar..268...50L"><span>Probing Pluto's underworld: Ice temperatures from microwave radiometry decoupled from <span class="hlt">surface</span> conditions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Leyrat, Cedric; Lorenz, Ralph D.; Le Gall, Alice</p> <p>2016-04-01</p> <p>Present models admit a wide range of 2015 <span class="hlt">surface</span> conditions at Pluto and Charon, where the atmospheric pressure may undergo dramatic seasonal variation and for which measurements are imminent from the New Horizons mission. One anticipated <span class="hlt">observation</span> is the microwave <span class="hlt">brightness</span> temperature, heretofore anticipated as indicating <span class="hlt">surface</span> conditions relevant to <span class="hlt">surface</span>-atmosphere equilibrium. However, drawing on recent experience with Cassini <span class="hlt">observations</span> at Iapetus and Titan, we call attention to the large electrical skin depth of outer Solar System materials such as methane, nitrogen or water ice, such that this <span class="hlt">observation</span> may indicate temperatures averaged over depths of several or tens of meters beneath the <span class="hlt">surface</span>. Using a seasonally-forced thermal model to determine microwave emission we predict that the southern hemisphere <span class="hlt">observations</span> (in polar night) of New Horizons in July 2015 will suggest effective temperatures of ∼40 K, reflecting deep heat buried over the last century of summer, even if the atmospheric pressure suggests that the <span class="hlt">surface</span> nitrogen frost point may be much lower.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=phi&pg=4&id=EJ838375','ERIC'); return false;" href="https://eric.ed.gov/?q=phi&pg=4&id=EJ838375"><span>Does Stevens's Power Law for <span class="hlt">Brightness</span> Extend to Perceptual <span class="hlt">Brightness</span> Averaging?</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Bauer, Ben</p> <p>2009-01-01</p> <p>Stevens's power law ([Psi][infinity][Phi][beta]) captures the relationship between physical ([Phi]) and perceived ([Psi]) magnitude for many stimulus continua (e.g., luminance and <span class="hlt">brightness</span>, weight and heaviness, area and size). The exponent ([beta]) indicates whether perceptual magnitude grows more slowly than physical magnitude ([beta] less…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22342269-tidally-modulated-eruptions-enceladus-cassini-iss-observations-models','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22342269-tidally-modulated-eruptions-enceladus-cassini-iss-observations-models"><span>Tidally modulated eruptions on Enceladus: Cassini ISS <span class="hlt">observations</span> and models</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Nimmo, Francis; Porco, Carolyn; Mitchell, Colin, E-mail: carolyn@ciclops.org</p> <p>2014-09-01</p> <p>We use images acquired by the Cassini Imaging Science Subsystem (ISS) to investigate the temporal variation of the <span class="hlt">brightness</span> and height of the south polar plume of Enceladus. The plume's <span class="hlt">brightness</span> peaks around the moon's apoapse, but with no systematic variation in scale height with either plume <span class="hlt">brightness</span> or Enceladus' orbital position. We compare our results, both alone and supplemented with Cassini near-infrared <span class="hlt">observations</span>, with predictions obtained from models in which tidal stresses are the principal control of the eruptive behavior. There are three main ways of explaining the <span class="hlt">observations</span>: (1) the activity is controlled by right-lateral strike slip motion;more » (2) the activity is driven by eccentricity tides with an apparent time delay of about 5 hr; (3) the activity is driven by eccentricity tides plus a 1:1 physical libration with an amplitude of about 0.°8 (3.5 km). The second hypothesis might imply either a delayed eruptive response, or a dissipative, viscoelastic interior. The third hypothesis requires a libration amplitude an order of magnitude larger than predicted for a solid Enceladus. While we cannot currently exclude any of these hypotheses, the third, which is plausible for an Enceladus with a subsurface ocean, is testable by using repeat imaging of the moon's <span class="hlt">surface</span>. A dissipative interior suggests that a regional background heat source should be detectable. The lack of a systematic variation in plume scale height, despite the large variations in plume <span class="hlt">brightness</span>, is plausibly the result of supersonic flow; the details of the eruption process are yet to be understood.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.473.4653K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.473.4653K"><span>Studying the ICM in clusters of galaxies via <span class="hlt">surface</span> <span class="hlt">brightness</span> fluctuations of the cosmic X-ray background</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kolodzig, Alexander; Gilfanov, Marat; Hütsi, Gert; Sunyaev, Rashid</p> <p>2018-02-01</p> <p>We study <span class="hlt">surface</span> <span class="hlt">brightness</span> fluctuations of the cosmic X-ray background (CXB) using Chandra data of XBOOTES. After masking out resolved sources we compute the power spectrum of fluctuations of the unresolved CXB for angular scales from {≈ } 2 arcsec to ≈3°. The non-trivial large-scale structure (LSS) signal dominates over the shot noise of unresolved point sources on angular scales above {˜ } 1 arcmin and is produced mainly by the intracluster medium (ICM) of unresolved clusters and groups of galaxies, as shown in our previous publication. The shot-noise-subtracted power spectrum of CXB fluctuations has a power-law shape with the slope of Γ = 0.96 ± 0.06. Their energy spectrum is well described by the redshifted emission spectrum of optically thin plasma with the best-fitting temperature of T ≈ 1.3 keV and the best-fitting redshift of z ≈ 0.40. These numbers are in good agreement with theoretical expectations based on the X-ray luminosity function and scaling relations of clusters. From these values we estimate the typical mass and luminosity of the objects responsible for CXB fluctuations, M500 ∼ 1013.6 M⊙ h-1 and L0.5-2.0 keV ∼ 1042.5 erg s-1. On the other hand, the flux-weighted mean temperature and redshift of resolved clusters are T ≈ 2.4 keV and z ≈ 0.23 confirming that fluctuations of unresolved CXB are caused by cooler (i.e. less massive) and more distant clusters, as expected. We show that the power spectrum shape is sensitive to the ICM structure all the way to the outskirts, out to ∼few × R500. We also searched for possible contribution of the warm-hot intergalactic medium (WHIM) to the <span class="hlt">observed</span> CXB fluctuations. Our results underline the significant diagnostic potential of the CXB fluctuation analysis in studying the ICM structure in clusters.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_25 --> <div class="footer-extlink text-muted" style="margin-bottom:1rem; text-align:center;">Some links on this page may take you to non-federal websites. Their policies may differ from this site.</div> </div><!-- container --> <footer><a id="backToTop" href="#top"> </a><nav><a id="backToTop" href="#top"> </a><ul class="links"><a id="backToTop" href="#top"> </a><li><a id="backToTop" href="#top"></a><a href="/sitemap.html">Site Map</a></li> <li><a href="/members/index.html">Members Only</a></li> <li><a href="/website-policies.html">Website Policies</a></li> <li><a href="https://doe.responsibledisclosure.com/hc/en-us" target="_blank">Vulnerability Disclosure Program</a></li> <li><a href="/contact.html">Contact Us</a></li> </ul> <div class="small">Science.gov is maintained by the U.S. Department of Energy's <a href="https://www.osti.gov/" target="_blank">Office of Scientific and Technical Information</a>, in partnership with <a href="https://www.cendi.gov/" target="_blank">CENDI</a>.</div> </nav> </footer> <script type="text/javascript"><!-- // var lastDiv = ""; function showDiv(divName) { // hide last div if (lastDiv) { document.getElementById(lastDiv).className = "hiddenDiv"; } //if value of the box is not nothing and an object with that name exists, then change the class if (divName && document.getElementById(divName)) { document.getElementById(divName).className = "visibleDiv"; lastDiv = divName; } } //--> </script> <script> /** * Function that tracks a click on an outbound link in Google Analytics. * This function takes a valid URL string as an argument, and uses that URL string * as the event label. */ var trackOutboundLink = function(url,collectionCode) { try { h = window.open(url); setTimeout(function() { ga('send', 'event', 'topic-page-click-through', collectionCode, url); }, 1000); } catch(err){} }; </script> <!-- Google Analytics --> <script> (function(i,s,o,g,r,a,m){i['GoogleAnalyticsObject']=r;i[r]=i[r]||function(){ (i[r].q=i[r].q||[]).push(arguments)},i[r].l=1*new Date();a=s.createElement(o), m=s.getElementsByTagName(o)[0];a.async=1;a.src=g;m.parentNode.insertBefore(a,m) })(window,document,'script','//www.google-analytics.com/analytics.js','ga'); ga('create', 'UA-1122789-34', 'auto'); ga('send', 'pageview'); </script> <!-- End Google Analytics --> <script> showDiv('page_1') </script> </body> </html>